11.03.2015 Views

Diffusion in GaAs and other III-V Semiconductors - Stephen J. Pearton

Diffusion in GaAs and other III-V Semiconductors - Stephen J. Pearton

Diffusion in GaAs and other III-V Semiconductors - Stephen J. Pearton

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Diffusion</strong> <strong>in</strong> <strong>GaAs</strong><br />

<strong>and</strong> <strong>other</strong> <strong>III</strong>-V <strong>Semiconductors</strong><br />

10 Years of Research<br />

Editor:<br />

D.J. Fisher<br />

SCITEC PUBLICATIONS


Notes:<br />

Each item <strong>in</strong> this section of the volume beg<strong>in</strong>s with a graphical compilation of relevant diffusion data which<br />

have been reported dur<strong>in</strong>g the past decade. The plotted data are also tabulated as <strong>in</strong>dicated on the graph. In some<br />

cases, the tabulated data have been obta<strong>in</strong>ed by digitiz<strong>in</strong>g published graphs <strong>and</strong> the values may not correspond<br />

exactly with the author's unpublished raw data.<br />

Refers to table N<br />

3N Bulk <strong>Diffusion</strong> - Quantitative Data<br />

The migration of Ag from epitaxial layers <strong>and</strong> <strong>in</strong>to (111) samples of Si,<br />

dur<strong>in</strong>g anneal<strong>in</strong>g at temperatures of between 450 <strong>and</strong> 500C, was studied<br />

by means of secondary ion mass spectrometric depth profil<strong>in</strong>g. It was<br />

found that the diffusivities lay between 8 x 10 -16 <strong>and</strong> 1.6 x 10 -15 cm 2 /s<br />

(table N). These values were lower than were expected on the basis of<br />

previous data.<br />

T.C.Nason, G.R.Yang, K.H.Park, T.M.Lu: Journal of Applied Physics,<br />

1991, 70[3], 1392-6<br />

[446-91/92-027]<br />

Indicates volume <strong>and</strong> page number <strong>in</strong><br />

DDF where abstract first appeared


AlAs<br />

Ag<br />

AlAs/<strong>GaAs</strong>: Ag <strong>Diffusion</strong><br />

Various elements were diffused <strong>in</strong>to a superlattice structure at temperatures of between<br />

700 <strong>and</strong> 1000C. Their disorder<strong>in</strong>g effect upon the superlattice was assessed by us<strong>in</strong>g a<br />

small-angle polish<strong>in</strong>g method. The diffusion of Ag had no disorder<strong>in</strong>g effect upon the<br />

superlattice. The results were expla<strong>in</strong>ed <strong>in</strong> terms of the <strong>in</strong>terstitial-substitutional<br />

mechanism, <strong>and</strong> of the solubility of the given dopant <strong>in</strong> <strong>GaAs</strong>.<br />

H.P.Ho, I.Harrison, N.Baba-Ali, B.Tuck, M.Hen<strong>in</strong>i: Journal of Electronic Materials,<br />

1991, 20[9], 649-52<br />

[446-84/85-002]<br />

Al<br />

31 AlAs/<strong>GaAs</strong>: Al <strong>Diffusion</strong><br />

The <strong>in</strong>termix<strong>in</strong>g of superlattices was <strong>in</strong>vestigated as a function of the Si concentration<br />

follow<strong>in</strong>g anneal<strong>in</strong>g at temperatures rang<strong>in</strong>g from 500 to 900C. The superlattice samples<br />

were prepared by means of molecular beam epitaxy, <strong>and</strong> the near-surface layers were<br />

doped with Si to concentrations of between 2 x 10 17 <strong>and</strong> 5 x 10 18 /cm 3 . The Si <strong>and</strong> Al<br />

depth profiles were measured by means of secondary ion mass spectrometry. The<br />

diffusion length <strong>and</strong> activation energy of Al, as a function of Si dopant concentration,<br />

were deduced from the secondary ion mass spectrometry data. With<strong>in</strong> the above<br />

temperature range a s<strong>in</strong>gle activation energy, for Al diffusion, of about 4eV was observed<br />

(table 1). The Al diffusion coefficient <strong>in</strong>creased rapidly with Si concentration.<br />

P.Mei, H.W.Yoon, T.Venkatesan, S.A.Schwarz, J.P.Harbison: Applied Physics Letters,<br />

1987, 50[25], 1823-5<br />

[446-157/159-227]<br />

227


Al AlAs Au<br />

Table 1<br />

Diffusivity of Al <strong>in</strong> AlAs/<strong>GaAs</strong><br />

Si (/cm 3 ) Temperature (C) D (cm 2 /s)<br />

5 x 10 17 900 6.1 x 10 -17<br />

5 x 10 17 850 3.0 x 10 -17<br />

1 x 10 18 795 5.0 x 10 -17<br />

2 x 10 18 750 6.6 x 10 -17<br />

2 x 10 18 750 5.4 x 10 -17<br />

5 x 10 18 700 4.2 x 10 -17<br />

5 x 10 17 795 2.3 x 10 -18<br />

2 x 10 18 695 1.1 x 10 -17<br />

2 x 10 18 700 5.6 x 10 -18<br />

5 x 10 18 650 5.6 x 10 -18<br />

1 x 10 18 745 2.6 x 10 -18<br />

2 x 10 18 655 1.0 x 10 -18<br />

2 x 10 18 655 7.3 x 10 -19<br />

2 x 10 17 900 1.2 x 10 -17<br />

5 x 10 17 745 3.6 x 10 -19<br />

- 850 4.2 x 10 -20<br />

- 800 2.6 x 10 -20<br />

AlAs/<strong>GaAs</strong>: Al <strong>Diffusion</strong><br />

Enhanced layer <strong>in</strong>terdiffusion. <strong>in</strong> Te-doped (2 x 10 17 to 3 x 10 18 /cm 3 ) organometallic<br />

chemical vapor deposited superlattices was studied by us<strong>in</strong>g secondary ion mass<br />

spectrometry. It was found that, at temperatures rang<strong>in</strong>g from 800 to 1000C, the Al<br />

diffusion coefficient had an activation energy of 3eV <strong>and</strong> was approximately proportional<br />

to the Te content. In the case of Si-<strong>in</strong>duced mix<strong>in</strong>g, the activation energy for Al diffusion<br />

was 4.1eV <strong>and</strong> exhibited a power-law dependence upon the Si content.<br />

P.Mei, S.A.Schwarz, T.Venkatesan, C.L.Schwartz, E.Colas: Journal of Applied Physics,<br />

1989, 65[5], 2165-7<br />

[446-72/73-002]<br />

Au<br />

AlAs/<strong>GaAs</strong>: Au <strong>Diffusion</strong><br />

Various elements were diffused <strong>in</strong>to a superlattice structure at temperatures of between<br />

700 <strong>and</strong> 1000C. Their disorder<strong>in</strong>g effect upon the superlattice was assessed by us<strong>in</strong>g a<br />

small-angle polish<strong>in</strong>g method. The diffusion of Au had no disorder<strong>in</strong>g effect upon the<br />

228


Au AlAs Si<br />

superlattice. The results were expla<strong>in</strong>ed <strong>in</strong> terms of the <strong>in</strong>terstitial-substitutional<br />

mechanism, <strong>and</strong> of the solubility of the given dopant <strong>in</strong> <strong>GaAs</strong>.<br />

H.P.Ho, I.Harrison, N.Baba-Ali, B.Tuck, M.Hen<strong>in</strong>i: Journal of Electronic Materials,<br />

1991, 20[9], 649-52<br />

[446-84/85-002]<br />

Cu<br />

AlAs/<strong>GaAs</strong>: Cu <strong>Diffusion</strong><br />

Various elements were diffused <strong>in</strong>to a superlattice structure at temperatures of between<br />

700 <strong>and</strong> 1000C. Their disorder<strong>in</strong>g effect upon the superlattice was assessed by us<strong>in</strong>g a<br />

small-angle polish<strong>in</strong>g method. The diffusion of Cu had no disorder<strong>in</strong>g effect upon the<br />

superlattice. The results were expla<strong>in</strong>ed <strong>in</strong> terms of the <strong>in</strong>terstitial-substitutional<br />

mechanism, <strong>and</strong> of the solubility of the given dopant <strong>in</strong> <strong>GaAs</strong>.<br />

H.P.Ho, I.Harrison, N.Baba-Ali, B.Tuck, M.Hen<strong>in</strong>i: Journal of Electronic Materials,<br />

1991, 20[9], 649-52<br />

[446-84/85-002]<br />

Mn<br />

AlAs/<strong>GaAs</strong>: Mn <strong>Diffusion</strong><br />

Various elements were diffused <strong>in</strong>to a superlattice structure at temperatures of between<br />

700 <strong>and</strong> 1000C. Their disorder<strong>in</strong>g effect upon the superlattice was assessed by us<strong>in</strong>g a<br />

small-angle polish<strong>in</strong>g method. It was found that Mn <strong>in</strong>duced disorder<strong>in</strong>g of the<br />

superlattice. However, the disorder<strong>in</strong>g effect which arose from Mn diffusion could be<br />

entirely <strong>in</strong>hibited if the fraction of As <strong>in</strong> the diffusion source were considerably higher<br />

than that of Mn. This <strong>in</strong>hibition effect was related to the formation of MnAs or MnAs 2 .<br />

This left very little Mn, <strong>in</strong> the vapor phase, which was available for diffusion. The results<br />

were expla<strong>in</strong>ed <strong>in</strong> terms of the <strong>in</strong>terstitial-substitutional mechanism, <strong>and</strong> of the solubility<br />

of the given dopant <strong>in</strong> <strong>GaAs</strong>.<br />

H.P.Ho, I.Harrison, N.Baba-Ali, B.Tuck, M.Hen<strong>in</strong>i: Journal of Electronic Materials,<br />

1991, 20[9], 649-52<br />

[446-84/85-002]<br />

Si<br />

AlAs/Al<strong>GaAs</strong>P/<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The diffusion of Si <strong>III</strong> -Si V neutral pairs versus the diffusion of Si <strong>III</strong> -V <strong>III</strong> complexes <strong>in</strong> <strong>III</strong>-V<br />

crystals was considered with regard to experimental data which revealed the effect of Si<br />

diffusion upon the self-diffusion of column-<strong>III</strong> <strong>and</strong> column-V lattice atoms. Secondary<br />

ion mass spectroscopy was used to compare the enhanced diffusion of column-<strong>III</strong> or<br />

column-V atoms <strong>in</strong> various Si-diffused heterostructures which were closely lattice-<br />

229


Si AlAs Surface<br />

matched to <strong>GaAs</strong>. An enhancement of lattice atom self-diffusion, due to impurity<br />

diffusion, was found to occur predom<strong>in</strong>antly on the column-<strong>III</strong> lattice. The data supported<br />

the Si <strong>III</strong> -V <strong>III</strong> diffusion model <strong>and</strong> <strong>in</strong>dicated that the ma<strong>in</strong> native defects which<br />

accompanied Si diffusion were column-<strong>III</strong> vacancies. These diffused directly on the<br />

column-<strong>III</strong> sub-lattice.<br />

D.G.Deppe, W.E.Plano, J.E.Baker, N.Holonyak, M.J.Ludowise, C.P.Kuo, R.M.Fletcher,<br />

T.D.Osentowski, M.G.Craford: Applied Physics Letters, 1988, 53[22], 2211-3<br />

[446-64/65-157]<br />

Zn<br />

AlAs/<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The diffusion of Zn <strong>in</strong>to superlattices was studied by us<strong>in</strong>g transmission electron<br />

microscopy <strong>and</strong> secondary ion mass spectroscopy. It was found that micro-defects existed<br />

near to the Zn diffusion front. These defects were <strong>in</strong>terstitial dislocation loops. It was<br />

suggested that the diffusion of Zn <strong>in</strong>to the present materials was similar to Zn diffusion<br />

<strong>in</strong>to <strong>GaAs</strong>. This was considered to be evidence for an <strong>in</strong>terstitial mechanism for the<br />

enhancement of <strong>in</strong>terdiffusion.<br />

I.Harrison, H.P.Ho, B.Tuck, M.Hen<strong>in</strong>i, O.H.Hughes: Semiconductor Science <strong>and</strong><br />

Technology, 1989, 4[10], 841-6<br />

[446-72/73-002]<br />

AlAs/<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The effect of an As pressure upon the disorder<strong>in</strong>g effect of Zn diffusion <strong>in</strong>to superlattices<br />

was studied. It was found that the degree of disorder<strong>in</strong>g <strong>in</strong>creased when no excess As was<br />

added to the ampoule. It had previously been found that dislocation loops formed near to<br />

the Zn diffusion front. The same effect was observed here, except when Zn diffusion was<br />

carried out <strong>in</strong> the absence of excess As. The Zn penetration was found to be greatest<br />

when no excess As was added to the diffusion ampoule.<br />

I.Harrison, H.P.Ho, B.Tuck, M.Hen<strong>in</strong>i, O.H.Hughes: Semiconductor Science <strong>and</strong><br />

Technology, 1990, 5[6], 561-5<br />

[446-74-001]<br />

Surface <strong>Diffusion</strong><br />

Al<br />

AlAs: Al Surface <strong>Diffusion</strong><br />

Dur<strong>in</strong>g the molecular beam epitaxial growth of AlAs on the vic<strong>in</strong>al (100) surface of<br />

<strong>GaAs</strong>, reflection high-energy electron diffraction was used to measure the transition<br />

temperature between 2-dimensional nucleation <strong>and</strong> pure step propagation which occurred<br />

when sub-monolayer amounts of Sn were present on the surface. In the case of samples<br />

230


Surface AlAs Surface<br />

which were misoriented by 0.5º with respect to the [011] or [01¯1] direction, the transition<br />

temperature decreased by approximately 100C after the deposition of 0.6 of a monolayer<br />

of Sn. The presence of Sn <strong>in</strong>creased the surface mobility of Al adatoms on (100) AlAs<br />

surfaces; as <strong>in</strong>dicated by the anneal<strong>in</strong>g behavior of the AlAs surface at 600C.<br />

G.S.Petrich, A.M.Dabiran, P.I.Cohen: Applied Physics Letters, 1992, 61[2], 162-4<br />

[446-93/94-001]<br />

AlAs/<strong>GaAs</strong>: Al Surface <strong>Diffusion</strong><br />

A study was made of reflection high-energy electron diffraction specular-beam <strong>in</strong>tensity<br />

oscillations on vic<strong>in</strong>al (001)AlAs which had been grown onto <strong>GaAs</strong>(001) substrates that<br />

were misoriented by 2 or 3° towards [110], [010], or [¯110]. It was found that the<br />

temperature dependence of the oscillation behavior on vic<strong>in</strong>al surfaces was similar to that<br />

on <strong>GaAs</strong>(001) <strong>and</strong> InAs(001). Contrary to the case of <strong>GaAs</strong>(001), however, the surface<br />

reconstruction could not be kept constant dur<strong>in</strong>g the growth-mode transition <strong>and</strong> it was<br />

therefore difficult to analyze AlAs(001) data <strong>in</strong> as much detail as that for <strong>GaAs</strong>(001).<br />

Nevertheless, from the similarity between them it was estimated that the effective surface<br />

migration barrier for Al adatoms on AlAs(001) was about 1.74eV.<br />

T.Shitara, J.H.Neave, B.A.Joyce: Applied Physics Letters, 1993, 62[14], 1658-60<br />

[446-106/107-007]<br />

Ga<br />

AlAs: Ga Surface <strong>Diffusion</strong><br />

Dur<strong>in</strong>g the molecular beam epitaxial growth of AlAs on the vic<strong>in</strong>al (100) surface of<br />

<strong>GaAs</strong>, reflection high-energy electron diffraction was used to measure the transition<br />

temperature between 2-dimensional nucleation <strong>and</strong> pure step propagation which occurred<br />

when sub-monolayer amounts of Sn were present on the surface. In the case of samples<br />

which were misoriented by 0.5º with respect to the [011] or [01¯1] direction, the transition<br />

temperature decreased by approximately 100C after the deposition of 0.6 of a monolayer<br />

of Sn. This <strong>in</strong>dicated that the Ga mobility had <strong>in</strong>creased.<br />

G.S.Petrich, A.M.Dabiran, P.I.Cohen: Applied Physics Letters, 1992, 61[2], 162-4<br />

[446-93/94-001]<br />

-miscellaneous<br />

AlAs: Surface <strong>Diffusion</strong><br />

An <strong>in</strong>vestigation was made of surface k<strong>in</strong>etics, dur<strong>in</strong>g metalorganic vapor-phase epitaxial<br />

growth, by means of high-vacuum scann<strong>in</strong>g tunnell<strong>in</strong>g microscopic observations of 2-<br />

dimensional nuclei <strong>and</strong> denuded zones. Monte Carlo simulations were carried out which<br />

were based upon the solid-on-solid model. Two-dimensional nucleus densities were used to<br />

deduce that the surface diffusion coefficient of AlAs was equal to 1.5 x 10 -7 cm 2 /s at 530C.<br />

The activation energy for migration was estimated to be 0.80eV. The 2-dimensional nucleus<br />

size <strong>in</strong> the [110] direction was about twice that <strong>in</strong> the [¯110] direction.<br />

231


Surface AlAs General<br />

This anisotropy was attributed ma<strong>in</strong>ly to a difference <strong>in</strong> the lateral stick<strong>in</strong>g probabilities<br />

between steps along [¯110] <strong>and</strong> those along [110]. The ratio of the stick<strong>in</strong>g probabilities<br />

was estimated to be greater than 3:1. The denuded zone widths on the upper terraces were<br />

some 2 times wider than those on the lower terraces. This suggested that the stick<strong>in</strong>g<br />

probability at descend<strong>in</strong>g steps was 10 to 300 times larger than the probability at<br />

ascend<strong>in</strong>g steps.<br />

M.Kasu, N.Kobayashi: Journal of Crystal Growth, 1997, 170, 246-50<br />

[446-141/142-093]<br />

AlAs: Surface <strong>Diffusion</strong><br />

The mechanisms of molecular beam epitaxy were <strong>in</strong>vestigated by grow<strong>in</strong>g <strong>and</strong> analyz<strong>in</strong>g<br />

the shapes of facet structures which consisted of an (001) top surface <strong>and</strong> two (111)B side<br />

surfaces. The diffusion of Al was found to be almost negligible; regardless of the As flux.<br />

By analyz<strong>in</strong>g the shape of the facet, the diffusion length of Al on a (001) surface was<br />

estimated to be about 0.02µ at 580C.<br />

S.Koshiba, Y.Nakamura, M.Tsuchiya, H.Noge, H.Kano, Y.Nagamune, T.Noda,<br />

H.Sakaki: Journal of Applied Physics, 1994, 76[7], 4138-44<br />

[446-117/118-159]<br />

AlAs/<strong>GaAs</strong>: Surface <strong>Diffusion</strong><br />

After deposit<strong>in</strong>g 1/6 of a monolayer of AlAs onto a very flat <strong>GaAs</strong> (001) surface by<br />

means of metalorganic vapor-phase epitaxy, a study was made of AlAs 2-dimensional<br />

nuclei by means of high-vacuum scann<strong>in</strong>g tunnell<strong>in</strong>g microscopy. The AlAs 2-<br />

dimensional nuclei elongated <strong>in</strong> the [110] direction, like <strong>GaAs</strong>. The density of AlAs 2-<br />

dimensional nuclei <strong>in</strong> the saturation region was 5 x 10 10 /cm 2 at 580C. The saturated AlAs<br />

2-dimensional nucleus density decreased with <strong>in</strong>creas<strong>in</strong>g temperature. On the basis of the<br />

saturated AlAs 2-dimensional nucleus densities, the surface diffusion coefficient of AlAs<br />

on <strong>GaAs</strong> was estimated to be 1.5 x 10 -7 cm 2 /s at 530C. This was an order of magnitude<br />

lower than that of <strong>GaAs</strong> on <strong>GaAs</strong>.<br />

M.Kasu, N.Kobayashi: Applied Physics Letters, 1995, 67[19], 2842-4<br />

[446-125/126-111]<br />

General<br />

AlAs/<strong>GaAs</strong>: Self-<strong>Diffusion</strong><br />

Cation self-diffusion <strong>in</strong> superlattices was exam<strong>in</strong>ed <strong>in</strong> terms of the activation enthalpy. It<br />

was suggested that cation diffusion should be mediated by As-antisite po<strong>in</strong>t defects, via<br />

the use of (As)antisite-rich materials or As-rich diffusion sources. It was also proposed<br />

that (As)antisite-mediated cation diffusion should exhibit a characteristic activation<br />

enthalpy of about 2.5eV under extr<strong>in</strong>sic conditions. Published data on <strong>in</strong>terdiffusion <strong>in</strong><br />

superlattices revealed a Fermi level dependence of the activation enthalpy. On this basis,<br />

232


General AlAs Interdiffusion<br />

it was concluded that the As-antisite defect was responsible for p-type impurity-enhanced<br />

cation self-diffusion.<br />

H.Iguchi: Japanese Journal of Applied Physics, 1989, 28[12], L2115-8<br />

[446-74-001]<br />

Interdiffusion<br />

AlAs/<strong>GaAs</strong>: Interdiffusion<br />

It was noted that undoped superlattices which had been grown at low temperatures<br />

underwent marked <strong>in</strong>terface <strong>in</strong>termix<strong>in</strong>g upon <strong>in</strong>creas<strong>in</strong>g the anneal<strong>in</strong>g temperature up to<br />

900C. Quantum conf<strong>in</strong>ement shifts which were caused by the <strong>in</strong>termix<strong>in</strong>g of lowtemperature<br />

re-grown <strong>and</strong> normal superlattices were studied by us<strong>in</strong>g electro-modulation<br />

spectroscopy. The effective activation energy for <strong>in</strong>termix<strong>in</strong>g <strong>in</strong> the low-temperature<br />

superlattices dur<strong>in</strong>g isochronal post-growth anneal<strong>in</strong>g (30s) was found to be 0.32eV. This<br />

value was anomalously lower than that for superlattices that were grown at normal<br />

temperatures. Roughen<strong>in</strong>g of the <strong>in</strong>terfaces, due to As precipitates, was associated with<br />

the <strong>in</strong>termix<strong>in</strong>g.<br />

I.Lahiri, D.D.Nolte, J.C.P.Chang, J.M.Woodall, M.R.Melloch: Applied Physics Letters,<br />

1995, 67[9], 1244-6<br />

[446-125/126-111]<br />

AlAs/<strong>GaAs</strong>: Interdiffusion<br />

The dopant-<strong>in</strong>duced <strong>in</strong>termix<strong>in</strong>g of Al <strong>and</strong> Ga <strong>in</strong> as-grown short-period superlattices was<br />

studied by means of atomic resolution cross-sectional scann<strong>in</strong>g tunnell<strong>in</strong>g microscopy. In<br />

the case of Si-doped n-type AlAs/<strong>GaAs</strong> short-period superlattices, the <strong>in</strong>termix<strong>in</strong>g<br />

<strong>in</strong>creased with <strong>in</strong>creas<strong>in</strong>g Si concentration (0 to 5 x 10 18 /cm 3 ). In the case of Be-doped p-<br />

type AlAs/<strong>GaAs</strong> short-period superlattices, no <strong>in</strong>termix<strong>in</strong>g of Al <strong>and</strong> Ga was observed;<br />

regardless of the Be concentration (0 to 5 x 10 18 /cm 3 ).<br />

J.F.Zheng, M.Salmeron, E.R.Weber: Solid State Communications, 1995, 93[5], 419-23<br />

[446-119/120-187]<br />

233


(Al,Ga)As<br />

Al<br />

Al<strong>GaAs</strong>/AlAs: Al <strong>Diffusion</strong><br />

Undoped superlattice structures were grown, with or without the presence of 120 Sn<br />

implants, by us<strong>in</strong>g molecular beam epitaxy. They were then annealed under Si 3 N 4 , SiO 2<br />

or encapsulant films. It was found that an enhancement of the Al-Ga <strong>in</strong>terdiffusion<br />

coefficient occurred under the Si 3 N 4 <strong>and</strong> SiO 2 films, due to the <strong>in</strong>-diffusion of Si from the<br />

films. The enhancement was greater dur<strong>in</strong>g diffusion of the Sn implant. Intermix<strong>in</strong>g<br />

enhancement was attributed to the operation of the Fermi effect. Beneath the WN x film,<br />

<strong>in</strong>terdiffusion was suppressed even <strong>in</strong> the presence of the Sn dopant.<br />

E.L.Allen, C.J.Pass, M.D.Deal, J.D.Plummer, V.F.K.Chia: Applied Physics Letters, 1991,<br />

59[25], 3252-4<br />

[446-84/85-006]<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Al <strong>Diffusion</strong><br />

Data were presented which showed that the Al-Ga <strong>in</strong>terdiffusion coefficient for an<br />

Al x Ga 1-x As-<strong>GaAs</strong> quantum-well heterostructure or a superlattice was highly dependent<br />

upon the crystal encapsulation conditions. The activation energy for Al-Ga <strong>in</strong>terdiffusion,<br />

<strong>and</strong> thus layer-disorder<strong>in</strong>g, was smaller (about 3.5eV) for dielectric encapsulated samples<br />

than after capless anneal<strong>in</strong>g (about 4.7eV). The <strong>in</strong>terdiffusion coefficient for Si 3 N 4 -<br />

capped samples was almost an order of magnitude smaller than for the cases of capless or<br />

SiO 2 -capped samples at temperatures of between 800 <strong>and</strong> 875C. As well as the type of<br />

encapsulant, the encapsulation geometry (stripes or capped stripes) was important<br />

because of stra<strong>in</strong> effects. These were a major source of anisotropic Al-Ga <strong>in</strong>terdiffusion.<br />

L.J.Guido, N.Holonyak, K.C.Hsieh, R.W.Kaliski, W.E.Plano, R.D.Burnham,<br />

R.L.Thornton, J.E.Epler, T.L.Paoli: Journal of Applied Physics, 1987, 61[4], 1372-9<br />

[446-60-002]<br />

234


Al (Al,Ga)As Be<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Al <strong>Diffusion</strong><br />

Photolum<strong>in</strong>escence spectroscopy was used to determ<strong>in</strong>e the temperature <strong>and</strong><br />

compositional dependence of the <strong>in</strong>terdiffusion of Al <strong>and</strong> Ga <strong>in</strong> (Al,Ga)As/<strong>GaAs</strong><br />

superlattices. The position of the b<strong>and</strong>-to-b<strong>and</strong> lum<strong>in</strong>escence <strong>in</strong> the superlattices was<br />

measured before <strong>and</strong> after thermal anneal<strong>in</strong>g. The diffusion equation was solved for a<br />

fixed value of the diffusion coefficient <strong>in</strong> order to establish the potential profile of the<br />

superlattice structure after anneal<strong>in</strong>g. A solution of the Schröd<strong>in</strong>ger equation, where the<br />

electron or hole wave function was exp<strong>and</strong>ed as a Fourier series, was used to determ<strong>in</strong>e<br />

the position of the superlattice b<strong>and</strong> edges before <strong>and</strong> after anneal<strong>in</strong>g <strong>and</strong> thus deduce the<br />

expected lum<strong>in</strong>escence peak positions. The value of the coefficient which yielded a<br />

calculated shift which was <strong>in</strong> agreement with the measured shift <strong>in</strong> the lum<strong>in</strong>escence was<br />

taken to be the actual value of the <strong>in</strong>terdiffusion coefficient. For structures consist<strong>in</strong>g of<br />

<strong>GaAs</strong> wells <strong>and</strong> Al x Ga 1-x As barriers, where x was 1 or 0.3, the <strong>in</strong>terdiffusion process was<br />

characterized by an activation energy of 6.0eV <strong>and</strong> a value of 4 x 10 -19 cm 2 /s at 850C.<br />

When x was equal to 0.7, the <strong>in</strong>terdiffusion was characterized by an activation energy of<br />

4.0eV <strong>and</strong> a value of 7 x 10 -18 cm 2 /s at 850C.<br />

J.C.Lee, T.E.Schles<strong>in</strong>ger, T.F.Kuech: Journal of Vacuum Science <strong>and</strong> Technology B,<br />

1987, 5[4], 1187-90<br />

[446-55/56-002]<br />

Be<br />

Al<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

A close relationship between Be surface segregation <strong>and</strong> diffusion, <strong>in</strong> molecular beam<br />

epitaxial Al<strong>GaAs</strong> layers which were heavily doped with Be, was analyzed with<strong>in</strong> the<br />

framework of a thermodynamic approach to segregation effects.. The effect of growth<br />

parameters (excess As pressure, substrate temperature, growth rate) <strong>and</strong> dopant level<br />

upon the likelihood of Be segregation layer formation was considered.<br />

S.V.Ivanov, P.S.Kopev, N.N.Ledentsov: Journal of Crystal Growth, 1991, 108[3-4], 661-<br />

9<br />

[446-81/82-002]<br />

GaAlAs: Be <strong>Diffusion</strong><br />

A study was made of the contributions of segregation, diffusion <strong>and</strong> aggregation to the<br />

broaden<strong>in</strong>g of d-doped planes of Be <strong>in</strong> Ga 0.67 Al 0.33 As. It was found that sharp spikes of<br />

Be could be obta<strong>in</strong>ed for sheet densities which were below 10 13 /cm 2 <strong>and</strong> for growth<br />

temperatures of 500C or less. At higher temperatures or densities, segregation or<br />

concentration-dependent rapid diffusion could occur; thus caus<strong>in</strong>g significant spread<strong>in</strong>g<br />

even dur<strong>in</strong>g growth. The co-deposition of Si <strong>and</strong> Be markedly reduced this broaden<strong>in</strong>g.<br />

235


Be (Al,Ga)As Be<br />

J.J.Harris, J.B.Clegg, R.B.Beall, J.Castagné, K.Woodbridge, C.Roberts: Journal of<br />

Crystal Growth, 1991, 111[1-4], 239-45<br />

[446-91/92-001]<br />

GaAlAs: Be <strong>Diffusion</strong><br />

A molecular beam epitaxial technique was developed <strong>in</strong> order to suppress Be diffusion by<br />

<strong>in</strong>corporat<strong>in</strong>g In <strong>in</strong>to a GaAlAs epilayer. The diffusion coefficients of Be-doped<br />

In x (Ga 0.9 Al 0.1 ) 1-x As, grown at 600C, decreased from 10 -14 to 2 x 10 -15 cm 2 /s when the InAs<br />

mole fraction, x, was <strong>in</strong>creased from 0 to 0.07. This <strong>in</strong>dicated that compressive stresses <strong>in</strong><br />

the epilayer, caused by the <strong>in</strong>corporation of In, played an important role <strong>in</strong> suppress<strong>in</strong>g<br />

Be diffusion.<br />

T.Tomioka, T.Fujii, H.Ishikawa, S.Sasa, A.Endoh, Y.Bamba, K.Ishii, Y.Kataoka:<br />

Japanese Journal of Applied Physics, 1990, 29[5], L716-9<br />

[446-76/77-001]<br />

GaAlAs: Be <strong>Diffusion</strong><br />

The suppression of Be diffusion <strong>in</strong> molecular beam epitaxial Ga 0.9 Al 0.1 As was reported<br />

here for the first time. It was achieved by <strong>in</strong>corporat<strong>in</strong>g In <strong>in</strong>to the epilayer. The m<strong>in</strong>imum<br />

Be diffusion coefficient <strong>in</strong> In-doped layers with a carrier concentration of 7 x 10 19 /cm 3<br />

<strong>and</strong> an InAs mole fraction of 0.07, which had been grown at 600C, was equal to about 2 x<br />

10 -15 cm 2 /s. This value was 5 times smaller than that which was observed <strong>in</strong> the absence<br />

of In. The photolum<strong>in</strong>escence <strong>in</strong>tensity of the layers decreased markedly <strong>in</strong> In x (Al,Ga) 1-x<br />

when x was greater than 0.05. This behavior was attributed to a crystal degradation which<br />

resulted from the presence of misfit dislocations.<br />

T.Fujii, T.Tomioka, H.Ishikawa, S.Sasa, A.Endoh, Y.Bamba, K.Ishii, Y.Kataoka: Journal<br />

of Vacuum Science <strong>and</strong> Technology B, 1990, 8[2], 154-6<br />

[446-74-003]<br />

GaAlAs: Be <strong>Diffusion</strong><br />

The migration of ion-implanted Be was studied as a function of Al concentration <strong>and</strong><br />

anneal<strong>in</strong>g temperature <strong>and</strong> was compared with its diffusivity <strong>in</strong> <strong>GaAs</strong>. The behavior of Be<br />

<strong>in</strong> Al<strong>GaAs</strong> was similar to that <strong>in</strong> <strong>GaAs</strong>, <strong>and</strong> it even exhibited the anomalous characteristic<br />

of <strong>in</strong>creased redistribution with decreas<strong>in</strong>g temperature. The results could be described<br />

by:<br />

Ga 0.8 Al 0.2 As: D(cm 2 /s) = 1.8 x 10 -9 exp[-0.90(eV)/kT]<br />

Ga 0.6 Al 0.4 As: D(cm 2 /s) = 3.3 x 10 -9 exp[-0.84(eV)/kT]<br />

The diffusivity of Be appeared to <strong>in</strong>crease with Al content. This was suggested to be due<br />

to an <strong>in</strong>crease <strong>in</strong> the bond strength of matrix atoms upon add<strong>in</strong>g Al. This prevented the<br />

easy transfer of Be from <strong>in</strong>terstitial to substitutional sites. An over-saturation of Be<br />

<strong>in</strong>terstitials could also expla<strong>in</strong> the persistence of anomalous diffusion <strong>in</strong> Al<strong>GaAs</strong> with<br />

respect to the anneal<strong>in</strong>g temperature. The results were expla<strong>in</strong>ed <strong>in</strong> terms of a<br />

substitutional-<strong>in</strong>terstitial diffusion mechanism, the relative amounts of <strong>in</strong>terstitial <strong>and</strong><br />

236


Be (Al,Ga)As Be<br />

substitutional Be, <strong>and</strong> the relative difficulty of mov<strong>in</strong>g from an <strong>in</strong>terstitial to a<br />

substitutional site.<br />

C.C.Lee, M.D.Deal, J.C.Bravman: Applied Physics Letters, 1995, 66[3], 355-7<br />

[446-123/124-160]<br />

Al<strong>GaAs</strong>/AlAs: Be <strong>Diffusion</strong><br />

Dopant diffusion was <strong>in</strong>vestigated via depth-profil<strong>in</strong>g us<strong>in</strong>g secondary ion mass<br />

spectrometric analysis of heterostructures which conta<strong>in</strong>ed Be-doped short-period Al x Ga 1-<br />

xAs superlattices, where x ranged from 0.3 to 0.8, that had been grown by means of<br />

molecular beam epitaxy. Out-diffusion of Be <strong>in</strong>to the undoped <strong>GaAs</strong> layers was observed<br />

only at a substrate temperature of 660C, when the Be concentration was 2 x 10 18 /cm 3 . At<br />

a dopant concentration of 2 x 10 19 /cm 3 , a marked <strong>in</strong>crease <strong>in</strong> diffusion occurred at all<br />

growth temperatures. The solubility limits of Be were 10 19 /cm 3 at x = 0.6, <strong>and</strong> 2 x<br />

10 18 /cm 3 at x = 0.8. Secondary ion mass spectrometry profiles revealed that the amount of<br />

diffused Be <strong>in</strong> the active region was twice as high <strong>in</strong> samples with a th<strong>in</strong> (450 to 600nm)<br />

p-type cladd<strong>in</strong>g layer.<br />

A.Gaymann, M.Maier, W.Bronner, N.Gruen, K.Koehler: Materials Science <strong>and</strong><br />

Eng<strong>in</strong>eer<strong>in</strong>g B, 1997, 44[1-3], 12-5<br />

[446-157/159-237]<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

It was po<strong>in</strong>ted out that the characteristic features of Be diffusion <strong>in</strong> <strong>GaAs</strong> substrates <strong>and</strong><br />

<strong>GaAs</strong>/Al<strong>GaAs</strong> superlattices could be expla<strong>in</strong>ed <strong>in</strong> terms of a kick-out mechanism <strong>in</strong><br />

which the doubly positively charged Ga self-<strong>in</strong>terstitial governed Ga self-diffusion. Such<br />

characteristics <strong>in</strong>cluded much lower diffusivities of Be under out-diffusion conditions<br />

than under <strong>in</strong>-diffusion conditions. It was found that the Long<strong>in</strong>i mechanism was able to<br />

expla<strong>in</strong> most of the features.<br />

S.Yu, T.Y.Tan, U.Gösele: Journal of Applied Physics, 1991, 69[6], 3547-65<br />

[446-86/87-009]<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

The effect of substrate orientation upon Be transport dur<strong>in</strong>g <strong>GaAs</strong> molecular beam<br />

epitaxial growth was evaluated by means of secondary ion mass spectrometry <strong>and</strong><br />

measurements of the current-voltage characteristics of Al<strong>GaAs</strong>/<strong>GaAs</strong> heterojunction<br />

bipolar transistors. The Be dop<strong>in</strong>g level was between 2 x 10 19 <strong>and</strong> 9 x 10 19 /cm 3 . The Be<br />

transport which was observed for the conventional (100) orientation <strong>in</strong>creased rapidly<br />

upon <strong>in</strong>creas<strong>in</strong>g the growth temperature from 530 to 630C. However, with a substrate<br />

misorientation away from (100) <strong>and</strong> towards (111)A, Be transport decreased at 630C <strong>and</strong><br />

reached a m<strong>in</strong>imum value for the (311)A orientation. The maximum current ga<strong>in</strong>, of<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong> heterojunction bipolar transistors which had been grown at 560C, was<br />

equal to 264 for the (411)A orientation <strong>and</strong> 3 for the (100) orientation. It was concluded<br />

that this confirmed the applicability of substrate orientations <strong>other</strong> than the conventional<br />

(100) one for obta<strong>in</strong><strong>in</strong>g a sharp Be profile.<br />

237


Be (Al,Ga)As D<br />

K.Mochizuki, S.Goto, T.Mishima, C.Kusano: Japanese Journal of Applied Physics, 1992,<br />

31[1-11], 3495-9<br />

[446-99/100-059]<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

A secondary ion mass spectrometry <strong>in</strong>vestigation was made of Be diffusion, dur<strong>in</strong>g the<br />

molecular beam epitaxial growth of graded-<strong>in</strong>dex separate conf<strong>in</strong>ement heterostructure<br />

laser structures. In the case of growth at 700C, it was found that Be from the p-type<br />

Al<strong>GaAs</strong> cladd<strong>in</strong>g layer diffused <strong>in</strong>to the quantum well <strong>and</strong> beyond. As a result, the p-n<br />

junction was displaced from the heterojunction. The extent of Be diffusion was found to<br />

depend upon the dopants <strong>in</strong> the graded-<strong>in</strong>dex regions which adjo<strong>in</strong>ed the <strong>GaAs</strong> active<br />

layer. When the graded-<strong>in</strong>dex segments were left undoped, Be diffused through the entire<br />

p-side graded-<strong>in</strong>dex region, the quantum well active region, <strong>and</strong> a significant portion of<br />

the n-side graded-<strong>in</strong>dex region. However, when the graded-<strong>in</strong>dex regions were doped<br />

with Be <strong>and</strong> Si on the p-side <strong>and</strong> n-side, respectively, the displacement of the p-n junction<br />

which was caused by Be diffusion was significantly reduced. Upon assum<strong>in</strong>g that Be<br />

diffused from a constant surface source <strong>and</strong> <strong>in</strong>to an n-type layer, as a s<strong>in</strong>gly charged<br />

<strong>in</strong>terstitial donor, the present analysis predicted that <strong>in</strong>creas<strong>in</strong>g the dop<strong>in</strong>g of the n-type<br />

layer would retard Be diffusion; whereas <strong>in</strong>creas<strong>in</strong>g the dop<strong>in</strong>g of the p-type layer would<br />

enhance it. Upon <strong>in</strong>clud<strong>in</strong>g the electric field of the p-n junction <strong>in</strong> the model, peaks <strong>and</strong><br />

<strong>in</strong>flections were predicted which resembled those that were observed <strong>in</strong> experimental<br />

secondary ion mass spectroscopy profiles. It was concluded that, because of Be-related O<br />

contam<strong>in</strong>ation <strong>and</strong> Be diffusion <strong>in</strong> the p-side graded-<strong>in</strong>dex region, the presence of Be<br />

should be avoided on the p side. However, Si additions to the n side were expected to be<br />

beneficial as they m<strong>in</strong>imized Be diffusion <strong>and</strong> p-n junction displacement.<br />

V.Swam<strong>in</strong>athan, N.Ch<strong>and</strong>, M.Geva, P.J.Anthony, A.S.Jordan: Journal of Applied<br />

Physics, 1992, 72[10], 4648-54<br />

[446-106/107-015]<br />

D<br />

GaAlAs: D <strong>Diffusion</strong><br />

The diffusion of D <strong>in</strong> Si-doped Al x Ga 1-x As was studied for x-values of up to 0.30. It was<br />

found that, for x = 0, the diffusion profile could be closely fitted by us<strong>in</strong>g an erfc<br />

function. When the x-value was greater than 0.055, the profiles exhibited a plateau that<br />

was followed by a sharp decrease. It was suggested that, <strong>in</strong> Si-doped samples, the D<br />

behaved like a deep acceptor with a level, H -/0 , which was slightly resonant <strong>in</strong> the<br />

conduction b<strong>and</strong> of <strong>GaAs</strong>. It appeared as a localized state, for x-values above 0.07, as the<br />

b<strong>and</strong>-gap energy <strong>in</strong>creased. In this region, the H - species became dom<strong>in</strong>ant <strong>and</strong> were<br />

trapped on the positively charged donors dur<strong>in</strong>g diffusion.<br />

J.Chevallier, B.Machayekhi, C.M.Grattepa<strong>in</strong>, R.Rahbi, B.Theys: Physical Review B,<br />

1992, 45[15], 8803-6<br />

[446-86/87-001]<br />

238


Ga (Al,Ga)As Ga<br />

Ga<br />

32 Al<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

The <strong>in</strong>termix<strong>in</strong>g of Al<strong>GaAs</strong>-based <strong>in</strong>terfaces was enhanced by capp<strong>in</strong>g wafers with a<br />

layer of SiO 2 . By assum<strong>in</strong>g that this enhancement resulted from the <strong>in</strong>troduction of<br />

additional Ga vacancies <strong>in</strong>to the sample, it was possible to estimate the temperaturedependent<br />

equilibrium Ga vacancy diffusivity. Experiments were performed <strong>in</strong> which<br />

SiO 2 -capped quantum-well samples were annealed at temperatures rang<strong>in</strong>g from 800 to<br />

1025C. The calculated photolum<strong>in</strong>escence shifts were compared with the measured<br />

spectra <strong>and</strong> a relationship, for the Ga vacancy diffusivity, of the form:<br />

D (cm 2 /s) = 0.962 exp[-2.72(eV)/kT]<br />

(table 2) was obta<strong>in</strong>ed. By us<strong>in</strong>g this relationship, the equilibrium Ga vacancy<br />

concentration could be estimated via Monte Carlo simulation. The resultant expression<br />

was: C (/cm 3 ) = 1.25 x 10 31 exp[-3.28(eV)/kT].<br />

K.B.Kahen, D.L.Peterson, G.Rajeswaran, D.J.Lawrence: Applied Physics Letters, 1989,<br />

55[7], 651-3<br />

[446-70/71-103]<br />

Table 2<br />

Diffusivity of Ga Vacancies <strong>in</strong> Al<strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

1030 2.8 x 10 -11<br />

1005 1.7 x 10 -11<br />

950 6.3 x 10 -12<br />

900 2.3 x 10 -12<br />

Al<strong>GaAs</strong>/AlAs: Ga <strong>Diffusion</strong><br />

Undoped superlattice structures were grown, with or without the presence of 120 Sn<br />

implants, by us<strong>in</strong>g molecular beam epitaxy. They were then annealed under Si 3 N 4 , SiO 2<br />

or encapsulant films. It was found that an enhancement of the Al-Ga <strong>in</strong>terdiffusion<br />

coefficient occurred under the Si 3 N 4 <strong>and</strong> SiO 2 films, due to the <strong>in</strong>-diffusion of Si from the<br />

films. The enhancement was greater dur<strong>in</strong>g diffusion of the Sn implant. In both cases,<br />

<strong>in</strong>termix<strong>in</strong>g enhancement was attributed to the operation of the Fermi effect. Beneath the<br />

WN x film, <strong>in</strong>terdiffusion was suppressed even <strong>in</strong> the presence of the Sn dopant.<br />

E.L.Allen, C.J.Pass, M.D.Deal, J.D.Plummer, V.F.K.Chia: Applied Physics Letters, 1991,<br />

59[25], 3252-4<br />

[446-84/85-006]<br />

239


Ga (Al,Ga)As Ga<br />

33 Al<strong>GaAs</strong>/<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

Photolum<strong>in</strong>escence spectroscopy was used to determ<strong>in</strong>e the temperature <strong>and</strong><br />

compositional dependence of the <strong>in</strong>terdiffusion of Al <strong>and</strong> Ga <strong>in</strong> (Al,Ga)As/<strong>GaAs</strong><br />

superlattices. The position of the b<strong>and</strong>-to-b<strong>and</strong> lum<strong>in</strong>escence <strong>in</strong> the superlattices was<br />

measured before <strong>and</strong> after thermal anneal<strong>in</strong>g. The diffusion equation was solved for a<br />

fixed value of the diffusion coefficient <strong>in</strong> order to establish the potential profile of the<br />

superlattice structure after anneal<strong>in</strong>g. A solution of the Schröd<strong>in</strong>ger equation, where the<br />

electron or hole wave function was exp<strong>and</strong>ed as a Fourier series, was used to determ<strong>in</strong>e<br />

the position of the superlattice b<strong>and</strong> edges before <strong>and</strong> after anneal<strong>in</strong>g <strong>and</strong> thus deduce the<br />

expected lum<strong>in</strong>escence peak positions. The value of the coefficient which yielded a<br />

calculated shift which was <strong>in</strong> agreement with the measured shift <strong>in</strong> the lum<strong>in</strong>escence was<br />

taken to be the actual value of the <strong>in</strong>terdiffusion coefficient. For structures consist<strong>in</strong>g of<br />

<strong>GaAs</strong> wells <strong>and</strong> Al x Ga 1-x As barriers, where x was 1 or 0.3, the <strong>in</strong>terdiffusion process was<br />

characterized by an activation energy of 6.0eV <strong>and</strong> a value of 4 x 10 -19 cm 2 /s at 850C.<br />

When x was equal to 0.7, the <strong>in</strong>terdiffusion was characterized by an activation energy of<br />

4.0eV <strong>and</strong> a value of 7 x 10 -18 cm 2 /s at 850C (table 3).<br />

J.C.Lee, T.E.Schles<strong>in</strong>ger, T.F.Kuech: Journal of Vacuum Science <strong>and</strong> Technology B,<br />

1987, 5[4], 1187-90<br />

[446-55/56-002]<br />

Table 3<br />

Interdiffusivity (Al-Ga) <strong>in</strong> Al x Ga 1-x As/<strong>GaAs</strong><br />

x Temperature (C) D (cm 2 /s)<br />

0.3 905 1.1 x 10 -17<br />

0.3 875 3.4 x 10 -18<br />

0.3 865 8.2 x 10 -19<br />

0.3 850 4.0 x 10 -19<br />

0.3 820 1.1 x 10 -19<br />

0.7 820 2.4 x 10 -18<br />

0.7 790 1.1 x 10 -18<br />

0.7 775 4.6 x 10 -19<br />

0.7 750 1.3 x 10 -19<br />

0.7 750 1.1 x 10 -19<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

It was po<strong>in</strong>ted out that the characteristic features of Be <strong>and</strong> Zn diffusion <strong>in</strong> <strong>GaAs</strong><br />

substrates <strong>and</strong> <strong>GaAs</strong>/Al<strong>GaAs</strong> superlattices could be expla<strong>in</strong>ed <strong>in</strong> terms of a kick-out<br />

mechanism <strong>in</strong> which the doubly positively charged Ga self-<strong>in</strong>terstitial governed Ga selfdiffusion.<br />

It was found that the Long<strong>in</strong>i mechanism was able to expla<strong>in</strong> most of these<br />

features. However, the predictions of the Long<strong>in</strong>i mechanism with regard to Ga self-<br />

240


Ga (Al,Ga)As In<br />

diffusion disagreed with experimental observations of the effect of superlattice<br />

disorder<strong>in</strong>g.<br />

S.Yu, T.Y.Tan, U.Gösele: Journal of Applied Physics, 1991, 69[6], 3547-65<br />

[446-86/87-009]<br />

H<br />

Al<strong>GaAs</strong>: H <strong>Diffusion</strong><br />

<strong>Diffusion</strong> experiments were performed on samples of Si-doped Al x Ga 1-x As epitaxial<br />

layers, with x-values which ranged from 0 to 0.30, as a function of the Si dop<strong>in</strong>g level<br />

<strong>and</strong> the diffusion temperature. For each composition, calculated H diffusion profiles<br />

which had been deduced by us<strong>in</strong>g Mathiot's model were fitted to the experimental<br />

profiles. It was assumed that H behaved as a deep acceptor, <strong>and</strong> that H o <strong>and</strong> H - were the<br />

diffus<strong>in</strong>g species. The trapp<strong>in</strong>g of H - by Si + donors, <strong>and</strong> their acceleration by an electric<br />

field, were <strong>in</strong>corporated <strong>in</strong>to the model. As well as the diffusion coefficient of H, <strong>and</strong> the<br />

dissociation constant of the SiH complexes, the model provided for a compositional<br />

dependence of the H acceptor level <strong>in</strong> Al<strong>GaAs</strong> alloys. It was concluded that the H<br />

acceptor level was localized <strong>in</strong> the b<strong>and</strong>-gap of the present Al<strong>GaAs</strong> alloys, <strong>and</strong> deepened<br />

below the Γ conduction b<strong>and</strong> as x <strong>in</strong>creased.<br />

B.Machayekhi, R.Rahbi, B.Theys, M.Miloche, J.Chevallier: Materials Science Forum,<br />

1994, 143-147, 951-6<br />

[446-113/114-001]<br />

GaAlAs: H <strong>Diffusion</strong><br />

Layers of material, which was doped with various group-VI donors (S, Se, Te), were<br />

exposed to H plasma. By us<strong>in</strong>g secondary ion mass spectroscopy it was shown that, as <strong>in</strong><br />

the case of Si-doped materials, the diffusivity of H depended strongly upon the AlAs<br />

content. Electronic measurements <strong>in</strong>dicated that, after H diffusion, the electron<br />

concentration systematically decreased while their mobility <strong>in</strong>creased; thus demonstrat<strong>in</strong>g<br />

the passivation of the group-VI donors by H.<br />

B.Theys, B.Machayekhi, J.Chevallier, K.Somogyi, K.Zahraman, P.Gibart, M.Miloche:<br />

Journal of Applied Physics, 1995, 77[7], 3186-93<br />

[446-121/122-052]<br />

In<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: In <strong>Diffusion</strong><br />

Data were presented on the disorder<strong>in</strong>g of an Al<strong>GaAs</strong>/<strong>GaAs</strong> laser structure us<strong>in</strong>g In solid<br />

sources. By us<strong>in</strong>g the <strong>in</strong>dependent or comb<strong>in</strong>ed diffusion of Si <strong>and</strong> In from th<strong>in</strong>-film<br />

sources, it was deduced that In had a higher diffusion coefficient than Si <strong>and</strong> led to a<br />

similar degree of impurity-<strong>in</strong>duced disorder<strong>in</strong>g. The degree of <strong>in</strong>dex guid<strong>in</strong>g was tested<br />

by mak<strong>in</strong>g excess-loss measurements <strong>in</strong> s<strong>in</strong>gle-mode raised-cos<strong>in</strong>e s-bends. It was found<br />

241


In (Al,Ga)As Mn<br />

that structures which were patterned by SiO 2 /In disorder<strong>in</strong>g suffered excess losses which<br />

were similar to those <strong>in</strong> structures that were patterned with SiO 2 . An 0.26mm transition<br />

length for 3dB loss was measured for 1000nm-wide guides with 0.1mm guide offsets.<br />

This corresponded to a lateral <strong>in</strong>dex of refraction difference of between 0.8 <strong>and</strong> 1.0%.<br />

There was no evidence for an <strong>in</strong>creased l<strong>in</strong>ear loss due to the presence of a dilute In<strong>GaAs</strong><br />

alloy at a measurement wavelength of 870nm.<br />

T.K.Tang, J.J.Alwan, C.M.Herz<strong>in</strong>ger, T.M.Cockerill, A.Crook, T.A.DeTemple,<br />

J.J.Coleman, J.E.Baker: Applied Physics Letters, 1991, 59[22], 2880-2<br />

[446-91/92-005]<br />

Mg<br />

Al<strong>GaAs</strong>: Mg <strong>Diffusion</strong><br />

Layer samples were diffused, at 785C, with Mg from an As-saturated Ga solution that<br />

conta<strong>in</strong>ed 0.1wt%Mg. Secondary ion mass spectrometry <strong>and</strong> differential Hall effect<br />

measurements revealed that the depth profile consisted of a high-Mg concentration region<br />

close to the surface, <strong>and</strong> a lower-concentration plateau with<strong>in</strong> the sample. The diffusion<br />

of Mg <strong>in</strong>to <strong>GaAs</strong>, Al 0.5 Ga 0.5 As <strong>and</strong> Al 0.7 Ga 0.3 As for 0.33h resulted <strong>in</strong> diffusion fronts at<br />

about 0.002, 0.004 <strong>and</strong> 0.006mm from the surface, respectively. The depth, for a fixed<br />

hole concentration, was proportional to the square root of the diffusion time <strong>in</strong> both <strong>GaAs</strong><br />

<strong>and</strong> Ga 0.65 Al 0.35 As.<br />

S.Mukai, Y.Kaneko, T.Nukui, M.Mori, M.Watanabe, H.Itoh, H.Yajima: Japanese Journal<br />

of Applied Physics, 1989, 28[1], L1-3<br />

[446-64/65-158]<br />

Al<strong>GaAs</strong>: Mg <strong>Diffusion</strong><br />

An <strong>in</strong>vestigation was made of dual Mg/F or Mg/Ar implantation. It was found that the<br />

dual implantation suppressed Mg diffusion, but degraded the electrical properties. This<br />

was more apparent <strong>in</strong> material with lower Al contents. The Ar dual implantation<br />

suppressed Mg diffusion <strong>and</strong> improved the electrical properties of material with a high Al<br />

content. It was suggested that Mg-F bonds formed as a result of F dual implantation <strong>and</strong><br />

that successive anneal<strong>in</strong>g suppressed Mg diffusion <strong>and</strong> disturbed Mg activation. The<br />

extensive radiation damage which was caused by Ar dual implantation caused Mg atoms<br />

to occupy lattice sites <strong>in</strong> Al<strong>GaAs</strong> with a high Al content.<br />

N.Hara, H.Suehiro, S.Kuroda: Materials Science Forum, 1995, 196-201, 1943-8<br />

[446-127/128-107]<br />

Mn<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Mn <strong>Diffusion</strong><br />

The diffusion of Mn was carried out <strong>in</strong> sealed quartz ampoules, us<strong>in</strong>g 4 types of Mn<br />

source. These were: solid crystall<strong>in</strong>e Mn gra<strong>in</strong>s, Mn 3 As, MnAs, <strong>and</strong> Mn th<strong>in</strong> films on<br />

242


Mn (Al,Ga)As Si<br />

<strong>GaAs</strong> substrates. It was found that only MnAs led to the formation of a smooth <strong>GaAs</strong><br />

surface <strong>and</strong> a uniform dopant distribution. In the case of the <strong>other</strong> sources, <strong>in</strong>teractions<br />

between the source materials <strong>and</strong> the substrate gave rise to poor surface morphologies <strong>and</strong><br />

<strong>in</strong>homogeneous distributions. In the case of diffusion at 800C, surface p-type carrier<br />

concentrations of the order of 10 20 /cm 3 were obta<strong>in</strong>ed. The diffusion profiles which were<br />

determ<strong>in</strong>ed by us<strong>in</strong>g capacitance-voltage techniques resembled those which were<br />

obta<strong>in</strong>ed for Zn diffusion. It was suggested that a substitutional-<strong>in</strong>terstitial mechanism<br />

was the predom<strong>in</strong>ant one. It was also noted that layer disorder<strong>in</strong>g could be produced <strong>in</strong><br />

Al<strong>GaAs</strong>-<strong>GaAs</strong> superlattices by Mn impurities.<br />

C.H.Wu, K.C.Hsieh, G.E.Höfler, N.El-Ze<strong>in</strong>, N.Holonyak: Applied Physics Letters, 1991,<br />

59[10], 1224-6<br />

[446-84/85-007]<br />

O<br />

Al<strong>GaAs</strong>: O <strong>Diffusion</strong><br />

A layer of SiO 2 , deposited by sputter<strong>in</strong>g, was used as a diffusion source for O impurities,<br />

as well as a source of Ga vacancies which enhanced impurity diffusion <strong>and</strong> permitted<br />

reductions to be made <strong>in</strong> the required anneal<strong>in</strong>g temperatures <strong>and</strong> times. A self-aligned<br />

native oxide of an Al<strong>GaAs</strong> cladd<strong>in</strong>g layer was used to form a Zn diffusion mask <strong>and</strong><br />

dielectric layer.<br />

R.S.Burton, T.E.Schles<strong>in</strong>ger, D.J.Holmgren, S.C.Smith, R.D.Burnham: Journal of<br />

Applied Physics, 1993, 73[4], 2015-8<br />

[446-106/107-008]<br />

Si<br />

Al<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Data were presented which showed that the native oxide that could form on Al x Ga 1-x As<br />

conf<strong>in</strong><strong>in</strong>g layers (where x was greater than 0.7) on Al y Ga 1-y As/Al z Ga 1-z As superlattices or<br />

quantum-well heterostructures (where y was greater than z) served as an effective barrier<br />

to Si impurity diffusion. It thus impeded impurity-<strong>in</strong>duced layer disorder<strong>in</strong>g. High-quality<br />

native oxide was produced by the conversion of high x-value Al x Ga 1-x As conf<strong>in</strong><strong>in</strong>g layers<br />

(which could be grown on a variety of heterostructures) via H 2 O vapor oxidation (at<br />

temperatures of more than 400C) <strong>in</strong> N 2 carrier gas.<br />

J.M.Dallesasse, N.Holonyak, N.El-Ze<strong>in</strong>, T.A.Richard, F.A.Kish, A.R.Sugg,<br />

R.D.Burnham, S.C.Smith: Applied Physics Letters, 1991, 58[9], 974-6<br />

[446-81/82-002]<br />

Al<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The diffusion <strong>and</strong> drift of Si were studied by means of capacitance-voltage measurements.<br />

These revealed that low substrate temperatures, dur<strong>in</strong>g growth via molecular beam<br />

243


Si (Al,Ga)As Si<br />

epitaxy, were required <strong>in</strong> order to achieve Dirac d-like dopant profiles. It was further<br />

shown theoretically that the r<strong>and</strong>om Poisson distribution, which was usually assumed for<br />

dopant distributions <strong>in</strong> semiconductors, should be modified at high dopant concentrations.<br />

This was because of repulsive <strong>in</strong>teractions between the impurities.<br />

E.F.Schubert, C.W.Tu, R.F.Kopf, J.M.Kuo, L.M.Lunardi: Applied Physics Letters, 1989,<br />

54[25], 2592-4<br />

[446-70/71-103]<br />

Al<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The self-aligned diffusion of Si was studied. It was found that the use of a Si film for the<br />

diffusion led to major problems of morphology degradation <strong>and</strong> dopant contam<strong>in</strong>ation<br />

dur<strong>in</strong>g Si diffusion. A method which <strong>in</strong>volved both a SiO 2 encapsulant <strong>and</strong> a sputtered Si<br />

film source (Si diffusion) or mask (Zn diffusion) was <strong>in</strong>vestigated. The optimum<br />

thicknesses of the Si <strong>and</strong> SiO 2 films were 18 <strong>and</strong> 55nm, respectively.<br />

W.X.Zou, S.Corz<strong>in</strong>e, G.A.Vawter, J.L.Merz, L.A.Coldren, E.L.Hu: Journal of Applied<br />

Physics, 1988, 64[4], 1855-8<br />

[446-72/73-003]<br />

Al<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Data were presented which demonstrated that the surface encapsulant <strong>and</strong> As 4 overpressure<br />

strongly affected Si diffusion <strong>in</strong> Al x Ga 1-x As, <strong>and</strong> were important parameters <strong>in</strong><br />

impurity-<strong>in</strong>duced layer disorder<strong>in</strong>g. An <strong>in</strong>crease <strong>in</strong> the As 4 over-pressure resulted <strong>in</strong> a<br />

decrease <strong>in</strong> the diffusion depth for Al x Ga 1-x As. In addition, the b<strong>and</strong>-edge exciton was<br />

observed <strong>in</strong> absorption on an Al x Ga 1-x As-<strong>GaAs</strong> superlattice that was diffused with Si <strong>and</strong><br />

was converted to bulk crystal Al y Ga 1-y As via impurity-<strong>in</strong>duced layer disorder<strong>in</strong>g. The data<br />

<strong>in</strong>dicated that the Si diffusion process <strong>and</strong> the properties of the diffused material were<br />

different for <strong>GaAs</strong> <strong>and</strong> Al x Ga 1-x As-<strong>GaAs</strong> superlattices which were converted <strong>in</strong>to uniform<br />

Al y Ga 1-y As (where y was between 0 <strong>and</strong> 1) via impurity-<strong>in</strong>duced layer disorder<strong>in</strong>g with<br />

amphoteric dopant Si.<br />

L.J.Guido, W.E.Plano, D.W.Nam, N.Holonyak, J.E.Baker, R.D.Burnham, P.Gavrilovic:<br />

Journal of Electronic Materials, 1988, 17[1], 53-6<br />

[446-60-004]<br />

Al<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The migration of Si <strong>in</strong>to Al x Ga 1-x As from a sputtered Si film was described, where x<br />

ranged from 0 to 0.4. It was shown that both the diffusion rate <strong>and</strong> the surface Si<br />

concentration decreased with <strong>in</strong>creas<strong>in</strong>g Al mole fraction. The diffusion behavior of Si<br />

was expla<strong>in</strong>ed <strong>in</strong> terms of the b<strong>in</strong>d<strong>in</strong>g energy of the Al-As bond <strong>and</strong> of the disorder of the<br />

mixed crystal.<br />

E.Omura, X.S.Wu, G.A.Vawter, E.L.Hu, L.A.Coldren, J.L.Merz: Applied Physics<br />

Letters, 1987, 50[5], 265-6<br />

[446-55/56-001]<br />

244


Si (Al,Ga)As Si<br />

Al<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

A new method for self-aligned Si-Zn diffusion was described. In this method, closed-tube<br />

Si diffusion was carried out by us<strong>in</strong>g a sputtered SiN x film. Then, Zn diffusion which was<br />

self-aligned to the Si diffusion w<strong>in</strong>dow was carried out by re-us<strong>in</strong>g the SiN x film as a<br />

mask. The key factor was that the SiN x film should have the correct refractive <strong>in</strong>dex<br />

profile.<br />

W.X.Zou, R.Boudreau, H.T.Han, T.Bowen, S.S.Shi, D.S.L.Mui, J.L.Merz: Journal of<br />

Applied Physics, 1995, 77[12], 6244-6<br />

[446-121/122-045]<br />

Al<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

A layer of SiO 2 , deposited by sputter<strong>in</strong>g, was used as a diffusion source for Si impurities,<br />

as well as a source of Ga vacancies which enhanced impurity diffusion <strong>and</strong> permitted<br />

reductions to be made <strong>in</strong> the required anneal<strong>in</strong>g temperatures <strong>and</strong> times. A self-aligned<br />

native oxide of an Al<strong>GaAs</strong> cladd<strong>in</strong>g layer was used to form a Zn diffusion mask <strong>and</strong><br />

dielectric layer.<br />

R.S.Burton, T.E.Schles<strong>in</strong>ger, D.J.Holmgren, S.C.Smith, R.D.Burnham: Journal of<br />

Applied Physics, 1993, 73[4], 2015-8<br />

[446-106/107-008]<br />

GaAlAs: Si <strong>Diffusion</strong><br />

The mechanism of Si diffusion <strong>in</strong> Ga 0.7 Al 0.3 As was studied by us<strong>in</strong>g photolum<strong>in</strong>escence<br />

<strong>and</strong> secondary ion mass spectrometry, <strong>and</strong> transmission electron microscopy across the<br />

corner of a wedge-shaped sample. The diffusion source was a grown-<strong>in</strong> highly Si-doped<br />

layer. It was deduced that Frenkel defects (column-<strong>III</strong> vacancies <strong>and</strong> <strong>in</strong>terstitials) were<br />

generated with<strong>in</strong> the highly Si-doped region. The column-<strong>III</strong> <strong>in</strong>terstitials rapidly diffused<br />

towards the surface, where they reacted with the column-<strong>III</strong> vacancies which were<br />

generated at the surface dur<strong>in</strong>g anneal<strong>in</strong>g <strong>in</strong> a gaseous As ambient. This caused a<br />

supersaturation, of column-<strong>III</strong> vacancies <strong>in</strong> the Si-doped region, which drove Si diffusion.<br />

Anneal<strong>in</strong>g <strong>in</strong> vacuum reduced the supersaturation of column-<strong>III</strong> vacancies, <strong>and</strong> thus<br />

decreased Si diffusion. A predom<strong>in</strong>ant Si-donor plus column-<strong>III</strong> vacancy complex<br />

emission b<strong>and</strong> was found <strong>in</strong> spectra from the Si-diffused region. The results supported the<br />

concept of a vacancy-assisted mechanism for Si diffusion <strong>and</strong> impurity-<strong>in</strong>duced<br />

disorder<strong>in</strong>g.<br />

L.Pavesi, N.H.Ky, J.D.Ganière, F.K.Re<strong>in</strong>hart, N.Baba-Ali, I.Harrison, B.Tuck, M.Hen<strong>in</strong>i:<br />

Journal of Applied Physics, 1992, 71[5], 2225-37<br />

[446-86/87-002]<br />

GaAlAs: Si <strong>Diffusion</strong><br />

The effect of the substrate temperature, dur<strong>in</strong>g molecular beam epitaxial growth, upon the<br />

migration of Si atoms <strong>in</strong> d-doped or planar-doped Ga 0.75 Al 0.25 As was <strong>in</strong>vestigated by<br />

us<strong>in</strong>g secondary ion mass spectrometry. For substrate temperatures of 580 to 640C, the Si<br />

spread over about 35nm <strong>in</strong> d-doped Ga 0.75 Al 0.25 As. For substrate temperatures below<br />

245


Si (Al,Ga)As Si<br />

580C, the measured width of the Si profile was limited by the resolution of the secondary<br />

ion mass spectrometer. Magneto-transport measurements were also performed <strong>in</strong> order to<br />

determ<strong>in</strong>e dopant spread<strong>in</strong>g. The Si migration which was measured by means of<br />

secondary ion mass spectrometry was <strong>in</strong> qualitative agreement with the transport results.<br />

However, the secondary ion mass spectrometry data <strong>in</strong>dicated larger Si areal densities.<br />

Two mechanisms, auto-compensation <strong>and</strong> electron localization by a DX center, were<br />

believed to be responsible for the latter observations.<br />

A.M.Lanzillotto, M.Santos, M.Shayegan: Applied Physics Letters, 1989, 55[14], 1445-7<br />

[446-72/73-002]<br />

GaAlAs: Si <strong>Diffusion</strong><br />

A study was made of the contributions of segregation, diffusion <strong>and</strong> aggregation to the<br />

broaden<strong>in</strong>g of d-doped planes of Si <strong>in</strong> Ga 0.67 Al 0.33 As. It was found that sharp spikes of Si<br />

could be obta<strong>in</strong>ed for sheet densities which were below 10 13 /cm 2 <strong>and</strong> for growth<br />

temperatures of 500C or less. At higher temperatures or densities, segregation or<br />

concentration-dependent rapid diffusion could occur; thus caus<strong>in</strong>g significant spread<strong>in</strong>g<br />

even dur<strong>in</strong>g growth. The co-deposition of Si <strong>and</strong> Be markedly reduced this broaden<strong>in</strong>g.<br />

J.J.Harris, J.B.Clegg, R.B.Beall, J.Castagné, K.Woodbridge, C.Roberts: Journal of<br />

Crystal Growth, 1991, 111[1-4], 239-45<br />

[446-91/92-001]<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Under growth conditions which were optimized so as to give the best transport with<br />

normal-side dop<strong>in</strong>g, the migration of the Si dopant towards the <strong>in</strong>verted <strong>in</strong>terface dur<strong>in</strong>g<br />

growth was the ma<strong>in</strong> reason for a reduced <strong>in</strong>verted well mobility. This discovery<br />

permitted the preparation of modulation-doped <strong>in</strong>verted quantum wells of unprecedented<br />

quality.<br />

L.Pfeiffer, E.F.Schubert, K.W.West, C.W.Magee: Applied Physics Letters, 1991, 58[20],<br />

2258-60<br />

[446-84/85-007]<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The migration of Si dur<strong>in</strong>g the metal-organic vapor-phase epitaxial growth of laser<br />

structures was studied by means of secondary ion mass spectroscopy. The migration<br />

process was found to depend ma<strong>in</strong>ly upon the Si concentration <strong>in</strong> the Al<strong>GaAs</strong> layer; for<br />

both silane <strong>and</strong> disilane dop<strong>in</strong>g gases. Above a critical concentration of about 3 x<br />

10 18 /cm 3 , Si migrated <strong>in</strong>to the nom<strong>in</strong>ally undoped <strong>GaAs</strong> layer. This shift <strong>in</strong> the Si front<br />

became even more pronounced when the <strong>GaAs</strong> layer was grown at a lower rate than that<br />

of the Al<strong>GaAs</strong> layer. The Si depth profile had the same gradient as the Al depth profile;<br />

even <strong>in</strong> layers with a large shift of the Si front. Migration appeared to occur preferentially<br />

towards the growth front. It was concluded that the process was governed not only by<br />

diffusion, but also by surface k<strong>in</strong>etics. The effect of Si migration upon the threshold<br />

246


Si (Al,Ga)As Si<br />

current density of broad-area lasers was significant only for a large shift of the Si front<br />

<strong>in</strong>to the active <strong>GaAs</strong> layer.<br />

E.Veuhoff, E.Baumeister, R.Treichler: Journal of Crystal Growth, 1988, 93, 650-5<br />

[446-64/65-159]<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Data were presented which demonstrated that the surface encapsulant <strong>and</strong> As 4 overpressure<br />

strongly affected Si diffusion <strong>in</strong> <strong>GaAs</strong> <strong>and</strong> Al x Ga 1-x As, <strong>and</strong> were important<br />

parameters <strong>in</strong> impurity-<strong>in</strong>duced layer disorder<strong>in</strong>g. An <strong>in</strong>crease <strong>in</strong> the As 4 over-pressure<br />

resulted <strong>in</strong> an <strong>in</strong>crease <strong>in</strong> the diffusion depth <strong>in</strong> the case of <strong>GaAs</strong>, <strong>and</strong> a decrease <strong>in</strong> the<br />

diffusion depth for Al x Ga 1-x As. In addition, the b<strong>and</strong>-edge exciton was observed <strong>in</strong><br />

absorption on an Al x Ga 1-x As-<strong>GaAs</strong> superlattice that was diffused with Si <strong>and</strong> was<br />

converted to bulk crystal Al y Ga 1-y As via impurity-<strong>in</strong>duced layer disorder<strong>in</strong>g. In contrast,<br />

the exciton was not observed dur<strong>in</strong>g absorption on <strong>GaAs</strong> that was diffused with Si, <strong>in</strong><br />

spite of the high degree of compensation. The data <strong>in</strong>dicated that the Si diffusion process<br />

<strong>and</strong> the properties of the diffused material were different for <strong>GaAs</strong> <strong>and</strong> Al x Ga 1-x As-<strong>GaAs</strong><br />

superlattices which were converted <strong>in</strong>to uniform Al y Ga 1-y As (where y was between 0 <strong>and</strong><br />

1) via impurity-<strong>in</strong>duced layer disorder<strong>in</strong>g with amphoteric dopant Si.<br />

L.J.Guido, W.E.Plano, D.W.Nam, N.Holonyak, J.E.Baker, R.D.Burnham, P.Gavrilovic:<br />

Journal of Electronic Materials, 1988, 17[1], 53-6<br />

[446-60-004]<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Data were presented which showed that dislocations <strong>and</strong> Si diffusion promoted<br />

accelerated layer disorder<strong>in</strong>g of quantum well heterostructures which were grown on<br />

<strong>GaAs</strong>/Si substrates by metalorganic chemical vapor deposition. The accelerated impurity<strong>in</strong>duced<br />

layer disorder<strong>in</strong>g was more extreme at temperatures greater than 900C, <strong>and</strong> was<br />

virtually non-existent at temperatures below 775C.<br />

W.E.Plano, D.W.Nam, K.C.Hsieh, L.J.Guido, F.A.Kish, A.R.Sugg, N.Holonyak,<br />

R.J.Matyi, H.Shichijo: Applied Physics Letters, 1989, 55[19], 1993-5<br />

[446-72/73-005]<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Data were presented on the disorder<strong>in</strong>g of an Al<strong>GaAs</strong>/<strong>GaAs</strong> laser structure us<strong>in</strong>g In solid<br />

sources. By us<strong>in</strong>g the <strong>in</strong>dependent or comb<strong>in</strong>ed diffusion of Si <strong>and</strong> In from th<strong>in</strong>-film<br />

sources, it was deduced that In had a higher diffusion coefficient than Si <strong>and</strong> led to a<br />

similar degree of impurity-<strong>in</strong>duced disorder<strong>in</strong>g. The degree of <strong>in</strong>dex guid<strong>in</strong>g was tested<br />

by mak<strong>in</strong>g excess-loss measurements <strong>in</strong> s<strong>in</strong>gle-mode raised-cos<strong>in</strong>e s-bends. It was found<br />

that structures which were patterned by SiO 2 /In disorder<strong>in</strong>g suffered excess losses which<br />

were similar to those <strong>in</strong> structures that were patterned with SiO 2 . An 0.26mm transition<br />

length for 3dB loss was measured for 1000nm-wide guides with 0.1mm guide offsets.<br />

This corresponded to a lateral <strong>in</strong>dex of refraction difference of between 0.8 <strong>and</strong> 1.0%.<br />

247


Si (Al,Ga)As Sn<br />

There was no evidence for an <strong>in</strong>creased l<strong>in</strong>ear loss due to the presence of a dilute In<strong>GaAs</strong><br />

alloy at a measurement wavelength of 870nm.<br />

T.K.Tang, J.J.Alwan, C.M.Herz<strong>in</strong>ger, T.M.Cockerill, A.Crook, T.A.DeTemple,<br />

J.J.Coleman, J.E.Baker: Applied Physics Letters, 1991, 59[22], 2880-2<br />

[446-91/92-005]<br />

GaAlAs/<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The Si migration <strong>and</strong> impurity-<strong>in</strong>duced layer <strong>in</strong>termix<strong>in</strong>g from a buried impurity source<br />

were studied by means of transmission electron microscopic <strong>and</strong> secondary ion mass<br />

spectroscopic studies of isolated Si-doped <strong>GaAs</strong> layers <strong>in</strong> an undoped Ga 0.6 Al 0.4 As/<strong>GaAs</strong><br />

superlattice, <strong>and</strong> by photolum<strong>in</strong>escence measurements of Si-doped quantum wells with<br />

undoped Ga 0.6 Al 0.4 As barriers. In annealed samples, the Si profiles suggested the<br />

occurrence of a Si diffusion process which <strong>in</strong>volved multiply ionized column-<strong>III</strong><br />

vacancies. The width of the resultant Si profile, <strong>and</strong> the spatial extent <strong>and</strong> completeness<br />

of <strong>in</strong>termix<strong>in</strong>g, depended strongly upon the <strong>in</strong>itial Si concentration <strong>in</strong> the doped layer.<br />

K.J.Beern<strong>in</strong>k, R.L.Thornton, G.B.Anderson, M.A.Emanuel: Applied Physics Letters,<br />

1995, 66[19], 2522-4<br />

[446-121/122-053]<br />

Sn<br />

GaAlAs: Sn <strong>Diffusion</strong><br />

A study was made of the contributions of segregation, diffusion <strong>and</strong> aggregation to the<br />

broaden<strong>in</strong>g of d-doped planes of Sn <strong>in</strong> Ga 0.67 Al 0.33 As. It was found that the Sn planes<br />

were severely broadened by all 3 processes. At higher temperatures or densities,<br />

segregation or concentration-dependent rapid diffusion could occur; thus caus<strong>in</strong>g<br />

significant spread<strong>in</strong>g even dur<strong>in</strong>g growth.<br />

J.J.Harris, J.B.Clegg, R.B.Beall, J.Castagné, K.Woodbridge, C.Roberts: Journal of<br />

Crystal Growth, 1991, 111[1-4], 239-45<br />

[446-91/92-001]<br />

Al<strong>GaAs</strong>/AlAs: Sn <strong>Diffusion</strong><br />

Undoped superlattice structures were grown, with or without the presence of 120 Sn<br />

implants, by us<strong>in</strong>g molecular beam epitaxy. They were then annealed under Si 3 N 4 , SiO 2<br />

or encapsulant films. It was found that an enhancement of the Al-Ga <strong>in</strong>terdiffusion<br />

coefficient occurred under the Si 3 N 4 <strong>and</strong> SiO 2 films, due to the <strong>in</strong>-diffusion of Si from the<br />

films. The enhancement was greater dur<strong>in</strong>g diffusion of the Sn implant. In both cases,<br />

<strong>in</strong>termix<strong>in</strong>g enhancement was attributed to the operation of the Fermi effect. Beneath the<br />

WN x film, <strong>in</strong>terdiffusion was suppressed even <strong>in</strong> the presence of the Sn dopant.<br />

E.L.Allen, C.J.Pass, M.D.Deal, J.D.Plummer, V.F.K.Chia: Applied Physics Letters, 1991,<br />

59[25], 3252-4<br />

[446-84/85-006]<br />

248


Te (Al,Ga)As Zn<br />

Te<br />

Al<strong>GaAs</strong>: Te <strong>Diffusion</strong><br />

Monte Carlo simulation was used to model the enhanced disorder<strong>in</strong>g of Al<strong>GaAs</strong>-based<br />

<strong>in</strong>terfaces <strong>in</strong> the presence of high concentrations of Te atoms. The model was based upon<br />

the experimental f<strong>in</strong>d<strong>in</strong>g that the thermal <strong>in</strong>terdiffusion process was similar to the selfdiffusion<br />

of Ga <strong>in</strong> <strong>GaAs</strong>. The model agreed well with experimental data for both Ga selfdiffusion<br />

<strong>and</strong> for <strong>in</strong>termix<strong>in</strong>g. The <strong>in</strong>termix<strong>in</strong>g was found to be caused by the enhanced<br />

solubility of Ga vacancy acceptors <strong>in</strong> the presence of donor Te atoms, <strong>and</strong> not by<br />

diffusion of the Te atoms. The activation energy for the process was found to be about<br />

2.7eV.<br />

K.B.Kahen: Applied Physics Letters, 1988, 53[21], 2071-3<br />

Zn<br />

[446-64/65-158]<br />

Al<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

A new method for self-aligned Si-Zn diffusion was described. In this method, closed-tube<br />

Si diffusion was carried out by us<strong>in</strong>g a sputtered SiN x film. Then, Zn diffusion which was<br />

self-aligned to the Si diffusion w<strong>in</strong>dow was carried out by re-us<strong>in</strong>g the SiN x film as a<br />

mask. The key factor was that the SiN x film should have the correct refractive <strong>in</strong>dex<br />

profile.<br />

W.X.Zou, R.Boudreau, H.T.Han, T.Bowen, S.S.Shi, D.S.L.Mui, J.L.Merz: Journal of<br />

Applied Physics, 1995, 77[12], 6244-6<br />

[446-121/122-045]<br />

Al<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

Data were presented which showed that the native oxide that could form on Al x Ga 1-x As<br />

conf<strong>in</strong><strong>in</strong>g layers (where x was greater than 0.7) on Al y Ga 1-y As/Al z Ga 1-z As superlattices or<br />

quantum-well heterostructures (where y was greater than z) served as an effective barrier<br />

to Zn impurity diffusion. It thus impeded impurity-<strong>in</strong>duced layer disorder<strong>in</strong>g. Highquality<br />

native oxide was produced by the conversion of high x-value Al x Ga 1-x As<br />

conf<strong>in</strong><strong>in</strong>g layers (which could be grown on a variety of heterostructures) via H 2 O vapor<br />

oxidation (at temperatures of more than 400C) <strong>in</strong> N 2 carrier gas.<br />

J.M.Dallesasse, N.Holonyak, N.El-Ze<strong>in</strong>, T.A.Richard, F.A.Kish, A.R.Sugg,<br />

R.D.Burnham, S.C.Smith: Applied Physics Letters, 1991, 58[9], 974-6<br />

[446-81/82-002]<br />

Al<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The self-aligned diffusion of Zn was studied. It was found that the use of a Si film for the<br />

diffusion led to major problems of morphology degradation <strong>and</strong> dopant contam<strong>in</strong>ation<br />

249


Zn (Al,Ga)As Zn<br />

dur<strong>in</strong>g Si diffusion. A method which <strong>in</strong>volved both a SiO 2 encapsulant <strong>and</strong> a sputtered Si<br />

film source (Si diffusion) or mask (Zn diffusion) was <strong>in</strong>vestigated. The optimum<br />

thicknesses of the Si <strong>and</strong> SiO 2 films were 18 <strong>and</strong> 55nm, respectively.<br />

W.X.Zou, S.Corz<strong>in</strong>e, G.A.Vawter, J.L.Merz, L.A.Coldren, E.L.Hu: Journal of Applied<br />

Physics, 1988, 64[4], 1855-8<br />

[446-72/73-003]<br />

Al<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The diffusion of Zn was studied by us<strong>in</strong>g liquid-phase epitaxy methods, <strong>and</strong> Si-doped n-<br />

type substrate material. The measurements were carried out at 850C, <strong>and</strong> dopant<br />

concentrations which ranged from 10 18 to 10 19 /cm 3 were <strong>in</strong>troduced. It was found that the<br />

Zn concentration <strong>in</strong> the solid depended upon the square root of the atomic fraction of Zn<br />

<strong>in</strong> the liquid. The diffusivity was dom<strong>in</strong>ated by the <strong>in</strong>terstitial-substitutional process, <strong>and</strong><br />

exhibited a cubic dependence upon the Zn content. The Zn <strong>in</strong>terstitial was ma<strong>in</strong>ly doublyionized<br />

Zn i 2+ . It was noted that Al played the role of a catalyst <strong>in</strong> the diffusion process.<br />

The Zn diffusion coefficient <strong>in</strong> Al 0.85 Ga 0.15 As was some 4 times greater than that <strong>in</strong><br />

<strong>GaAs</strong>.<br />

C.Algora, G.L.Araujo, A.Marti: Journal of Applied Physics, 1990, 68[6], 2723-30<br />

[446-86/87-002]<br />

Al<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The use of th<strong>in</strong> Si films for the selective-area diffusion of Si <strong>and</strong> Zn was described. It was<br />

found that Si films behaved as ideal masks for Zn diffusion at temperatures below 750C.<br />

Ideal lateral Zn diffusion profiles were also observed when us<strong>in</strong>g these films; regardless<br />

of the stress at the <strong>in</strong>terface.<br />

G.A.Vawter, E.Omura, X.S.Wu, J.L.Merz, L.Coldren, E.Hu: Journal of Applied Physics,<br />

1988, 63[11], 5541-7<br />

[446-72/73-003]<br />

Al<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The state of Zn diffusion at the hetero-<strong>in</strong>terface of 660nm double-hetero light-emitt<strong>in</strong>g<br />

diodes was <strong>in</strong>vestigated by us<strong>in</strong>g the electron beam-<strong>in</strong>duced current method. The p-n<br />

junction penetrated towards the n-cladd<strong>in</strong>g layer, as a result of Zn diffusion, when the<br />

carrier concentration of the p-active layer was greater than 10 18 /cm 3 . The distance<br />

between the hetero-<strong>in</strong>terface <strong>and</strong> the p-n junction was related to the optical output <strong>and</strong><br />

modulation b<strong>and</strong>-width of a light-emitt<strong>in</strong>g diode. The dependence of the Zn effective<br />

diffusion coefficient upon the carrier concentration of the p-active layer <strong>and</strong> n-cladd<strong>in</strong>g<br />

layer was <strong>in</strong>vestigated. It was concluded that a suitable growth temperature for the active<br />

layer was about 850C.<br />

M.Yoneda, Y.Nakamura, A.Tsushi, K.Ichimura: Japanese Journal of Applied Physics,<br />

1993, 32[1-9A], 3770-4<br />

[446-109/110-025]<br />

250


Zn (Al,Ga)As Zn<br />

GaAlAs: Zn <strong>Diffusion</strong><br />

The entry of Zn <strong>in</strong>to Ga 0.7 Al 0.3 As <strong>and</strong> s<strong>in</strong>gle heterostructures was studied. It was found<br />

that the depth of the diffusion front was proportional to the square root of the diffusion<br />

time. In the case of heterostructures, the Ga 0.7 Al 0.3 As layer thickness modified the<br />

relationship by decreas<strong>in</strong>g the junction depth to an extent which was some multiple of the<br />

layer thickness. The relationship could be used to predict diffusion fronts <strong>in</strong> double<br />

heterostructures.<br />

H.J.Yoo, Y.S.Kwon: Journal of Electronic Materials, 1988, 17[4], 337-9<br />

[446-62/63-203]<br />

GaAlAs: Zn <strong>Diffusion</strong><br />

Samples of Ga 0.62 Al 0.38 As were diffused with Zn, via a 200 to 300nm protective ZrO 2<br />

layer. The diffusion depth exhibited a square-root time dependence. The absolute<br />

diffusivity values depended slightly upon the diffusion conditions. The layer had<br />

essentially no effect upon the carrier concentration profile or the activation energy.<br />

J.E.Bisberg, A.K.Ch<strong>in</strong>, F.P.Dabkowski: Journal of Applied Physics, 1990, 67[3], 1347-51<br />

[446-74-003]<br />

GaAlAs: Zn <strong>Diffusion</strong><br />

Samples of Ga 0.7 Al 0.3 As were grown onto a Si substrate <strong>and</strong> were diffused with Zn from a<br />

ZnAs 2 source. It was found that the Zn diffusivity was greater <strong>in</strong> these layers than <strong>in</strong><br />

layers which were grown on a <strong>GaAs</strong> substrate. The effective diffusion coefficient was<br />

related to the defect density <strong>in</strong> the GaAlAs, <strong>and</strong> to the diffusion depth. A simple model<br />

showed that the diffusivity along dislocations was some 5 times higher than that <strong>in</strong><br />

dislocation-free bulk material.<br />

S.Sakai, Y.Terauchi, N.Wada, Y.Sh<strong>in</strong>tani: Japanese Journal of Applied Physics, 1991,<br />

30[9A], 1942-3<br />

[446-84/85-011]<br />

GaAlAs: Zn <strong>Diffusion</strong><br />

A simple method for the open-tube diffusion of Zn from (ZnO) x (SiO 2 ) 1-x film sources,<br />

<strong>and</strong> <strong>in</strong>to Ga 0.8 Al 0.2 As was described. The oxide films were deposited by us<strong>in</strong>g metalorganic<br />

chemical vapor deposition. A capp<strong>in</strong>g layer of SiO 2 was deposited on top of the<br />

source films, <strong>and</strong> diffusion was carried out <strong>in</strong> flow<strong>in</strong>g N at 650C. <strong>Diffusion</strong> depths of<br />

between 200nm <strong>and</strong> several microns could be easily obta<strong>in</strong>ed. The diffusion front <strong>in</strong> n-<br />

type substrates was sharp. The dependence of the diffusion depth upon the source film<br />

composition (for x-values of 0.04 to 1) was determ<strong>in</strong>ed by us<strong>in</strong>g section<strong>in</strong>g methods.<br />

D.J.Lawrence, F.T.Smith, S.T.Lee: Journal of Applied Physics, 1991, 69[5], 3011-5<br />

[446-78/79-002]<br />

251


Zn (Al,Ga)As Zn<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The diffusivity of ion-implanted Zn was deduced from secondary ion mass spectrometry<br />

profiles. <strong>Diffusion</strong> anneal<strong>in</strong>g was carried for various times at 750C. It was found that the<br />

diffusivity of Zn was proportional to the square of the Zn concentration. This implied the<br />

existence of local thermal equilibrium. The absolute values were 200 times smaller than<br />

those which had been reported for gaseous-source Zn diffusion at 650C <strong>in</strong> <strong>GaAs</strong>. The<br />

superlattice disorder<strong>in</strong>g rate <strong>in</strong>creased with <strong>in</strong>creas<strong>in</strong>g Zn concentration <strong>and</strong> was<br />

attributed to the diffusion of positively charged <strong>in</strong>terstitials such as Ga n+ or Al n+ , where n<br />

was between 2 <strong>and</strong> 3.<br />

E.P.Zucker, A.Hashimoto, T.Fukunaga, N.Watanabe: Applied Physics Letters, 1989,<br />

54[6], 564-6<br />

[446-64/65-159]<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

It was recalled that previous work had <strong>in</strong>dicated that (Si 2 ) x (<strong>GaAs</strong>) 1-x could be formed<br />

with<strong>in</strong> the <strong>GaAs</strong> quantum well of an Al x Ga 1-x As-<strong>GaAs</strong> quantum well heterostructure. It<br />

was shown here that the Si concentration <strong>in</strong> the (Si 2 ) x (<strong>GaAs</strong>) 1-x layer greatly exceeded<br />

typical dop<strong>in</strong>g levels. The stability of the quantum well heterostructures was <strong>in</strong>vestigated<br />

with respect to thermal anneal<strong>in</strong>g <strong>and</strong> Zn impurity-<strong>in</strong>duced layer disorder<strong>in</strong>g. The results<br />

showed that the (Si 2 ) x (<strong>GaAs</strong>) 1-x alloy was stable with respect to thermal anneal<strong>in</strong>g unless<br />

a rich source of Ga vacancies was provided. Also, relatively low-temperature Zn diffusion<br />

greatly enhanced the disorder<strong>in</strong>g of the alloy layer.<br />

L.J.Guido, N.Holonyak, K.C.Hsieh, R.W.Kaliski, J.E.Baker, D.G.Deppe, R.D.Burnham,<br />

R.L.Thornton, T.L.Paoli: Journal of Electronic Materials, 1987, 16[1], 87-91<br />

[446-51/52-111]<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

It was po<strong>in</strong>ted out that the characteristic features of Zn diffusion <strong>in</strong> <strong>GaAs</strong> substrates <strong>and</strong><br />

<strong>GaAs</strong>/Al<strong>GaAs</strong> superlattices could be expla<strong>in</strong>ed <strong>in</strong> terms of a kick-out mechanism <strong>in</strong><br />

which the doubly positively charged Ga self-<strong>in</strong>terstitial governed Ga self-diffusion. Such<br />

characteristics <strong>in</strong>cluded a square-law dependence of the Zn diffusivity upon its own<br />

background concentration under Zn iso-concentration diffusion conditions, various Zn <strong>in</strong>diffusion<br />

profiles, much lower diffusivities of Zn under out-diffusion conditions than<br />

under <strong>in</strong>-diffusion conditions, <strong>and</strong> a huge enhancement of Zn <strong>in</strong>-diffusion dur<strong>in</strong>g<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong> superlattice disorder<strong>in</strong>g. It was found that the Long<strong>in</strong>i mechanism was able<br />

to expla<strong>in</strong> most of these features.<br />

S.Yu, T.Y.Tan, U.Gösele: Journal of Applied Physics, 1991, 69[6], 3547-65<br />

[446-86/87-009]<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

Data were presented on the reduction of layer <strong>in</strong>termix<strong>in</strong>g <strong>in</strong> quantum-well<br />

heterostructures, dur<strong>in</strong>g high-temperature anneal<strong>in</strong>g, by us<strong>in</strong>g an <strong>in</strong>itial low-temperature<br />

252


Zn (Al,Ga)As General<br />

block<strong>in</strong>g diffusion of Zn. Room-temperature photolum<strong>in</strong>escence measurements of the<br />

<strong>in</strong>crease <strong>in</strong> the lowest electron to heavy-hole transition energy <strong>in</strong> the quantum-wells were<br />

used to characterize the extent of layer <strong>in</strong>termix<strong>in</strong>g. Doped (C <strong>and</strong> Si) samples which had<br />

been annealed (850C, 12h) after a low-temperature block<strong>in</strong>g Zn diffusion (480C)<br />

exhibited reductions <strong>in</strong> energy shift from about 0.177eV, to as little as 0.018eV. Similar<br />

effects were observed, but to a lesser extent, <strong>in</strong> the case of undoped samples. The<br />

improved thermal stability was attributed to a Zn diffusion-<strong>in</strong>duced reduction <strong>in</strong> the<br />

number of column-<strong>III</strong> vacancies <strong>in</strong> the active layers. This was confirmed by secondaryion<br />

mass spectroscopy measurements.<br />

M.R.Krames, A.D.M<strong>in</strong>erv<strong>in</strong>i, E.I.Chen, N.Holonyak, J.E.Baker: Applied Physics Letters,<br />

1995, 67[13], 1859-61<br />

[446-125/126-112]<br />

GaAlAs/<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The migration of th<strong>in</strong> highly p-doped layers <strong>in</strong> s<strong>in</strong>gle <strong>and</strong> double heterostructures, grown<br />

us<strong>in</strong>g metalorganic vapor-phase epitaxy, was studied us<strong>in</strong>g capacitance-voltage etch<br />

profil<strong>in</strong>g <strong>and</strong> secondary ion mass spectrometry. It was deduced that the diffusivity of Zn<br />

<strong>in</strong> Ga 0.7 Al 0.3 As could be described by:<br />

D (cm 2 /s) = 1.5 x 10 -3 exp[-2.2(eV)/kT]<br />

for rapid thermal anneal<strong>in</strong>g. A model which was based upon an <strong>in</strong>terstitial cum<br />

substitutional diffusion mechanism, with certa<strong>in</strong> k<strong>in</strong>etic limitations, was successfully used<br />

to simulate the observed dopant concentration profiles. Markedly anomalous diffusion of<br />

Zn, from <strong>GaAs</strong> <strong>and</strong> <strong>in</strong>to highly n-doped GaAlAs, was found.<br />

N.Nordell, P.Ojala, W.H.Van, G.L<strong>and</strong>gren, M.K.L<strong>in</strong>narsson: Journal of Applied Physics,<br />

1990, 67[2], 778-86<br />

[446-74-004]<br />

- miscellaneous<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: <strong>Diffusion</strong><br />

A model was presented which accounted for the anomalous diffusion of p-type dopants<br />

dur<strong>in</strong>g the growth of bipolar transistors. The model was based upon Fermi-level p<strong>in</strong>n<strong>in</strong>g<br />

at the crystal surface dur<strong>in</strong>g epitaxial growth. This led to an <strong>in</strong>creased concentration of<br />

column-<strong>III</strong> <strong>in</strong>terstitial defects <strong>in</strong> heavily n-type Al<strong>GaAs</strong> or <strong>GaAs</strong>. The excess column-<strong>III</strong><br />

<strong>in</strong>terstitials which were generated <strong>in</strong> the n-type crystal then flowed <strong>in</strong>to the p + base region<br />

<strong>and</strong> led to a transfer of p-type impurity atoms from column-<strong>III</strong> lattice sites to <strong>in</strong>terstitial<br />

positions, via a kick-out mechanism. Once located <strong>in</strong> <strong>in</strong>terstitial positions, the impurity<br />

atoms diffused rapidly. The model was consistent with previously proposed mechanisms<br />

for both impurity diffusion <strong>and</strong> column-<strong>III</strong> self-diffusion.<br />

D.G.Deppe: Applied Physics Letters, 1990, 56[4], 370-2<br />

[446-74-004]<br />

253


Surface (Al,Ga)As Surface<br />

Surface <strong>Diffusion</strong><br />

Al<strong>GaAs</strong>: Surface <strong>Diffusion</strong><br />

It was found that surface migration was effectively enhanced by evaporat<strong>in</strong>g Ga or Al<br />

atoms onto a clean surface under an As-free atmosphere or low As pressure. This<br />

characteristic was exploited by alternately supply<strong>in</strong>g Ga <strong>and</strong>/or Al <strong>and</strong> As to the substrate<br />

surface <strong>in</strong> order to grow atomically-flat <strong>GaAs</strong>-Al<strong>GaAs</strong> hetero-<strong>in</strong>terfaces, <strong>and</strong> also to grow<br />

high-quality Al<strong>GaAs</strong> layers at very low temperatures. The migration characteristics of<br />

surface adatoms were <strong>in</strong>vestigated by us<strong>in</strong>g reflection high-energy electron diffraction<br />

measurements. It was found that differ<strong>in</strong>g growth mechanisms operated at high <strong>and</strong> low<br />

temperatures. Both mechanisms were expected to yield flat heterojunction <strong>in</strong>terfaces. By<br />

apply<strong>in</strong>g this method, <strong>GaAs</strong>-Al<strong>GaAs</strong> s<strong>in</strong>gle quantum-well structures could be grown at<br />

substrate temperatures of 200 <strong>and</strong> 300C, respectively.<br />

Y.Horikoshi, M.Kawashima, H.Yamaguchi: Japanese Journal of Applied. Physics, 1988,<br />

27[2], 169-79<br />

[446-60-001]<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Surface <strong>Diffusion</strong><br />

Ribbed crystals could be grown <strong>in</strong> a s<strong>in</strong>gle process<strong>in</strong>g step because the ribs were def<strong>in</strong>ed<br />

by the two non-grow<strong>in</strong>g (111)B surfaces which developed at each edge of (011) mesas on<br />

a patterned <strong>GaAs</strong> substrate dur<strong>in</strong>g the organometallic chemical vapor deposition of<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong> structures. The study revealed the importance of surface diffusionenhanced<br />

crystal growth when a growth surface was adjacent to a non-grow<strong>in</strong>g surface<br />

such as a (111)B facet. The magnitude of this effect suggested that the present deposition<br />

technique was well-suited to the growth of structures which were tapered <strong>in</strong> 3<br />

dimensions.<br />

E.Colas, A.Shahar, W.J.Toml<strong>in</strong>son: Applied Physics Letters, 1990, 56[10], 955-7<br />

[446-74-005]<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Surface <strong>Diffusion</strong><br />

Monte Carlo methods were used to model the adatom migration of cations <strong>in</strong> Al x Ga 1-x As,<br />

where x was equal to 0, 0.5 or 1, <strong>and</strong> the resultant atomic arrangements on a reconstructed<br />

As-stabilized <strong>GaAs</strong>(001) surface with an adatom coverage of up to 0.5. It was found that<br />

the cation adatom migration depended strongly upon the adatom coverage of the surface.<br />

R<strong>and</strong>omly imp<strong>in</strong>g<strong>in</strong>g cations occupied lattice sites on the As dimers at coverages of less<br />

than 0.1. As the coverage was <strong>in</strong>creased from 0.1 to 0.3, the imp<strong>in</strong>g<strong>in</strong>g cations migrated<br />

ma<strong>in</strong>ly along the miss<strong>in</strong>g dimer rows. At a coverage of more than 0.3, adatoms tended to<br />

favor lattice sites on the As dimers; <strong>in</strong>clud<strong>in</strong>g those with non-tetrahedral coord<strong>in</strong>ation. In<br />

the case of Al 0.5 Ga 0.5 As, lattice sites along the miss<strong>in</strong>g dimer were occupied ma<strong>in</strong>ly by Al<br />

adatoms, while those on As dimers were favored by Ga adatoms. This was because Al<br />

adatom migration was several times slower than Ga adatom migration. The resultant<br />

254


Surface (Al,Ga)As Interdiffusion<br />

atomic arrangements were expla<strong>in</strong>ed <strong>in</strong> terms of a coverage dependence of the migration<br />

potential.<br />

T.Ito, K.Shiraishi, T.Ohno: Applied Surface Science, 1994, 82-83, 208-13<br />

[446-121/122-047]<br />

Interdiffusion<br />

1.0E-14<br />

1.0E-15<br />

1.0E-16<br />

table 3<br />

table 4<br />

table 5<br />

table 6<br />

table 32<br />

D (cm 2 /s)<br />

1.0E-17<br />

1.0E-18<br />

1.0E-19<br />

1.0E-20<br />

7 8 9 10<br />

10 4 /T(K)<br />

Figure 1: Interdiffusivity <strong>in</strong> Al<strong>GaAs</strong>/<strong>GaAs</strong><br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Interdiffusion<br />

A systematic study was made of impurity-free Al-Ga <strong>in</strong>terdiffusion <strong>in</strong> superlattices which<br />

were sealed <strong>in</strong>to ampoules. Four structures were used, with superlattice periods that<br />

ranged from 9 to 52nm. Three ambients were explored: along the Ga-rich solidus, with no<br />

excess Ga or As <strong>in</strong> the evacuated ampoule, or with an excess As content which was less<br />

than that required to reach the As-rich solidus limit. In each of the ambients, the<br />

Arrhenius dependence of the Al-Ga <strong>in</strong>terdiffusion coefficient could be described by a<br />

s<strong>in</strong>gle activation energy at temperatures rang<strong>in</strong>g from 700 to 1050C. Excellent agreement<br />

was obta<strong>in</strong>ed for the Al-Ga <strong>in</strong>terdiffusion coefficients which were measured by us<strong>in</strong>g<br />

superlattices on Si-doped <strong>and</strong> undoped <strong>GaAs</strong> substrates. By normalization to a constant<br />

255


Interdiffusion (Al,Ga)As Interdiffusion<br />

As over-pressure of 1atm, the Ga- <strong>and</strong> As-rich activation energies were deduced to be<br />

3.26 <strong>and</strong> 4.91eV, respectively. These activation energies were <strong>in</strong> the range which was<br />

predicted for Al-Ga <strong>in</strong>terdiffusion, mediated by group-<strong>III</strong> vacancy second-nearest<br />

neighbor hopp<strong>in</strong>g. An <strong>in</strong>crease <strong>in</strong> energy which occurred upon go<strong>in</strong>g from Ga- to As-rich<br />

conditions was attributed to a shift <strong>in</strong> the Fermi-level position, towards the valence b<strong>and</strong>;<br />

with an <strong>in</strong>crease <strong>in</strong> the ionized group-<strong>III</strong> vacancy concentration.<br />

B.L.Olmsted, S.N.Houde-Walter: Applied Physics Letters, 1993, 63[4], 530-2<br />

[446-106/107-016]<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Interdiffusion<br />

Impurity-<strong>in</strong>duced layer disorder<strong>in</strong>g experiments were performed on quantum-well<br />

heterostructures which were heavily doped with C. The results showed that the presence<br />

of C retarded Al <strong>and</strong> Ga <strong>in</strong>terdiffusion, as compared with un-doped material.<br />

Interdiffusion <strong>in</strong> C-doped quantum-well heterostructures was not enhanced by the use of<br />

a Ga-rich rather than an As-rich anneal<strong>in</strong>g ambient. The data were <strong>in</strong>consistent with most<br />

Fermi-level effect models for layer disorder<strong>in</strong>g which did not <strong>in</strong>clude a chemical impurity<br />

dependence or sub-lattice dependence, <strong>and</strong> which did not take account of the possibility<br />

of <strong>in</strong>hibited Al <strong>and</strong> Ga <strong>in</strong>terdiffusion <strong>in</strong> extr<strong>in</strong>sic crystals.<br />

L.J.Guido, B.T.Cunn<strong>in</strong>gham, D.W.Nam, K.C.Hsieh, W.E.Plano, J.S.Major, E.J.Vesely,<br />

A.R.Sugg, N.Holonyak, G.E.Stillman: Journal of Applied Physics, 1990, 67[4], 2179-82<br />

[446-74-004]<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Interdiffusion<br />

Impurity-<strong>in</strong>duced layer disorder<strong>in</strong>g or <strong>in</strong>termix<strong>in</strong>g <strong>in</strong> quantum well heterostructures <strong>and</strong><br />

superlattices were reviewed. On the basis of the behavior of column-<strong>III</strong> vacancies <strong>and</strong><br />

<strong>in</strong>terstitials, suitable models for layer disorder<strong>in</strong>g were developed.<br />

D.G.Deppe, N.Holonyak: Journal of Applied Physics, 1988, 64[12], R93-113<br />

[446-72/73-005]<br />

34 Al<strong>GaAs</strong>/<strong>GaAs</strong>: Interdiffusion<br />

Data were presented which showed that the Al-Ga <strong>in</strong>terdiffusion coefficient for an<br />

Al x Ga 1-x As-<strong>GaAs</strong> quantum-well heterostructure or a superlattice was highly dependent<br />

upon the crystal encapsulation conditions (table 4). The activation energy for Al-Ga<br />

<strong>in</strong>terdiffusion, <strong>and</strong> thus layer-disorder<strong>in</strong>g, was smaller (about 3.5eV) for dielectric<br />

encapsulated samples than after capless anneal<strong>in</strong>g (about 4.7eV). The <strong>in</strong>terdiffusion<br />

coefficient for Si 3 N 4 -capped samples was almost an order of magnitude smaller than for<br />

the cases of capless or SiO 2 -capped samples at temperatures of between 800 <strong>and</strong> 875C.<br />

As well as the type of encapsulant, the encapsulation geometry (stripes or capped stripes)<br />

was important because of stra<strong>in</strong> effects. These were a major source of anisotropic Al-Ga<br />

<strong>in</strong>terdiffusion.<br />

L.J.Guido, N.Holonyak, K.C.Hsieh, R.W.Kaliski, W.E.Plano, R.D.Burnham,<br />

R.L.Thornton, J.E.Epler, T.L.Paoli: Journal of Applied Physics, 1987, 61[4], 1372-9<br />

[446-60-002]<br />

256


Interdiffusion (Al,Ga)As Interdiffusion<br />

Table 4<br />

Interdiffusivity (Al-Ga) <strong>in</strong> Al<strong>GaAs</strong>/<strong>GaAs</strong><br />

Conditions Temperature (C) D (cm 2 /s)<br />

capless 875 1.1 x 10 -17<br />

SiO 2 cap 875 9.6 x 10 -18<br />

capless 850 2.9 x 10 -18<br />

SiO 2 cap 850 2.9 x 10 -18<br />

SiO 2 cap 825 2.8 x 10 -18<br />

capless 825 1.0 x 10 -18<br />

Si 3 N 4 cap 875 2.2 x 10 -18<br />

Si 3 N 4 cap 850 7.8 x 10 -19<br />

SiO 2 cap 800 7.0 x 10 -19<br />

capless 800 4.5 x 10 -19<br />

Si 3 N 4 cap 825 3.8 x 10 -19<br />

Si 3 N 4 cap 800 2.2 x 10 -19<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Interdiffusion<br />

A model was presented which described the diffusion of Si <strong>in</strong>to <strong>GaAs</strong> from grown-<strong>in</strong><br />

dopant sources. The effects of background impurities upon Si diffusion <strong>and</strong> layer<br />

<strong>in</strong>terdiffusion <strong>in</strong> the present superlattices were also described. Epitaxial <strong>GaAs</strong> samples<br />

with alternat<strong>in</strong>g doped <strong>and</strong> undoped layers, <strong>and</strong> superlattices with Mg- or Si-doped layers,<br />

were studied. Various anneal<strong>in</strong>g conditions were used to study <strong>in</strong>teractions between<br />

grown-<strong>in</strong> impurities <strong>and</strong> native defects. A model which described impurity diffusion <strong>and</strong><br />

Al-Ga layer <strong>in</strong>terdiffusion was based upon the behavior of column-<strong>III</strong> vacancies (V <strong>III</strong> ) <strong>and</strong><br />

<strong>in</strong>terstitials (I <strong>III</strong> ), <strong>and</strong> the control of their contents. The results <strong>in</strong>dicated that n-type<br />

superlattices underwent enhanced layer <strong>in</strong>terdiffusion because of an <strong>in</strong>creased solubility<br />

of the V <strong>III</strong> defect. Enhanced layer <strong>in</strong>terdiffusion <strong>in</strong> p-type superlattices was attributed to<br />

an enhanced solubility of I <strong>III</strong> .<br />

D.G.Deppe, N.Holonyak, W.E.Plano, V.M.Robb<strong>in</strong>s, J.M.Dallesasse, K.C.Hsieh,<br />

J.E.Baker: Journal of Applied Physics, 1988, 64[4], 1838-44<br />

[446-72/73-006]<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Interdiffusion<br />

It was shown that donor diffusion <strong>and</strong> layer <strong>in</strong>termix<strong>in</strong>g were greatly enhanced <strong>in</strong> the<br />

presence of defects which were created by crystal overgrowth on locally laser-melted<br />

substrates. Accelerated defect <strong>and</strong> impurity-<strong>in</strong>duced layer disorder<strong>in</strong>g, <strong>and</strong> donor<br />

diffusion from a solid SiO 2 source, a Ge vapor source or a grown-<strong>in</strong> Se source were<br />

257


Interdiffusion (Al,Ga)As Interdiffusion<br />

observed <strong>in</strong> regions of high defect density. Enhanced donor diffusion <strong>and</strong> crystal selfdiffusion<br />

were attributed to an <strong>in</strong>creased density of column-<strong>III</strong> defects <strong>and</strong> dislocations.<br />

F.A.Kish, W.E.Plano, K.C.Hsieh, A.R.Sugg, N.Holonyak, J.E.Baker: Journal of Applied<br />

Physics, 1989, 66[12], 5821-5<br />

[446-74-005]<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Interdiffusion<br />

A study was made of the <strong>in</strong>-diffusion of various group-IV <strong>and</strong> group-VI n-type impurities.<br />

In all cases, the n-type dopants enhanced the Al-Ga <strong>in</strong>terdiffusion coefficient above that<br />

which could be attributed to an As over-pressure alone. The Si-<strong>in</strong>duced enhancement had<br />

previously been attributed to a change <strong>in</strong> Fermi-level position with dop<strong>in</strong>g <strong>and</strong> could<br />

therefore account for disorder<strong>in</strong>g by <strong>other</strong> n-type impurities. However, important<br />

differences were observed <strong>in</strong> the <strong>in</strong>terdiffusion characteristics that were <strong>in</strong>duced by Si or<br />

Ge, <strong>and</strong> by S or Se. The disorder<strong>in</strong>g was attributed to an enhancement of the group-<strong>III</strong><br />

vacancy concentration for each of these n-type impurities. This was also true of undoped<br />

crystals which were disordered by an As ambient alone at 855C.<br />

B.L.Olmsted, S.N.Houde-Walter: Applied Physics Letters, 1993, 62[13], 1516-8<br />

[446-106/107-016]<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Interdiffusion<br />

The As vapor pressure dependence of <strong>in</strong>terdiffusion <strong>in</strong> a hetero-<strong>in</strong>terface at high<br />

temperatures was studied by measur<strong>in</strong>g the wavelength shift of the photolum<strong>in</strong>escence <strong>in</strong><br />

a multi quantum well. It was found that <strong>in</strong>terdiffusion at a temperature of 850C was<br />

m<strong>in</strong>imized by an As pressure of 100torr <strong>and</strong> was enhanced at both lower <strong>and</strong> higher As<br />

pressures. A degradation of the photolum<strong>in</strong>escence <strong>in</strong>tensity was observed only at higher<br />

As pressures. These effects were attributed to the presence of excess Al <strong>and</strong> Ga<br />

vacancies, <strong>and</strong> their associated defects.<br />

A.Furuya, O.Wada, A.Takamori, H.Hashimoto: Japanese Journal of Applied Physics,<br />

1987, 26[6], L926-8<br />

[446-55/56-002]<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Interdiffusion<br />

Transmission electron microscopy <strong>and</strong> carrier concentration measurements were used to<br />

characterize the layer <strong>in</strong>terdiffusion mechanism of a Se-doped Al x Ga 1-x As-<strong>GaAs</strong><br />

superlattice dur<strong>in</strong>g high-temperature anneal<strong>in</strong>g. By vary<strong>in</strong>g the anneal<strong>in</strong>g environment<br />

<strong>and</strong> compar<strong>in</strong>g the results with similarly annealed un-doped superlattices <strong>and</strong> Mg-doped<br />

superlattices, it was found that layer <strong>in</strong>terdiffusion occurred via the <strong>in</strong>teraction of the Se<br />

impurity with native defects which were associated with As-rich conditions. The most<br />

likely c<strong>and</strong>idate was suggested to be the column-<strong>III</strong> vacancy.<br />

D.G.Deppe, N.Holonyak, K.C.Hsieh, P.Gavrilovic, W.Stutius, J.Williams. Applied<br />

Physics Letters, 1987, 51[8], 581-3<br />

[446-55/56-002]<br />

258


Interdiffusion (Al,Ga)As Interdiffusion<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Interdiffusion<br />

A study was made of the role of defect diffusion, from crystal surfaces, <strong>in</strong> the disorder<strong>in</strong>g<br />

of a multiple quantum well structure that was Si-doped dur<strong>in</strong>g molecular beam epitaxial<br />

growth. The distribution of the native defects was deduced from photolum<strong>in</strong>escence<br />

spectroscopic, secondary ion mass spectrometric, <strong>and</strong> electrochemical C-V profil<strong>in</strong>g data.<br />

No significant difference was observed between the Al-Ga <strong>in</strong>terdiffusion coefficients of<br />

Si-doped <strong>and</strong> undoped superlattices when they were annealed with excess Ga. This was<br />

attributed to the lack of a source of group-<strong>III</strong> vacancies. Only a small fraction of the<br />

enhancement which was predicted to result from Si dop<strong>in</strong>g was observed when excess As<br />

was used <strong>in</strong>stead. The largest Fermi-level enhancement was observed when no excess Ga<br />

or As was present <strong>in</strong> the evacuated ampoule. The results <strong>in</strong>dicated that the crystal surface<br />

was both source <strong>and</strong> s<strong>in</strong>k for the native defects which were known to mediate Al-Ga<br />

<strong>in</strong>terdiffusion. Significant electrical compensation of the donors was observed after<br />

anneal<strong>in</strong>g <strong>in</strong> both As-rich or Ga-rich ambients. This was attributed to ionized group-<strong>III</strong><br />

vacancy generation <strong>in</strong> the former case, <strong>and</strong> to Si atoms which moved from group-<strong>III</strong> to<br />

group-V sites <strong>in</strong> the latter case.<br />

B.L.Olmsted, S.N.Houde-Walter: Applied Physics Letters, 1993, 63[8], 1131-3<br />

[446-106/107-017]<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Interdiffusion<br />

Spatially resolved values of the Al/Ga <strong>in</strong>terdiffusion coefficient for p-i-n <strong>and</strong> n-i-p<br />

Al<strong>GaAs</strong>-<strong>GaAs</strong> device structures were found to be almost identical <strong>in</strong> magnitude, but<br />

varied with position (by a factor of 2) across a 1µ-thick multiple quantum well active<br />

region. These observations contrasted with theoretical predictions, given that the Fermi<br />

level to valence-b<strong>and</strong> energy separation changed by 0.7eV across the <strong>in</strong>tr<strong>in</strong>sic region, <strong>and</strong><br />

suggested that impurity-free layer disorder<strong>in</strong>g did not provide the necessary uniformity <strong>in</strong><br />

energy shift for photonic <strong>in</strong>tegrated circuit fabrication.<br />

S.Seshadri, L.J.Guido, P.Mitev: Applied Physics Letters, 1995, 67[4], 497-9<br />

[446-123/124-157]<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Interdiffusion<br />

Direct optical observations were made of diffusion-related deep levels that were<br />

associated with <strong>in</strong>terdiffusion <strong>in</strong> superlattice structures. Low-energy cathodolum<strong>in</strong>escence<br />

spectroscopy was used to <strong>in</strong>vestigate the formation <strong>and</strong> evolution of deep levels under<br />

various conditions. It was found that the spatial distribution of the deep levels was<br />

strongly related to the extent of superlattice <strong>in</strong>termix<strong>in</strong>g, as measured us<strong>in</strong>g secondary ion<br />

mass spectrometry <strong>and</strong> photolum<strong>in</strong>escence spectroscopy. The results strongly suggested<br />

that a larger <strong>in</strong>terdiffusion rate of the Si-<strong>in</strong>duced layer <strong>in</strong>termix<strong>in</strong>g was related to the<br />

formation of a deep level which was associated with an optical emission at 1.3eV.<br />

R.E.Viturro, B.L.Olmsted, S.N.Houde-Walter, G.W.Wicks: Journal of Vacuum Science<br />

<strong>and</strong> Technology B, 1991, 9[4], 2244-50<br />

[446-88/89-008]<br />

259


Interdiffusion (Al,Ga)As Interdiffusion<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Interdiffusion<br />

The effect of pressure <strong>and</strong> stoichiometry upon the Al-Ga <strong>in</strong>terdiffusion of undoped<br />

multiple quantum wells was <strong>in</strong>vestigated over the entire composition range of tile <strong>GaAs</strong><br />

solidus. The occurrence of a two orders of magnitude <strong>in</strong>crease <strong>in</strong> the <strong>in</strong>terdiffusion<br />

coefficient suggested that <strong>in</strong>terdiffusion <strong>in</strong> an <strong>in</strong>tr<strong>in</strong>sic crystal was mediated ma<strong>in</strong>ly by<br />

column-<strong>III</strong> vacancies over the whole solidus range. It was noted that the<br />

photolum<strong>in</strong>escence <strong>in</strong>tensity of the Ga-rich crystal was more than 3 orders of magnitude<br />

greater than that of the As-rich crystal.<br />

B.L.Olmsted, S.N.Houde-Walter: Applied Physics Letters, 1992, 60[3], 368-70<br />

[446-88/89-008]<br />

35 Al<strong>GaAs</strong>/<strong>GaAs</strong>: Interdiffusion<br />

The <strong>in</strong>terdiffusion of Ga <strong>and</strong> Al <strong>in</strong> Al<strong>GaAs</strong> alloys which were subjected to various<br />

anneal<strong>in</strong>g temperatures, times <strong>and</strong> environments was considered. The <strong>in</strong>terdiffusion<br />

coefficients (table 5) <strong>and</strong> activation energies were determ<strong>in</strong>ed by relat<strong>in</strong>g shifts <strong>in</strong> the<br />

photolum<strong>in</strong>escence peaks to calculated transition energies which were based upon an erf<br />

composition profile. It was noted that a Ga over-pressure reduced <strong>in</strong>terdiffusion whereas<br />

an As over-pressure <strong>in</strong>creased <strong>in</strong>terdiffusion. This was thought to be the first study of the<br />

effect of a Ga over-pressure upon the <strong>in</strong>terdiffusion of Al <strong>and</strong> Ga <strong>in</strong> Al<strong>GaAs</strong>.<br />

K.Y.Hsieh, Y.C.Lo, J.H.Lee, R.M.Kolbas: Institute of Physics Conference Series, 1989,<br />

96, 393-6<br />

[446-125/126-120]<br />

Table 5<br />

Interdiffusivity (Al-Ga) <strong>in</strong> Al<strong>GaAs</strong>/<strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

940 9.0 x 10 -18<br />

940 5.9 x 10 -18<br />

920 2.5 x 10 -18<br />

920 1.9 x 10 -18<br />

890 1.2 x 10 -18<br />

845 1.9 x 10 -19<br />

Al<strong>GaAs</strong>/<strong>GaAs</strong>: Interdiffusion<br />

Quantum-well heterostructures were annealed <strong>in</strong> an AsH 3 /H 2 atmosphere, <strong>and</strong> the use of<br />

photolum<strong>in</strong>escence spectroscopy revealed a uniform <strong>and</strong> reproducible <strong>in</strong>crease <strong>in</strong> the<br />

effective quantum-well b<strong>and</strong>-gap. The energy shift data <strong>in</strong>dicated that Al/Ga<br />

<strong>in</strong>terdiffusion occurred under non-equilibrium conditions. The activation energies varied<br />

from about 5.2eV, <strong>in</strong> the equilibrium case, to about 3.4eV <strong>in</strong> the non-equilibrium case.<br />

260


Interdiffusion (Al,Ga)As Interdiffusion<br />

S.Seshadri, L.J.Guido, T.S.Moise, J.C.Beggy, T.J.Cunn<strong>in</strong>gham, R.C.Barker, R.N.Sacks:<br />

Journal of Electronic Materials, 1992, 21[1], 33-8<br />

[446-93/94-004]<br />

GaAlAs/<strong>GaAs</strong>: Interdiffusion<br />

A study was made of Al-Ga <strong>in</strong>terdiffusion <strong>and</strong> C acceptor diffusion <strong>in</strong> C-doped<br />

Ga 0.6 Al 0.4 As/<strong>GaAs</strong> superlattices which had been annealed, under various ambient As 4<br />

pressure conditions, at temperatures rang<strong>in</strong>g from 825 to 960C. The superlattices were<br />

doped with C to an <strong>in</strong>itial acceptor concentration of about 2.9 x 10 19 /cm 3 . The Al-Ga<br />

<strong>in</strong>terdiffusion was found to be most predom<strong>in</strong>ant <strong>in</strong> Ga-rich anneal<strong>in</strong>g ambients. The<br />

<strong>in</strong>terdiffusivity values were about 2 orders of magnitude lower than those predicted by the<br />

Fermi-level effect model. In As-rich ambients, the <strong>in</strong>terdiffusion values were <strong>in</strong><br />

approximate agreement with those which were predicted by the Fermi-level effect model.<br />

By analyz<strong>in</strong>g measured hole concentration profiles, it was deduced that both C acceptor<br />

diffusion <strong>and</strong> reduction occurred dur<strong>in</strong>g anneal<strong>in</strong>g. Both the C acceptor diffusivity <strong>and</strong><br />

the C acceptor reduction coefficient data could be characterized approximately by a ¼-<br />

power dependence upon the As 4 pressure. These pressure dependences <strong>in</strong>dicated that C<br />

diffused via the <strong>in</strong>terstitialcy or <strong>in</strong>terstitial-substitutional mechanism, while hole reduction<br />

was governed by a C acceptor precipitation mechanism.<br />

H.M.You, T.Y.Tan, U.M.Gösele, S.T.Lee, G.E.Höfler, K.C.Hsieh, N.Holonyak: Journal<br />

of Applied Physics, 1993, 74[4], 2450-60<br />

[446-109/110-028]<br />

GaAlAs/<strong>GaAs</strong>: Interdiffusion<br />

Superlattices of Ga 0.7 Al 0.3 As/<strong>GaAs</strong> which had been grown by means of metalorganic<br />

chemical vapor deposition, <strong>and</strong> which were heavily doped with C us<strong>in</strong>g CCl 4 , were<br />

annealed (825C, 24h) under various ambients <strong>and</strong> encapsulants. Photolum<strong>in</strong>escence<br />

monitor<strong>in</strong>g at 1.7K was used to determ<strong>in</strong>e approximate <strong>in</strong>terdiffusion coefficients, D Al-Ga ,<br />

for various anneal<strong>in</strong>g conditions. For all of the encapsulants which were studied, D Al-Ga<br />

<strong>in</strong>creased with <strong>in</strong>creas<strong>in</strong>g As 4 pressure <strong>in</strong> the anneal<strong>in</strong>g ampoule. This result disagreed<br />

with trends which had been reported for Mg-doped crystals, <strong>and</strong> with the predictions of<br />

the charged po<strong>in</strong>t-defect (Fermi-level) model. It was noted that a Si 3 N 4 cap provided the<br />

most effective surface protection aga<strong>in</strong>st ambient-stimulated layer <strong>in</strong>terdiffusion (D Al-Ga =<br />

1.5 x 10 -19 to 3.9 x 10 -19 cm 2 /s). The most extensive layer <strong>in</strong>termix<strong>in</strong>g occurred for<br />

uncapped superlattices which were annealed <strong>in</strong> an As-rich ambient (D Al-Ga equal to about<br />

3.3 x 10 -18 cm 2 /s). These values were up to 40 times greater than those which had<br />

previously been reported for nom<strong>in</strong>ally undoped Al<strong>GaAs</strong>/<strong>GaAs</strong> superlattices. This<br />

implied that dop<strong>in</strong>g slightly enhanced layer <strong>in</strong>termix<strong>in</strong>g; but this was significantly less<br />

than that predicted by the Fermi-level effect. The discrepancies between the experimental<br />

data <strong>and</strong> the model were considered. Marked changes <strong>in</strong> the optical properties of the<br />

annealed superlattices, as a function of storage time at room temperature, were also<br />

reported. It was suggested that these changes might reflect a degraded thermal stability of<br />

the annealed crystals, due to lattice defects which were generated at high temperatures. It<br />

261


Interdiffusion (Al,Ga)As Interdiffusion<br />

was proposed that this was related to the failure to prepare buried heterostructure<br />

quantum-well lasers, via impurity-<strong>in</strong>duced layer disorder<strong>in</strong>g, <strong>in</strong> similar doped crystals.<br />

I.Szafranek, M.Szafranek, J.S.Major, B.T.Cunn<strong>in</strong>gham, L.J.Guido, N.Holonyak,<br />

G.E.Stillman: Journal of Electronic Materials, 1991, 20[6], 409-18<br />

[446-91/92-005]<br />

36 GaAlAs/<strong>GaAs</strong>: Interdiffusion<br />

The migration of Al <strong>and</strong> Ga <strong>in</strong> Ga 0.6 Al 0.4 As/<strong>GaAs</strong> quantum wells was <strong>in</strong>vestigated by<br />

measur<strong>in</strong>g the photolum<strong>in</strong>escence of samples which had been annealed at temperatures<br />

rang<strong>in</strong>g from 850 to 1065C; with <strong>and</strong> without a SiO 2 cap. At 1000C, under a SiO 2 cap,<br />

the Al-Ga <strong>in</strong>terdiffusion coefficient was at least 2 orders of magnitude higher for a<br />

GaAlAs/<strong>GaAs</strong> quantum well than for an InAlGaP/GaInP quantum well, with<strong>in</strong> the same<br />

sample. By compar<strong>in</strong>g the calculated photolum<strong>in</strong>escence shifts with measured values, an<br />

activation energy of 4.5eV was estimated (table 6) for Al-Ga <strong>in</strong>terdiffusion <strong>in</strong> a<br />

GaAlAs/<strong>GaAs</strong> quantum well under a SiO 2 cap.<br />

K.J.Beern<strong>in</strong>k, D.Sun, D.W.Treat, B.P.Bour: Applied Physics Letters, 1995, 66[26], 3597-<br />

9<br />

[446-125/126-120]<br />

Table 6<br />

Interdiffusivity (Al-Ga) <strong>in</strong> GaAlAs/<strong>GaAs</strong><br />

Conditions Temperature (C) D (cm 2 /s)<br />

RTA/SiO 2 1065 2.5 x 10 -15<br />

RTA/SiO 2 1025 1.8 x 10 -15<br />

RTA/SiO 2 1000 4.3 x 10 -16<br />

no SiO 2 1000 4.1 x 10 -17<br />

RTA/SiO 2 975 1.8 x 10 -16<br />

FA/SiO 2 925 3.0 x 10 -17<br />

FA/SiO 2 900 2.3 x 10 -17<br />

FA/SiO 2 870 4.7 x 10 -18<br />

FA/SiO 2 845 1.6 x 10 -18<br />

262


(Al,Ga,In)P<br />

Si<br />

InAlGaP/<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The diffusion of Si <strong>III</strong> -Si V neutral pairs versus the diffusion of Si <strong>III</strong> -V <strong>III</strong> complexes <strong>in</strong> <strong>III</strong>-V<br />

crystals was considered with regard to experimental data which revealed the effect of Si<br />

diffusion upon the self-diffusion of column-<strong>III</strong> <strong>and</strong> column-V lattice atoms. Secondary<br />

ion mass spectroscopy was used to compare the enhanced diffusion of column-<strong>III</strong> or<br />

column-V atoms <strong>in</strong> various Si-diffused heterostructures which were closely latticematched<br />

to <strong>GaAs</strong>. An enhancement of lattice atom self-diffusion, due to impurity<br />

diffusion, was found to occur predom<strong>in</strong>antly on the column-<strong>III</strong> lattice. The data supported<br />

the Si <strong>III</strong> -V <strong>III</strong> diffusion model <strong>and</strong> <strong>in</strong>dicated that the ma<strong>in</strong> native defects which<br />

accompanied Si diffusion were column-<strong>III</strong> vacancies. These diffused directly on the<br />

column-<strong>III</strong> sub-lattice.<br />

D.G.Deppe, W.E.Plano, J.E.Baker, N.Holonyak, M.J.Ludowise, C.P.Kuo, R.M.Fletcher,<br />

T.D.Osentowski, M.G.Craford: Applied Physics Letters, 1988, 53[22], 2211-3<br />

[446-64/65-157]<br />

Zn<br />

AlGaInP: Zn <strong>Diffusion</strong><br />

The behavior of Zn acceptors <strong>in</strong> p-type layers dur<strong>in</strong>g thermal anneal<strong>in</strong>g <strong>in</strong> gaseous AsH 3 -<br />

H 2 mixtures was <strong>in</strong>vestigated. It was found that the electrical activity of Zn acceptors was<br />

greatly affected by H atoms which orig<strong>in</strong>ated from the ars<strong>in</strong>e dur<strong>in</strong>g anneal<strong>in</strong>g. In<br />

addition, observations of atomic disorder<strong>in</strong>g suggested that H passivation of Zn acceptors<br />

suppressed atomic diffusion. The results could be consistently expla<strong>in</strong>ed by a simple H<br />

passivation model which <strong>in</strong>volved the term<strong>in</strong>ation of dangl<strong>in</strong>g bonds by H atoms.<br />

A.Ishibashi, M.Mannoh, I.Kidoguchi, Y.Ban, K.Ohnaka: Applied Physics Letters, 1994,<br />

65[10], 1275-7<br />

[446-119/120-188]<br />

263


Interdiffusion (Al,Ga,In)P|(Al,Ga)Sb Be<br />

Interdiffusion<br />

InAlGaP/GaInP: Interdiffusion<br />

The migration of Al <strong>and</strong> Ga <strong>in</strong> In 0.5 Al 0.3 Ga 0.2 P/Ga 0.6 In 0.4 P quantum wells was<br />

<strong>in</strong>vestigated by measur<strong>in</strong>g the photolum<strong>in</strong>escence of samples which had been annealed at<br />

temperatures rang<strong>in</strong>g from 850 to 1065C; with <strong>and</strong> without a SiO 2 cap. At 1000C, under a<br />

SiO 2 cap, the Al-Ga <strong>in</strong>terdiffusion coefficient was at least 2 orders of magnitude higher<br />

for a GaAlAs/<strong>GaAs</strong> quantum well, than for an InAlGaP/GaInP quantum well, with<strong>in</strong> the<br />

same sample.<br />

K.J.Beern<strong>in</strong>k, D.Sun, D.W.Treat, B.P.Bour: Applied Physics Letters, 1995, 66[26], 3597-<br />

9<br />

[446-125/126-120]<br />

(Al,Ga)Sb<br />

Be<br />

AlGaSb: Be <strong>Diffusion</strong><br />

The ion-implantation p-type dop<strong>in</strong>g of Al 0.75 Ga 0.25 Sb was studied. The surface<br />

morphology <strong>and</strong> electrical properties were shown, by us<strong>in</strong>g atomic force microscopy <strong>and</strong><br />

Hall measurements, to be degraded after rapid thermal anneal<strong>in</strong>g at 650C. The<br />

implantation of Be resulted <strong>in</strong> sheet hole concentrations which were twice those of the<br />

implanted acceptor dose (10 13 /cm 2 ) after 600C anneal<strong>in</strong>g. This was attributed to double<br />

acceptor or antisite defect formation. Implanted C acted as an acceptor, but also<br />

demonstrated an excess hole conduction which was attributed to implantation-<strong>in</strong>duced<br />

defects. Implanted Zn required a higher anneal<strong>in</strong>g temperature than did Be, <strong>in</strong> order to<br />

achieve 100% effective activation for a dose of 10 13 /cm 2 . It was suggested that this was<br />

probably the result of the greater implantation-<strong>in</strong>duced damage that was created by the<br />

264


Be (Al,Ga)Sb Mg<br />

heavier Zn ion. The secondary ion mass spectroscopy of as-implanted <strong>and</strong> annealed Be,<br />

Mg <strong>and</strong> C samples was studied. The diffusion of implanted Be (5 x 10 13 cm 2 , 45keV) was<br />

shown to exhibit an <strong>in</strong>verse dependence upon temperature. This was attributed to a<br />

substitutional-<strong>in</strong>terstitial diffusion mechanism. Implanted C (2.5 x 10 14 /cm 2 , 70keV)<br />

exhibited no redistribution, even after anneal<strong>in</strong>g at 650C.<br />

J.C.Zolper, J.F.Klem, A.J.Howard, M.J.Hafich: Journal of Applied Physics, 1996, 79[3],<br />

1365-70<br />

[446-131/132-168]<br />

Mg<br />

AlGaSb: Mg <strong>Diffusion</strong><br />

The ion-implantation p-type dop<strong>in</strong>g of Al 0.75 Ga 0.25 Sb was studied. The surface<br />

morphology <strong>and</strong> electrical properties were shown, by us<strong>in</strong>g atomic force microscopy <strong>and</strong><br />

Hall measurements, to be degraded after rapid thermal anneal<strong>in</strong>g at 650C. The<br />

implantation of Mg resulted <strong>in</strong> sheet hole concentrations which were twice those of the<br />

implanted acceptor dose (10 13 /cm 2 ) after 600C anneal<strong>in</strong>g. This was attributed to double<br />

acceptor or antisite defect formation. Implanted C acted as an acceptor, but also<br />

demonstrated an excess hole conduction which was attributed to implantation-<strong>in</strong>duced<br />

defects. Implanted Zn required a higher anneal<strong>in</strong>g temperature than did Mg, <strong>in</strong> order to<br />

achieve 100% effective activation for a dose of 10 13 /cm 2 . It was suggested that this was<br />

probably the result of the greater implantation-<strong>in</strong>duced damage that was created by the<br />

heavier Zn ion. The secondary ion mass spectroscopy of as-implanted <strong>and</strong> annealed Mg<br />

<strong>and</strong> C samples was studied. Implanted Mg (10 14 /cm 2 , 110keV) exhibited marked<br />

redistribution, <strong>and</strong> losses of up to 56% at the <strong>in</strong>terface, after 600C anneal<strong>in</strong>g. Implanted C<br />

(2.5 x 10 14 /cm 2 , 70keV) exhibited no redistribution, even after anneal<strong>in</strong>g at 650C.<br />

J.C.Zolper, J.F.Klem, A.J.Howard, M.J.Hafich: Journal of Applied Physics, 1996, 79[3],<br />

1365-70<br />

[446-131/132-168]<br />

265


(Al,In)As<br />

Be<br />

InAlAs: Be <strong>Diffusion</strong><br />

The behavior of implanted Be + ions was <strong>in</strong>vestigated dur<strong>in</strong>g rapid thermal anneal<strong>in</strong>g at<br />

temperatures of between 600 <strong>and</strong> 900C. It was found that the apparent activation energy<br />

for Be was equal to 0.43eV. Lower activation efficiencies of the dopant were found <strong>in</strong><br />

InAlAs, as compared with In<strong>GaAs</strong>. Anomalously low activation was detected for lowdose<br />

Be implants. The latter effect was attributed to a lack of vacant sites for the Be<br />

atoms to occupy. Extensive redistribution of the Be was observed after anneal<strong>in</strong>g (750C,<br />

10s).<br />

E.Hailemariam, S.J.<strong>Pearton</strong>, W.S.Hobson, H.S.Luftman, A.P.Perley: Journal of Applied<br />

Physics, 1992, 71[1], 215-20<br />

[446-86/87-038]<br />

Fe<br />

InAlAs: Fe <strong>Diffusion</strong><br />

S<strong>in</strong>gle (200keV) <strong>and</strong> multiple-energy Fe implantation of n-type material was carried out<br />

on In 0.52 Al 0.48 As at room temperature or 200C. Secondary ion mass spectrometry profiles<br />

revealed marked out-diffusion dur<strong>in</strong>g all of the rapid thermal anneal<strong>in</strong>g treatments which<br />

were used; regardless of the implantation temperature. The Fe implantation peaks which<br />

were observed after the anneal<strong>in</strong>g of <strong>other</strong> In-based compounds were not observed here.<br />

J.M.Mart<strong>in</strong>, R.K.Nadella, M.V.Rao, D.S.Simons, P.H.Chi, C.Caneau: Journal of<br />

Electronic Materials, 1993, 22[9], 1153-8<br />

[446-109/110-039]<br />

Si<br />

InAlAs: Si <strong>Diffusion</strong><br />

The behavior of implanted Si + ions was <strong>in</strong>vestigated dur<strong>in</strong>g rapid thermal anneal<strong>in</strong>g at<br />

temperatures of between 600 <strong>and</strong> 900C. The apparent activation energy for Si was equal<br />

266


Si (Al,In)As General<br />

to 0.58eV. Lower activation efficiencies of the dopant were found <strong>in</strong> InAlAs, as<br />

compared with In<strong>GaAs</strong>. The Si underwent no migration, even after anneal<strong>in</strong>g at 850C.<br />

E.Hailemariam, S.J.<strong>Pearton</strong>, W.S.Hobson, H.S.Luftman, A.P.Perley: Journal of Applied<br />

Physics, 1992, 71[1], 215-20<br />

[446-86/87-038]<br />

Ti<br />

InAlAs: Ti <strong>Diffusion</strong><br />

The Ti implantation of p-type material was carried out on In 0.52 Al 0.48 As at room<br />

temperature or 200C. The implanted Ti exhibited only slight <strong>in</strong>-diffusion <strong>and</strong> outdiffusion<br />

after both room temperature <strong>and</strong> 200C implantation. Rutherford back-scatter<strong>in</strong>g<br />

measurements of annealed samples which had been implanted at 200C revealed a crystal<br />

quality which was similar to that of virg<strong>in</strong> material. The resistivity of all of the samples<br />

after anneal<strong>in</strong>g was greater than 10 6 Ωcm.<br />

J.M.Mart<strong>in</strong>, R.K.Nadella, M.V.Rao, D.S.Simons, P.H.Chi, C.Caneau: Journal of<br />

Electronic Materials, 1993, 22[9], 1153-8<br />

[446-109/110-039]<br />

General<br />

InAlAs: <strong>Diffusion</strong><br />

It was shown that systematic variations <strong>in</strong> the experimental parameters could turn multilayers<br />

<strong>in</strong>to so-called microscopic laboratories for the study of po<strong>in</strong>t defects. In this way,<br />

the effects of composition, dop<strong>in</strong>g <strong>and</strong> stra<strong>in</strong> could be separated. It was also possible to<br />

determ<strong>in</strong>e the nature <strong>and</strong> charge state of the mediat<strong>in</strong>g defect. The present results showed<br />

that <strong>in</strong>terdiffusion <strong>in</strong> the present system was mediated by a double-acceptor vacancy-like<br />

defect. The activation energy, <strong>in</strong> the case of In 0.52 Al 0.48 As, was equal to 4eV. Its value<br />

varied to the extent of 0.051eV for every percent of stra<strong>in</strong>.<br />

F.H.Baumann, J.H.Huang, J.A.Rentschler, T.Y.Chang, A.Ourmazd: Physical Review<br />

Letters, 1994, 73[3], 448-51<br />

[446-115/116-141]<br />

267


AlN<br />

1.0E-07<br />

1.0E-08<br />

1.0E-09<br />

D (cm 2 /s)<br />

1.0E-10<br />

1.0E-11<br />

1.0E-12<br />

1.0E-13<br />

table 7<br />

table 8<br />

table 9<br />

1.0E-14<br />

1.0E-15<br />

1.0E-16<br />

4 5 6<br />

10 4 /T(K)<br />

Figure 2: Diffusivity of O <strong>in</strong> AlN<br />

D<br />

AlN: D <strong>Diffusion</strong><br />

The out-diffusion of H was studied, us<strong>in</strong>g 2 H plasma-treated (250 or 400C, 0.5h) or 2 H + -<br />

implanted samples, dur<strong>in</strong>g anneal<strong>in</strong>g at temperatures rang<strong>in</strong>g from 300 to 900C.<br />

268


D AlN O<br />

Secondary ion mass spectrometry was used to measure the resultant distributions. At<br />

concentrations that were greater than 10 20 /cm 3 , there was a near-surface (less than 0.3µ)<br />

region that was probably due to the formation of platelet defects. At concentrations of<br />

about 10 18 /cm 3 , a plateau region was present which extended throughout the film<br />

thickness of about 1µ. This was attributed to the pair<strong>in</strong>g of 2 H with po<strong>in</strong>t defects. In<br />

implanted samples, 2 H redistribution occurred <strong>in</strong> the same manner as the bulk population<br />

<strong>in</strong> plasma-treated material. The thermal stability of the D profile <strong>in</strong> the nitride was much<br />

higher than that <strong>in</strong> <strong>GaAs</strong> <strong>and</strong> similar compounds.<br />

R.G.Wilson, S.J.<strong>Pearton</strong>, C.R.Abernathy, J.M.Zavada: Journal of Vacuum Science <strong>and</strong><br />

Technology A, 1995, 13[3], 719-23<br />

[446-140-029]<br />

Table 7<br />

Diffusivity of O <strong>in</strong> AlN<br />

Temperature (C) D (cm 2 /s)<br />

1500 8.0 x 10 -16<br />

1600 3.5 x 10 -15<br />

1700 6.7 x 10 -15<br />

1800 1.3 x 10 -14<br />

1900 1.9 x 10 -14<br />

N<br />

AlN: N Permeation<br />

A method was described for the determ<strong>in</strong>ation of ion migration numbers on the basis of<br />

permeability data. The latter data were here obta<strong>in</strong>ed by us<strong>in</strong>g an emf technique. The<br />

permeability was equal to 8.66cm 2 /MPa-s. It was po<strong>in</strong>ted out that the present method<br />

could establish the presence of an ionic component of the conductivity for nitrides, <strong>and</strong> its<br />

importance could be estimated for cases where a dense ceramic could not be prepared.<br />

R.P.Lesunova, L.S.Karen<strong>in</strong>a, V.K.Gilderman, S.F.Palguev: Izvestiya Akademii Nauk<br />

SSSR - Neorganicheskie Materialy, 1989, 25[11], 1926-8. (Inorganic Materials, 1989,<br />

25[11], 1633-5)<br />

[446-76/77-131]<br />

O<br />

37 AlN: O <strong>Diffusion</strong><br />

Interdiffusion of N <strong>and</strong> O was <strong>in</strong>vestigated by means of electron energy-loss spectroscopy<br />

<strong>and</strong> transmission electron microscopy. <strong>Diffusion</strong> couples, Al 2 O 3 /AlN, were prepared by<br />

oxidiz<strong>in</strong>g AlN ceramics, <strong>and</strong> were annealed at temperatures rang<strong>in</strong>g from 1500 to 1900C<br />

269


O AlN O<br />

(table 7). The couples were encapsulated <strong>in</strong> a Ta ampoule <strong>in</strong> order to ensure an <strong>in</strong>ert<br />

atmosphere, <strong>and</strong> the O concentration profiles across the oxide/nitride <strong>in</strong>terface were<br />

measured by means of electron energy-loss spectroscopy. It was found that O/N<br />

<strong>in</strong>terdiffusion <strong>in</strong> AlN could be described by:<br />

D(cm 2 /s) = 1.5 x 10 -8 exp[-240(kJ/mol)/RT]<br />

The magnitude <strong>and</strong> temperature-dependence of the <strong>in</strong>terdiffusion were comparable to<br />

those which had been reported for <strong>other</strong> non-oxide ceramic materials. Under normal AlN<br />

s<strong>in</strong>ter<strong>in</strong>g conditions, the O/N <strong>in</strong>terdiffusion was too slow to provide an effective means<br />

for O removal from AlN gra<strong>in</strong>s.<br />

M.Sternitzke, G.Müller: Journal of the American Ceramic Society, 1994, 77[3], 737-42<br />

[446-123/124-267]<br />

Table 8<br />

Bulk Diffusivity of O <strong>in</strong> AlN<br />

Temperature (C) D (cm 2 /s)<br />

1600 3.43 x 10 -14<br />

1700 1.00 x 10 -13<br />

1800 4.27 x 10 -13<br />

1900 1.43 x 10 -12<br />

38,9 AlN: O Gra<strong>in</strong> Boundary <strong>Diffusion</strong><br />

The diffusion of O <strong>in</strong> commercial nitride substrates was studied by carry<strong>in</strong>g out<br />

<strong>in</strong>terdiffusion-type experiments. The diffusion of O with<strong>in</strong> AlN gra<strong>in</strong>s <strong>and</strong> along gra<strong>in</strong><br />

boundaries was <strong>in</strong>vestigated. By us<strong>in</strong>g as-received AlN substrates <strong>and</strong> an electron<br />

microprobe as an analytical tool, it was found that the rate of O diffusion along gra<strong>in</strong><br />

boundaries was strongly affected by the presence of impurities <strong>and</strong>/or <strong>other</strong> phases at<br />

these boundaries. The diffusion of O with<strong>in</strong> AlN gra<strong>in</strong>s was studied at temperatures of<br />

between 1600 <strong>and</strong> 1900C (table 8) <strong>in</strong> a flow<strong>in</strong>g N atmosphere by us<strong>in</strong>g secondary ion<br />

mass spectrometry to determ<strong>in</strong>e O concentration profiles. The chemical diffusion<br />

coefficient of O <strong>in</strong> the AlN lattice was described by:<br />

log[D(cm 2 /s)] = -1.68 - 427(kJ/mol)/2.303RT<br />

The O concentration profiles which were determ<strong>in</strong>ed by means of secondary ion mass<br />

spectrometry also revealed a contribution that arose from diffusion along gra<strong>in</strong> boundaries<br />

(table 9). It was therefore possible to determ<strong>in</strong>e values of the product of gra<strong>in</strong> boundary<br />

width <strong>and</strong> gra<strong>in</strong> boundary O diffusivity.<br />

H.Solmon, D.Rob<strong>in</strong>son, R.Dieckmann: Journal of the American Ceramic Society, 1994,<br />

77[11], 2841-8<br />

[446-123/124-268]<br />

270


Interdiffusion AlN Interdiffusion<br />

Interdiffusion<br />

AlN/Al 2 OC: Interdiffusion<br />

It was recalled that AlN <strong>and</strong> Al 2 OC were isostructural. The AlN-Al 2 OC system formed<br />

homogeneous solid solutions above 1950C. Interdiffusion was studied <strong>in</strong> the solidsolution<br />

regime <strong>in</strong> order to clarify differences <strong>in</strong> the k<strong>in</strong>etics of phase separation when<br />

samples were annealed at lower temperatures. The diffusion couples, (AlN) 0.7 (Al 2 OC) 0.3<br />

/(AlN) 0.3 (Al 2 OC) 0.7 , were prepared by hot press<strong>in</strong>g <strong>and</strong> were annealed at 2273K. It was<br />

found that the <strong>in</strong>terdiffusion coefficients <strong>in</strong> this system were much larger than those <strong>in</strong> the<br />

SiC-AlN system.<br />

Q.Tian, A.V.Virkar: Journal of the American Ceramic Society, 1996, 79[8], 2168-74<br />

[446-150/151-214]<br />

Table 9<br />

Gra<strong>in</strong> Boundary Diffusivity of O <strong>in</strong> AlN<br />

Temperature (C)<br />

D (cm 2 /s)<br />

1700 9.50 x 10 -10<br />

1800 1.59 x 10 -9<br />

1900 1.78 x 10 -8<br />

AlN/SiC: Interdiffusion<br />

It was recalled that AlN <strong>and</strong> 2H-type SiC were isostructural. The SiC-AlN system formed<br />

homogeneous solid solutions above 2000C. Interdiffusion was studied <strong>in</strong> the solidsolution<br />

regime <strong>in</strong> order to clarify differences <strong>in</strong> the k<strong>in</strong>etics of phase separation dur<strong>in</strong>g<br />

anneal<strong>in</strong>g at lower temperatures. <strong>Diffusion</strong> couples, (SiC) 0.3 (AlN) 0.7 /(SiC) 0.7 (AlN) 0.3 ,<br />

were prepared by hot press<strong>in</strong>g. Interdiffusion coefficients were measured at 2373, 2473<br />

<strong>and</strong> 2573K, <strong>and</strong> the correspond<strong>in</strong>g activation energy was estimated to be 632kJ/mol.<br />

Q.Tian, A.V.Virkar: Journal of the American Ceramic Society, 1996, 79[8], 2168-74<br />

[446-150/151-214]<br />

271


BN<br />

D<br />

BN: D <strong>Diffusion</strong><br />

Re-emission curves <strong>and</strong> thermodesorption spectra were measured for D ions which had<br />

been implanted <strong>in</strong>to this material. The thermodesorption spectra consisted of several<br />

peaks at temperatures rang<strong>in</strong>g from 400 to 1100K. The re-emission curves could be<br />

described by a simple mathematical model which <strong>in</strong>cluded the effects of diffusion,<br />

second-order thermodesorption <strong>and</strong> defect trapp<strong>in</strong>g. The recycl<strong>in</strong>g factor <strong>and</strong> defecttrapp<strong>in</strong>g<br />

factor were found to depend exponentially upon the temperature, at temperatures<br />

above 600K. They deviated from this behavior at room temperature. It was supposed that<br />

radiation-enhanced <strong>and</strong> thermally-activated processes predom<strong>in</strong>ated at room <strong>and</strong> high<br />

temperatures, respectively.<br />

A.A.Pisarev, V.M.Smirnov, S.K.Zhdanov, A.V.Varava, V.V.B<strong>and</strong>urko: Journal of<br />

Nuclear Materials, 1992, 187[3], 254-9<br />

[446-91/92-075]<br />

272


<strong>GaAs</strong><br />

1.0E-07<br />

1.0E-08<br />

1.0E-09<br />

1.0E-10<br />

D (cm 2 /s)<br />

1.0E-11<br />

1.0E-12<br />

1.0E-13<br />

1.0E-14<br />

1.0E-15<br />

1.0E-16<br />

1.0E-17<br />

1.0E-18<br />

1.0E-19<br />

1.0E-20<br />

Be (table 10)<br />

Cr (table 11)<br />

Fe (table 12)<br />

Ga (tables 13 to 15)<br />

Ge (table 16)<br />

H (tables 17 <strong>and</strong> 18)<br />

S (table 19)<br />

Si (tables 21 to 25)<br />

Sn (table 26)<br />

Zn (tables 27 to 31)<br />

6 7 8 9 10 11 12 13 14 15 16 17 18 19 20<br />

10 4 /T(K)<br />

Figure 3: Summary of Diffusivities of Various Elements <strong>in</strong> <strong>GaAs</strong><br />

Al<br />

<strong>GaAs</strong>: Al <strong>Diffusion</strong><br />

An <strong>in</strong>fra-red absorption spectroscopic <strong>in</strong>vestigation was made of the thermal diffusion of<br />

Al from monatomic Al layers which were embedded <strong>in</strong> a <strong>GaAs</strong> epitaxial film. After<br />

273


Al <strong>GaAs</strong> Al<br />

anneal<strong>in</strong>g, the absorption peak of the 2-dimensionally localized vibrational modes at<br />

358/cm (due to Al layers) decreased while the peak at 362/cm (due to isolated Al atoms)<br />

<strong>in</strong>creased. The 362/cm peak height was compared with the fraction of isolated Al atoms,<br />

as calculated by assum<strong>in</strong>g second-nearest neighbor hopp<strong>in</strong>g diffusion from a monatomic<br />

Al layer <strong>and</strong> <strong>in</strong>to the <strong>GaAs</strong> matrix. It was thus deduced that the diffusion coefficient of Al<br />

<strong>in</strong> <strong>GaAs</strong> was equal to 2 x 10 -19 cm 2 /s at 700C. It was concluded that this was a simple <strong>and</strong><br />

reliable method for the <strong>in</strong>vestigation of impurity diffusion <strong>in</strong> crystals.<br />

H.Ono. N.Ikarashi, T.Baba: Applied Physics Letters, 1995, 66[5], 601-3<br />

[446-121/122-053]<br />

<strong>GaAs</strong>/Al/<strong>GaAs</strong>: Al <strong>Diffusion</strong><br />

Theoretical <strong>and</strong> experimental aspects of the growth of heterostructures were <strong>in</strong>vestigated.<br />

In these heterostructures, <strong>GaAs</strong> was grown on top of the buried metal layer via migrationenhanced<br />

epitaxy at low temperatures (200 or 400C) <strong>in</strong> order to provide a k<strong>in</strong>etic barrier<br />

to the out-diffusion of Al dur<strong>in</strong>g super-layer growth. The crystall<strong>in</strong>ity <strong>and</strong> orientation of<br />

the Al film which was deposited on (100)<strong>GaAs</strong> at about 0C was studied by us<strong>in</strong>g<br />

transmission electron diffraction, dark-field imag<strong>in</strong>g, <strong>and</strong> X-ray diffraction methods. It<br />

was found that the Al was polycrystall<strong>in</strong>e, with a gra<strong>in</strong> size of about 6nm, <strong>and</strong> that the<br />

preferred growth orientation was (111). This could be textured <strong>in</strong> the plane, but oriented<br />

out of the plane. The quality of the <strong>GaAs</strong> super-layer, which was grown on top of the Al<br />

by means of migration-enhanced epitaxy, was very sensitive to the growth temperature. A<br />

layer which was grown at 400C had a good structural <strong>and</strong> optical quality, but was<br />

associated with considerable Al out-diffusion at the Al/<strong>GaAs</strong> hetero-<strong>in</strong>terface. At 200C,<br />

where the <strong>in</strong>terface had good structural <strong>in</strong>tegrity, the super-layer exhibited tw<strong>in</strong>n<strong>in</strong>g <strong>and</strong><br />

no lum<strong>in</strong>escence was observed.<br />

P.Bhattacharya, J.E.Oh, J.S<strong>in</strong>gh, D.Biswas, R.Clarke, W.Dos Passos, R.Merl<strong>in</strong>,<br />

N.Mestres, K.H.Chang, R.Gibala: Journal of Applied Physics, 1990, 67[8], 3700-5<br />

[446-78/79-030]<br />

<strong>GaAs</strong>/AlAs: Al <strong>Diffusion</strong><br />

The implantation of Be ions <strong>in</strong>to heterostructures at room temperature or liquid N<br />

temperatures was <strong>in</strong>vestigated. It was found that room-temperature implantation created<br />

dislocation loops at the first <strong>in</strong>terface; a distance which was far short of the maximum<br />

projected range. Implantation at low temperatures caused tw<strong>in</strong>n<strong>in</strong>g. The latter could be<br />

removed by anneal<strong>in</strong>g (900C, 1200s), without lead<strong>in</strong>g to the <strong>in</strong>terdiffusion of Al. The<br />

presence of dislocation networks tended to enhance <strong>in</strong>termix<strong>in</strong>g. The Be concentrations<br />

were sufficiently low to prevent Be-<strong>in</strong>duced <strong>in</strong>termix<strong>in</strong>g.<br />

S.Mitra: Semiconductor Science <strong>and</strong> Technology, 1990, 5[11], 1138-40<br />

[446-76/77-017]<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Al <strong>Diffusion</strong><br />

The effect of room-temperature electron irradiation upon <strong>in</strong>terdiffusion at quantum-well<br />

<strong>in</strong>terfaces was <strong>in</strong>vestigated by us<strong>in</strong>g low-temperature cathodolum<strong>in</strong>escence spectroscopy.<br />

It was found that <strong>in</strong>terdiffusion was enhanced by the presence of defects which were<br />

274


Al <strong>GaAs</strong> As<br />

generated by irradiation with a 400keV electron beam. After irradiation at room<br />

temperature to doses of between about 1.5 x 10 17 <strong>and</strong> 2.5 x 10 17 /cm 2 , followed by rapid<br />

thermal anneal<strong>in</strong>g (900C, 60s), an <strong>in</strong>terdiffusion length of 0.3 to 0.5nm was found. The<br />

resultant damage tended to saturate with <strong>in</strong>creas<strong>in</strong>g irradiation dose. The formation of<br />

defect clusters at high doses limited the degree of defect <strong>in</strong>troduction, <strong>and</strong> therefore the<br />

extent of <strong>in</strong>terdiffusion at the <strong>in</strong>terface.<br />

Y.J.Li, M.Tsuchiya, P.M.Petroff: Applied Physics Letters, 1990, 57[5], 472-4<br />

[446-76/77-018]<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Al <strong>Diffusion</strong><br />

An <strong>in</strong>vestigation was made of the proton-implantation enhanced <strong>in</strong>termix<strong>in</strong>g of quantum<br />

wells, for H + doses which ranged from 5 x 10 13 to 10 16 /cm 2 . The implantation of 20keV<br />

H + , followed by high-temperature rapid thermal anneal<strong>in</strong>g, led to the enhanced diffusion<br />

of Al <strong>in</strong>to the <strong>GaAs</strong> quantum well. Shifts <strong>in</strong> the electron heavy-hole recomb<strong>in</strong>ation<br />

energies, due to compositional changes, were observed by us<strong>in</strong>g room-temperature<br />

cathodolum<strong>in</strong>escence methods. <strong>Diffusion</strong> lengths of more than 2nm were deduced from<br />

the energy shifts <strong>in</strong> a 5nm well, <strong>and</strong> were found to depend upon the implanted dose <strong>and</strong><br />

the anneal<strong>in</strong>g time. It was suggested that this was to be expected if the enhanced<br />

<strong>in</strong>terdiffusion was caused by defects which were <strong>in</strong>troduced by implantation.<br />

G.F.Red<strong>in</strong>bo, H.G.Craighead, J.M.Hong: Journal of Applied Physics, 1993, 74[5], 3099-<br />

102<br />

[446-106/107-079]<br />

As<br />

<strong>GaAs</strong>: As <strong>Diffusion</strong><br />

Atomistic thermodynamic calculations were made of the energetics of self-diffusion. An<br />

assessment of the activation enthalpy of the saddle-po<strong>in</strong>t configuration of various modes<br />

of vacancy self-diffusion <strong>in</strong>dicated that second-nearest neighbor hopp<strong>in</strong>g was the most<br />

energetically favorable mechanism, if vacancies were available <strong>in</strong> equilibrium<br />

concentrations. An assessment of the activation entropy <strong>in</strong>dicated that normal diffusion<br />

pre-exponential factors, of the order of 10 -5 to 0.1cm 2 /s, were consistent with vacancy<br />

self-diffusion via second-nearest neighbor hopp<strong>in</strong>g. It was proposed that self-diffusion<br />

which was characterized by pre-exponential factors of the order of 10 7 to 10 8 cm 2 /s, <strong>and</strong><br />

activation energies of the order of 6eV, <strong>in</strong>volved processes <strong>in</strong> which surface vacancy<br />

generation was <strong>in</strong>hibited <strong>and</strong> self-diffusion was mediated by Frenkel pair generation.<br />

J.F.Wager: Journal of Applied Physics, 1991, 69[5], 3022-31<br />

[446-78/79-011]<br />

<strong>GaAs</strong>: As <strong>Diffusion</strong><br />

The growth behavior <strong>and</strong> mechanisms of epitaxial lateral overgrowth of <strong>GaAs</strong> on (001)<br />

<strong>GaAs</strong> substrates were <strong>in</strong>vestigated. It was found that the lateral growth exhibited a strong<br />

dependence upon the orientation of the seed. Lateral growth was slowest when the aligned<br />

275


As <strong>GaAs</strong> As<br />

seed was oriented <strong>in</strong> the [100], [110], [010], [¯110], or equivalent, directions. However, it<br />

<strong>in</strong>creased sharply when the seed was tilted away from these orientations. Monomolecular<br />

growth steps on the surface, with a strong contrast <strong>and</strong> a high lateral spatial resolution,<br />

were successfully observed by us<strong>in</strong>g Nomarski differential <strong>in</strong>terference contrast<br />

microscopy plus image process<strong>in</strong>g. Their average propagation velocity was estimated to<br />

be between 0.003 <strong>and</strong> 0.03mm/s. It was found that the surface of the epitaxial lateral<br />

overgrowth layer was extremely flat at the atomic scale. The experimental results<br />

<strong>in</strong>dicated that vertical growth was governed ma<strong>in</strong>ly by the propagation of steps that<br />

orig<strong>in</strong>ated from their source. On the <strong>other</strong> h<strong>and</strong>, lateral growth was limited by the<br />

diffusion of As <strong>in</strong> Ga solution until (111)- or (001)-type facets appeared at the lateral<br />

growth front. The growth then became limited by k<strong>in</strong>etic processes. It was concluded that<br />

substrates of (001)<strong>GaAs</strong>, as well as (111)<strong>GaAs</strong>, were suitable for obta<strong>in</strong><strong>in</strong>g a very smooth<br />

epitaxial lateral overgrowth layer with a large ratio of the lateral growth width to the<br />

vertical growth thickness.<br />

S.Zhang, T.Nish<strong>in</strong>aga: Journal of Crystal Growth, 1990, 99, 292-6<br />

[446-76/77-007]<br />

<strong>GaAs</strong>: As <strong>Diffusion</strong><br />

The chemical reactions <strong>and</strong> Schottky-barrier characteristics of W(200nm)/Si(0 to<br />

2.5nm)/<strong>GaAs</strong> contacts when annealed at 800C were <strong>in</strong>vestigated. The Si <strong>in</strong>terfacial layers<br />

<strong>and</strong> W films were sputter-deposited onto chemically etched <strong>GaAs</strong> substrates. The<br />

W/Si/<strong>GaAs</strong> diodes clearly exhibited the same Schottky-barrier characteristics as those of<br />

(WSi 0.6 )/<strong>GaAs</strong> diodes. By us<strong>in</strong>g secondary ion mass spectrometry, the Si layer was found<br />

to suppress As atom diffusion from <strong>GaAs</strong> substrates <strong>and</strong> <strong>in</strong>to W films dur<strong>in</strong>g anneal<strong>in</strong>g<br />

(800C, 1h). A reduction <strong>in</strong> natively oxidized <strong>GaAs</strong> surfaces was also observed <strong>in</strong> the<br />

<strong>in</strong>itial stages of Si layer deposition by X-ray photo-emission spectroscopy. These results<br />

suggested that the Si layer elim<strong>in</strong>ated native oxides from <strong>GaAs</strong> surfaces, result<strong>in</strong>g <strong>in</strong><br />

tungsten-silicide/<strong>GaAs</strong> <strong>in</strong>timate contact formation at the <strong>in</strong>terface. The Si obstructed the<br />

diffusion paths of As atoms at W gra<strong>in</strong> boundaries with W-Si-O ternary compounds.<br />

Y.Kuriyama, S.Ohfuji, J.Nagano: Journal of Applied Physics, 1987, 62[4], 1318-23<br />

[446-55/56-005]<br />

<strong>GaAs</strong>: As <strong>Diffusion</strong><br />

The <strong>in</strong>-diffusion of As vacancies, <strong>and</strong> their <strong>in</strong>teraction with the mid-gap electron trap,<br />

EL2, dur<strong>in</strong>g unprotected <strong>and</strong> proximity high-temperature anneal<strong>in</strong>g was modelled. By<br />

fitt<strong>in</strong>g exist<strong>in</strong>g data, it was found that the diffusive capture of V As by EL2 was <strong>in</strong>hibited<br />

by a large (greater than 1eV) repulsive barrier of unknown orig<strong>in</strong>. When taken together<br />

with <strong>other</strong> published results, the model <strong>in</strong>dicated that the diffusivity of V As was described<br />

by:<br />

D(cm 2 /s) = 0.004 exp[-1.8(eV)/kT]<br />

However, this value was thought to be uncerta<strong>in</strong> by at least an order of magnitude. A new<br />

discovery was the existence of a strong repulsive barrier between V As <strong>and</strong> EL2. This<br />

<strong>in</strong>hibited their <strong>in</strong>teraction, <strong>and</strong> was of the order of 1.2eV. When <strong>in</strong>terpreted as be<strong>in</strong>g a<br />

Coulomb barrier, it required that EL2 (or a portion of EL2) should carry a positive<br />

charge.<br />

276


As <strong>GaAs</strong> As<br />

In order to keep EL2 neutral, there would then have to be a compensat<strong>in</strong>g acceptor <strong>in</strong><br />

EL2.<br />

K.M.Luken, R.A.Morrow: Journal of Applied Physics, 1996, 79[3], 1388-90<br />

[446-134/135-124]<br />

<strong>GaAs</strong>: As <strong>Diffusion</strong><br />

Quantitative determ<strong>in</strong>ations were made of the contribution which As self-<strong>in</strong>terstitials<br />

made to the As self-diffusion coefficient. Values of the As self-<strong>in</strong>terstitial contribution<br />

were deduced from S <strong>in</strong>-diffusion profiles, which were simulated on the basis of a kickout<br />

mechanism.<br />

M.Uematsu, P.Werner, M.Schultz, T.Y.Tan, U.M.Gösele: Applied Physics Letters, 1995,<br />

67[19], 2863-5<br />

[446-125/126-120]<br />

<strong>GaAs</strong>: As <strong>Diffusion</strong><br />

Samples with a 100nm Co over-layer, which had been subjected to rapid thermal<br />

anneal<strong>in</strong>g (400 to 650C, 60s), were analyzed by us<strong>in</strong>g mass <strong>and</strong> energy dispersive recoil<br />

spectrometry. Separate characterizations of the C, O, Co, Ga, <strong>and</strong> As depth distributions<br />

were carried out. It was found that As migrated to the surface at anneal<strong>in</strong>g temperatures<br />

which were higher than 450C. The composition at various depths was determ<strong>in</strong>ed at a<br />

number of temperatures. On the basis of Arrhenius plots, the apparent activation energies<br />

were estimated to be equal to about 0.6eV for phase formation <strong>and</strong> equal to 1.3eV for<br />

diffusion. The X-ray diffraction data <strong>in</strong>dicated that CoGa <strong>and</strong> CoAs were present <strong>in</strong> all of<br />

the annealed samples. Scann<strong>in</strong>g electron microscopy showed that the surface was<br />

reticulated after heat treatment, <strong>and</strong> that gra<strong>in</strong> growth occurred at higher temperatures.<br />

M.Hult, H.J.Whitlow, M.Ostl<strong>in</strong>g, M.Andersson, Y.Andersson, I.L<strong>in</strong>deberg, K.Stähl:<br />

Journal of Applied Physics, 1994, 75[2], 835-43<br />

[446-117/118-165]<br />

<strong>GaAs</strong>: As <strong>Diffusion</strong><br />

An <strong>in</strong>vestigation was made of near-b<strong>and</strong>edge photolum<strong>in</strong>escence from semi-<strong>in</strong>sulat<strong>in</strong>g<br />

crystals after they had been annealed <strong>in</strong> wafer or bulk form. The results, with respect to<br />

uniformity after anneal<strong>in</strong>g, were <strong>in</strong> agreement with previous data. The 1.360eV emission<br />

b<strong>and</strong> which was seen <strong>in</strong> annealed crystals <strong>and</strong> which had been assumed to imply that a<br />

V As -related rapid-diffusion process was the mechanism which was responsible for the<br />

anneal<strong>in</strong>g-<strong>in</strong>duced uniformity, was shown to be unconnected with it. The <strong>in</strong>volvement of<br />

V As <strong>in</strong> the b<strong>and</strong> was questioned. From literature data, it was estimated that the diffusion<br />

coefficient of V As (1.2 x 10 -12 at 1050C, 5.8 x 10 -14 at 850C <strong>and</strong> 6.7 x 10 -16 cm 2 /s at 650C)<br />

was too low to permit bulk equilibrium <strong>and</strong> uniformity via vacancy diffusion from the<br />

surface at the anneal<strong>in</strong>g temperatures which were used. It was concluded that local<br />

rearrangement of defects was a viable mechanism for produc<strong>in</strong>g uniformity dur<strong>in</strong>g postgrowth<br />

anneal<strong>in</strong>g.<br />

V.Swam<strong>in</strong>athan, R.Caruso, S.J.<strong>Pearton</strong>: Journal of Applied Physics, 1988, 63[6], 2164-7<br />

[446-157/58-273]<br />

277


As <strong>GaAs</strong> B<br />

<strong>GaAs</strong>[l]: As <strong>Diffusion</strong><br />

The As concentration profiles ahead of a crystal <strong>in</strong>terface, when advanc<strong>in</strong>g <strong>in</strong>to a Ga-rich<br />

solution, were determ<strong>in</strong>ed (dur<strong>in</strong>g the electro-epitaxial growth of layers) by us<strong>in</strong>g<br />

computer simulation techniques. The effects of Peltier heat<strong>in</strong>g or cool<strong>in</strong>g, <strong>and</strong> of<br />

electromigration, dur<strong>in</strong>g growth were <strong>in</strong>corporated. The growth velocity <strong>in</strong> the absence or<br />

presence of convection, due to the Peltier effect <strong>and</strong> to electromigration, was calculated<br />

under various conditions. It was observed that there was a transition, <strong>in</strong> the movement of<br />

As atoms towards the crystal <strong>in</strong>terface, from smooth <strong>and</strong> orderly to turbulent <strong>and</strong> wavy as<br />

the <strong>in</strong>tensity of the electric field <strong>in</strong>creased dur<strong>in</strong>g electro-epitaxial growth.<br />

R.S.Q.Fareed, R.Dhanasekaran, P.Ramasamy: Journal of Applied Physics, 1994, 75[8],<br />

3953-8<br />

[446-117/118-165]<br />

Au<br />

<strong>GaAs</strong>: Au <strong>Diffusion</strong><br />

Rutherford back-scatter<strong>in</strong>g spectrometry <strong>and</strong> X-ray photoelectron spectroscopy were used<br />

to <strong>in</strong>vestigate compositional changes, <strong>in</strong> th<strong>in</strong> metal-semiconductor systems, which were<br />

caused by Ar + <strong>and</strong> N + ion bombardment or by anneal<strong>in</strong>g. The <strong>in</strong>vestigation was carried<br />

out on contacts of Ni-Au-Ge on <strong>GaAs</strong>, as well as on irradiated <strong>and</strong> non-irradiated Au-<br />

<strong>GaAs</strong>. The structures were bombarded with Ar + <strong>and</strong> N + ions to doses of between 10 14 <strong>and</strong><br />

3 x 10 16 /cm 2 . The experimental results <strong>in</strong>dicated that the <strong>in</strong>terdiffusion of Au <strong>and</strong> Ga<br />

atoms depended upon the bombardment dose. Upon anneal<strong>in</strong>g the samples, block<strong>in</strong>g of<br />

<strong>in</strong>terdiffusion was observed <strong>in</strong> Au-<strong>GaAs</strong> structures (which had been deposited on 50keV<br />

Ar + ion-irradiated <strong>and</strong> pre-treated <strong>GaAs</strong>) at certa<strong>in</strong> radiation-defect concentrations. This<br />

behavior was attributed to Au-Ga bond formation, <strong>and</strong> appeared to depend upon Ga<br />

<strong>in</strong>terstitial atoms.<br />

L.B.Guoba, A.A.Vitkauskas, J.V.Kameneckas, V.R.Sargünas, A.P.Sakalas: Physica<br />

Status Solidi A, 1989, 111[2], 507-13<br />

[446-64/65-162]<br />

B<br />

<strong>GaAs</strong>: B <strong>Diffusion</strong><br />

The <strong>in</strong>troduction of B <strong>in</strong>to As sites, as deduced from the strength of vibrational modes at<br />

601.7 <strong>and</strong> 628.3/cm, was studied as a function of the fluence of 2MeV electrons.<br />

Simultaneous monitor<strong>in</strong>g of the strength of the Is-2p electronic transitions of neutral<br />

shallow acceptors, <strong>and</strong> of the neutral 0.078eV acceptor <strong>and</strong> its s<strong>in</strong>gly-ionized 0.203eV<br />

level, provided accurate data on the position of the Fermi level dur<strong>in</strong>g irradiation. The<br />

results were <strong>in</strong>consistent with previous models for the location of B on As sites. A model<br />

278


B <strong>GaAs</strong> Be<br />

was proposed which was based upon the onset of enhanced B diffusion when the Fermi<br />

level lay above 0.078eV.<br />

W.J.Moore, R.L.Hawk<strong>in</strong>s: Journal of Applied Physics, 1988, 63[12], 5699-702<br />

[446-72/73-010]<br />

Be<br />

310 <strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

Spatial localisation of Be <strong>in</strong> d-doped material, with<strong>in</strong> a few lattice constants (less than<br />

2nm), was achieved at low growth temperatures for Be concentrations of less than<br />

10 14 /cm 2 ; as revealed by capacitance-voltage profiles <strong>and</strong> secondary ion mass<br />

spectroscopy. At high growth temperatures, <strong>and</strong> at higher Be concentrations, significant<br />

spread<strong>in</strong>g of the dopants occurred <strong>and</strong> was expla<strong>in</strong>ed <strong>in</strong> terms of Fermi-level p<strong>in</strong>n<strong>in</strong>g<strong>in</strong>duced<br />

segregation, repulsive Coulomb <strong>in</strong>teractions of dopants, <strong>and</strong> diffusion. The<br />

highest Be concentration which was obta<strong>in</strong>ed at low growth temperatures exceeded 2 x<br />

10 20 /cm 3 , <strong>and</strong> was limited by repulsive dopant <strong>in</strong>teractions. It was shown that the<br />

repulsive Coulomb <strong>in</strong>teraction resulted <strong>in</strong> a correlated non-r<strong>and</strong>om dopant distribution. It<br />

was deduced that the diffusivity of Be (table 10) could be described by:<br />

D (cm 2 /s) = 0.00002 exp[-1.95(eV)/kT]<br />

The present values were much lower than those previously reported.<br />

E.F.Schubert, J.M.Kuo, R.F.Kopf, H.S.Luftman, L.C.Hopk<strong>in</strong>s, N.J.Sauer: Journal of<br />

Applied Physics, 1990, 67[4], 1969-72<br />

[446-157/159-279]<br />

Table 10<br />

Diffusivity of Be <strong>in</strong> <strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

1005 4.7 x 10 -13<br />

945 1.5 x 10 -13<br />

895 6.0 x 10 -14<br />

800 1.4 x 10 -14<br />

700 1.7 x 10 -15<br />

600 8.6 x 10 -16<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

It was po<strong>in</strong>ted out that Be was one of the ma<strong>in</strong> p-type dopants which were used for the<br />

fabrication of devices that were based upon <strong>GaAs</strong> or related <strong>III</strong>-V materials. The element<br />

dissolved substitutionally on the group-<strong>III</strong> sub-lattice, <strong>and</strong> diffused via a kick-out<br />

mechanism which <strong>in</strong>volved group-<strong>III</strong> self-<strong>in</strong>terstitials. Non-equilibrium concentrations of<br />

these self-<strong>in</strong>terstitials had a marked effect upon the diffusivity of Be. Various situations<br />

were considered <strong>in</strong> which non-equilibrium po<strong>in</strong>t defects played a role <strong>in</strong> Be diffusion.<br />

279


Be <strong>GaAs</strong> Be<br />

These <strong>in</strong>cluded the <strong>in</strong>-diffusion of such dopants from an external source, the diffusion of<br />

grown-<strong>in</strong> dopants, self-<strong>in</strong>terstitial generation by Fermi-level surface p<strong>in</strong>n<strong>in</strong>g, <strong>and</strong><br />

recomb<strong>in</strong>ation-enhanced Be diffusion dur<strong>in</strong>g device operation. It was noted that the<br />

diffusion behavior of C, which was found on the group-V sub-lattice of <strong>GaAs</strong>, was much<br />

less sensitive to non-equilibrium po<strong>in</strong>t defects. It was therefore used to replace Be as a p-<br />

type dopant.<br />

M.Uematsu, K.Wada, U.Gösele: Applied Physics A, 1992, 55[4], 301-12<br />

[446-93/94-008]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

A study was made of the contributions of segregation, diffusion <strong>and</strong> aggregation to the<br />

broaden<strong>in</strong>g of d-doped planes of Be. It was found that sharp spikes of Be could be<br />

obta<strong>in</strong>ed for sheet densities which were below 10 13 /cm 2 <strong>and</strong> for growth temperatures of<br />

500C or less. At higher temperatures or densities, segregation or concentration-dependent<br />

rapid diffusion could occur; thus caus<strong>in</strong>g significant spread<strong>in</strong>g even dur<strong>in</strong>g growth. The<br />

co-deposition of Si <strong>and</strong> Be markedly reduced this broaden<strong>in</strong>g.<br />

J.J.Harris, J.B.Clegg, R.B.Beall, J.Castagné, K.Woodbridge, C.Roberts: Journal of<br />

Crystal Growth, 1991, 111[1-4], 239-45<br />

[446-91/92-001]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

The slow positron technique was used to study undoped <strong>and</strong> Be-doped samples, <strong>and</strong><br />

thereby determ<strong>in</strong>e the effect of Be upon the creation <strong>and</strong> migration of Ga vacancies, V Ga ,<br />

dur<strong>in</strong>g anneal<strong>in</strong>g. It was deduced that a V Ga mono-vacancy which was created <strong>in</strong> Bedoped<br />

material resulted <strong>in</strong> an enhanced Coulombic <strong>in</strong>teraction between an As vacancy,<br />

V As , <strong>and</strong> a Be acceptor, Be Ga . In the case of undoped material, the formation of divacancies,<br />

V Ga -V As , predom<strong>in</strong>ated. The migration length of the vacancies was shorter <strong>in</strong><br />

Be-doped material than <strong>in</strong> undoped material. It was therefore suggested that Ga<br />

<strong>in</strong>terstitials, I Ga , existed <strong>in</strong> the Be-diffused layer <strong>and</strong> <strong>in</strong>teracted with V Ga which were<br />

<strong>in</strong>troduced from the surface. It was suggested that a kick-out mechanism governed Be<br />

diffusion <strong>in</strong> this material.<br />

J.L.Lee, L.Wei, S.Tanigawa, M.Kawabe: Journal of Applied Physics, 1991, 69[9], 6364-3<br />

[446-86/87-009]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

Recomb<strong>in</strong>ation-enhanced impurity diffusion was observed for the first time <strong>in</strong> Be-doped<br />

<strong>GaAs</strong>. It was found that Be diffusion under forward bias was enhanced by a factor of<br />

about 10 15 at room temperature, <strong>and</strong> that the activation energy for diffusion decreased<br />

from 1.8eV for thermal diffusion:<br />

D(cm 2 /s) = 8.3 x 10 -7 exp[-1.8(eV)/kT]<br />

to 0.6eV under recomb<strong>in</strong>ation-enhanced conditions:<br />

D(cm 2 /s) = 8.7 x 10 -11 exp[-0.59(eV)/kT]<br />

280


Be <strong>GaAs</strong> Be<br />

M.Uematsu, K.Wada: Applied Physics Letters, 1991, 58[18], 2015-7<br />

[446-84/85-013]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

A close relationship between Be surface segregation <strong>and</strong> diffusion, <strong>in</strong> molecular beam<br />

epitaxial <strong>GaAs</strong> layers which were heavily doped with Be, was analyzed with<strong>in</strong> the<br />

framework of a thermodynamic approach to segregation effects. Good agreement between<br />

the theoretical <strong>and</strong> experimental results suggested that the ma<strong>in</strong> cause of extremely fast<br />

Be <strong>in</strong>-diffusion <strong>in</strong> Be-doped <strong>GaAs</strong>, <strong>and</strong> the deterioration of its surface morphology <strong>and</strong><br />

lum<strong>in</strong>escence properties, was Be surface segregation. This resulted <strong>in</strong> near-surface solidphase<br />

layer enrichment with Be, as compared with bulk Be-doped <strong>GaAs</strong>. The effect of<br />

growth parameters (excess As pressure, substrate temperature, growth rate) <strong>and</strong> dopant<br />

level upon the likelihood of Be segregation layer formation was considered.<br />

S.V.Ivanov, P.S.Kopev, N.N.Ledentsov: Journal of Crystal Growth, 1991, 108[3-4], 661-<br />

9<br />

[446-81/82-002]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

The effect of the substrate orientation upon Be transport dur<strong>in</strong>g <strong>GaAs</strong> molecular beam<br />

epitaxy was studied by means of secondary ion mass spectrometry. The substrates were<br />

misoriented from (100) towards (111)A, <strong>and</strong> epitaxial growth was performed at 630C for<br />

Be dopant contents of between 5 x 10 19 <strong>and</strong> 7 x 10 19 /cm 3 . Surface segregation <strong>and</strong><br />

anomalous diffusion similarly depended upon the substrate orientation. In the case of the<br />

(311)A orientation, Be transport was sharply reduced from its value for the conventional<br />

(100) orientation. The results were expla<strong>in</strong>ed qualitatively by consider<strong>in</strong>g the effect of<br />

atomic steps upon the grow<strong>in</strong>g surface.<br />

K.Mochizuki, S.Goto, C.Kusano: Applied Physics Letters, 1991, 58[25], 2939-41<br />

[446-81/82-009]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

The out-diffusion of implanted Be was found to be identical after capless or capped<br />

(nitride or oxide) rapid thermal anneal<strong>in</strong>g at temperatures of 900 to 1000C. It depended<br />

upon the Be dose <strong>and</strong> its proximity to the surface. Out-diffusion was more pronounced<br />

when the Be implant was shallow (less than 100nm) <strong>and</strong>/or the Be + dose was high<br />

(greater than 10 15 /cm 2 ). It was demonstrated that Be out-diffusion was driven by the<br />

presence of a highly damaged surface layer. Auger results <strong>in</strong>dicated the formation of a<br />

BeO x compound at the surface of a high-dose (10 16 /cm 2 ) Be-implanted sample that was<br />

subjected to capless rapid thermal anneal<strong>in</strong>g (1000C, 1s). It appeared that BeO x formation<br />

occurred when the out-diffused Be <strong>in</strong>teracted with native Ga/As oxides dur<strong>in</strong>g anneal<strong>in</strong>g.<br />

All of the Be which rema<strong>in</strong>ed <strong>in</strong> the <strong>GaAs</strong>, after rapid thermal anneal<strong>in</strong>g at temperatures<br />

above 900C for 2s, was electrically active.<br />

H.Baratte, D.K.Sadana, J.P.De Souza, P.E.Hallali, R.G.Schad, M.Norcott, F.Cardone:<br />

Journal of Applied Physics, 1990, 67[10], 6589-91<br />

[446-78/79-012]<br />

281


Be <strong>GaAs</strong> Be<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

High depth-resolution secondary ion mass spectrometry profil<strong>in</strong>g was used to <strong>in</strong>vestigate<br />

the broaden<strong>in</strong>g of d-doped planes of Be <strong>in</strong> material which had been prepared by us<strong>in</strong>g<br />

molecular beam epitaxial methods. It was confirmed that concentration-dependent<br />

diffusion was the predom<strong>in</strong>ant broaden<strong>in</strong>g process for Be at growth temperatures of less<br />

than 600C. By <strong>in</strong>corporat<strong>in</strong>g Si atoms <strong>in</strong>to the same plane, it was shown that the<br />

broaden<strong>in</strong>g could be completely <strong>in</strong>hibited. This suggested that the rapid diffusion process<br />

resulted from mutual repulsion between the Be Ga - ions, <strong>and</strong> was prevented by the reverse<br />

field which arose from Si Ga + ions or by the formation of low-mobility Si Ga + -Be Ga<br />

-<br />

complexes. The rapid diffusion of Si as Si Ga -Si As pairs was also reduced. The latter was<br />

attributed to a Fermi-level effect, with compensation by Be tend<strong>in</strong>g to reduce the<br />

probability of Si As formation. The surface segregation of Si was unaffected, whereas that<br />

of Be was reduced. This <strong>in</strong>dicated that the surface fields which existed dur<strong>in</strong>g growth<br />

contributed to the behavior of Be, but not to that of Si.<br />

J.J.Harris, J.B.Clegg, R.B.Beall, J.Castagné: Semiconductor Science <strong>and</strong> Technology,<br />

1990, 5[7], 785-8<br />

[446-76/77-007]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

The redistribution of Be implants dur<strong>in</strong>g post-implantation anneal<strong>in</strong>g was studied <strong>in</strong> order<br />

to evaluate the effect of implantation damage upon the diffusion process. The Be implants<br />

exhibited only uniform concentration-dependent diffusion, unlike the rapid up-hill<br />

diffusion which was observed <strong>in</strong> the peak of Mg implants. This difference was expla<strong>in</strong>ed<br />

by <strong>in</strong>vok<strong>in</strong>g a substitutional-<strong>in</strong>terstitial diffusion mechanism <strong>and</strong> by perform<strong>in</strong>g computer<br />

simulations of damage-generated po<strong>in</strong>t defects. In the up-hill diffusion region, the dopants<br />

diffused from areas of excess <strong>in</strong>terstitial concentration towards areas of excess vacancy<br />

concentration. A critical po<strong>in</strong>t defect concentration was necessary <strong>in</strong> order to <strong>in</strong>itiate uphill<br />

diffusion. This behavior could be <strong>in</strong>duced, <strong>in</strong> the case of Be implants, by coimplant<strong>in</strong>g<br />

with a heavier element such as Ar.<br />

H.G.Rob<strong>in</strong>son, M.D.Deal, D.A.Stevenson: Applied Physics Letters, 1990, 56[6], 554-6<br />

[446-74-009]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

The behavior of implanted Be was studied by anneal<strong>in</strong>g samples which had been<br />

implanted with low or high Be doses. The high-dose (10 14 /cm 2 ) samples exhibited an<br />

<strong>in</strong>crease <strong>in</strong> diffusion upon <strong>in</strong>creas<strong>in</strong>g the anneal<strong>in</strong>g temperature from 700 to 900C.<br />

However, the low-dose (2 x 10 13 /cm 2 ) samples exhibited a decrease <strong>in</strong> diffusivity as the<br />

temperature <strong>in</strong>creased. The temperature dependence <strong>in</strong> the low-dose case could be<br />

reversed by the co-implantation of B (10 14 /cm 2 ). This behavior was expla<strong>in</strong>ed <strong>in</strong> terms of<br />

282


Be <strong>GaAs</strong> Be<br />

the substitutional-<strong>in</strong>terstitial diffusion mechanism <strong>and</strong> the relative concentrations of<br />

<strong>in</strong>terstitial <strong>and</strong> substitutional Be atoms <strong>in</strong> the various cases.<br />

M.D.Deal, H.G.Rob<strong>in</strong>son: Applied Physics Letters, 1989, 55[10], 996-9<br />

[446-72/73-010]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

Layers of Be-doped material were grown at 300C by us<strong>in</strong>g the migration enhanced<br />

epitaxy technique. The layers exhibited essentially no electrical activation. Rapid thermal<br />

anneal<strong>in</strong>g of the layers resulted <strong>in</strong> a mobility <strong>and</strong> hole concentration which were<br />

comparable to those of conventional molecular beam epitaxial layers which were grown<br />

at 600C. Secondary ion mass spectroscopy showed that Be diffusion <strong>in</strong> annealed<br />

migration enhanced epitaxial layers was much smaller than that <strong>in</strong> conventional molecular<br />

beam epitaxial layers; especially highly-doped ones. Raman spectroscopy <strong>and</strong> 4K<br />

photolum<strong>in</strong>escence experiments were also performed. It was concluded that the migration<br />

enhanced epitaxy method could replace the conventional molecular beam epitaxy method<br />

<strong>in</strong> applications which required a high hole concentration <strong>and</strong> little diffusion.<br />

B.Tadayon, S.Tadayon, W.J.Schaff, M.G.Spencer, G.L.Harris, P.J.Tasker, C.E.C.Wood,<br />

L.F.Eastman: Applied Physics Letters, 1989, 55[1], 59-61<br />

[446-70/71-106]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

Abrupt Be dop<strong>in</strong>g profiles were obta<strong>in</strong>ed by means of organometallic vapor phase<br />

epitaxy. Secondary ion mass spectroscopy was used to study the anneal<strong>in</strong>g behavior of<br />

profiles for Be concentrations of 2 x 10 18 /cm 3 . The diffusion fronts were non-Gaussian<br />

<strong>and</strong> abrupt. Estimates of the diffusion coefficient of Be were obta<strong>in</strong>ed by assum<strong>in</strong>g a<br />

quadratic concentration dependence. The Be diffusion coefficient was equal to about 10 -<br />

15 cm 2 /s at 825C. This was at least an order of magnitude lower than that reported for Zn<br />

profiles which were grown by means of organometallic vapor phase epitaxy. In addition,<br />

anomalous surface tail<strong>in</strong>g growth was observed. This was very similar to that which was<br />

reported to occur dur<strong>in</strong>g Be dop<strong>in</strong>g via molecular beam epitaxy.<br />

M.J.Tejwani, H.Kanber, B.M.Pa<strong>in</strong>e, J.M.Whelan: Applied Physics Letters, 1988, 53[24],<br />

2411-3<br />

[446-64/65-162]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

A review was presented of self-diffusion mechanisms <strong>and</strong> dop<strong>in</strong>g-enhanced superlattice<br />

disorder<strong>in</strong>g. With regard to the <strong>in</strong>fluence of Be p-type dopants, the Fermi level effect had<br />

to be considered; together with dopant diffusion-<strong>in</strong>duced Ga self-<strong>in</strong>terstitial<br />

supersaturation or undersaturation. In accord with its effect upon superlattice disorder<strong>in</strong>g,<br />

Be diffusion appeared to be governed by the kick-out mechanism. It was concluded that<br />

dislocations <strong>in</strong> this material <strong>and</strong> <strong>in</strong> <strong>other</strong> <strong>III</strong>-V compounds were only moderately efficient<br />

s<strong>in</strong>ks or sources for po<strong>in</strong>t defects.<br />

T.Y.Tan, U.Gösele: Materials Science <strong>and</strong> Eng<strong>in</strong>eer<strong>in</strong>g, 1988, B1, 47-65<br />

[446-62/63-208]<br />

283


Be <strong>GaAs</strong> Be<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

The diffusion profiles of buried Be dopant which had been implanted by us<strong>in</strong>g a focussed<br />

ion beam were determ<strong>in</strong>ed after anneal<strong>in</strong>g. The diffusion coefficient of the Be was<br />

determ<strong>in</strong>ed by fitt<strong>in</strong>g the results of computer calculations. It was found that the diffusion<br />

coefficient of the Be was enhanced by excess <strong>in</strong>terstitial Be. The Be diffusion profiles<br />

exp<strong>and</strong>ed upon anneal<strong>in</strong>g at 850C. The diffusion coefficient of Si which had been<br />

<strong>in</strong>troduced by us<strong>in</strong>g a focussed ion beam was undetectably small when compared with<br />

that of Be at 850C.<br />

T.Morita, J.Kobayashi, T.Takamori, A.Takamori, E.Miyauchi, H.Hashimoto: Japanese<br />

Journal of Applied Physics, 1987, 26[8], 1324-7<br />

[446-55/56-005]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

Models were presented for the distribution profiles of Be <strong>in</strong> ion-doped layers after<br />

implantation <strong>and</strong> anneal<strong>in</strong>g. The possibility of predict<strong>in</strong>g the mean free path of Be <strong>in</strong> <strong>III</strong>-<br />

V compounds was considered. The effect of defect-impurity <strong>in</strong>teractions upon Be<br />

diffusion was also exam<strong>in</strong>ed. It was found that a flux of impurities towards the surface<br />

occurred which was not diffusive <strong>in</strong> nature.<br />

G.I.Koltsov, V.V.Makarov, S.J.Yurchuk: Fizika i Tekhnika Poluprovodnikov, 1996,<br />

30[10], 1907-16 (<strong>Semiconductors</strong>, 1996, 30[10], 996-1000)<br />

[446-148/149-171]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

Atomic-resolution images of Be delta-doped layers were obta<strong>in</strong>ed by means of crosssectional<br />

scann<strong>in</strong>g tunnell<strong>in</strong>g microscopy. In the case of samples which had been grown<br />

at 480C, it was observed that the dop<strong>in</strong>g layer width for concentrations of up to 10 13 /cm 2<br />

was less than 1nm. At higher dopant concentrations, it was found that the dopant layer<br />

thickness <strong>in</strong>creased markedly with dopant concentration. It was suggested that this<br />

broaden<strong>in</strong>g of the dopant layer at high dopant concentrations was due to Coulombic<br />

repulsion between <strong>in</strong>dividual Be ions. The effect of Coulombic repulsion could also be<br />

observed <strong>in</strong> the spatial distribution of the dopant atoms <strong>in</strong> the plane of the dopant layer.<br />

P.M.Koenraad, M.B.Johnson, H.W.M.Salem<strong>in</strong>k, W.C.Van der Vleuten, J.H.Wolter:<br />

Materials Science <strong>and</strong> Eng<strong>in</strong>eer<strong>in</strong>g B, 1995, 35[1-3], 485-8<br />

[446-136/137-109]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

A study was made of the role that was played by the wafer surface <strong>in</strong> the transient<br />

diffusion of Be. Samples were doped dur<strong>in</strong>g molecular beam epitaxial growth, <strong>and</strong> were<br />

annealed (900C, 0.25 or 2h) under 2 different caps. In some of the annealed samples, the<br />

dopant was <strong>in</strong>itially located near to the surface. In <strong>other</strong> samples, the dopant was <strong>in</strong>itially<br />

located <strong>in</strong> a buried layer. Both types of sample were analyzed by means of secondary ion<br />

mass spectrometry. It was found that variations <strong>in</strong> the diffusion behavior under the<br />

various experimental conditions could all be qualitatively expla<strong>in</strong>ed <strong>in</strong> terms of a model<br />

284


Be <strong>GaAs</strong> Be<br />

which took account of 3 important effects. These were the transient evolution of po<strong>in</strong>t<br />

defect populations, the <strong>in</strong>jection of Ga vacancies by an oxide cap, <strong>and</strong> the efficiency of<br />

the surface <strong>in</strong> restor<strong>in</strong>g po<strong>in</strong>t-defect equilibrium.<br />

Y.M.Haddara, M.D.Deal, J.C.Bravman: Applied Physics Letters, 1996, 68[14], 1939-41<br />

[446-134/135-124]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

The migration of ion-implanted Be was studied as a function of Al concentration. The<br />

behavior of Be <strong>in</strong> Al<strong>GaAs</strong> was similar to that <strong>in</strong> <strong>GaAs</strong>, <strong>and</strong> it even exhibited the<br />

anomalous characteristic of <strong>in</strong>creased redistribution with decreas<strong>in</strong>g temperature. The<br />

results could be described by:<br />

Ga 0.8 Al 0.2 As: D(cm 2 /s) = 1.8 x 10 -9 exp[-0.90(eV)/kT]<br />

Ga 0.6 Al 0.4 As: D(cm 2 /s) = 3.3 x 10 -9 exp[-0.84(eV)/kT]<br />

<strong>GaAs</strong>: D(cm 2 /s) = 8.9 x 10 -10 exp[-1.00(eV)/kT]<br />

The diffusivity of Be appeared to <strong>in</strong>crease with Al content. This was suggested to be due<br />

to an <strong>in</strong>crease <strong>in</strong> the bond strength of matrix atoms upon add<strong>in</strong>g Al. This prevented the<br />

easy transfer of Be from <strong>in</strong>terstitial to substitutional sites. An over-saturation of Be<br />

<strong>in</strong>terstitials could also expla<strong>in</strong> the persistence of anomalous diffusion <strong>in</strong> Al<strong>GaAs</strong> with<br />

respect to the anneal<strong>in</strong>g temperature. The results were expla<strong>in</strong>ed <strong>in</strong> terms of a<br />

substitutional-<strong>in</strong>terstitial diffusion mechanism, the relative amounts of <strong>in</strong>terstitial <strong>and</strong><br />

substitutional Be, <strong>and</strong> the relative difficulty of mov<strong>in</strong>g from an <strong>in</strong>terstitial to a<br />

substitutional site.<br />

C.C.Lee, M.D.Deal, J.C.Bravman: Applied Physics Letters, 1995, 66[3], 355-7<br />

[446-123/124-160]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

The diffusion of implanted Be <strong>in</strong> liquid-encapsulated Czochralski material was modelled<br />

by us<strong>in</strong>g a computer simulation. The so-called plus-one approach to defect generation<br />

after implantation, as well as the assumed existence of local Ga <strong>in</strong>terstitial s<strong>in</strong>ks, were<br />

successfully used to simulate a high Be diffusivity, up-hill diffusion <strong>and</strong> a time-dependent<br />

Bi diffusivity. The fast diffusion of implanted Be could be simulated by us<strong>in</strong>g the same<br />

<strong>in</strong>tr<strong>in</strong>sic Bi diffusivity as that used <strong>in</strong> simulat<strong>in</strong>g the slow diffusion of molecular beam<br />

epitaxially grown-<strong>in</strong> Be. Account was taken of the roles which were played by extended<br />

defects <strong>and</strong> non-equilibrium Ga po<strong>in</strong>t defects <strong>in</strong> affect<strong>in</strong>g the anomalous diffusion<br />

behavior of implanted Be.<br />

J.C.Hu, M.D.Deal, J.D.Plummer: Journal of Applied Physics, 1995, 78[3], 1606-13<br />

[446-123/124-161]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

<strong>Diffusion</strong> was studied <strong>in</strong> samples of molecular beam epitaxial material with grown-<strong>in</strong> Be.<br />

The diffusion profiles of samples which had been annealed under various conditions were<br />

determ<strong>in</strong>ed by us<strong>in</strong>g secondary ion mass spectrometry, <strong>and</strong> a computer simulation was<br />

used to analyze the experimental results <strong>and</strong> extract diffusion parameters. The Be<br />

285


Be <strong>GaAs</strong> Be<br />

diffusion profiles exhibited k<strong>in</strong>ks, <strong>and</strong> a time-dependent diffusivity, which were<br />

successfully simulated. It was deduced that the <strong>in</strong>tr<strong>in</strong>sic Be diffusivity was described by:<br />

D(cm 2 /s) = 0.17 exp[-3.39(eV)/kT]<br />

J.C.Hu, M.D.Deal, J.D.Plummer: Journal of Applied Physics, 1995, 78[3], 1595-605<br />

[446-123/124-161]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

Heavily-doped polycrystall<strong>in</strong>e material which had been grown by molecular beam epitaxy<br />

<strong>and</strong> metalorganic chemical vapor deposition was studied by us<strong>in</strong>g secondary ion mass<br />

spectrometry <strong>and</strong> Hall measurements. It was found that Be rapidly diffused <strong>in</strong>to the<br />

undoped buffer layer at a growth temperature of 450C. The concentration-depth profiles<br />

of Be <strong>in</strong> Al<strong>GaAs</strong>/<strong>GaAs</strong> heterojunction bipolar transistor layers <strong>in</strong>dicated that Be diffused<br />

ma<strong>in</strong>ly along gra<strong>in</strong> boundaries.<br />

K.Mochizuki, T.Nakamura: Applied Physics Letters, 1994, 65[16], 2066-8<br />

[446-119/120-191]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

Implantation (1.5 x 10 14 /cm 2 ) of 30keV Be <strong>in</strong>to (x11)A-oriented semi-<strong>in</strong>sulat<strong>in</strong>g <strong>GaAs</strong><br />

substrates (where x took values of up to 4) was carried out. For comparison, (110)- <strong>and</strong><br />

(100)-oriented substrates were also implanted. It was found that the <strong>in</strong>-diffusion of Be <strong>in</strong><br />

(311)A-oriented substrates was lower than that <strong>in</strong> (100) material.<br />

M.V.Rao, H.B.Dietrich, P.B.Kle<strong>in</strong>, A.Fathimulla, D.S.Simons, P.H.Chi: Journal of<br />

Applied Physics, 1994, 75[12], 7774-8<br />

[446-117/118-166]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

A study was made of the diffusion of Be <strong>in</strong> δ-doped layers of (111)A- or (100)-type, <strong>and</strong><br />

of the evaporation of As atoms from the surfaces. It was found that the diffusion of<br />

dopants <strong>in</strong> (111)A layers was slower than <strong>in</strong> (100), regardless of the presence of As<br />

vacancies. On the <strong>other</strong> h<strong>and</strong>, diffusion <strong>in</strong> (100) layers was enhanced by the presence of<br />

As vacancies. It was noted that As atoms on the (111)A surface did not evaporate easily,<br />

as compared with those on the (100) surface.<br />

A.Sh<strong>in</strong>oda, T.Yamamoto, M.Inai, T.Takebe, T.Watanabe: Japanese Journal of Applied<br />

Physics, 1993, 32[2-10A], L1374-6<br />

[446-115/116-116]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

Abnormal out-diffusion from heavily Be-doped material, prepared by molecular beam<br />

epitaxy, was found to be <strong>in</strong>itiated by a decrease <strong>in</strong> the lattice constant of the p + epilayer.<br />

From double-crystal X-ray spectra, Van der Pauw measurements, photolum<strong>in</strong>escence data,<br />

<strong>and</strong> <strong>in</strong>fra-red absorption spectra for Be-doped material with various dopant concentrations,<br />

it was deduced that there existed a critical dop<strong>in</strong>g concentration (2.6 x 10 19 /cm 3 ) beyond<br />

which the lattice constant of the epilayer began to decrease, <strong>and</strong> Be out-<br />

286


Be <strong>GaAs</strong> Be<br />

diffusion <strong>in</strong>to the substrate was significantly enhanced. It was suggested that the tensile<br />

stress on the epilayer resulted <strong>in</strong> abnormal Be out-diffusion. The absorption coefficient,<br />

<strong>in</strong> the 8 to 10µ region, of Be-doped material with a carrier concentration of 8.3 x<br />

10 19 /cm 3 was found to be about 10 4 /cm.<br />

B.D.Liu, T.H.Shieh, M.Y.Wu, T.C.Chang, S.C.Lee, H.H.L<strong>in</strong>: Journal of Applied Physics,<br />

1992, 72[7], 2767-72<br />

[446-106/107-033]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

Samples of n-type <strong>and</strong> p-type d-doped material, grown by molecular beam epitaxy <strong>and</strong><br />

with quite high doses of Be, were <strong>in</strong>vestigated by means of transmission electron<br />

microscopy. The magnitude of the doses ranged from half a monolayer to 2 monolayers.<br />

The microscopic structures of the d-doped regions <strong>and</strong> of the adjacent epilayers were<br />

observed directly. The effect of impurity spread<strong>in</strong>g upon the hetero-<strong>in</strong>terfaces <strong>and</strong><br />

superlattices was studied. The Be atoms which were present <strong>in</strong> Be d-doped samples<br />

spread over a wide region <strong>and</strong> caused rough hetero-<strong>in</strong>terfaces <strong>and</strong> wavy superlattices to<br />

form. The spread<strong>in</strong>g of Be was attributed to segregation <strong>and</strong> diffusion which occurred<br />

dur<strong>in</strong>g growth. Stack<strong>in</strong>g faults were found <strong>in</strong> d-doped samples when they were grown at<br />

low temperatures. Their presence was attributed to local stra<strong>in</strong>s that were caused by<br />

heavy dop<strong>in</strong>g.<br />

D.G.Liu, J.C.Fan, C.P.Lee, K.H.Chang, D.C.Liou: Journal of Applied Physics, 1993,<br />

73[2], 608-14<br />

[446-106/107-034]<br />

<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

A study was made of recomb<strong>in</strong>ation-enhanced impurity diffusion <strong>in</strong> Be-doped material.<br />

An <strong>in</strong>vestigation of tunnel diodes revealed that Be diffusion under forward bias was<br />

enhanced by a factor of about 10 15 at room temperature, <strong>and</strong> that the activation energy for<br />

diffusion was reduced from 1.8eV for thermal diffusion to 0.6eV for recomb<strong>in</strong>ationenhanced<br />

impurity diffusion. In the latter process for Be, the energy which was related to<br />

m<strong>in</strong>ority carrier <strong>in</strong>jection at the recomb<strong>in</strong>ation center encouraged the annihilation of the<br />

recomb<strong>in</strong>ation center, where a po<strong>in</strong>t defect which enhanced the Be diffusion was<br />

generated.<br />

M.Uematsu, K.Wada: Materials Science Forum, 1992, 83-87, 1551-6<br />

[446-99/100-063]<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

The active layers of s<strong>in</strong>gle quantum-well separate conf<strong>in</strong>ement heterostructure lasers<br />

which had been grown by means of molecular beam epitaxy were <strong>in</strong>vestigated by us<strong>in</strong>g<br />

photolum<strong>in</strong>escence absorption spectroscopy, secondary ion mass spectroscopy,<br />

capacitance-voltage profil<strong>in</strong>g <strong>and</strong> laser threshold current measurements. It was found that<br />

a significant amount of Be diffusion occurred under normal growth conditions. Large<br />

concentrations of Be <strong>in</strong> the quantum well were correlated with the lack of an exciton<br />

feature <strong>in</strong> the absorption spectrum. The amount of Be <strong>in</strong> the active region was reduced by<br />

287


Be <strong>GaAs</strong> C<br />

reduc<strong>in</strong>g the Be concentration <strong>and</strong> by decreas<strong>in</strong>g the growth temperature <strong>in</strong> the upper<br />

cladd<strong>in</strong>g region of the laser.<br />

G.E.Kohnke, M.W.Koch, C.E.C.Wood, G.W.Wicks: Applied Physics Letters, 1995,<br />

66[21], 2786-8<br />

[446-121/122-062]<br />

<strong>GaAs</strong>/GaAlAs: Be <strong>Diffusion</strong><br />

Heterostructures of <strong>GaAs</strong>/Ga 0.7 Al 0.3 As, which conta<strong>in</strong>ed Zn <strong>and</strong> Se as <strong>in</strong>tr<strong>in</strong>sic p <strong>and</strong> n<br />

dopants, were subjected to comb<strong>in</strong>ed Be <strong>and</strong> O implantation. Rapid thermal anneal<strong>in</strong>g<br />

then resulted <strong>in</strong> the redistribution of Be. The Se dopant profile rema<strong>in</strong>ed essentially<br />

unchanged. The atomic profile Be could be related to the microscopic defect<br />

distributions. A change <strong>in</strong> the photolum<strong>in</strong>escence spectrum, due to over-compensation of<br />

the n-doped <strong>GaAs</strong> <strong>and</strong> GaAlAs layers, was observed <strong>and</strong> the correspond<strong>in</strong>g signals which<br />

were associated with Be were identified.<br />

T.Humer-Hager, R.Treichler, P.Wurz<strong>in</strong>ger, H.Tews, P.Zwicknagl: Journal of Applied<br />

Physics, 1989, 66[1], 181-6<br />

[446-74-026]<br />

C<br />

<strong>GaAs</strong>: C <strong>Diffusion</strong><br />

The effects of background dop<strong>in</strong>g, surface encapsulation, <strong>and</strong> an As 4 over-pressure upon<br />

C diffusion were studied by anneal<strong>in</strong>g samples which had 100nm p-type C dop<strong>in</strong>g spikes<br />

with<strong>in</strong> 0.00lmm layers of undoped (n - ), Se-doped (n + ) <strong>and</strong> Mg-doped (p + ) material. The<br />

layers were grown via low-pressure metalorganic chemical vapor deposition, us<strong>in</strong>g CCl 4<br />

as the dopant source. Two different As 4 over-pressure conditions were <strong>in</strong>vestigated.<br />

These were those of an equilibrium P As (no excess As), <strong>and</strong> of a pressure, P As , of about<br />

2.5atm. For each As 4 over-pressure condition, both capless <strong>and</strong> Si 3 N 4 -capped samples of<br />

the n - , n + <strong>and</strong> p + crystals were simultaneously annealed (825C, 24h). Secondary-ion mass<br />

spectroscopy was used to measure atomic C depth profiles. The C diffusion coefficient<br />

was always low, but depended upon the background dop<strong>in</strong>g. It was highest <strong>in</strong> Mg-doped<br />

(p + ) material <strong>and</strong> was lowest <strong>in</strong> Se-doped (n + ) material. The effect of Si 3 N 4 surface<br />

encapsulation <strong>and</strong> P As upon C diffusion was m<strong>in</strong>imal.<br />

B.T.Cunn<strong>in</strong>gham, L.J.Guido, J.E.Baker, J.S.Major, N.Holonyak, G.E.Stillman: Applied<br />

Physics Letters, 1989, 55[7], 687-9<br />

[446-70/71-106]<br />

<strong>GaAs</strong>: C <strong>Diffusion</strong><br />

Atomic <strong>and</strong> carrier concentration profiles <strong>in</strong> C-implanted material were measured. The<br />

300keV C-ion implantation was carried out to a dose of 1.0 x 10 14 /cm 2 . The C<br />

concentration profiles which were revealed by secondary ion mass spectrometric<br />

measurements were found to be <strong>in</strong> good agreement with profiles which were predicted by<br />

288


C <strong>GaAs</strong> Cd<br />

Monte Carlo simulations. The implanted C did not diffuse greatly dur<strong>in</strong>g anneal<strong>in</strong>g at<br />

900C because the diffusion coefficient was less than 4 x 10 -16 cm 2 /s for ion-implanted C.<br />

Therefore, a shallow carrier concentration profile was found after anneal<strong>in</strong>g. The<br />

activation efficiency was equal to 17% at the surface (with a depth of less than 0.47µ).<br />

However, the efficiency was as low as 4% <strong>in</strong> deeper regions. This was attributed to the<br />

suppression of activation by the precipitation of C after anneal<strong>in</strong>g.<br />

T.Hara, S.Takeda, A.Mochizuki, H.Oikawa, A.Higashisaka, H.Kohzu: Japanese Journal<br />

of Applied Physics, 1995, 34[2-8B], L1020-3<br />

[446-125/126-120]<br />

<strong>GaAs</strong>: C <strong>Diffusion</strong><br />

Heavily-doped polycrystall<strong>in</strong>e material which had been grown by molecular beam epitaxy<br />

<strong>and</strong> metalorganic chemical vapor deposition was studied by us<strong>in</strong>g secondary ion mass<br />

spectrometry <strong>and</strong> Hall measurements. It was found that C diffusion was negligible, even<br />

dur<strong>in</strong>g post-growth anneal<strong>in</strong>g at 800C. However, anneal<strong>in</strong>g <strong>in</strong>creased the resistivity of C-<br />

doped <strong>GaAs</strong>, <strong>and</strong> this was suggested to be due to a change <strong>in</strong> the occupation site<br />

preference of C atoms from As sites.<br />

K.Mochizuki, T.Nakamura: Applied Physics Letters, 1994, 65[16], 2066-8<br />

[446-119/120-191]<br />

<strong>GaAs</strong>: C <strong>Diffusion</strong><br />

First-pr<strong>in</strong>ciples estimates were made of the dop<strong>in</strong>g efficiency <strong>and</strong> diffusion mechanism of<br />

C. The C acceptor which occupied an As site was found to be the most stable, <strong>and</strong> was<br />

responsible for a high dop<strong>in</strong>g efficiency. However, the hole concentration saturated at<br />

about 10 20 /cm 3 , due to compensation by donors such as [100] split <strong>in</strong>terstitial (CC) [100]<br />

complexes. A mechanism was proposed, for C diffusion that was accompanied by the<br />

formation <strong>and</strong> dissociation of the (CC) [100] complex, <strong>in</strong> which the activation energy was<br />

lower than that for atom diffusion.<br />

B.H.Cheong, K.J.Chang: Physical Review B, 1994, 49[24], 17436-9<br />

Cd<br />

[446-115/116-116]<br />

<strong>GaAs</strong>: Cd <strong>Diffusion</strong><br />

Low p-type surface concentrations were <strong>in</strong>troduced at high temperatures by us<strong>in</strong>g a Ga-<br />

Cd alloy as a diffusion source. Concentration profiles were determ<strong>in</strong>ed by us<strong>in</strong>g<br />

electrochemical profil<strong>in</strong>g techniques. The resultant profiles were of erfc-type. It was<br />

found that the surface concentration of the carriers was reduced if diffusion was carried<br />

out by us<strong>in</strong>g Ga-Cd alloys which conta<strong>in</strong>ed less than 1at%Cd. Results which could be<br />

described by:<br />

D(cm 2 /s) = 1.10 x 10 -13 exp[-2.12(eV)/kT]<br />

were found when pure Cd was used as a diffusion source, together with a 5nm SiO 2 overlayer.<br />

When Ga-1at%Cd alloy was used as a source, at temperatures of 800 <strong>and</strong> 850C, the<br />

289


Cd <strong>GaAs</strong> Co<br />

diffusivities were 8.2 x 10 -15 <strong>and</strong> 2.54 x 10 -14 cm 2 /s, respectively; thus suggest<strong>in</strong>g that the<br />

diffusivity could be described by:<br />

D(cm 2 /s) = 1.29 x 10 -14 exp[-2.17(eV)/kT]<br />

The surface concentration could be further reduced to 10 17 /cm 3 if an 0.1at%Cd alloy was<br />

used. In this case, the diffusion coefficient at 850C was 1.45 x 10 -14 cm 2 /s.<br />

D.K.Gautam, Y.Nakano, K.Tada: Japanese Journal of Applied Physics, 1991, 30[6],<br />

1176-80<br />

[446-84/85-013]<br />

<strong>GaAs</strong>: Cd <strong>Diffusion</strong><br />

The removal of damage after heavy implantation with 111m Cd <strong>and</strong> 111 In was <strong>in</strong>vestigated<br />

by us<strong>in</strong>g perturbed angular correlation <strong>and</strong> Hall techniques. After implantation at 90K,<br />

<strong>and</strong> subsequent anneal<strong>in</strong>g, the removal of structural disorder <strong>in</strong> the vic<strong>in</strong>ity of the 111 In<br />

probe atom was observed at about 300K. The anneal<strong>in</strong>g behavior, at temperatures rang<strong>in</strong>g<br />

from 500 to 1100K, of <strong>GaAs</strong> which had been implanted with 111m Cd <strong>and</strong> 111 In was<br />

<strong>in</strong>vestigated as a function of total implantation dose. After anneal<strong>in</strong>g at 600K, some of the<br />

Cd probe atoms were located <strong>in</strong> a slightly perturbed environment while the rema<strong>in</strong>der<br />

were <strong>in</strong> a heavily perturbed one. Anneal<strong>in</strong>g at temperatures above 900K led to the outdiffusion<br />

of Cd which was located <strong>in</strong> heavily perturbed sites, <strong>and</strong> electrical activation<br />

occurred. In contrast to Cd, all of the In probe atoms were located <strong>in</strong> a slightly perturbed<br />

environment <strong>and</strong> no In was lost by out-diffusion. These differences were expla<strong>in</strong>ed <strong>in</strong><br />

terms of extended defects <strong>and</strong> their <strong>in</strong>teractions with probe atoms.<br />

W.Pfeiffer, M.Deicher, R.Kalish, R.Keller, R.Magerle, N.Moriya, P.Pross, H.Skudlik,<br />

T.Wichert, H.Wolf: Materials Science Forum, 1992, 83-87, 1481-6<br />

[446-99/100-063]<br />

Co<br />

<strong>GaAs</strong>: Co <strong>Diffusion</strong><br />

Interfacial reactions between th<strong>in</strong> Co films <strong>and</strong> monocrystall<strong>in</strong>e <strong>GaAs</strong> substrates were<br />

studied by us<strong>in</strong>g Auger electron spectroscopic, transmission electron microscopic, <strong>and</strong> X-<br />

ray diffraction methods. The <strong>in</strong>teraction began at about 325C, with the formation of a<br />

ternary phase (probably Co 2 <strong>GaAs</strong>) which grew <strong>in</strong> a highly oriented manner with respect<br />

to the (001) substrate; with a lattice mismatch of about -10%. The reaction k<strong>in</strong>etics were<br />

studied <strong>and</strong> were found to be diffusion-controlled; with an activation energy of 0.7eV.<br />

The Co was deduced to be the predom<strong>in</strong>ant diffus<strong>in</strong>g species. The oriented ternary phase<br />

coexisted with r<strong>and</strong>omly oriented CoGa <strong>and</strong> CoAs at temperatures of between 325 <strong>and</strong><br />

500C while, at higher temperatures, only the b<strong>in</strong>ary compounds prevailed.<br />

M.Genut, M.Eizenberg: Applied Physics Letters, 1987, 50[19], 1358-60<br />

[446-51/52-116]<br />

290


Cr <strong>GaAs</strong> Cr<br />

Cr<br />

311 <strong>GaAs</strong>: Cr <strong>Diffusion</strong><br />

The diffusion of Cr S resulted from the rapid migration of Cr i , <strong>and</strong> their subsequent<br />

change-over so as to occupy Ga sites (or vice versa); a typical substitutional-<strong>in</strong>terstitial<br />

diffusion process. It was noted that there were 2 ways <strong>in</strong> which the Cr i -Cr s change-over<br />

could occur. One <strong>in</strong>volved a kick-out mechanism <strong>in</strong> which Ga self-<strong>in</strong>terstitials took part,<br />

<strong>and</strong> the <strong>other</strong> <strong>in</strong>volved a dissociative mechanism <strong>in</strong> which Ga vacancies took part. It was<br />

observed that the Cr s <strong>in</strong>-diffusion profiles had a characteristic shape which revealed the<br />

predom<strong>in</strong>ance of a kick-out mechanism, whereas the Cr s out-diffusion profiles were erfshaped;<br />

thus reflect<strong>in</strong>g the predom<strong>in</strong>ance of the dissociative mechanism. An <strong>in</strong>tegrated<br />

substitutional-<strong>in</strong>terstitial diffusion mechanism, which took account of kick-out <strong>and</strong><br />

dissociative mechanisms, was here used to analyze Cr diffusion results. It was confirmed<br />

that the kick-out mechanism governed Cr <strong>in</strong>-diffusion, while the dissociative mechanism<br />

governed Cr out-diffusion. The parameters which were used to fit exist<strong>in</strong>g experimental<br />

results provided quantitative <strong>in</strong>formation on the Ga self-<strong>in</strong>terstitial contribution to the Ga<br />

self-diffusion coefficient. The values which were obta<strong>in</strong>ed (table 11) were consistent with<br />

those of a study of Zn diffusion <strong>in</strong> <strong>GaAs</strong>, <strong>and</strong> with available experimental data on Al-Ga<br />

<strong>in</strong>terdiffusion coefficients.<br />

S.Yu, T.Y.Tan, U.Gösele: Journal of Applied Physics, 1991, 70[9], 4827-36<br />

[446-93/94-008]<br />

Table 11<br />

Diffusivity of Cr <strong>in</strong> <strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

990 8.1 x 10 -18<br />

890 2.6 x 10 -18<br />

790 1.1 x 10 -19<br />

<strong>GaAs</strong>: Cr <strong>Diffusion</strong><br />

It was po<strong>in</strong>ted out that one of the potential advantages of rapid thermal anneal<strong>in</strong>g, as<br />

compared with conventional furnace anneal<strong>in</strong>g, was a reduced implanted dopant <strong>and</strong><br />

background impurity diffusion. Here, the migration of Cr dur<strong>in</strong>g the anneal<strong>in</strong>g of Crdoped<br />

semi-<strong>in</strong>sulat<strong>in</strong>g material implanted with 100keV Si + ions to a dose of 7 x 10 12 /cm 2<br />

was measured us<strong>in</strong>g secondary ion mass spectrometry. Un-capped rapid thermal<br />

anneal<strong>in</strong>g (860 or 930C, 1 to 60s) was <strong>in</strong>vestigated <strong>and</strong> its effect was compared with that<br />

of capless furnace anneal<strong>in</strong>g (0.5h). It was deduced that, dur<strong>in</strong>g rapid thermal anneal<strong>in</strong>g,<br />

Cr migration was marked <strong>and</strong> exhibited a strong time-temperature dependence.<br />

H.Kanber, J.M.Whelan: Journal of the Electrochemical Society, 1987, 134[10], 2596-9<br />

[446-55/56-005]<br />

291


Cr <strong>GaAs</strong> Cu<br />

<strong>GaAs</strong>: Cr <strong>Diffusion</strong><br />

The migration of Cr was studied by us<strong>in</strong>g a new photolum<strong>in</strong>escence method which<br />

<strong>in</strong>volved measurement of the Cr-related lum<strong>in</strong>escence <strong>in</strong>tensity. The form of the <strong>in</strong>tensity<br />

profiles after thermal anneal<strong>in</strong>g was expla<strong>in</strong>ed <strong>in</strong> terms of the substitutional-<strong>in</strong>terstitial<br />

dissociative mechanism. It was observed that the redistributed profiles had an abnormal<br />

peak, <strong>in</strong> the near-surface region of wafers, which was annealed out by heat treatment at<br />

temperatures above 900C for 6h. This was attributed to the <strong>in</strong>-diffusion of As vacancies<br />

from the surface. The Cr-related lum<strong>in</strong>escence was studied as a function of the As<br />

pressure dur<strong>in</strong>g wafer anneal<strong>in</strong>g. The data showed that the <strong>in</strong>-depth profiles could be<br />

understood <strong>in</strong> terms of the diffusion of Cr <strong>and</strong> As vacancies. The lum<strong>in</strong>escence center<br />

was deduced to be a Cr Ga -V As complex.<br />

J.T.Hsu, T.Nish<strong>in</strong>o, Y.Hamakawa: Japanese Journal of Applied Physics, 1987, 26[5],<br />

685-9<br />

[446-61-066]<br />

Cu<br />

<strong>GaAs</strong>: Cu <strong>Diffusion</strong><br />

The diffusion of Cu <strong>in</strong> Si-doped material was studied. Photo-<strong>in</strong>duced current transient<br />

spectroscopic techniques were used to identify deep levels <strong>in</strong> Si-doped Cu-compensated<br />

material. The effect of capp<strong>in</strong>g the deposited Cu layer dur<strong>in</strong>g the diffusion of Cu <strong>in</strong> Sidoped<br />

material was studied by obta<strong>in</strong><strong>in</strong>g photo-<strong>in</strong>duced current transient spectra at<br />

various depths from the sample surface. A concentration gradient of the energy levels of<br />

Cu-associated complexes was found to exist <strong>in</strong>to the depth of the diffused sample. This<br />

was to be expected, because the formation of the V As Cu Ga V As complex, which was<br />

considered to be responsible for the Cu B level, was favored by an <strong>in</strong>crease <strong>in</strong> the<br />

concentration of V As .<br />

L.M.Thomas, V.K.Lakdawala: Defect <strong>and</strong> <strong>Diffusion</strong> Forum, 1993, 95-98, 931-6<br />

[446-95/98-931]<br />

<strong>GaAs</strong>: Cu <strong>Diffusion</strong><br />

Experiments were performed on polished plates of Te-doped material. Irradiation with 15<br />

to 150keV protons was carried out at 300K to doses of between 10 16 <strong>and</strong> 10 17 /cm 2 . It was<br />

found that the impurity profiles did not depend upon whether the diffusion source was<br />

deposited before or after irradiation. The penetration depth <strong>in</strong> samples which were<br />

irradiated with 15keV protons was greater than that <strong>in</strong> samples which were irradiated with<br />

150keV protons. It was suggested that this was because the low-energy ions generated<br />

more defects at depths of between 50 <strong>and</strong> 100nm.<br />

V.N.Abrosimova, V.V.Kozlovskii, N.N.Korobkov, V.N.Lomasov: Izvestiya Akademii<br />

Nauk SSSR - Neorganicheskie Materialy, 1990, 26[3], 488-91. (Inorganic Materials,<br />

1990, 26[3], 411-4)<br />

[446-84/85-013]<br />

292


Cu <strong>GaAs</strong> D<br />

<strong>GaAs</strong>: Cu <strong>Diffusion</strong><br />

It was recalled that a Cu-related peak at about 1.35eV was generally observed <strong>in</strong> the lowtemperature<br />

photolum<strong>in</strong>escence spectra of epitaxial layers. Samples were treated <strong>in</strong> Cusaturated<br />

aqueous KOH solutions at room temperature, <strong>and</strong> it was shown that Cu could be<br />

deposited onto the semiconductor surface when a source of metallic Cu was present <strong>in</strong> the<br />

KOH solution. The results suggested that Cu could diffuse <strong>in</strong>to the semiconductor, even<br />

at room temperature.<br />

K.Somogyi, D.N.Korbutyak, L.N.Lashkevich, I.Pozsgai: Physica Status Solidi A, 1989,<br />

114[2], 635-42<br />

[446-74-009]<br />

<strong>GaAs</strong>: Cu <strong>Diffusion</strong><br />

The diffusion of Cu <strong>in</strong> semi-<strong>in</strong>sulat<strong>in</strong>g liquid-encapsulated Czochralski-type material at<br />

800C was studied by us<strong>in</strong>g photolum<strong>in</strong>escence, photo-etch<strong>in</strong>g, secondary ion mass<br />

spectroscopic, <strong>and</strong> temperature-dependent Hall techniques. The results <strong>in</strong>dicated a<br />

diffusivity of 4.5 x 10 -6 cm 2 /s. It was deduced that diffusion occurred ma<strong>in</strong>ly by<br />

substitution on lattice sites. The Cu migrated preferentially along the walls of the<br />

dislocation cells.<br />

S.Griehl, M.Herms, J.Klöber, J.R.Niklas, W.Siegel: Applied Physics Letters, 1996,<br />

69[12], 1767-9<br />

[446-138/139-076]<br />

<strong>GaAs</strong>: Cu <strong>Diffusion</strong><br />

The diffusion of Cu <strong>in</strong>to undoped material reduced the concentration of the EL6 <strong>and</strong> EL2<br />

deep-donors <strong>and</strong> created a deep-donor level at about 0.66eV, <strong>in</strong> addition to the wellknown<br />

acceptor levels at 0.15 <strong>and</strong> 0.44eV; as revealed by deep-level transient<br />

spectroscopic <strong>and</strong> temperature-dependent Hall measurements. An analysis which was<br />

based upon these observations, <strong>and</strong> a charge-balance equation, was carried out <strong>in</strong> order to<br />

underst<strong>and</strong> the compensation process at various stages. The possible identity of the deepdonor<br />

which was responsible for the semi-<strong>in</strong>sulat<strong>in</strong>g properties was considered, as was<br />

the anomalous transport behavior <strong>in</strong> the highly-compensated samples.<br />

B.H.Yang, H.P.Gislason: Materials Science Forum, 1995, 196-201, 713-8<br />

[446-127/128-118]<br />

D<br />

<strong>GaAs</strong>: D <strong>Diffusion</strong><br />

The diffusion of D <strong>in</strong> Si-doped <strong>GaAs</strong> was studied. It was found that the diffusion profile<br />

could be closely fitted by us<strong>in</strong>g an erfc function. It was suggested that, <strong>in</strong> Si-doped<br />

samples, the D behaved like a deep acceptor with a level, H -/0 , which was slightly<br />

resonant <strong>in</strong> the conduction b<strong>and</strong>.<br />

293


D <strong>GaAs</strong> D<br />

J.Chevallier, B.Machayekhi, C.M.Grattepa<strong>in</strong>, R.Rahbi, B.Theys: Physical Review B,<br />

1992, 45[15], 8803-6<br />

[446-86/87-001]<br />

<strong>GaAs</strong>: D <strong>Diffusion</strong><br />

The diffusion depth <strong>and</strong> total amount of D which was <strong>in</strong>corporated dur<strong>in</strong>g exposure to a<br />

plasma was found to depend markedly upon the conductivity type of the surface. Thus, a<br />

shallow n + layer <strong>in</strong>hibited D <strong>in</strong>-diffusion <strong>in</strong> a manner which was consistent with the<br />

suggestion that the species had a level <strong>in</strong> the upper half of the <strong>GaAs</strong> b<strong>and</strong>-gap. The deactivation<br />

of donors <strong>and</strong> acceptors by D was then the result of various chemical reactions<br />

which were based upon its differ<strong>in</strong>g charge state <strong>in</strong> n-type <strong>and</strong> p-type material. Thus, Zn<br />

acceptors exhibited a re-activation energy of 1.6eV. This was less than the typical donor<br />

value (2.1eV).<br />

S.J.<strong>Pearton</strong>, W.C.Dautremont-Smith, J.Lopata, C.W.Tu, C.R.Abernathy: Physical Review<br />

B, 1987, 36[8], 4260-4<br />

[446-55/56-006]<br />

<strong>GaAs</strong>: D <strong>Diffusion</strong><br />

The dynamics of D <strong>in</strong> Zn-doped material were <strong>in</strong>vestigated by us<strong>in</strong>g anelastic relaxation<br />

techniques. It was suggested that the most likely configuration was for D to be trapped by<br />

substitutional Zn, although it was also possible that D was trapped at a Ga vacancy.<br />

Relaxation of D occurred at about 20K <strong>in</strong> the kHz range, <strong>and</strong> had the highest rate which<br />

had yet been found for a H isotope <strong>in</strong> a semiconductor. The shape of the curves of elastic<br />

energy loss versus temperature <strong>in</strong>dicated that the D reorientation was strongly quantized.<br />

G.Cannelli, R.Cantelli, F.Cordero, E.Giov<strong>in</strong>e, F.Trequattr<strong>in</strong>i, M.Capizzi, A.Frova: Solid<br />

State Communications, 1996, 98[10], 873-7<br />

[446-136/137-109]<br />

<strong>GaAs</strong>: D <strong>Diffusion</strong><br />

The D diffusion profiles <strong>in</strong> material which was doped with various group-II (Mg, Zn, Cd)<br />

or group-IV (C, Ge) acceptors had similar characteristics; even though the neutralization<br />

of acceptors at 300K was not always efficient. Conductivity <strong>and</strong> Hall-effect<br />

measurements were used to study the electrical characteristics of hydrogenated p-type<br />

epilayers. The temperature dependences of the free carrier concentration <strong>and</strong> hole<br />

mobility, before <strong>and</strong> after hydrogenation, showed that the neutralization of acceptors by<br />

atomic H led to the elim<strong>in</strong>ation of the shallow acceptor states. Infra-red absorption l<strong>in</strong>es<br />

that were associated with H-acceptor complexes were observed for all of the acceptors,<br />

except Mg. It was established that the microscopic structure of H-acceptor complexes<br />

depended upon the acceptor site <strong>in</strong> the lattice.<br />

R.Rahbi, B.Pajot, J.Chevallier, A.Marbeuf, R.C.Logan, M.Gav<strong>and</strong>: Journal of Applied<br />

Physics, 1993, 73[4], 1723-31<br />

[446-106/107-034]<br />

294


D <strong>GaAs</strong> Fe<br />

<strong>GaAs</strong>: D <strong>Diffusion</strong><br />

It was recalled that H diffusion <strong>in</strong> <strong>III</strong>-V semiconductors usually led to a reduction <strong>in</strong> the<br />

active dopant concentration, <strong>and</strong> to an <strong>in</strong>crease <strong>in</strong> the free carrier mobility. It was<br />

considered that this neutralization of the dopants was a result of the formation of a<br />

complex which <strong>in</strong>cluded H <strong>and</strong> the dopant atom. The microscopic structure was deduced<br />

from a detailed analysis of the <strong>in</strong>fra-red local vibrational modes of H <strong>and</strong> the dopants.<br />

Modell<strong>in</strong>g of D diffusion profiles Zn-doped material <strong>in</strong>dicated the presence of a H donor<br />

level at 1.1eV below the conduction b<strong>and</strong> m<strong>in</strong>imum. The D diffusion behavior <strong>in</strong> Sidoped<br />

<strong>GaAs</strong> or Si-doped Al x Ga 1-x As <strong>in</strong>dicated that the acceptor level of H was slightly<br />

resonant <strong>in</strong> the <strong>GaAs</strong> conduction b<strong>and</strong>, <strong>and</strong> became localized <strong>in</strong> the b<strong>and</strong> gap of alloys<br />

when x was greater than 0.08.<br />

J.Chevallier, B.Pajot: Materials Science Forum, 1992, 83-87, 539-50<br />

Fe<br />

[446-99/100-064]<br />

312 <strong>GaAs</strong>: Fe <strong>Diffusion</strong><br />

The Fe was diffused from a spun-on glass film, <strong>and</strong> onto n-type wafers, at temperatures<br />

of between 700 <strong>and</strong> 900C (table 12). The diffusivities, as determ<strong>in</strong>ed by us<strong>in</strong>g junctiondepth<br />

<strong>and</strong> conductivity techniques, could be expla<strong>in</strong>ed <strong>in</strong> terms of a model which<br />

assumed the existence of exhaustible diffusion sources. It was found that the diffusivity<br />

was described by:<br />

D(cm 2 /s) = 1000 exp[-2.7(eV)/kT]<br />

with<strong>in</strong> the above temperature range.<br />

J.Ohsawa, H.Kak<strong>in</strong>oki, H.Ikeda, M.Migitaka: Journal of the Electrochemical Society,<br />

1990, 137[8], 2608-11<br />

[446-76/77-008]<br />

Table 12<br />

Diffusivity of Fe <strong>in</strong> <strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

905 2.2 x 10 -9<br />

805 1.4 x 10 -10<br />

705 1.2 x 10 -11<br />

<strong>GaAs</strong>: Fe <strong>Diffusion</strong><br />

Secondary ion mass spectroscopic results revealed that the use of spun-on film diffusion<br />

could produce very flat Fe profiles whose concentrations of 10 15 to 10 17 /cm 3 were<br />

consistent with the solubility of Fe at diffusion temperatures of 650 to 900C. The surface<br />

accumulation region was far smaller than that which was produced by the use of<br />

295


Fe <strong>GaAs</strong> Ga<br />

conventional techniques. It was suggested that there was a relationship between the<br />

solubility <strong>and</strong> the equilibrium concentration of Ga antisite defects.<br />

J.Ohsawa, M.Nakamura, Y.Nekado, M.Migitaka, N.Tsuchida: Japanese Journal of<br />

Applied Physics, 1995, 34[2-5B], L600-2<br />

[446-123/124-162]<br />

1.0E-09<br />

1.0E-10<br />

1.0E-11<br />

1.0E-12<br />

D (cm 2 /s)<br />

1.0E-13<br />

1.0E-14<br />

1.0E-15<br />

1.0E-16<br />

1.0E-17<br />

table 13<br />

table 14<br />

table 15<br />

1.0E-18<br />

1.0E-19<br />

7 8 9 10 11 12<br />

10 4 /T(K)<br />

Figure 4: Diffusivity of Ga <strong>in</strong> <strong>GaAs</strong><br />

Ga<br />

<strong>GaAs</strong>: Fe <strong>Diffusion</strong><br />

The structural properties of samples which had been implanted with 150 or 400keV Fe, to<br />

doses of between 10 12 <strong>and</strong> 10 15 /cm 2 , were studied. The depth distributions of the implants<br />

were compared before <strong>and</strong> after anneal<strong>in</strong>g with, or without, a Si 3 N 4 cap. Rutherford backscatter<strong>in</strong>g,<br />

X-ray double-crystal diffractometry, <strong>and</strong> secondary ion mass spectroscopy<br />

results <strong>in</strong>dicated that Fe was markedly redistributed <strong>in</strong> all of the materials dur<strong>in</strong>g<br />

anneal<strong>in</strong>g. On the <strong>other</strong> h<strong>and</strong>, Ti did not redistribute at all. The driv<strong>in</strong>g force for the<br />

redistribution of Fe was thought to be not classical diffusion, but reaction with<br />

implantation-<strong>in</strong>duced defects <strong>and</strong> stoichiometric imbalances. The defect chemistry of as-<br />

296


Ga <strong>GaAs</strong> Ga<br />

implanted arsenides was found to be fundamentally different to that of as-implanted<br />

phosphides s<strong>in</strong>ce, <strong>in</strong> the latter case, the mass ratio of the constituents was much greater<br />

<strong>and</strong> the specific energy for amorphization was much lower.<br />

H.Ullrich, A.Knecht, D.Bimberg, H.Kräutle, W.Schlaak: Journal of Applied Physics,<br />

1992, 72[8], 3514-21<br />

[446-106/107-034]<br />

<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

A review was presented of progress <strong>in</strong> the underst<strong>and</strong><strong>in</strong>g of the mechanisms of Ga selfdiffusion<br />

<strong>and</strong> impurity diffusion <strong>in</strong> <strong>GaAs</strong>, <strong>and</strong> of the disorder<strong>in</strong>g of <strong>GaAs</strong>/Al<strong>GaAs</strong><br />

superlattices. The self-diffusion of Ga, <strong>and</strong> Al-Ga <strong>in</strong>terdiffusion, under <strong>in</strong>tr<strong>in</strong>sic <strong>and</strong> n-<br />

dop<strong>in</strong>g conditions were governed by triply negatively charged group-<strong>III</strong> sub-lattice<br />

vacancies (V Ga 3- ) while, under heavy p-dop<strong>in</strong>g conditions, they were probably governed<br />

by the doubly positively charged self-<strong>in</strong>terstitial, I Ga 2+ . The <strong>GaAs</strong>/Al<strong>GaAs</strong> superlattice<br />

disorder<strong>in</strong>g enhancement which was observed under n-dop<strong>in</strong>g by Si or Te was attributed<br />

to the Fermi-level effect, which <strong>in</strong>creased the V Ga 3- concentration. An elusive disorder<strong>in</strong>g<br />

enhancement under p-dop<strong>in</strong>g by Zn or Be was attributed to the comb<strong>in</strong>ed effects of the<br />

Fermi level, which <strong>in</strong>creased the I Ga 2+ concentration, <strong>and</strong> to dopant <strong>in</strong>-diffusion or outdiffusion<br />

<strong>in</strong>duced I Ga 2+ supersaturation or undersaturation, respectively. In parallel with<br />

the Ga self-diffusion mechanism <strong>in</strong> <strong>GaAs</strong>, diffusion of the Si donor atoms which<br />

occupied Ga sites was also governed ma<strong>in</strong>ly by V Ga 3- . Meanwhile, Si acceptor atoms<br />

which occupied As sites (a small fraction of the total) diffused via a negatively charged<br />

As sub-lattice po<strong>in</strong>t-defect species. The <strong>in</strong>terstitial-substitutional p-type dopants, Zn <strong>and</strong><br />

Be, diffused via the kick-out mechanism. Their diffusion <strong>in</strong>duced I Ga 2+ supersaturation<br />

<strong>and</strong> undersaturation, respectively, under <strong>in</strong>-diffusion <strong>and</strong> out-diffusion conditions.<br />

T.Y.Tan, U.Gösele, S.Yu: Critical Reviews <strong>in</strong> Solid State <strong>and</strong> Materials Science, 1991,<br />

17[1], 47-106<br />

[446-157/58-293]<br />

<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

Self-diffusion was <strong>in</strong>vestigated by apply<strong>in</strong>g secondary ion mass spectroscopic techniques<br />

to heterostructures which consisted of molecular beam epitaxial layers that conta<strong>in</strong>ed one<br />

or more stable isotopes of host crystal elements. Intermix<strong>in</strong>g of the stable isotopes<br />

between epilayers of various isotopic compositions constituted near-ideal self-diffusion<br />

conditions; free from the complications of impurities, stra<strong>in</strong>, electric fields <strong>and</strong> surfaces.<br />

When diffusion was <strong>in</strong>vestigated by us<strong>in</strong>g 69 <strong>GaAs</strong>/ 71 <strong>GaAs</strong> isotope heterostructures, the<br />

Ga self-diffusion coefficient <strong>in</strong> <strong>in</strong>tr<strong>in</strong>sic <strong>GaAs</strong> under As-rich ambient could be described<br />

by:<br />

D (cm 2 /s) = 43 exp[-4.24(eV)/kT]<br />

over 6 orders of magnitude at temperatures between 800 <strong>and</strong> 1225C. This suggested that a<br />

s<strong>in</strong>gle defect mechanism controlled the process.<br />

E.E.Haller, L.Wang: Defect <strong>and</strong> <strong>Diffusion</strong> Forum, 1997, 143-147, 1067-78<br />

[446-143/147-1067]<br />

297


Ga <strong>GaAs</strong> Ga<br />

<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

The diffusion of Cd <strong>in</strong>to <strong>GaAs</strong> s<strong>in</strong>gle crystals was <strong>in</strong>vestigated at temperatures rang<strong>in</strong>g<br />

from 804 to 1201C. The penetration profiles which were measured by us<strong>in</strong>g secondary<br />

ion mass spectroscopy <strong>and</strong> spread<strong>in</strong>g-resistance profil<strong>in</strong>g techniques were modelled<br />

numerically on the basis of the kick-out diffusion mechanism. This permitted estimates to<br />

be made of the Ga self-diffusivity, as mediated by doubly positively charged Ga self<strong>in</strong>terstitials,<br />

I Ga 2+ . The Ga self-diffusivities which were found for As-rich <strong>and</strong> As-poor<br />

ambients were consistent. Under an As vapor pressure of 1atm, <strong>and</strong> under electronically<br />

<strong>in</strong>tr<strong>in</strong>sic conditions, the I Ga 2+ -mediated Ga self-diffusion data could be described by:<br />

D (cm 2 /s) = 3.5 x 10 4 exp[-5.74(eV)/kT]<br />

G.Bösker, N.A.Stolwijk, U.Södervall, W.Jäger: Defect <strong>and</strong> <strong>Diffusion</strong> Forum, 1997, 143-<br />

147, 1109-16<br />

[446-143/147-1109]<br />

<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

The effect of triply negatively charged Ga vacancies (V Ga 3- ) <strong>and</strong> doubly positively<br />

charged Ga self-<strong>in</strong>terstitials (I Ga 2+ ) upon the self-diffusivity of Ga was studied. Under<br />

thermal equilibrium <strong>and</strong> <strong>in</strong>tr<strong>in</strong>sic conditions, the contribution of V Ga 3- was characterized<br />

by an activation enthalpy of 6eV <strong>in</strong> As-rich crystals <strong>and</strong> of 7.52eV <strong>in</strong> Ga-rich crystals.<br />

The contribution of I Ga 2+ was characterized by an activation enthalpy of 4.89eV <strong>in</strong> the<br />

case of As-rich crystals, <strong>and</strong> of 3.37eV <strong>in</strong> the case of Ga-rich crystals.<br />

T.Y.Tan, S.Yu, U.Gösele: Journal of Applied Physics, 1991, 70[9], 4823-6<br />

[446-91/92-007]<br />

<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

A review of previous data <strong>in</strong>dicated that the self-diffusion of Ga <strong>in</strong>volved either the<br />

vacancy or the <strong>in</strong>terstitialcy mechanism. A switch from one to the <strong>other</strong> could occur, due<br />

to variations <strong>in</strong> the Fermi level or the As pressure. The self-diffusion appeared to be<br />

associated with Ga vacancies with a charge of -3, <strong>in</strong> the case of n-type material, or with<br />

Ga <strong>in</strong>terstitials with a charge of 2 <strong>in</strong> the case of p-type material. In the case of <strong>in</strong>tr<strong>in</strong>sic<br />

material, a commonly observed V-form of the Ga diffusivity versus As pressure plot was<br />

suggested to be consistent with the 2 diffusion mechanisms. It was noted that it was a<br />

simple matter to predict the near-equilibrium Ga diffusivity on the basis of a s<strong>in</strong>gle<br />

<strong>in</strong>tr<strong>in</strong>sic Ga diffusion coefficient. A model was developed which l<strong>in</strong>ked the changeover <strong>in</strong><br />

Ga diffusivity <strong>and</strong> electron concentration at high donor concentrations.<br />

R.M.Cohen: Journal of Electronic Materials, 1991, 20[6], 425-30<br />

[446-88/89-014]<br />

<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

Calculations were made of the absolute formation energies of native defects <strong>in</strong> this<br />

compound. It was found that the formation energy, <strong>and</strong> thus the equilibrium concentration<br />

of the defects, depended strongly upon the atomic chemical potentials of As <strong>and</strong> Ga, as<br />

298


Ga <strong>GaAs</strong> Ga<br />

well as upon the electron chemical potential. Hence, the Ga vacancy concentration<br />

changed by more than 10 orders of magnitude as the chemical potentials of As <strong>and</strong> Ga<br />

varied between the thermodynamically allowed limits. It was deduced that the rate of selfdiffusion<br />

depended markedly upon the surface anneal<strong>in</strong>g conditions.<br />

S.B.Zhang, J.E.Northrup: Physical Review Letters, 1991, 67[17], 2339-42<br />

[446-88/89-014]<br />

<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

The effects of rapid thermal process<strong>in</strong>g, upon material with various thicknesses of SiO 2<br />

encapsulant, were studied by us<strong>in</strong>g capacitance-voltage, secondary ion mass<br />

spectroscopic, <strong>and</strong> X-ray photo-electron spectroscopic methods. The process<strong>in</strong>g was<br />

carried out at 760 or 910C, for a period of 9s. The results <strong>in</strong>dicated that a decrease <strong>in</strong><br />

carrier concentration was related to Ga out-diffusion through the SiO 2 . The decrease <strong>in</strong><br />

carrier concentration was attributed to the formation of V Ga -Si Ga complex defects (selfactivated<br />

centers). At 760C, the amount of Ga out-diffusion was larger <strong>in</strong> samples with a<br />

thick SiO 2 coat<strong>in</strong>g. At 910C, the amount of Ga out-diffusion was larger <strong>in</strong> samples with a<br />

th<strong>in</strong> SiO 2 coat<strong>in</strong>g. This behavior was expla<strong>in</strong>ed by assum<strong>in</strong>g the operation of 2 different<br />

types of driv<strong>in</strong>g force. These were <strong>in</strong>terfacial thermal stresses, <strong>and</strong> an <strong>in</strong>terfacial reaction<br />

between <strong>GaAs</strong> <strong>and</strong> SiO 2 . It was noted that <strong>in</strong>terfacial thermal stresses enhanced Ga outdiffusion<br />

at 760C, whereas <strong>in</strong>terfacial reactions enhanced such out-diffusion at 910C.<br />

M.Katayama, Y.Tokuda, Y.Inoue, A.Usami, T.Wada: Journal of Applied Physics, 1991,<br />

69[6], 3541-6<br />

[446-86/87-010]<br />

<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

The absolute formation energies of native defects were calculated. It was found that the<br />

formation energies, <strong>and</strong> thus the equilibrium concentrations of the defects, depended<br />

strongly upon the atomic chemical potentials of As <strong>and</strong> Ga as well as upon the electron<br />

chemical potential. Thus, the Ga vacancy concentration changed by more than 10 orders<br />

of magnitude as the chemical potentials of As <strong>and</strong> Ga varied over the thermodynamically<br />

possible range. It was concluded that the rate of self-diffusion depended markedly upon<br />

the surface anneal<strong>in</strong>g conditions, <strong>and</strong> that impurity-enhanced self-diffusion <strong>in</strong>volved<br />

V Ga 3- under As-rich n-type conditions, or Ga i 3 under Ga-rich p-type conditions.<br />

S.B.Zhang, J.E.Northrup: Physical Review Letters, 1991, 67[17], 2339-42<br />

[446-84/85-014]<br />

<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

Atomistic thermodynamic calculations were made of the energetics of self-diffusion. An<br />

assessment of the activation enthalpy of the saddle-po<strong>in</strong>t configuration of various modes<br />

of vacancy self-diffusion <strong>in</strong>dicated that second-nearest neighbor hopp<strong>in</strong>g was the most<br />

energetically favorable mechanism, if vacancies were available <strong>in</strong> equilibrium<br />

concentrations. An assessment of the activation entropy <strong>in</strong>dicated that normal diffusion<br />

pre-exponential factors, of the order of 10 -5 to 0.1cm 2 /s, were consistent with vacancy<br />

299


Ga <strong>GaAs</strong> Ga<br />

self-diffusion via second-nearest neighbor hopp<strong>in</strong>g. It was proposed that self-diffusion<br />

which was characterized by pre-exponential factors of the order of 10 7 to 10 8 cm 2 /s, <strong>and</strong><br />

activation energies of the order of 6eV, <strong>in</strong>volved processes <strong>in</strong> which surface vacancy<br />

generation was <strong>in</strong>hibited <strong>and</strong> self-diffusion was mediated by Frenkel pair generation.<br />

J.F.Wager: Journal of Applied Physics, 1991, 69[5], 3022-31<br />

[446-78/79-011]<br />

<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

Rutherford back-scatter<strong>in</strong>g spectrometry <strong>and</strong> X-ray photoelectron spectroscopy were used<br />

to <strong>in</strong>vestigate compositional changes, <strong>in</strong> th<strong>in</strong> metal-semiconductor systems, which were<br />

caused by Ar + <strong>and</strong> N + ion bombardment or by anneal<strong>in</strong>g. The <strong>in</strong>vestigation was carried<br />

out on contacts of Ni-Au-Ge on <strong>GaAs</strong>, as well as on irradiated <strong>and</strong> non-irradiated Au-<br />

<strong>GaAs</strong>. The structures were bombarded with Ar + <strong>and</strong> N + ions to doses of between 10 14 <strong>and</strong><br />

3 x 10 16 /cm 2 . The experimental results <strong>in</strong>dicated that the <strong>in</strong>terdiffusion of Au <strong>and</strong> Ga<br />

atoms depended upon the bombardment dose. Upon anneal<strong>in</strong>g the samples, block<strong>in</strong>g of<br />

<strong>in</strong>terdiffusion was observed <strong>in</strong> Au-<strong>GaAs</strong> structures (which had been deposited on 50keV<br />

Ar + ion-irradiated <strong>and</strong> pre-treated <strong>GaAs</strong>) at certa<strong>in</strong> radiation-defect concentrations. This<br />

behavior was attributed to Au-Ga bond formation, <strong>and</strong> appeared to depend upon Ga<br />

<strong>in</strong>terstitial atoms.<br />

L.B.Guoba, A.A.Vitkauskas, J.V.Kameneckas, V.R.Sargünas, A.P.Sakalas: Physica<br />

Status Solidi A, 1989, 111[2], 507-13<br />

[446-64/65-162]<br />

<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

A review was presented of self-diffusion mechanisms <strong>and</strong> dop<strong>in</strong>g-enhanced superlattice<br />

disorder<strong>in</strong>g. It was concluded that Ga diffusion <strong>in</strong>volved triply negatively charged Ga<br />

vacancies, under <strong>in</strong>tr<strong>in</strong>sic <strong>and</strong> n-doped conditions. With regard to the <strong>in</strong>fluence of p-type<br />

dopants, the Fermi level effect had to be considered; together with dopant diffusion<strong>in</strong>duced<br />

Ga self-<strong>in</strong>terstitial supersaturation or undersaturation. The self-diffusion of Ga <strong>in</strong><br />

heavily p-doped material was governed by positively charged Ga self-<strong>in</strong>terstitials. It was<br />

concluded that dislocations <strong>in</strong> this material <strong>and</strong> <strong>in</strong> <strong>other</strong> <strong>III</strong>-V compounds were only<br />

moderately efficient s<strong>in</strong>ks or sources for po<strong>in</strong>t defects.<br />

T.Y.Tan, U.Gösele: Materials Science <strong>and</strong> Eng<strong>in</strong>eer<strong>in</strong>g, 1988, B1, 47-65<br />

[446-62/63-208]<br />

<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

New <strong>in</strong>terdiffusion data on <strong>GaAs</strong>/AlAs superlattices led to the conclusion that Ga selfdiffusion<br />

<strong>in</strong> <strong>GaAs</strong> <strong>in</strong>volved triply negatively charged Ga vacancies under <strong>in</strong>tr<strong>in</strong>sic <strong>and</strong> n-<br />

doped conditions. The mechanism of Si-enhanced superlattice disorder<strong>in</strong>g was the Fermilevel<br />

effect, which <strong>in</strong>creased the concentrations of the charged po<strong>in</strong>t defect species. With<br />

respect to the effect of Be <strong>and</strong> Zn p-type dopants, the Fermi level effect had to be<br />

considered together with dopant diffusion-<strong>in</strong>duced Ga self-<strong>in</strong>terstitial supersaturation or<br />

300


Ga <strong>GaAs</strong> Ga<br />

undersaturation. The self-diffusion of Ga under heavy p-dop<strong>in</strong>g conditions was governed<br />

by positively charged Ga self-<strong>in</strong>terstitials.<br />

T.Y.Tan, U.Gösele: Applied Physics Letters, 1988, 52[15], 1240-2<br />

[446-62/63-209]<br />

<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

The chemical reactions <strong>and</strong> Schottky-barrier characteristics of W(200nm)/Si(0 to<br />

2.5nm)/<strong>GaAs</strong> contacts when annealed at 800C were <strong>in</strong>vestigated. The Si <strong>in</strong>terfacial layers<br />

<strong>and</strong> W films were sputter-deposited onto chemically etched <strong>GaAs</strong> substrates. The<br />

W/Si/<strong>GaAs</strong> diodes clearly exhibited the same Schottky-barrier characteristics as those of<br />

(WSi 0.6 )/<strong>GaAs</strong> diodes. By us<strong>in</strong>g secondary ion mass spectrometry, the Si layer was found<br />

to suppress Ga atom diffusion from <strong>GaAs</strong> substrates <strong>in</strong>to W films dur<strong>in</strong>g anneal<strong>in</strong>g<br />

(800C, 1h). A reduction <strong>in</strong> natively oxidized <strong>GaAs</strong> surfaces was also observed <strong>in</strong> the<br />

<strong>in</strong>itial stages of Si layer deposition by X-ray photo-emission spectroscopy. These results<br />

suggested that the Si layer elim<strong>in</strong>ated native oxides from <strong>GaAs</strong> surfaces, result<strong>in</strong>g <strong>in</strong><br />

tungsten-silicide/<strong>GaAs</strong> <strong>in</strong>timate contact formation at the <strong>in</strong>terface. The Si obstructed the<br />

diffusion paths of Ga atoms at W gra<strong>in</strong> boundaries with W-Si-O ternary compounds.<br />

Y.Kuriyama, S.Ohfuji, J.Nagano: Journal of Applied Physics, 1987, 62[4], 1318-23<br />

[446-55/56-005]<br />

<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

First-pr<strong>in</strong>ciples molecular dynamics simulations were used to <strong>in</strong>vestigate the predom<strong>in</strong>ant<br />

migration mechanism of the Ga vacancy, as well as its free energy of formation <strong>and</strong> the<br />

rate constant for Ga self-diffusion. The results suggested that the vacancy migrated via<br />

second-nearest neighbor hopp<strong>in</strong>g. The calculated diffusion constant was <strong>in</strong> good<br />

agreement with the experimental value which was deduced by us<strong>in</strong>g 69 <strong>GaAs</strong>/ 71 <strong>GaAs</strong><br />

heterostructures. However, the predictions differed considerably from the results which<br />

had been obta<strong>in</strong>ed by perform<strong>in</strong>g <strong>in</strong>terdiffusion experiments on GaAlAs/<strong>GaAs</strong><br />

heterostructures.<br />

M.Bockstedte, M.Scheffler: Zeitschrift für Physikalische Chemie, 1997, 200[1-2], 195-<br />

207<br />

[446-157/159-301]<br />

313 <strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

The diffusion of implanted Zn was studied, at anneal<strong>in</strong>g temperatures of between 625 <strong>and</strong><br />

850C, by means of secondary ion mass spectrometry. A substitutional-<strong>in</strong>terstitial<br />

diffusion mechanism was proposed <strong>in</strong> order to expla<strong>in</strong> how deviations of the local Ga<br />

<strong>in</strong>terstitial concentration, from its equilibrium value, regulated Zn diffusion. It was found<br />

that it was possible to simulate both box-shaped profiles, that resulted from hightemperature<br />

anneal<strong>in</strong>g, <strong>and</strong> k<strong>in</strong>k-<strong>and</strong>-tail profiles which resulted from lower-temperature<br />

anneal<strong>in</strong>g. The simulation data permitted the determ<strong>in</strong>ation of Arrhenius relationships.<br />

The equilibrium Ga <strong>in</strong>terstitial concentration was described by:<br />

C (/cm 3 ) = 7.98 x 10 30 exp[-3.47(eV)/kT]<br />

while the Ga <strong>in</strong>terstitial diffusion coefficient (table 13) was described by:<br />

D (cm 2 /s) = 0.4384 exp[-2.14(eV)/kT]<br />

301


Ga <strong>GaAs</strong> Ga<br />

M.P.Chase, M.D.Deal, J.D.Plummer: Journal of Applied Physics, 1997, 81[4], 1670-6<br />

[446-148/149-172]<br />

Table 13<br />

Diffusivity of Ga Self-Interstitials <strong>in</strong> <strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

850 9.8 x 10 -11<br />

800 3.9 x 10 -11<br />

750 2.2 x 10 -11<br />

675 1.8 x 10 -12<br />

<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

Scann<strong>in</strong>g electron microscopy was used to observe the √19 x √19 <strong>and</strong> (1 x 1) HT<br />

reconstructions, <strong>and</strong> their transitions, on (111)B vic<strong>in</strong>al surfaces under an As pressure.<br />

These reconstructions were observed <strong>in</strong> dark <strong>and</strong> bright contrast, respectively. Dur<strong>in</strong>g the<br />

transition, √19 x √19 doma<strong>in</strong>s began to develop from macro-step edges onto lower (1 x<br />

1) HT reconstructed terraces, while (1 x 1) HT doma<strong>in</strong>s began to develop from the macrostep<br />

edges onto the upper √19 x √19 reconstructed terraces. The transition diagram for the<br />

surface coverage of doma<strong>in</strong>s exhibited hysteresis. S<strong>in</strong>ce Ga diffusion, As <strong>in</strong>corporation or<br />

re-evaporation were enhanced dur<strong>in</strong>g the transitions, marked step-bunch<strong>in</strong>g with rough<br />

macro-step edges was observed.<br />

H.W.Ren, M.Tanaka, T.Nish<strong>in</strong>aga: Applied Physics Letters, 1996, 69[4], 565-7<br />

[446-136/137-109]<br />

314 <strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

Isotopically controlled heterostructures were used to study Ga self-diffusion by us<strong>in</strong>g<br />

secondary-ion mass spectrometry. This approach produced a near-ideal r<strong>and</strong>om walk<br />

situation that was free of perturbations aris<strong>in</strong>g from electric fields, mechanical stresses, or<br />

chemical potentials. It was found that the Ga self-diffusion coefficient <strong>in</strong> <strong>in</strong>tr<strong>in</strong>sic<br />

material (table 14) could be described by:<br />

D(cm 2 /s) = 43 exp[-4.24(eV)/kT]<br />

over 6 orders of magnitude, at temperatures of between 800 <strong>and</strong> 1225C, under As-rich<br />

conditions. No significant dop<strong>in</strong>g effects were observed <strong>in</strong> samples with substrates that<br />

were doped with Te up to 4 x 10 17 /cm 3 or with Zn up to 10 19 /cm 3 .<br />

L.Wang, L.Hsu, E.E.Haller, J.W.Erickson, A.Fischer, K.Eberl, M.Cardona: Physical<br />

Review Letters, 1996, 76[13], 2342-5<br />

[446-134/135-125]<br />

<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

Impurity <strong>and</strong> self-diffusion mechanisms, <strong>and</strong> the nature of the associated po<strong>in</strong>t defects,<br />

were considered with regard to the Ga sub-lattice. Analyses of dop<strong>in</strong>g-enhanced<br />

AlAs/<strong>GaAs</strong> superlattice disorder<strong>in</strong>g data <strong>and</strong> impurity diffusion data led to the conclusion<br />

302


Ga <strong>GaAs</strong> Ga<br />

that, under thermal equilibrium <strong>and</strong> <strong>in</strong>tr<strong>in</strong>sic conditions, the triply negatively-charged Ga<br />

vacancy (V 3- Ga ) governed Ga self-diffusion <strong>and</strong> Al-Ga <strong>in</strong>terdiffusion <strong>in</strong> As-rich crystals,<br />

while the doubly positively-charged Ga self-<strong>in</strong>terstitial (I 2+ Ga ) predom<strong>in</strong>ated <strong>in</strong> Ga-rich<br />

3-<br />

2+<br />

crystals. With sufficient dop<strong>in</strong>g, V Ga predom<strong>in</strong>ated <strong>in</strong> n-type crystals, while I Ga<br />

predom<strong>in</strong>ated <strong>in</strong> p-type crystals; regardless of the composition. The V 3- Ga species also<br />

contributed to the diffusion of the ma<strong>in</strong> donor species, Si, while I 2+ Ga also governed the<br />

diffusion of the ma<strong>in</strong> acceptor species, Zn <strong>and</strong> Be, via the kick-out mechanism. The<br />

3-<br />

thermal equilibrium concentration of V Ga was found to exhibit a temperature<br />

<strong>in</strong>dependence, or even a small negative temperature dependence. That is, when the<br />

temperature was lowered, the equilibrium concentration of V 3- Ga was either unchanged or<br />

slightly <strong>in</strong>creased. This behavior was consistent with many experimental results.<br />

T.Y.Tan: Materials Chemistry <strong>and</strong> Physics, 1995, 40[4], 245-52<br />

Table 14<br />

Diffusivity of Ga <strong>in</strong> <strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

1230 3.3 x 10 -13<br />

1125 2.4 x 10 -14<br />

1095 1.1 x 10 -14<br />

1075 6.5 x 10 -15<br />

1025 2.0 x 10 -15<br />

975 2.6 x 10 -16<br />

950 2.4 x 10 -16<br />

925 3.8 x 10 -17<br />

900 2.8 x 10 -17<br />

845 5.3 x 10 -18<br />

795 6.9 x 10 -19<br />

[446-125/126-121]<br />

<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

<strong>Diffusion</strong> was studied <strong>in</strong> samples of molecular beam epitaxial material with grown-<strong>in</strong> Be.<br />

The diffusion profiles of samples which had been annealed under various conditions were<br />

determ<strong>in</strong>ed by us<strong>in</strong>g secondary ion mass spectrometry, <strong>and</strong> a computer simulation was<br />

used to analyze the experimental results <strong>and</strong> extract diffusion parameters. It was deduced<br />

that the Ga <strong>in</strong>terstitial diffusivity was described by:<br />

D(cm 2 /s) = 6.4 x 10 -5 exp[-1.28(eV)/kT]<br />

while the equilibrium concentration of Ga <strong>in</strong>terstitials was described by:<br />

C(/cm 3 ) = 4.7 x 10 28 exp[-3.25(eV)/kT]<br />

J.C.Hu, M.D.Deal, J.D.Plummer: Journal of Applied Physics, 1995, 78[3], 1595-605<br />

[446-123/124-161]<br />

303


Ga <strong>GaAs</strong> Ga<br />

<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

Samples with a 100nm Co over-layer, which had been subjected to rapid thermal<br />

anneal<strong>in</strong>g (400 to 650C, 60s), were analyzed by us<strong>in</strong>g mass <strong>and</strong> energy dispersive recoil<br />

spectrometry. Separate characterizations of the C, O, Co, Ga, <strong>and</strong> As depth distributions<br />

were carried out. It was found that Ga migrated to the surface at anneal<strong>in</strong>g temperatures<br />

which were higher than 450C. In samples which were annealed at 650C, clear enrichment<br />

of Ga with<strong>in</strong> the outer 35nm was observed. The composition at various depths was<br />

determ<strong>in</strong>ed at a number of temperatures. On the basis of Arrhenius plots, the apparent<br />

activation energies were estimated to be equal to about 0.6eV for phase formation <strong>and</strong><br />

equal to 1.3eV for diffusion. The X-ray diffraction data <strong>in</strong>dicated that CoGa <strong>and</strong> CoAs<br />

were present <strong>in</strong> all of the annealed samples. Scann<strong>in</strong>g electron microscopy showed that<br />

the surface was reticulated after heat treatment, <strong>and</strong> that gra<strong>in</strong> growth occurred at higher<br />

temperatures.<br />

M.Hult, H.J.Whitlow, M.Ostl<strong>in</strong>g, M.Andersson, Y.Andersson, I.L<strong>in</strong>deberg, K.Stähl:<br />

Journal of Applied Physics, 1994, 75[2], 835-43<br />

[446-117/118-165]<br />

<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

It was recalled that Ga vacancies were believed to mediate Ga self-diffusion, <strong>and</strong> the<br />

diffusion of substitutional impurities which resided on the Ga sub-lattice. First-pr<strong>in</strong>ciples<br />

calculations were presented for the vacancy-mediated diffusion of Ga. It was shown that a<br />

DX-like mechanism facilitated the migration of lattice-site atoms <strong>in</strong>to the <strong>in</strong>terstitial<br />

region, <strong>and</strong> that the dangl<strong>in</strong>g bonds of a second-nearest neighbor vacancy assisted<br />

migration through the <strong>in</strong>terstitial region. Due to these 2 mechanisms, vacancy-assisted<br />

diffusion of Ga occurred with a low-energy barrier.<br />

J.Dabrowski, J.E.Northrup: Physical Review B, 1994, 49[20], 14286-9<br />

[446-115/116-117]<br />

<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

The thermal equilibrium concentrations of the various negatively charged Ga vacancy<br />

species were calculated. The triply negatively charged Ga vacancy, V Ga 3- , was studied <strong>in</strong><br />

particular because it dom<strong>in</strong>ated Ga self-diffusion <strong>and</strong> Ga/Al <strong>in</strong>terdiffusion under <strong>in</strong>tr<strong>in</strong>sic<br />

<strong>and</strong> n-dop<strong>in</strong>g conditions, as well as the diffusion of Si donor atoms which occupied Ga<br />

sites. Under strong n-dop<strong>in</strong>g conditions, the thermal equilibrium V Ga 3- concentration was<br />

found to exhibit a temperature <strong>in</strong>dependence or a negative temperature dependence. That<br />

is, the concentration was unchanged, or <strong>in</strong>creased, as the temperature was decreased. This<br />

was contrary to the normal po<strong>in</strong>t defect behavior, <strong>in</strong> which the po<strong>in</strong>t defect thermal<br />

equilibrium concentration decreased as the temperature was lowered. The observed<br />

behavior expla<strong>in</strong>ed a number of known experimental results, <strong>and</strong> required either that V Ga<br />

3-<br />

should atta<strong>in</strong> its thermal equilibrium concentration at the onset of each experiment, or<br />

304


Ga <strong>GaAs</strong> Ga<br />

required the operation of mechanisms which <strong>in</strong>volved po<strong>in</strong>t defect non-equilibrium<br />

phenomena.<br />

T.Y.Tan, H.M.You, U.M.Gösele: Applied Physics A, 1993, 56[3], 249-58<br />

[446-111/112-050]<br />

315 <strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

Undoped 69 <strong>GaAs</strong>/ 71 <strong>GaAs</strong> isotope superlattice structures were molecular beam epitaxially<br />

deposited onto n-type <strong>GaAs</strong> substrates which had been Si-doped to about 3 x 10 18 /cm 3 .<br />

They were then used to study Ga self-diffusion <strong>in</strong> <strong>GaAs</strong> (table 15). At temperatures<br />

rang<strong>in</strong>g from 850 to 960C, secondary ion mass spectrometric data <strong>in</strong>dicated an activation<br />

enthalpy of 4eV for Ga self-diffusion. This value was larger than those previously found<br />

for Ga self-diffusion <strong>and</strong> Al-Ga <strong>in</strong>terdiffusion under thermal equilibrium <strong>and</strong> <strong>in</strong>tr<strong>in</strong>sic<br />

conditions; which were characterized by an activation enthalpy of 6eV. Secondary ion<br />

mass spectroscopic, capacitance-voltage, <strong>and</strong> transmission electron microscopic data<br />

showed that the as-grown superlattice layers were <strong>in</strong>tr<strong>in</strong>sic. They became p-type, with<br />

hole concentrations of about 2 x 10 17 /cm 3 , after anneal<strong>in</strong>g. This occurred because the<br />

layers conta<strong>in</strong>ed C. Dislocations were also present, at a density of 10 6 to 10 7 /cm 2 .<br />

However, the factor which was responsible for the larger Ga self-diffusivity values which<br />

were observed here appeared to be Si out-diffusion from the substrate. Such out-diffusion<br />

decreased the electron concentration <strong>in</strong> the substrate, <strong>and</strong> caused the release of Ga<br />

vacancies <strong>in</strong>to the superlattice layers, where they became supersaturated. This Ga<br />

vacancy supersaturation led to enhanced Ga self-diffusion <strong>in</strong> the superlattice layers.<br />

T.Y.Tan, H.M.You, S.Yu, U.M.Gösele, W.Jäger, D.W.Boer<strong>in</strong>ger, F.Zypman, R.Tsu,<br />

S.T.Lee: Journal of Applied Physics, 1992, 72[11], 5206-12<br />

[446-106/107-036]<br />

Table 15<br />

Diffusivity of Ga <strong>in</strong> <strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

960 4.8 x 10 -16<br />

930 1.3 x 10 -16<br />

905 5.6 x 10 -17<br />

875 2.7 x 10 -17<br />

850 8.6 x 10 -18<br />

965 3.1 x 10 -16<br />

935 1.0 x 10 -16<br />

905 4.5 x 10 -17<br />

875 2.0 x 10 -17<br />

850 6.9 x 10 -18<br />

305


Ga <strong>GaAs</strong> Ga<br />

<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

The lateral diffusion of sources dur<strong>in</strong>g the selective growth of metalorganic vapor-phase<br />

epitaxial Si-doped layers was analyzed. The diffusion lengths of Ga species were deduced<br />

from the carrier concentration profiles which were measured by us<strong>in</strong>g Raman<br />

spectroscopy <strong>and</strong> thickness profil<strong>in</strong>g. On the basis of these diffusion lengths, it was<br />

speculated that the effective diffusion material was monomethyl Ga. It was suggested that<br />

there was no difference between ars<strong>in</strong>e <strong>and</strong> tertiary butyl ars<strong>in</strong>e, as diffusion sources.<br />

N.Hara, K.Shi<strong>in</strong>a, T.Ohori, K.Kasai, J.Komeno: Journal of Applied Physics, 1993, 74[2],<br />

1327-30<br />

[446-106/107-036]<br />

<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

The formation energy of Si donors, acceptors, <strong>and</strong> defect complexes were calculated. The<br />

equilibrium concentrations of native defects <strong>and</strong> Si-defect complexes were deduced from<br />

these energies, as was the total solubility of Si. The calculated equilibrium solubility limit<br />

for Si was <strong>in</strong> good agreement with experimental data. The (Si Ga -V Ga ) 2- complex occurred<br />

at relatively high concentrations under As-rich conditions, <strong>and</strong> could therefore mediate Si<br />

<strong>and</strong> Ga diffusion. It was concluded that the donor-vacancy complex was an important<br />

compensation mechanism <strong>in</strong> heavily doped <strong>GaAs</strong>.<br />

J.E.Northrup, S.B.Zhang: Physical Review B, 1993, 47[11], 6791-4<br />

[446-106/107-036]<br />

<strong>GaAs</strong>/AlAs: Ga <strong>Diffusion</strong><br />

The implantation of Be ions <strong>in</strong>to heterostructures at room temperature or liquid N<br />

temperatures was <strong>in</strong>vestigated. It was found that room-temperature implantation created<br />

dislocation loops at the first <strong>in</strong>terface; a distance which was far short of the maximum<br />

projected range. Implantation at low temperatures caused tw<strong>in</strong>n<strong>in</strong>g. The latter could be<br />

removed by anneal<strong>in</strong>g (900C, 1200s), without lead<strong>in</strong>g to the <strong>in</strong>terdiffusion of Ga. The<br />

presence of dislocation networks tended to enhance <strong>in</strong>termix<strong>in</strong>g. The Be concentrations<br />

were sufficiently low to prevent Be-<strong>in</strong>duced <strong>in</strong>termix<strong>in</strong>g.<br />

S.Mitra: Semiconductor Science <strong>and</strong> Technology, 1990, 5[11], 1138-40<br />

[446-76/77-017]<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

The effect of room-temperature electron irradiation upon <strong>in</strong>terdiffusion at quantum-well<br />

<strong>in</strong>terfaces was <strong>in</strong>vestigated by us<strong>in</strong>g low-temperature cathodolum<strong>in</strong>escence spectroscopy.<br />

It was found that <strong>in</strong>terdiffusion was enhanced by the presence of defects which were<br />

generated by irradiation with a 400keV electron beam. After irradiation at room<br />

temperature to doses of between about 1.5 x 10 17 <strong>and</strong> 2.5 x 10 17 /cm 2 , followed by rapid<br />

thermal anneal<strong>in</strong>g (900C, 60s), an <strong>in</strong>terdiffusion length of 0.3 to 0.5nm was found. The<br />

resultant damage tended to saturate with <strong>in</strong>creas<strong>in</strong>g irradiation dose. The formation of<br />

306


Ga <strong>GaAs</strong> Ge<br />

defect clusters at high doses limited the degree of defect <strong>in</strong>troduction, <strong>and</strong> therefore the<br />

extent of <strong>in</strong>terdiffusion at the <strong>in</strong>terface.<br />

Y.J.Li, M.Tsuchiya, P.M.Petroff: Applied Physics Letters, 1990, 57[5], 472-4<br />

[446-76/77-018]<br />

<strong>GaAs</strong>/Si(O,N): Ga <strong>Diffusion</strong><br />

The out-diffusion of Ga atoms from a <strong>GaAs</strong> substrate, <strong>and</strong> <strong>in</strong>to a SiO x N y encapsulat<strong>in</strong>g<br />

film, dur<strong>in</strong>g anneal<strong>in</strong>g was studied by us<strong>in</strong>g secondary ion mass spectrometry. The<br />

concentration of Ga atoms which was detected with<strong>in</strong> the encapsulant, when annealed at<br />

850C, was found to <strong>in</strong>crease with an <strong>in</strong>creas<strong>in</strong>g O content <strong>in</strong> the encapsulant. This<br />

behavior could be correlated with changes <strong>in</strong> the concentration of the EL5 electron trap<br />

(E c - E T = 0.42eV); as detected by us<strong>in</strong>g deep-level transient spectroscopy. It was<br />

concluded that the generation of EL5 traps dur<strong>in</strong>g anneal<strong>in</strong>g was due to Ga out-diffusion.<br />

M.Kuzuhara, T.Nozaki, T.Kamejima: Journal of Applied Physics, 1989, 66[12], 5833-6<br />

[446-74-029]<br />

Ge<br />

316 <strong>GaAs</strong>: Ge <strong>Diffusion</strong><br />

This dopant diffused extensively after implantation <strong>and</strong> long-term anneal<strong>in</strong>g. The results<br />

could be expla<strong>in</strong>ed by assum<strong>in</strong>g that the diffusivity depended upon the square of the<br />

electron concentration. The dopant diffusion was affected by the presence of implantation<br />

damage; the higher the concentration of extended defects, the slower was the diffusivity<br />

as compared with the values for conventional diffusion from a solid source. If the sample<br />

was amorphized dur<strong>in</strong>g implantation, extended defects did not form <strong>and</strong> the diffusivity of<br />

the ion was very close to that <strong>in</strong> material which had been diffused from a solid source.<br />

When amorphization did not occur, extended defects formed after implantation, <strong>and</strong><br />

diffusion was <strong>in</strong>hibited; especially after low doses, <strong>in</strong> the short term, or at low<br />

temperatures. The higher the density of extended defects, the greater was the suppression<br />

of diffusion. No time-dependence was observed. It was concluded that the results (table<br />

16) were consistent with a diffusion mechanism <strong>in</strong> which the mobile species was the<br />

donor that was coupled with a charged Ga vacancy. The equilibrium vacancy<br />

concentration was thought to be suppressed by the presence of extended defects <strong>and</strong>/or<br />

excess Ga <strong>in</strong>terstitials which resulted from implantation.<br />

E.L.Allen, J.J.Murray, M.D.Deal, J.D.Plummer, K.S.Jones, W.S.Rubart: Journal of the<br />

Electrochemical Society, 1991, 138[11], 3440-9<br />

[446-84/85-016]<br />

<strong>GaAs</strong>: Ge <strong>Diffusion</strong><br />

The behavior of Ge which was pulse-diffused <strong>in</strong>to high-purity epitaxial <strong>GaAs</strong> from a th<strong>in</strong><br />

elemental source, us<strong>in</strong>g rapid thermal process<strong>in</strong>g, was <strong>in</strong>vestigated. A comparison of<br />

secondary ion mass spectrometry <strong>and</strong> differential Hall effect measurements showed that<br />

the resultant n + -doped layers were highly compensated. In contrast to the case where Ge<br />

307


Ge <strong>GaAs</strong> Ge<br />

was <strong>in</strong>troduced dur<strong>in</strong>g crystal growth or by ion implantation, Ge Ga donors were not<br />

compensated by Ge As acceptors when Ge was pulse-diffused <strong>in</strong>to <strong>GaAs</strong>.<br />

Photolum<strong>in</strong>escence spectroscopy showed that Ge Ga donors were compensated by V Ga<br />

acceptors rather than by Ge As acceptors. At low diffusion temperatures, Ga vacancies<br />

were formed as Ga rapidly diffused <strong>in</strong>to the Ge layer. These vacancies suppressed the<br />

formation of As vacancies <strong>and</strong> thus Ge As acceptors. At higher diffusion temperatures,<br />

Ge Ga -V Ga complexes were formed more rapidly than V Ga acceptors. Low-temperature<br />

Hall effect measurements suggested that these complexes were neutral. The formation of<br />

complexes, at the expense of isolated Ge Ga donors <strong>and</strong> V Ga acceptors, was related to<br />

diffusion temperatures which exceeded 865C (the Ge-<strong>GaAs</strong> liquidus) <strong>and</strong> was expla<strong>in</strong>ed<br />

by a marked <strong>in</strong>crease <strong>in</strong> the V Ga concentration <strong>in</strong> the near-surface region, due to Ga<br />

dissolution at the Ge-Ga-As liquidus.<br />

C.W.Farley, T.S.Kim, S.D.Lester, B.G.Streetman, J.M.Anthony: Journal of the<br />

Electrochemical Society, 1987, 134[11], 2888-92<br />

[446-60-003]<br />

Table 16<br />

Diffusivity of Implanted Ge <strong>in</strong> <strong>GaAs</strong><br />

Dose (/cm 2 ) Temperature (C) D (cm 2 /s)<br />

1 x 10 14 1000 4.8 x 10 -14<br />

1 x 10 14 900 7.8 x 10 -15<br />

1 x 10 14 900 2.9 x 10 -15<br />

1 x 10 14 900 1.8 x 10 -15<br />

1 x 10 14 900 8.5 x 10 -16<br />

1 x 10 14 900 5.0 x 10 -16<br />

5 x 10 15 850 6.4 x 10 -16<br />

1 x 10 15 850 4.8 x 10 -16<br />

1 x 10 14 870 3.0 x 10 -16<br />

5 x 10 15 750 7.6 x 10 -17<br />

1 x 10 15 750 1.8 x 10 -17<br />

1 x 10 14 870 8.9 x 10 -17<br />

5 x 10 14 750 6.6 x 10 -18<br />

5 x 10 14 850 4.9 x 10 -17<br />

5 x 10 14 750 4.8 x 10 -18<br />

5 x 10 13 850 1.0 x 10 -17<br />

1 x 10 14 950 2.5 x 10 -17<br />

1 x 10 14 950 1.5 x 10 -16<br />

5 x 10 15 950 1.4 x 10 -15<br />

5 x 10 14 950 2.0 x 10 -15<br />

1 x 10 15 950 2.9 x 10 -15<br />

1 x 10 15 950 3.8 x 10 -15<br />

308


Ge <strong>GaAs</strong> H<br />

<strong>GaAs</strong>: Ge <strong>Diffusion</strong><br />

Implantation (3 x 10 13 /cm 2 ) of 200keV Ge <strong>in</strong>to (x11)A-oriented semi-<strong>in</strong>sulat<strong>in</strong>g <strong>GaAs</strong><br />

substrates (where x took values of up to 4) was carried out. For comparison, (110)- <strong>and</strong><br />

(100)-oriented substrates were also implanted. No <strong>in</strong>-diffusion of Ge was observed after<br />

anneal<strong>in</strong>g substrates of any orientation.<br />

M.V.Rao, H.B.Dietrich, P.B.Kle<strong>in</strong>, A.Fathimulla, D.S.Simons, P.H.Chi: Journal of<br />

Applied Physics, 1994, 75[12], 7774-8<br />

[446-117/118-166]<br />

1.0E-07<br />

1.0E-08<br />

1.0E-09<br />

D (cm 2 /s)<br />

1.0E-10<br />

1.0E-11<br />

1.0E-12<br />

table 17<br />

table 18<br />

1.0E-13<br />

13 14 15 16 17 18 19 20<br />

10 4 /T(K)<br />

Figure 5: Diffusivity of H <strong>in</strong> <strong>GaAs</strong><br />

H<br />

317 <strong>GaAs</strong>: H <strong>Diffusion</strong><br />

It was recalled that the diffusion of neutral H atoms <strong>in</strong>to a semiconductor was<br />

accompanied by their b<strong>in</strong>d<strong>in</strong>g <strong>in</strong>to molecules. When the thermal dissociation of molecules<br />

could be ignored, <strong>and</strong> for times which were sufficiently long to establish an equilibrium<br />

309


H <strong>GaAs</strong> H<br />

state for the b<strong>in</strong>d<strong>in</strong>g of H <strong>in</strong>to complexes with impurity atoms, the formation of molecules<br />

was the ma<strong>in</strong> process which determ<strong>in</strong>ed the steady-state profile of atomic H which<br />

formed near to the crystal surface. By analyz<strong>in</strong>g secondary ion mass spectrometry data <strong>in</strong><br />

terms of a model for this process, it was estimated that the diffusivity of H was equal to<br />

1.4 x 10 -10 cm 2 /s at 360C <strong>and</strong> equal to 5.5 x 10 -11 cm 2 /s at 390C. When comb<strong>in</strong>ed with<br />

an<strong>other</strong> result, it was deduced that the overall data (table 17) could be described by:<br />

D (cm 2 /s) = 0.0385 exp[-1.13(eV)/kT]<br />

N.S.Rytova: Fizika i Tekhnika Poluprovodnikov, 1991, 25[6], 1078-80 (Soviet Physics -<br />

<strong>Semiconductors</strong>, 1991, 25[6], 650-1)<br />

Table 17<br />

Diffusivity of H <strong>in</strong> <strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

390 5.2 x 10 -11<br />

360 1.2 x 10 -10<br />

250 1.0 x 10 -12<br />

<strong>GaAs</strong>: H <strong>Diffusion</strong><br />

Layers of material, which was doped with various group-VI donors (S, Se, Te), were<br />

exposed to H plasma. Electronic measurements <strong>in</strong>dicated that, after H diffusion, the<br />

electron concentration systematically decreased while their mobility <strong>in</strong>creased; thus<br />

demonstrat<strong>in</strong>g the passivation of the group-VI donors by H.<br />

B.Theys, B.Machayekhi, J.Chevallier, K.Somogyi, K.Zahraman, P.Gibart, M.Miloche:<br />

Journal of Applied Physics, 1995, 77[7], 3186-93<br />

[446-121/122-052]<br />

<strong>GaAs</strong>: H <strong>Diffusion</strong><br />

The ability of charged H to drift <strong>in</strong> this material was used to <strong>in</strong>vestigate the behavior of<br />

H. It was concluded that diffusion with<strong>in</strong> the above temperature range was entirely traplimited,<br />

<strong>and</strong> exhibited no dependence upon the diffusivity of free H. By drift<strong>in</strong>g H away<br />

from the donors <strong>in</strong> hydrogenated n-type <strong>GaAs</strong>, reactivation of the passivated donors could<br />

be studied. Thermal dissociation of the donor-H complex obeyed first-order k<strong>in</strong>etics, with<br />

a dissociation energy of 1.52eV.<br />

A.W.R.Leitch, T.Zundel, T.Prescha, J.Weber: Materials Science Forum, 1992, 83-87, 21-<br />

6<br />

[446-93/94-009]<br />

<strong>GaAs</strong>: H <strong>Diffusion</strong><br />

First-pr<strong>in</strong>ciples calculations were made of the properties of atomic <strong>and</strong> molecular H <strong>in</strong><br />

pure bulk material. The results <strong>in</strong>dicated that the H penetrated <strong>in</strong> atomic form. With<strong>in</strong> the<br />

sample, atomic H tended to form H 2 molecules on tetrahedral sites. These were deep<br />

energy wells for H 2 . The H 2 * defect, which consisted of a H atom <strong>in</strong> a bond-center site<br />

310


H <strong>GaAs</strong> H<br />

<strong>and</strong> a H atom <strong>in</strong> an adjacent tetrahedral position, had a higher energy than H 2 but was a<br />

lower-energy barrier to diffusion. Isolated H could be present as a metastable species.<br />

The stable charge state of isolated H was calculated as a function of the Fermi energy.<br />

The results suggested that H behaved as a negative-U defect. Consequently, isolated H<br />

was expected to be present only as a charged species (positively charged <strong>in</strong> p-doped<br />

samples, negatively charged <strong>in</strong> undoped <strong>and</strong> n-doped samples). The conclusions were<br />

compared with experimental results <strong>and</strong> with the results of calculations for H <strong>in</strong> <strong>other</strong><br />

semiconductors. The ma<strong>in</strong> features of H behavior <strong>in</strong> <strong>GaAs</strong> were quite similar to those for<br />

Si.<br />

L.Pavesi, P.Giannozzi: Physical Review B, 1992, 46[8], 4621-9<br />

[446-93/94-009]<br />

<strong>GaAs</strong>: H <strong>Diffusion</strong><br />

It was po<strong>in</strong>ted out that Se donors <strong>in</strong> n-type material were passivated by exposure to a H<br />

plasma. Thermal reactivation of the passivated donors was <strong>in</strong>vestigated at temperatures<br />

rang<strong>in</strong>g from 154 to 191C. By us<strong>in</strong>g the electric field of a reverse-biased Schottky diode<br />

structure, the drift of thermally dissociated H as a negatively charged species was<br />

demonstrated. The thermal dissociation of the SeH complex obeyed first-order k<strong>in</strong>etics,<br />

with a dissociation energy of 1.52eV.<br />

A.W.R.Leitch, T.Prescha, J.Weber: Physical Review B, 1991, 44[3], 1375-8<br />

[446-84/85-016]<br />

<strong>GaAs</strong>: H <strong>Diffusion</strong><br />

It was shown that the exponential depth profiles which were sometimes observed dur<strong>in</strong>g<br />

H diffusion <strong>in</strong>to semiconductors, such as Si <strong>and</strong> <strong>GaAs</strong>, could be expla<strong>in</strong>ed by <strong>in</strong>clud<strong>in</strong>g a<br />

term (<strong>in</strong> the diffusion equation) that described the multiple trapp<strong>in</strong>g of H at an impurity. It<br />

was shown that the effective dimensionality of the r<strong>and</strong>om walk <strong>in</strong> the present case was<br />

<strong>in</strong>f<strong>in</strong>ite.<br />

D.A.Tulch<strong>in</strong>sky, J.W.Corbett, J.T.Borenste<strong>in</strong>, S.J.<strong>Pearton</strong>: Physical Review B, 1990,<br />

42[18], 11881-3<br />

[446-81/82-010]<br />

<strong>GaAs</strong>: H <strong>Diffusion</strong><br />

The concentration versus depth profiles of carriers, electrically active defects, <strong>and</strong> D,<br />

after exposure to a H plasma (or molecular H), were fitted by us<strong>in</strong>g a simple diffusion<br />

model which <strong>in</strong>volved second-order reactions. It was found that the activation energy for<br />

H diffusion, <strong>and</strong> the dissociation energies of H-defect complexes, depended upon the H<br />

concentration. There was no molecular H formation <strong>and</strong> there were no fast-diffus<strong>in</strong>g H<br />

species away from the near-surface region. Atomic H could <strong>in</strong>-diffuse <strong>and</strong> passivate EL2<br />

defects when semi-<strong>in</strong>sulat<strong>in</strong>g material was annealed at a high temperature <strong>in</strong> a molecular<br />

H ambient.<br />

R.A.Morrow: Journal of Applied Physics, 1989, 66[7], 2973-9<br />

[446-74-010]<br />

311


H <strong>GaAs</strong> H<br />

<strong>GaAs</strong>: H <strong>Diffusion</strong><br />

Wafers of n-type material were exposed to a capacitively coupled radio-frequency H<br />

plasma, with power densities of between 0.01 <strong>and</strong> 0.2W/cm 2 at 260C. The properties of<br />

the layers were <strong>in</strong>vestigated by us<strong>in</strong>g deep-level transient spectroscopy <strong>and</strong> capacitancevoltage<br />

methods. As well as the neutralization of Si donors by <strong>in</strong>-diffused H atoms, it was<br />

found that there was a modification of the deep-level transient spectroscopy spectra after<br />

hydrogenation. At two radio-frequency power densities of less than 0.IW/cm 2 , deep levels<br />

<strong>in</strong> the orig<strong>in</strong>al material were passivated. The results also <strong>in</strong>dicated that the complexes<br />

which were present were either electrically <strong>in</strong>active or were located deep with<strong>in</strong> the<br />

energy b<strong>and</strong>-gap. At power densities of more than 0.1W/cm 2 , two new deep states<br />

appeared at 0.41 <strong>and</strong> 0.55eV below the conduction b<strong>and</strong>. These levels were attributed to a<br />

large number of defects which were situated <strong>in</strong> the near-surface region of n-type Si-doped<br />

material after H plasma exposure. It was suggested that the trapp<strong>in</strong>g of H at these defects<br />

was probably responsible for the observed accumulation of H <strong>in</strong> the near-surface region.<br />

A.Jalil, A.Heurtel, Y.Marfa<strong>in</strong>g, J.Chevallier: Journal of Applied Physics, 1989, 66[12],<br />

5854-61<br />

[446-74-010]<br />

<strong>GaAs</strong>: H <strong>Diffusion</strong><br />

The depth profiles of 60 <strong>and</strong> 100keV protons which were implanted to fluences of 10 16 or<br />

10 17 /cm 2 at room temperature were determ<strong>in</strong>ed by us<strong>in</strong>g ion beam techniques. The H<br />

profiles were measured as a function of anneal<strong>in</strong>g temperatures of up to 820K. It was<br />

found that the redistribution of implanted H depended upon the migration of<br />

implantation-<strong>in</strong>duced defects. The migration of H-defect complexes was described by the<br />

expression:<br />

D(cm 2 /s) = 2 x 10 5 exp[-2.16(eV)/kT]<br />

J.Räisänen, J.Ke<strong>in</strong>onen, V.Karttunen, I.Koponen: Journal of Applied Physics, 1988,<br />

64[5], 2334-6<br />

[446-72/73-010]<br />

318 <strong>GaAs</strong>: H <strong>Diffusion</strong><br />

Profiles which reflected the passivation of electrically active <strong>and</strong> recomb<strong>in</strong>ation-active<br />

centers by atomic H <strong>in</strong> s<strong>in</strong>gle crystals were <strong>in</strong>vestigated by us<strong>in</strong>g secondary ion mass<br />

spectrometry <strong>and</strong> microcathodolum<strong>in</strong>escence methods. The former data <strong>in</strong>dicated that the<br />

profiles were the same for all of the samples, <strong>and</strong> consisted of an 0.0005mm surface<br />

region with a H content of 10 20 /cm 3 <strong>and</strong> a low diffusion rate. This was followed by a<br />

diffusion tail with a much higher diffusion rate. In this tail region, the activation energy<br />

for H diffusion was 0.83eV at temperatures of between 200 <strong>and</strong> 500C (table 18), <strong>and</strong> the<br />

diffusivity was 6.7 x 10 -9 cm 2 /s at 400C. It was proposed that the results confirmed the<br />

suggested migration of neutral <strong>in</strong>terstitial H. The <strong>in</strong>troduction of atomic H completely<br />

suppressed a microcathodolum<strong>in</strong>escence <strong>in</strong>homogeneity which was associated with<br />

dislocations <strong>in</strong> the semi-<strong>in</strong>sulat<strong>in</strong>g material.<br />

312


H <strong>GaAs</strong> H<br />

E.M.Omelyanovskii, A.V.Pakhomov, A.J.Polyakov, A.V.Govorkov, O.M.Borod<strong>in</strong>a,<br />

A.S.Bruk: Fizika i Tekhnika Poluprovodnikov, 1988, 22[7], 1203-7. (Soviet Physics -<br />

<strong>Semiconductors</strong>, 1988, 22[7], 763-5)<br />

[446-62/63-210]<br />

Table 18<br />

Diffusivity of H <strong>in</strong> <strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

480 2.3 x 10 -8<br />

425 5.1 x 10 -8<br />

445 1.4 x 10 -8<br />

445 1.0 x 10 -8<br />

420 1.2 x 10 -8<br />

410 7.7 x 10 -9<br />

410 6.2 x 10 -9<br />

390 2.2 x 10 -9<br />

365 2.8 x 10 -9<br />

365 1.9 x 10 -9<br />

320 2.2 x 10 -9<br />

320 9.4 x 10 -10<br />

290 6.0 x 10 -10<br />

295 2.7 x 10 -10<br />

255 2.3 x 10 -10<br />

255 1.3 x 10 -10<br />

<strong>GaAs</strong>: H <strong>Diffusion</strong><br />

Experiments were performed on buried Si-doped epilayers under hydrostatic pressure. It<br />

was found that the D diffusion profile <strong>in</strong> n-type Si-doped samples depended upon the<br />

hydrostatic pressure, <strong>and</strong> consisted of a plateau that was followed by a steep progressive<br />

decrease as the pressure was <strong>in</strong>creased. This was expla<strong>in</strong>ed as be<strong>in</strong>g due to an <strong>in</strong>creas<strong>in</strong>g<br />

importance of the trapp<strong>in</strong>g-detrapp<strong>in</strong>g process of H - on Si + donors dur<strong>in</strong>g H diffusion.<br />

This <strong>in</strong>crease was attributed to a deepen<strong>in</strong>g of the H acceptor level, with respect to the<br />

bottom of the Γ conduction b<strong>and</strong>, as the hydrostatic pressure was <strong>in</strong>creased.<br />

D.Machayekhi, J.Chevallier, B.Theys, J.M.Besson, G.Weill, G.Syfosse: Solid State<br />

Communications, 1996, 100[12], 821-4<br />

[446-141/142-098]<br />

<strong>GaAs</strong>: H <strong>Diffusion</strong><br />

The results of <strong>in</strong>vestigations of transitions among the sites <strong>and</strong> charge states of muonium<br />

were summarized. A model was developed which accounted for all of the major features<br />

313


H <strong>GaAs</strong> H<br />

which were observed. Its validity for a wide range of dopant concentrations, us<strong>in</strong>g a<br />

s<strong>in</strong>gle set of parameters, reflected its predictive strength. The near equality of the energy<br />

parameters for muonium, as compared with those which were available for H, strongly<br />

implied that the results for muonium dynamic behavior should be applicable to H, with<br />

very little change. The model could be applied to all tetrahedrally coord<strong>in</strong>ated<br />

semiconductors, with few modifications, <strong>and</strong> served as a basis for the underst<strong>and</strong><strong>in</strong>g of<br />

muonium dynamics <strong>and</strong> therefore for the behavior of H <strong>in</strong> <strong>GaAs</strong> <strong>and</strong> <strong>other</strong> materials.<br />

Differences <strong>in</strong> H properties could be understood by exam<strong>in</strong><strong>in</strong>g material-specific<br />

deviations from the basic model.<br />

R.L.Lichti, C.Schwab, T.L.Estle: Materials Science Forum, 1995, 196-201, 831-6<br />

[446-127/128-119]<br />

<strong>GaAs</strong>: H <strong>Diffusion</strong><br />

The diffusivity of H + was directly determ<strong>in</strong>ed by means of capacitance transient<br />

measurements. The H was released from donor-H complexes, <strong>in</strong> the space-charge layer of<br />

Schottky diodes on n-type material, by the pulsed-laser <strong>in</strong>jection of m<strong>in</strong>ority carriers.<br />

Capacitance-voltage measurements revealed a recovery of donor dopants after <strong>in</strong>jection.<br />

This demonstrated m<strong>in</strong>ority-carrier enhanced dissociation of a donor-H complex.<br />

Capacitance transients which were recorded dur<strong>in</strong>g the migration of H + were analyzed <strong>in</strong><br />

order to obta<strong>in</strong> diffusivities at near to room temperature. At 320K, the diffusivity was<br />

equal to 10 -12 cm 2 /s, to with<strong>in</strong> a factor of 2. An Arrhenius plot of the migration timeconstant<br />

yielded an activation energy for H + diffusion of about 0.66eV.<br />

N.M.Johnson, C.Herr<strong>in</strong>g, D.Bour: Physical Review B, 1993, 48[24], 18308-11<br />

[446-115/116-117]<br />

<strong>GaAs</strong>: H <strong>Diffusion</strong><br />

<strong>Diffusion</strong> experiments were performed on samples of Si-doped Al x Ga 1-x As epitaxial<br />

layers, with x-values which ranged from 0 to 0.30, as a function of the Si dop<strong>in</strong>g level<br />

<strong>and</strong> the diffusion temperature. For each composition, calculated H diffusion profiles<br />

which had been deduced by us<strong>in</strong>g Mathiot's model were fitted to the experimental<br />

profiles. It was assumed that H behaved as a deep acceptor, <strong>and</strong> that H o <strong>and</strong> H - were the<br />

diffus<strong>in</strong>g species. The trapp<strong>in</strong>g of H - by Si + donors, <strong>and</strong> their acceleration by an electric<br />

field, were <strong>in</strong>corporated <strong>in</strong>to the model. As well as the diffusion coefficient of H, <strong>and</strong> the<br />

dissociation constant of the SiH complexes, the model provided for a compositional<br />

dependence of the H acceptor level <strong>in</strong> Al<strong>GaAs</strong> alloys. Extrapolation of the H acceptor<br />

level to x = 0 gave its value for <strong>GaAs</strong>. This level was located at 0.10eV below the Γ<br />

conduction b<strong>and</strong> of <strong>GaAs</strong>.<br />

B.Machayekhi, R.Rahbi, B.Theys, M.Miloche, J.Chevallier: Materials Science Forum,<br />

1994, 143-147, 951-6<br />

[446-113/114-001]<br />

<strong>GaAs</strong>: H <strong>Diffusion</strong><br />

The depth profiles of diffused H <strong>in</strong> n-type samples were determ<strong>in</strong>ed by means of<br />

secondary ion mass spectroscopy. Specimens with Si donor concentrations which ranged<br />

314


H <strong>GaAs</strong> In<br />

from 10 17 to 5 x 10 18 /cm 3 were exposed to monatomic D from a remote microwave<br />

plasma at temperatures of between 250 <strong>and</strong> 310C. The profiles clearly revealed that D<br />

accumulation occurred up to a concentration that was almost equal to that of the donors<br />

over a depth which depended upon the temperature <strong>and</strong> time of hydrogenation <strong>and</strong> upon<br />

the donor concentration. A plateau <strong>in</strong> the H concentration was attributed to a trapp<strong>in</strong>glimited<br />

migration of H which was dom<strong>in</strong>ated by the formation of SiH complexes via the<br />

Coulombic attraction of Si + <strong>and</strong> H - . The analysis of profiles <strong>in</strong> the tail region beyond the<br />

plateau yielded <strong>in</strong>dependent estimates of the thermal dissociation rate for the SiH<br />

complex, <strong>and</strong> a lower limit of about 0.45eV for the b<strong>in</strong>d<strong>in</strong>g of Si + <strong>and</strong> H - <strong>in</strong>to SiH.<br />

G.Roos, N.M.Johnson, C.Herr<strong>in</strong>g, J.Walker: Materials Science Forum, 1994, 143-147,<br />

933-8<br />

[446-113/114-011]<br />

In<br />

<strong>GaAs</strong>: In <strong>Diffusion</strong><br />

An InAs monolayer was grown between <strong>GaAs</strong> layers by us<strong>in</strong>g the migration-enhanced<br />

epitaxy method. The surface chemical characteristics dur<strong>in</strong>g growth were <strong>in</strong>vestigated by<br />

means of reflection high-energy electron diffraction. When the substrate temperature was<br />

equal to about 500C, the oscillation amplitude of the reflected electron beam after the<br />

growth of one monolayer of InAs vanished dur<strong>in</strong>g the growth of <strong>GaAs</strong> over more than 20<br />

atomic layers. High-resolution secondary ion mass spectroscopic analysis of the<br />

fabricated structures <strong>in</strong>dicated that an anomalous distribution of In atoms, with a gradual<br />

slope towards the growth direction, occurred when the substrate temperature was 500C.<br />

The experimental results were expla<strong>in</strong>ed <strong>in</strong> terms of the replacement of In atoms, <strong>in</strong> the<br />

InAs monolayer, by newly deposited Ga atoms.<br />

H.Yamaguchi, Y.Horikoshi: Japanese Journal of Applied Physics, 1989, 28[11], L2010-2<br />

[446-72/73-011]<br />

<strong>GaAs</strong>: In <strong>Diffusion</strong><br />

The diffusion of In markers at 900C was measured <strong>in</strong> undoped <strong>and</strong> Te-doped epilayers<br />

which had been prepared by us<strong>in</strong>g organometallic vapor-phase epitaxy. It was found that<br />

the diffusivity was a l<strong>in</strong>ear function of electron concentrations rang<strong>in</strong>g from 2 x 10 17 to<br />

1.5 x 10 19 /cm 3 . It was concluded that the results were consistent with the <strong>in</strong>terdiffusion of<br />

AlAs <strong>and</strong> <strong>GaAs</strong> superlattices. Also, the In <strong>and</strong> Al diffusivities at 900C were essentially<br />

identical; with<strong>in</strong> experimental error. The results strongly suggested that group-<strong>III</strong><br />

<strong>in</strong>terdiffusion <strong>in</strong> this material was controlled by a Ga vacancy with a charge of -1.<br />

W.M.Li, R.M.Cohen, D.S.Simons, P.H.Chi: Applied Physics Letters, 1997, 70[25], 3392-<br />

4<br />

[446-152-0315]<br />

315


In <strong>GaAs</strong> Li<br />

<strong>GaAs</strong>/InAs: In <strong>Diffusion</strong><br />

The segregation <strong>and</strong> <strong>in</strong>terdiffusion of In atoms <strong>in</strong> <strong>GaAs</strong>/InAs/<strong>GaAs</strong> heterostructures were<br />

<strong>in</strong>vestigated by us<strong>in</strong>g secondary ion mass spectroscopy. When the InAs was grown <strong>in</strong> the<br />

layer-by-layer growth mode, with no dislocations, the segregation of In atoms became<br />

marked with <strong>in</strong>creas<strong>in</strong>g growth temperature. However, segregation was observed even at<br />

the relatively low growth temperature of 400C dur<strong>in</strong>g molecular beam epitaxy. It was<br />

found that segregation was markedly enhanced by dislocations which were near to the<br />

hetero-<strong>in</strong>terface when thick InAs layers were grown <strong>in</strong> a 3-dimensional isl<strong>and</strong> growth<br />

manner. Interdiffusion of In atoms towards the growth direction occurred after thermal<br />

anneal<strong>in</strong>g, <strong>and</strong> could be assisted by vacancies which propagated from the film surface <strong>and</strong><br />

<strong>in</strong>to the epilayer. It became apparent that <strong>in</strong>terdiffusion was effectively suppressed by<br />

<strong>in</strong>sert<strong>in</strong>g a th<strong>in</strong> AlAs layer <strong>in</strong>to the <strong>GaAs</strong> cap layer.<br />

T.Kawai, H.Yonezu, Y.Ogasawara, D.Saito, K.Pak: Journal of Applied Physics, 1993,<br />

74[3], 1770-5<br />

[446-106/107-085]<br />

<strong>GaAs</strong>/Si: In <strong>Diffusion</strong><br />

Samples of <strong>GaAs</strong>, which were encapsulated with th<strong>in</strong> films of amorphous Si at 450C,<br />

were annealed at temperatures of up to 1050C. The resultant poly-Si/<strong>GaAs</strong> <strong>in</strong>terfaces<br />

were <strong>in</strong>vestigated by us<strong>in</strong>g secondary ion mass spectroscopy, Rutherford back-scatter<strong>in</strong>g<br />

spectrometry, <strong>and</strong> transmission electron microscopy. Little or no <strong>in</strong>terdiffusion was<br />

detected at undoped Si/<strong>GaAs</strong> <strong>in</strong>terfaces. The diffusion of dopants such as InP was<br />

detected. An enhanced diffusivity of In <strong>in</strong>to <strong>GaAs</strong> was attributed to the diffusion of po<strong>in</strong>t<br />

defects which were created by the diffusion of As <strong>and</strong> Ga <strong>in</strong>to the encapsulant. It was<br />

deduced that the In diffusivities <strong>in</strong> <strong>GaAs</strong> at doped polycrystall<strong>in</strong>e Si <strong>in</strong>terfaces were<br />

enhanced by factors of about 10000.<br />

K.L.Kavanagh, C.W.Magee, J.Sheets, J.W.Meyer: Journal of Applied Physics, 1988,<br />

64[4], 1845-54<br />

[446-72/73-027]<br />

Li<br />

<strong>GaAs</strong>: Li <strong>Diffusion</strong><br />

Strong photolum<strong>in</strong>escence b<strong>and</strong>s, at 1.34 <strong>and</strong> 1.45eV, were observed after the Li<br />

diffusion of various samples of n-type material at temperatures of between 400 <strong>and</strong> 600C.<br />

The photolum<strong>in</strong>escence b<strong>and</strong>s were homogeneous throughout the bulk of the samples.<br />

They were never observed <strong>in</strong> orig<strong>in</strong>ally p-type or semi-<strong>in</strong>sulat<strong>in</strong>g material. A hole at E v +<br />

0.17eV was detected by means of deep-level transient spectroscopic measurements, <strong>and</strong><br />

an acceptor level at E v + 0.14eV was revealed by temperature-dependent Hall<br />

measurements of n-type material which had been converted to p-type by Li diffusion. It<br />

316


Li <strong>GaAs</strong> Mg<br />

was suggested that Li Ga -D acceptors were connected with compensation <strong>and</strong><br />

photolum<strong>in</strong>escence b<strong>and</strong>s.<br />

B.H.Yang, T.Egilsson, S.Kristjánsson, J.Pétursson, H.P.Gislason: Materials Science<br />

Forum, 1994, 143-147, 839-44<br />

[446-113/114-012]<br />

<strong>GaAs</strong>: Li <strong>Diffusion</strong><br />

The correlation between the electrical conductivity <strong>and</strong> photolum<strong>in</strong>escence spectra of a<br />

wide range of n-type, p-type, <strong>and</strong> semi-<strong>in</strong>sulat<strong>in</strong>g material which had been diffused with<br />

Li was <strong>in</strong>vestigated. It was found that Li diffusion compensated both n-type <strong>and</strong> p-type<br />

material, <strong>and</strong> had a marked effect upon the photolum<strong>in</strong>escence properties of the samples.<br />

The photolum<strong>in</strong>escence b<strong>and</strong>s which were observed, between 1.47 <strong>and</strong> 1.49eV <strong>in</strong> Zndoped<br />

<strong>GaAs</strong>, were related to the compensation of Zn acceptors. The photolum<strong>in</strong>escence<br />

b<strong>and</strong>s, which were observed at 1.43 to 1.45eV <strong>in</strong> all of the samples, were attributed to a<br />

Li-related acceptor complex which gave rise to the p-type conductivity which was<br />

observed <strong>in</strong> all of the samples after full compensation.<br />

H.P.Gislason, B.Yang, I.S.Haukson, J.T.Gudmundsson, M.L<strong>in</strong>narsson, E.Janzén:<br />

Materials Science Forum, 1992, 83-87, 985-90<br />

[446-99/100-064]<br />

Mg<br />

<strong>GaAs</strong>: Mg <strong>Diffusion</strong><br />

The migration of th<strong>in</strong> highly p-doped layers <strong>in</strong> s<strong>in</strong>gle <strong>and</strong> double heterostructures, grown<br />

us<strong>in</strong>g metalorganic vapor-phase epitaxy, was studied by us<strong>in</strong>g capacitance-voltage etch<br />

profil<strong>in</strong>g <strong>and</strong> secondary ion mass spectrometry. It was deduced that the diffusivity of Mg<br />

<strong>in</strong> <strong>GaAs</strong> could be described by:<br />

D (cm 2 /s) = 6.5 x 10 -5 exp[-1.8(eV)/kT]<br />

for rapid thermal anneal<strong>in</strong>g, while the diffusivity could be described by:<br />

D (cm 2 /s) = 1.9 x 10 -7 exp[-1.5(eV)/kT]<br />

for furnace anneal<strong>in</strong>g. A model which was based upon an <strong>in</strong>terstitial cum substitutional<br />

diffusion mechanism, with certa<strong>in</strong> k<strong>in</strong>etic limitations, was successfully used to simulate<br />

the observed dopant concentration profiles.<br />

N.Nordell, P.Ojala, W.H.Van Berlo, G.L<strong>and</strong>gren, M.K.L<strong>in</strong>narsson: Journal of Applied<br />

Physics, 1990, 67[2], 778-86<br />

[446-74-004]<br />

<strong>GaAs</strong>: Mg <strong>Diffusion</strong><br />

The transient up-hill diffusion of Mg dur<strong>in</strong>g anneal<strong>in</strong>g at 900C was computer-simulated.<br />

It was assumed that diffusion occurred via the substitutional-<strong>in</strong>terstitial mechanism, with<br />

excess <strong>in</strong>terstitials <strong>and</strong> vacancies be<strong>in</strong>g produced by implantation; thus caus<strong>in</strong>g the<br />

abnormal diffusion behavior. The substitutional-<strong>in</strong>terstitial mechanism was shown to be<br />

mathematically equivalent to an exist<strong>in</strong>g <strong>in</strong>terstitial-dopant pair diffusion model. This<br />

317


Mg <strong>GaAs</strong> Mg<br />

permitted the programme (a Si process simulator that <strong>in</strong>cluded dopant/po<strong>in</strong>t-defect<br />

<strong>in</strong>teractions) to be used to model up-hill diffusion if suitable diffusivity <strong>and</strong> defect<br />

parameters were <strong>in</strong>cluded. The profiles of excess <strong>in</strong>terstitials <strong>and</strong> vacancies which were<br />

produced by the implantation process were deduced from Boltzmann transport equation<br />

calculations. It was found that transient up-hill diffusion could be accurately simulated;<br />

with the dopant diffus<strong>in</strong>g, from regions with excess <strong>in</strong>terstitials, towards the surface or<br />

towards regions with excess vacancies. When the defect concentrations had returned to<br />

their steady-state levels, via diffusion, recomb<strong>in</strong>ation, or capture by s<strong>in</strong>ks, normal<br />

concentration-dependent diffusion <strong>in</strong>to the substrate occurred.<br />

H.G.Rob<strong>in</strong>son, M.D.Deal, G.Amaratunga. P.B.Griff<strong>in</strong>, D.A.Stevenson, J.D.Plummer:<br />

Journal of Applied Physics, 1992, 71[6], 2615-23<br />

[446-86/87-010]<br />

<strong>GaAs</strong>: Mg <strong>Diffusion</strong><br />

It was noted that Mg implants <strong>in</strong> this material exhibited 2 types of diffusion dur<strong>in</strong>g<br />

anneal<strong>in</strong>g. These were up-hill diffusion <strong>in</strong> the implantation peak, <strong>and</strong> concentrationdependent<br />

diffusion <strong>in</strong>to the bulk. The up-hill diffusion predom<strong>in</strong>ated at short times <strong>and</strong><br />

low temperatures, while the concentration-dependent diffusion was predom<strong>in</strong>ant at long<br />

times <strong>and</strong> high temperatures. By study<strong>in</strong>g implants that had been annealed at temperatures<br />

where no up-hill diffusion occurred, diffused profiles could be modelled <strong>and</strong> an<br />

expression for the Mg diffusivity could be obta<strong>in</strong>ed. The activation energy for this<br />

process was 1.77eV. The results of Fermi level experiments showed that the diffusivity<br />

was hole-dependent rather than concentration dependent. The hole-dependent exponent<br />

was equal to unity for Mg which was implanted <strong>in</strong>to semi-<strong>in</strong>sulat<strong>in</strong>g substrates, but could<br />

change to 2 at high hole concentrations.<br />

H.G.Rob<strong>in</strong>son, M.D.Deal, D.A.Stevenson: Applied Physics Letters, 1991, 58[24], 2800-2<br />

[446-81/82-010]<br />

<strong>GaAs</strong>: Mg <strong>Diffusion</strong><br />

The redistribution of Mg implants dur<strong>in</strong>g post-implantation anneal<strong>in</strong>g was studied <strong>in</strong><br />

order to evaluate the effect of implantation damage upon the diffusion process. Rapid uphill<br />

diffusion was observed <strong>in</strong> the peak of Mg implants. This was expla<strong>in</strong>ed by <strong>in</strong>vok<strong>in</strong>g a<br />

substitutional-<strong>in</strong>terstitial diffusion mechanism <strong>and</strong> by perform<strong>in</strong>g computer simulations of<br />

damage-generated po<strong>in</strong>t defects. In the up-hill diffusion region, the dopants diffused from<br />

areas of excess <strong>in</strong>terstitial concentration towards areas of excess vacancy concentration.<br />

A critical po<strong>in</strong>t defect concentration was necessary <strong>in</strong> order to <strong>in</strong>itiate up-hill diffusion.<br />

H.G.Rob<strong>in</strong>son, M.D.Deal, D.A.Stevenson: Applied Physics Letters, 1990, 56[6], 554-6<br />

[446-74-009]<br />

<strong>GaAs</strong>: Mg <strong>Diffusion</strong><br />

Un-capped wafers were annealed <strong>in</strong> an AsH 3 -N 2 atmosphere for 10s, at temperatures of<br />

up to 1100C. No surface decomposition occurred under an AsH 3 partial pressure of<br />

12.5torr. The present method was used to activate implanted Mg. Measurements of the<br />

sheet resistance of the annealed layers, as a function of the anneal<strong>in</strong>g temperature,<br />

318


Mg <strong>GaAs</strong> Mn<br />

revealed a m<strong>in</strong>imum at a temperature of about 930C. At higher temperatures, the<br />

diffusion of Mg became significant. Part of the Mg accumulated at the surface <strong>and</strong><br />

diffused out. The <strong>in</strong>ternal diffusion of Mg at high temperatures depended upon the AsH 3<br />

pressure dur<strong>in</strong>g anneal<strong>in</strong>g.<br />

H.Tews, R.Neumann, A.Hoepfner, S.Gisdakis: Journal of Applied Physics, 1990, 67[6],<br />

2857-61<br />

[446-74-010]<br />

<strong>GaAs</strong>: Mg <strong>Diffusion</strong><br />

The surface of material which had been implanted with Mg ions (100keV, 10 15 /cm 2 ) was<br />

encapsulated with As-doped amorphous hydrogenated Si to a thickness of about 80nm. It<br />

was found that the sheet carrier concentration <strong>in</strong> thermally annealed samples <strong>in</strong>creased<br />

with <strong>in</strong>creas<strong>in</strong>g concentration of As <strong>in</strong> the encapsulant. After anneal<strong>in</strong>g (800C, 0.25h), a<br />

sheet carrier concentration of about 3 x 10 14 /cm 2 was observed <strong>in</strong> samples which were<br />

capped with films that were doped with 2 x 10 20 /cm 3 of As. It was noted that the<br />

diffusivity of the implanted Mg was retarded upon <strong>in</strong>creas<strong>in</strong>g the concentration of As <strong>in</strong><br />

the amorphous hydrogenated Si encapsulant.<br />

K.Yokota, M.Sakaguchi, H.Inohara, H.Nakanishi, S.Tamura, Y.Hor<strong>in</strong>o, A.Chayahara,<br />

M.Satho, K.Hira, H.Takano, M.Kumagaya: Solid-State Electronics, 1994, 37[1], 9-15<br />

[446-113/114-012]<br />

Mn<br />

<strong>GaAs</strong>: Mn <strong>Diffusion</strong><br />

The diffusion of Mn was carried out <strong>in</strong> sealed quartz ampoules, us<strong>in</strong>g 4 types of Mn<br />

source. These were: solid crystall<strong>in</strong>e Mn gra<strong>in</strong>s, Mn 3 As, MnAs, <strong>and</strong> Mn th<strong>in</strong> films on<br />

<strong>GaAs</strong> substrates. It was found that only MnAs led to the formation of a smooth <strong>GaAs</strong><br />

surface <strong>and</strong> a uniform dopant distribution. In the case of the <strong>other</strong> sources, <strong>in</strong>teractions<br />

between the source materials <strong>and</strong> the substrate gave rise to poor surface morphologies <strong>and</strong><br />

<strong>in</strong>homogeneous distributions. In the case of diffusion at 800C, surface p-type carrier<br />

concentrations of the order of 10 20 /cm 3 were obta<strong>in</strong>ed. The diffusion profiles which were<br />

determ<strong>in</strong>ed by us<strong>in</strong>g capacitance-voltage techniques resembled those which were<br />

obta<strong>in</strong>ed for Zn diffusion. It was suggested that a substitutional-<strong>in</strong>terstitial mechanism<br />

was the predom<strong>in</strong>ant one.<br />

C.H.Wu, K.C.Hsieh, G.E.Höfler, N.El-Ze<strong>in</strong>, N.Holonyak: Applied Physics Letters, 1991,<br />

59[10], 1224-6<br />

[446-84/85-007]<br />

<strong>GaAs</strong>: Mn <strong>Diffusion</strong><br />

Plates of Sn-doped n-type material were coated with 54 Mn, annealed (1100C, 4h) <strong>and</strong><br />

analyzed by us<strong>in</strong>g alternate gr<strong>in</strong>d<strong>in</strong>g <strong>and</strong> activity measurements. The diffusion profiles<br />

could be described by an erfc function, <strong>and</strong> the diffusivity at the anneal<strong>in</strong>g temperature<br />

was estimated to be 4.3 x 10 -10 cm 2 /s.<br />

319


Mn <strong>GaAs</strong> Ni<br />

E.A.Skoryat<strong>in</strong>a, R.S.Malkovich: Fizika i Tekhnika Poluprovodnikov, 1989, 23[1], 164-6.<br />

(Soviet Physics - <strong>Semiconductors</strong>, 1989, 23[1], 101-2)<br />

[446-70/71-106]<br />

<strong>GaAs</strong>: Mn <strong>Diffusion</strong><br />

It was po<strong>in</strong>ted out that one of the potential advantages of rapid thermal anneal<strong>in</strong>g, as<br />

compared with conventional furnace anneal<strong>in</strong>g, was a reduced implanted dopant <strong>and</strong><br />

background impurity diffusion. Here, the migration of Mn dur<strong>in</strong>g the anneal<strong>in</strong>g of Crdoped<br />

semi-<strong>in</strong>sulat<strong>in</strong>g material implanted with 100keV Si + ions to a dose of 7 x 10 12 /cm 2<br />

was measured by us<strong>in</strong>g secondary ion mass spectrometry. Un-capped rapid thermal<br />

anneal<strong>in</strong>g (860 or 930C, 1 to 60s) was <strong>in</strong>vestigated <strong>and</strong> its effect was compared with that<br />

of capless furnace anneal<strong>in</strong>g (0.5h). The migration of Mn was undetectable for rapid<br />

thermal anneal<strong>in</strong>g times which were shorter than 60s, but dom<strong>in</strong>ated 0.5h furnace<br />

anneals.<br />

H.Kanber, J.M.Whelan: Journal of the Electrochemical Society, 1987, 134[10], 2596-9<br />

[446-55/56-005]<br />

<strong>GaAs</strong>: Mn <strong>Diffusion</strong><br />

A study was made of the effect of an As over-pressure upon Mn diffusion. The sources of<br />

Mn <strong>in</strong>cluded solid Mn th<strong>in</strong> films, which were deposited directly onto the <strong>GaAs</strong> substrate,<br />

<strong>and</strong> Mn vapor from pure solid Mn, MnAs, or Mn 3 As. When a solid Mn film was used as<br />

the diffusion source, a non-uniform dopant distribution <strong>and</strong> a poor surface morphology<br />

was obta<strong>in</strong>ed. This was due to reaction between the Mn film <strong>and</strong> the <strong>GaAs</strong> matrix. The<br />

degraded surface consisted of a layer of polycrystall<strong>in</strong>e cubic alloy with a lattice constant<br />

of almost 0.84nm, <strong>and</strong> with a composition that was close to Ga 2 Mn; with a small amount<br />

of As. Of the rema<strong>in</strong><strong>in</strong>g diffusion sources, only MnAs consistently produced a uniform<br />

dop<strong>in</strong>g distribution <strong>and</strong> a smooth surface morphology. By diffusion at 800C, a uniform<br />

surface hole carrier concentration as high as 10 20 /cm 3 could be obta<strong>in</strong>ed by us<strong>in</strong>g MnAs<br />

as a source. The As over-pressure was found to alter drastically the Mn diffusion profile<br />

<strong>and</strong> Mn, like Zn, could diffuse <strong>in</strong>terstitial-substitutionally. Vapor from Mn or Mn 3 As<br />

degraded the surface. However, Mn 3 As degraded the surface more rapidly. A sufficiently<br />

high As over-pressure completely suppressed surface degradation.<br />

C.H.Wu, K.C.Hsieh: Journal of Applied Physics, 1992, 72[12], 5642-8<br />

[446-106/107-037]<br />

Ni<br />

<strong>GaAs</strong>: Ni <strong>Diffusion</strong><br />

The redistribution of Ni <strong>in</strong> a Ni/<strong>GaAs</strong> contact specimen was studied by us<strong>in</strong>g neutron<br />

activation <strong>and</strong> section<strong>in</strong>g techniques at temperatures of between 360 <strong>and</strong> 870K. It was<br />

concluded that <strong>in</strong>terdiffusion of the components took place <strong>in</strong> an elastic stra<strong>in</strong> field which<br />

was generated by differences <strong>in</strong> the lattice parameters <strong>and</strong> thermal expansion coefficients<br />

of Ni, <strong>GaAs</strong> <strong>and</strong> <strong>in</strong>termetallic compounds which formed. At anneal<strong>in</strong>g temperatures below<br />

570K, reactive diffusion of Ni took place, with an activation energy of 0.51eV. The<br />

320


Ni <strong>GaAs</strong> O<br />

formation of micro-cracks <strong>in</strong> the surface layers of the microcrystall<strong>in</strong>e <strong>GaAs</strong> led to<br />

diffusion with an activation energy of 0.25eV. At anneal<strong>in</strong>g temperatures that were<br />

greater than 670K the <strong>in</strong>ternal electric field, <strong>and</strong> cluster formation, markedly affected the<br />

distribution of the components.<br />

W.A.Uskov, A.B.Fedotov, E.A.Erofeeva, A.I.Rodionov, D.T.Dzhumakulov: Izvestiya<br />

Akademii Nauk SSSR - Neorganicheskie Materialy, 1987, 23[2], 186-9. (Inorganic<br />

Materials, 1987, 23[2], 163-5)<br />

[446-55/56-006]<br />

O<br />

<strong>GaAs</strong>: O <strong>Diffusion</strong><br />

A new technique was developed <strong>in</strong> order to study atomic movements dur<strong>in</strong>g ultra-violet<br />

laser-enhanced <strong>and</strong> low-temperature (below 400C) thermal oxidation. This method<br />

comb<strong>in</strong>ed a classical marker technique with low-energy ion-scatter<strong>in</strong>g spectroscopy.<br />

Dur<strong>in</strong>g the formation of th<strong>in</strong> (about 1nm) oxide layers, the marker rema<strong>in</strong>ed on the oxide<br />

surface. This <strong>in</strong>dicated that oxidation occurred, at the <strong>GaAs</strong>/oxide <strong>in</strong>terface, via the<br />

diffusion of an O species. This differed from the oxidation of metals such as Ni <strong>and</strong> Cu,<br />

where the use of the same technique confirmed earlier observations that oxidation<br />

occurred at the oxide/ambient <strong>in</strong>terface. The diffusion of a metal species resulted <strong>in</strong> the<br />

marker be<strong>in</strong>g buried dur<strong>in</strong>g oxidation of the metal surfaces.<br />

M.T.Schmidt, Z.Wu, C.F.Yu, R.M.Osgood: Surface Science, 1990, 226[1-2], 199-205<br />

[446-74-011]<br />

<strong>GaAs</strong>: O <strong>Diffusion</strong><br />

The thermal surface oxidation of samples <strong>in</strong> dry O 2 <strong>and</strong> <strong>in</strong> ambient air was <strong>in</strong>vestigated at<br />

temperatures rang<strong>in</strong>g from 400 to 530C. The studies were carried out by us<strong>in</strong>g clean<br />

polished (111) substrates. It was found that monocrystall<strong>in</strong>e wafers oxidized parabolically<br />

<strong>in</strong> dry O 2 at temperatures of 400 to 450C. L<strong>in</strong>ear oxide growth occurred, at temperatures<br />

rang<strong>in</strong>g from 480 to 530C, <strong>in</strong> both <strong>in</strong> dry O 2 <strong>and</strong> <strong>in</strong> ambient air. Wagner <strong>and</strong> Grimley-<br />

Trapnell models for metal oxidation were used to identify the growth k<strong>in</strong>etics. The rate<br />

constants for parabolic <strong>and</strong> l<strong>in</strong>ear oxidation were temperature-dependent, <strong>and</strong> satisfied<br />

Arrhenius relationships.<br />

J.Kucera, K.Navratil: Th<strong>in</strong> Solid Films, 1990, 191[2], 211-20<br />

[446-78/79-012]<br />

<strong>GaAs</strong>: O <strong>Diffusion</strong><br />

Published depth profiles of D <strong>in</strong> n-type <strong>and</strong> p-type material, after extended anneal<strong>in</strong>g <strong>in</strong> D<br />

at 500C, were suggested to reflect the <strong>in</strong>-diffusion of a native defect (perhaps V As ) <strong>and</strong> an<br />

impurity (perhaps O); both of which were tagged with D. It was deduced that the<br />

diffusivity of the tagged impurity was equal to 4 x 10 -14 cm 2 /s at 500C. The diffusivity of<br />

321


O <strong>GaAs</strong> S<br />

the tagged native defect was deduced to be equal to 3 x 10 -15 cm 2 /s <strong>in</strong> n-type material, <strong>and</strong><br />

to be equal to about 8 x 10 -15 cm 2 /s <strong>in</strong> p-type material.<br />

R.A.Morrow: Applied Physics Letters, 1990, 57[3], 276-8<br />

[446-76/77-009]<br />

<strong>GaAs</strong>: O <strong>Diffusion</strong><br />

Infra-red measurements of the concentrations of H-B pairs, which formed <strong>in</strong> B-doped Si<br />

that had been heated <strong>in</strong> H 2 gas <strong>and</strong> quenched from temperatures of between 900 <strong>and</strong><br />

1300C, led to a new determ<strong>in</strong>ation of the H solubility:<br />

[H](/cm 3 ) =9.1 x 10 21 exp[-1.80(eV)/kT]<br />

There was some evidence that H 2 molecules also formed. The presence of H led to an<br />

enhancement of the diffusivity of O impurities at temperatures below 500C. Suggestions<br />

that H was present <strong>in</strong> as-grown Czochralski Si were supported by the observation of H-C<br />

complexes, us<strong>in</strong>g photolum<strong>in</strong>escence techniques. The analysis of the structure of a H<br />

complex, by means of <strong>in</strong>fra-red vibrational spectroscopy, was illustrated for the case of<br />

the H-C As pair <strong>in</strong> <strong>GaAs</strong>.<br />

R.C.Newman: Philosophical Transactions of the Royal Society A, 1995, 350[1693], 215-<br />

26<br />

[446-119/120-192]<br />

P<br />

<strong>GaAs</strong>/Si: P <strong>Diffusion</strong><br />

Samples of <strong>GaAs</strong>, which were encapsulated with th<strong>in</strong> films of amorphous Si at 450C,<br />

were annealed at temperatures of up to 1050C. The resultant poly-Si/<strong>GaAs</strong> <strong>in</strong>terfaces<br />

were <strong>in</strong>vestigated by us<strong>in</strong>g secondary ion mass spectroscopy, Rutherford back-scatter<strong>in</strong>g<br />

spectrometry, <strong>and</strong> transmission electron microscopy. Little or no <strong>in</strong>terdiffusion was<br />

detected at undoped Si/<strong>GaAs</strong> <strong>in</strong>terfaces. The diffusion of dopants such P was detected.<br />

An enhanced diffusivity of P <strong>in</strong>to <strong>GaAs</strong> was attributed to the diffusion of po<strong>in</strong>t defects<br />

which were created by the diffusion of As <strong>and</strong> Ga <strong>in</strong>to the encapsulant. It was deduced<br />

that the P diffusivities <strong>in</strong> <strong>GaAs</strong> at doped polycrystall<strong>in</strong>e Si <strong>in</strong>terfaces were enhanced by<br />

factors of about 10000.<br />

K.L.Kavanagh, C.W.Magee, J.Sheets, J.W.Meyer: Journal of Applied Physics, 1988,<br />

64[4], 1845-54<br />

[446-72/73-027]<br />

S<br />

<strong>GaAs</strong>: S <strong>Diffusion</strong><br />

Experiments were performed on polished plates of Te-doped material. Irradiation with 15<br />

to 150keV protons was carried out at 300K to doses of between 10 16 <strong>and</strong> 10 17 /cm 2 . It was<br />

found that the impurity profiles did not depend upon whether the diffusion source was<br />

322


S <strong>GaAs</strong> S<br />

deposited before or after irradiation. The penetration depth <strong>in</strong> samples which were<br />

irradiated with 15keV protons was greater than that <strong>in</strong> samples which were irradiated with<br />

150keV protons. It was suggested that this was because the low-energy ions generated<br />

more defects at depths of between 50 <strong>and</strong> 100nm.<br />

V.N.Abrosimova, V.V.Kozlovskii, N.N.Korobkov, V.N.Lomasov: Izvestiya Akademii<br />

Nauk SSSR - Neorganicheskie Materialy, 1990, 26[3], 488-91. (Inorganic Materials,<br />

1990, 26[3], 411-4)<br />

[446-84/85-013]<br />

Table 19<br />

Diffusivity of Implanted S <strong>in</strong> <strong>GaAs</strong><br />

120keV S + (/cm 2 ) Temperature (C) D (cm 2 /s)<br />

1 x 10 15 1000 1.4 x 10 -12<br />

1 x 10 15 950 9.6 x 10 -13<br />

1 x 10 15 900 7.1 x 10 -13<br />

1 x 10 15 850 5.4 x 10 -13<br />

5 x 10 14 1000 9.9 x 10 -13<br />

5 x 10 14 950 7.9 x 10 -13<br />

5 x 10 14 900 4.9 x 10 -13<br />

5 x 10 14 850 3.0 x 10 -13<br />

1 x 10 14 1000 5.0 x 10 -13<br />

1 x 10 14 950 3.4 x 10 -13<br />

1 x 10 14 900 1.3 x 10 -13<br />

5 x 10 13 1000 2.6 x 10 -13<br />

5 x 10 13 950 1.4 x 10 -13<br />

319,20 <strong>GaAs</strong>: S <strong>Diffusion</strong><br />

Samples were implanted with 120keV S + ions to doses of between 3 x 10 13 <strong>and</strong> 10 15 /cm 2<br />

(table 19). They were capped with an 80nm-thick film of amorphous hydrogenated Si,<br />

<strong>in</strong>to which As was doped to a concentration of 2 x 10 20 /cm 3 . The samples were then<br />

annealed <strong>in</strong> Ar gas (850 to 1000C, 0.25h). It was found that the diffusivity of S (table 20)<br />

could be described by the expression:<br />

D = D m [KQ 2 /(1 + KQ 2 )]<br />

where K was a constant, Q was the implantation dose, <strong>and</strong> D m was the diffusivity of a<br />

mobile complex.<br />

M.Sakaguchi, K.Yokota, A.Shiomi, K.Hirai, H.Takano, M.Kumagai: Japanese Journal of<br />

Applied Physics, 1996, 35[1-8], 4203-8<br />

[446-138/139-077]<br />

323


S <strong>GaAs</strong> S<br />

<strong>GaAs</strong>: S <strong>Diffusion</strong><br />

Samples were implanted with 500keV S ions, through 120nm-thick films of amorphous<br />

hydrogenated Si, to give a concentration of 2 x 10 20 /cm 3 . They were then annealed (700<br />

to 1000C, 0.25h, Ar). It was found that the diffusivity of the S decreased with <strong>in</strong>creas<strong>in</strong>g<br />

implantation dose.<br />

M.Sakaguchi, K.Yokota, A.Shiomi, H.Mori, A.Chayahara, Y.Fujii, K.Hirai, H.Takano,<br />

M.Kumagai: Japanese Journal of Applied Physics, 1996, 35[1-3], 1624-9<br />

[446-136/137-110]<br />

Table 20<br />

<strong>Diffusion</strong> Parameters for Implanted S <strong>in</strong> <strong>GaAs</strong><br />

Dose (/cm 2 ) D o (cm 2 /s) E (eV)<br />

1 x 10 15 2.0 x 10 -9 0.8<br />

5 x 10 14 9.0 x 10 -9 1.0<br />

1 x 10 14 4.2 x 10 -7 1.5<br />

5 x 10 13 5.5 x 10 -7 1.6<br />

<strong>GaAs</strong>: S <strong>Diffusion</strong><br />

An <strong>in</strong>vestigation was made of the composition <strong>and</strong> thermal stability of ultra-violet, ozoneoxidized,<br />

<strong>and</strong> P 2 S 5 /(NH 4 ) 2 S-treated (100) surfaces. In particular, X-ray photo-electron<br />

spectroscopy <strong>and</strong> Auger electron spectroscopy were used to probe the oxide <strong>and</strong> <strong>in</strong>terface<br />

at room temperature <strong>and</strong> after anneal<strong>in</strong>g at various temperatures. The room temperature<br />

data <strong>in</strong>dicated that S was buried between the oxide over-layer <strong>and</strong> the <strong>GaAs</strong> substrate.<br />

This oxide conta<strong>in</strong>ed various As <strong>and</strong> Ga bond<strong>in</strong>g configurations which, after moderate<br />

anneal<strong>in</strong>g, were transformed <strong>in</strong>to more thermally stable phases, such as As 2 O 3 <strong>and</strong> Ga 2 O 3 .<br />

Complete desorption of the oxide occurred after anneal<strong>in</strong>g at 600C. Heat<strong>in</strong>g the sample to<br />

495C caused some S to diffuse towards the oxide surface, while anneal<strong>in</strong>g at higher<br />

temperatures led to S diffusion <strong>in</strong>to the <strong>GaAs</strong> substrate. Even after complete desorption of<br />

the O, a small amount of S rema<strong>in</strong>ed <strong>in</strong> the <strong>GaAs</strong> lattice.<br />

M.J.Chester, T.Jach, J.A.Dagata: Journal of Vacuum Science <strong>and</strong> Technology A, 1993,<br />

11[3], 474-80<br />

[446-111/112-050]<br />

<strong>GaAs</strong>/Si: S <strong>Diffusion</strong><br />

The depth distribution of S near to a Si/<strong>GaAs</strong>(110) <strong>in</strong>terface was measured by us<strong>in</strong>g<br />

particle-<strong>in</strong>duced X-ray emission techniques, together with Rutherford back-scatter<strong>in</strong>g<br />

spectrometry. Ozone oxidation <strong>and</strong> HF etch<strong>in</strong>g were used to remove layers. The<br />

measurements revealed the presence of a half-monolayer of S on H 2 S x -passivated <strong>GaAs</strong><br />

(110) surfaces. Upon deposit<strong>in</strong>g 1.5nm of Si onto S-passivated <strong>GaAs</strong> (110), the total<br />

amount of S was found to rema<strong>in</strong> constant as compared to that before Si deposition.<br />

324


S <strong>GaAs</strong> Si<br />

However, no oriented S-Ga bonds were revealed by X-ray absorption measurements, <strong>and</strong><br />

the depth profiles revealed that S atoms diffused <strong>in</strong>to both the <strong>GaAs</strong> substrate <strong>and</strong> the Si<br />

hetero-layer.<br />

H.Xia, W.N.Lennard, L.J.Huang, W.M.Lau, J.M.Baribeau, D.L<strong>and</strong>heer: Journal of<br />

Applied Physics, 1996, 80[8], 4354-7<br />

[446-138/139-130]<br />

Se<br />

<strong>GaAs</strong>/Si: Se <strong>Diffusion</strong><br />

The surface of material which had been implanted with 100keV Se ions was encapsulated<br />

with As-doped amorphous hydrogenated Si to a thickness of about 80nm. Crystallization<br />

of the encapsulant upon anneal<strong>in</strong>g at 1000C was h<strong>in</strong>dered by dop<strong>in</strong>g with As atoms to a<br />

concentration of 2 x10 20 /cm 3 . However, the encapsulant could be crystallized when the<br />

concentration of the As dopant atoms was lowered. The crystallized encapsulant had<br />

many gra<strong>in</strong> boundaries, <strong>and</strong> the diffusion rate of impurities was thereby <strong>in</strong>creased. The<br />

<strong>GaAs</strong> surface decomposed markedly via the boundary of the Si/<strong>GaAs</strong> structure, <strong>and</strong> As<br />

<strong>and</strong> Ga vacancies were produced at the <strong>GaAs</strong> surface. A number of Si atoms also diffused<br />

<strong>in</strong>to the <strong>GaAs</strong> crystal. An As-vacancy rich region formed near to the surface, <strong>and</strong> the Ga<br />

vacancy diffused <strong>in</strong>to the <strong>GaAs</strong> because the diffusivity of the As vacancy <strong>in</strong> <strong>GaAs</strong> was<br />

much lower than that of the Ga vacancy. The Si atoms gave a flat profile, with a steep<br />

slope at the surface, because of the distribution of the vacancies. That is, the Si atoms<br />

preferentially occupied vacant As sites <strong>in</strong> the V As -rich region or vacant Ga sites <strong>in</strong> the<br />

V Ga region. A series of reactions,<br />

Se As + + (Si Ga + + Si As - ) + V Ga - - (Se As + + Si As - ) + (Si Ga + + V Ga - ) →<br />

(Si As - + Se As + ) + (V Ga - + Si Ga + )<br />

<strong>in</strong>creased the diffusion rate of the implanted Se atoms. Thus, the activation efficiency<br />

improved with <strong>in</strong>creas<strong>in</strong>g concentration of As atoms <strong>in</strong> the encapsulant, <strong>and</strong> the diffusion<br />

rate was reduced because the content of Si atoms <strong>and</strong> vacancies <strong>in</strong> <strong>GaAs</strong> was decreased.<br />

K.Yokota, K.Nishida, A.Yutani, S.Tamura, Y.Hor<strong>in</strong>o, A.Chayahara, M.Satho, K.Hirai,<br />

H.Takano, M.Kumagaya: Japanese Journal of Applied Physics, 1993, 32[1-10], 4418-24<br />

[446-115/116-131]<br />

Si<br />

321 <strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The diffusion of Si was studied <strong>in</strong> <strong>GaAs</strong> which had been implanted with 40keV<br />

30 Si + ions to a dose of 10 16 /cm 2 . The Si concentration profiles were determ<strong>in</strong>ed by means<br />

of secondary-ion mass spectrometry <strong>and</strong> nuclear resonance broaden<strong>in</strong>g techniques, <strong>and</strong><br />

the defect distributions were determ<strong>in</strong>ed by us<strong>in</strong>g the Rutherford back-scatter<strong>in</strong>g<br />

spectrometry channell<strong>in</strong>g technique. The implanted samples were subjected to anneal<strong>in</strong>g<br />

<strong>in</strong> Ar at temperatures rang<strong>in</strong>g from 650 to 850C (table 21). Two <strong>in</strong>dependent Si diffusion<br />

325


Si <strong>GaAs</strong> Si<br />

mechanisms were observed. It was found that concentration-<strong>in</strong>dependent diffusion, seen<br />

as a broaden<strong>in</strong>g of the <strong>in</strong>itial implanted distribution, was very slow <strong>and</strong> was attributed to<br />

Si atoms that diffused <strong>in</strong>terstitially. A concentration-dependent diffusivity with low<br />

solubility, which extended deeply <strong>in</strong>to the sample, was quantitatively expla<strong>in</strong>ed <strong>in</strong> terms<br />

of diffusion, via vacancies, of Si atoms on the Ga <strong>and</strong> As sub-lattices. <strong>Diffusion</strong><br />

coefficients, together with the carrier concentration as a function of Si concentration,<br />

were obta<strong>in</strong>ed at various temperatures. The concentration-<strong>in</strong>dependent diffusion of Si was<br />

described by:<br />

D (cm 2 /s) = 1.23 x 10 -7 exp[-1.72(eV)/kT]<br />

The solid solubility of Si <strong>in</strong> <strong>GaAs</strong> was determ<strong>in</strong>ed, <strong>and</strong> an exponential temperature<br />

dependence was observed. An estimate was made of the numbers of Si atoms which<br />

resided on the Ga <strong>and</strong> As sites, <strong>and</strong> the number of Si Ga + -Si As - pairs was deduced. The<br />

<strong>in</strong>tr<strong>in</strong>sic diffusivities via neutral Ga vacancy complexes, triply negatively charged As<br />

vacancy complexes <strong>and</strong> triply negatively charged Ga vacancy complexes were:<br />

D (cm 2 /s) = 3.74 x 10 -3 exp[-2.60(eV)/kT]<br />

D (cm 2 /s) = 4.67 x 10 -5 exp[-2.74(eV)/kT]<br />

<strong>and</strong><br />

D (cm 2 /s) = 5.92 x 10 -8 exp[-2.28(eV)/kT]<br />

respectively.<br />

T.Ahlgren, J.Likonen, J.Slotte, J.Räisänen, M.Rajatora, J.Ke<strong>in</strong>onen: Physical Review B,<br />

1997, 56[8], 4597-603<br />

[446-157/159-326]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The d-doped samples were grown by us<strong>in</strong>g molecular beam epitaxial methods, <strong>and</strong> were<br />

analyzed by us<strong>in</strong>g high-resolution secondary ion mass spectrometry. It was found that<br />

there was a marked difference between the profiles which were produced from samples<br />

that had been doped to a surface density of less than 1.3 x 10 13 /cm 2 (where all of the Si<br />

was <strong>in</strong>corporated on Ga sites), <strong>and</strong> highly-doped samples (where auto-compensation<br />

occurred). All of the samples were grown at a nom<strong>in</strong>al temperature of 580C, <strong>and</strong> all of<br />

the doped planes showed some degree of broaden<strong>in</strong>g. A computer model for a 2-step<br />

diffusion process was developed, <strong>and</strong> this predicted a set of diffusion coefficients for<br />

lightly-doped samples. The diffusion coefficient which was associated with the postdeposition<br />

growth of these lightly-doped samples was about 4.2 x 10 -17 cm 2 /s. Because of<br />

their complicated profiles, more highly-doped samples were modelled by us<strong>in</strong>g a<br />

graphical technique. This revealed the presence of a much larger diffusion coefficient,<br />

which was tentatively attributed to the diffusion of Si as nearest-neighbor pairs.<br />

H.C.Nutt, R.S.Smith, M.Towers, P.K.Rees, D.J.James: Journal of Applied Physics, 1991,<br />

70[2], 821-6<br />

[446-93/94-010]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Two Si-doped specimens, which had been grown by us<strong>in</strong>g molecular beam epitaxy<br />

techniques, were used to study Si diffusion at temperatures rang<strong>in</strong>g from 700 to 950C.<br />

326


Si <strong>GaAs</strong> Si<br />

Each specimen structure consisted of well-characterized regions which ranged <strong>in</strong> type<br />

from undoped semi-<strong>in</strong>sulat<strong>in</strong>g to heavily Si-doped (4 x 10 19 /cm 3 ). In one structure, the Si<br />

dop<strong>in</strong>g <strong>in</strong>creased step-wise from the surface to the semi-<strong>in</strong>sulat<strong>in</strong>g substrate. The <strong>other</strong><br />

structure was grown <strong>in</strong> the reverse fashion, with the maximum Si dop<strong>in</strong>g situated near to<br />

the surface. Anneal<strong>in</strong>g was carried out after encapsulat<strong>in</strong>g the samples with a plasmaenhanced<br />

chemical vapor deposited nitride or oxide layer. Both structures exhibited<br />

almost identical diffusion behaviors which were best modelled by us<strong>in</strong>g an electrondependent<br />

diffusion model. A least-squares fit to both sets of data showed that the Si<br />

diffusivity could be described by:<br />

D(cm 2 /s) = 60.1 exp[-3.9(eV)/kT](n/n i ) 2<br />

This diffusion behavior was <strong>in</strong>dependent of the encapsulat<strong>in</strong>g material which was used,<br />

<strong>and</strong> of the proximity of the dopant to the surface region. The results <strong>in</strong>dicated the<br />

existence of a Fermi level dependent diffusion behavior which was governed by<br />

compensat<strong>in</strong>g acceptor charged po<strong>in</strong>t defects, V Ga m- . These were a consequence of Si<br />

dop<strong>in</strong>g dur<strong>in</strong>g molecular beam epitaxial growth. On the basis of the observed electron<br />

concentration dependence of the diffusivity, <strong>and</strong> assum<strong>in</strong>g the operation of a simple Ga<br />

vacancy diffusion mechanism, these compensat<strong>in</strong>g vacancies were suggested to be at least<br />

doubly negatively charged.<br />

J.J.Murray, M.D.Deal, E.L.Allen, D.A.Stevenson, S.Nozaki: Journal of the<br />

Electrochemical Society, 1992, 137[7], 2037-41<br />

[446-93/94-011]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

A study was made of the contributions of segregation, diffusion <strong>and</strong> aggregation to the<br />

broaden<strong>in</strong>g of d-doped planes of Si. It was found that sharp spikes of Si could be obta<strong>in</strong>ed<br />

for sheet densities which were below 10 13 /cm 2 <strong>and</strong> for growth temperatures of 500C or<br />

less. At higher temperatures or densities, segregation or concentration-dependent rapid<br />

diffusion could occur; thus caus<strong>in</strong>g significant spread<strong>in</strong>g even dur<strong>in</strong>g growth. The codeposition<br />

of Si <strong>and</strong> Be markedly reduced this broaden<strong>in</strong>g.<br />

J.J.Harris, J.B.Clegg, R.B.Beall, J.Castagné, K.Woodbridge, C.Roberts: Journal of<br />

Crystal Growth, 1991, 111[1-4], 239-45<br />

[446-91/92-001]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The mechanism of Si diffusion was studied by us<strong>in</strong>g photolum<strong>in</strong>escence <strong>and</strong> secondary<br />

ion mass spectrometry, <strong>and</strong> transmission electron microscopy across the corner of a<br />

wedge-shaped sample. The diffusion source was a grown-<strong>in</strong> highly Si-doped layer. It was<br />

deduced that Frenkel defects (column-<strong>III</strong> vacancies <strong>and</strong> <strong>in</strong>terstitials) were generated<br />

with<strong>in</strong> the highly Si-doped region. The column-<strong>III</strong> <strong>in</strong>terstitials rapidly diffused towards<br />

the surface, where they reacted with the column-<strong>III</strong> vacancies which were generated at the<br />

surface dur<strong>in</strong>g anneal<strong>in</strong>g <strong>in</strong> a gaseous As ambient. This caused a supersaturation, of<br />

column-<strong>III</strong> vacancies <strong>in</strong> the Si-doped region, which drove Si diffusion. Anneal<strong>in</strong>g <strong>in</strong><br />

vacuum reduced the supersaturation of column-<strong>III</strong> vacancies, <strong>and</strong> thus decreased Si<br />

diffusion. A predom<strong>in</strong>ant Si-donor plus column-<strong>III</strong> vacancy complex emission b<strong>and</strong> was<br />

327


Si <strong>GaAs</strong> Si<br />

found <strong>in</strong> spectra from the Si-diffused region. The results supported the concept of a<br />

vacancy-assisted mechanism for Si diffusion <strong>and</strong> impurity-<strong>in</strong>duced disorder<strong>in</strong>g.<br />

L.Pavesi, N.H.Ky, J.D.Ganière, F.K.Re<strong>in</strong>hart, N.Baba-Ali, I.Harrison, B.Tuck, M.Hen<strong>in</strong>i:<br />

Journal of Applied Physics, 1992, 71[5], 2225-37<br />

[446-86/87-002]<br />

1.0E-11<br />

D (cm 2 /s)<br />

1.0E-12<br />

1.0E-13<br />

1.0E-14<br />

1.0E-15<br />

1.0E-16<br />

table 21<br />

table 22<br />

table 23<br />

table 24<br />

table 25<br />

1.0E-17<br />

1.0E-18<br />

1.0E-19<br />

7 8 9 10 11 12 13 14<br />

10 4 /T(K)<br />

Figure 6: Diffusivity of Si <strong>in</strong> <strong>GaAs</strong><br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The diffusion of Si was <strong>in</strong>vestigated by us<strong>in</strong>g sources from various tie-triangle regions <strong>in</strong><br />

the Si-Ga-As phase diagram. This ensured constant chemical potentials for the 3<br />

components under is<strong>other</strong>mal conditions. The Si profiles were determ<strong>in</strong>ed by us<strong>in</strong>g<br />

secondary ion mass spectrometry, <strong>and</strong> were found to be very different for the various<br />

types of source. The Ga-Si-<strong>GaAs</strong> tie-triangle source produced p-type Si dop<strong>in</strong>g with a<br />

concentration-<strong>in</strong>dependent diffusion coefficient. A neutral As or Ga vacancy was thought<br />

to be the predom<strong>in</strong>ant mobile defect under these conditions. The use of As-rich sources<br />

from 2 tie-triangle regions, or a Si-<strong>GaAs</strong> tie-l<strong>in</strong>e source, produced Si donor diffusion with<br />

a concentration-dependent diffusion behavior. The concentration-dependent diffusion<br />

328


Si <strong>GaAs</strong> Si<br />

coefficients of donor Si for As-rich source diffusion were related to the net ionized donor<br />

concentration, <strong>and</strong> exhibited 3 different regions. These were an <strong>in</strong>tr<strong>in</strong>sic regime, an<br />

<strong>in</strong>termediate regime, <strong>and</strong> a saturation regime. It was proposed that Ga vacancies were<br />

responsible for donor diffusion; with a V 0 -<br />

Ga <strong>and</strong>/or V Ga mechanism for the <strong>in</strong>tr<strong>in</strong>sic<br />

regime <strong>and</strong> a V - Ga -related mechanism for the extr<strong>in</strong>sic <strong>and</strong> saturation regimes. The use of<br />

a Si-<strong>GaAs</strong> tie-l<strong>in</strong>e source produced 2 branch-type profiles which were <strong>in</strong>termediate<br />

between the As-rich <strong>and</strong> Ga-rich diffusion cases.<br />

K.H.Lee, D.A.Stevenson, M.D.Deal: Journal of Applied Physics, 1990, 68[8], 4008-13<br />

[446-86/87-011]<br />

Table 21<br />

Diffusivity of Si <strong>in</strong> <strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

850 2.5 x 10 -15<br />

800 7.8 x 10 -16<br />

750 4.0 x 10 -16<br />

700 2.3 x 10 -16<br />

650 3.7 x 10 -17<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The effect of impurities upon the creation of Ga vacancies <strong>in</strong> Si-doped material, grown on<br />

a Be-doped epilayer by molecular beam epitaxy, was <strong>in</strong>vestigated by means of slow<br />

positron beam measurements. It was found that dop<strong>in</strong>g with Si enhanced the creation of<br />

Ga vacancies. The results supported a theoretical model <strong>in</strong> which the creation of Ga<br />

vacancies was expla<strong>in</strong>ed <strong>in</strong> terms of a change <strong>in</strong> the-Fermi level position due to Si<br />

dop<strong>in</strong>g. It was also suggested that Si atoms diffused as a neutral complex, Si Ga -V Ga , rather<br />

than as Si Ga -Si As . A change <strong>in</strong> the S-parameter distribution at the <strong>in</strong>terface between Sidoped<br />

<strong>and</strong> Be-doped regions was expla<strong>in</strong>ed <strong>in</strong> terms of a so-called Be carry-forward<br />

effect which occurred dur<strong>in</strong>g the growth of Si-doped <strong>GaAs</strong> on a Be-doped epilayer.<br />

J.L.Lee, L.Wei, S.Tanigawa, M.Kawabe: Journal of Applied Physics, 1990, 68[11], 5571-<br />

5<br />

[446-86/87-012]<br />

322 <strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

This dopant diffused extensively after implantation <strong>and</strong> long-term anneal<strong>in</strong>g. The results<br />

could be expla<strong>in</strong>ed by assum<strong>in</strong>g that the diffusivity depended upon the square of the<br />

electron concentration. The dopant diffusion was affected by the presence of implantation<br />

damage; the higher the concentration of extended defects, the slower be<strong>in</strong>g the diffusivity<br />

as compared with the values for conventional diffusion from a solid source. If the sample<br />

was amorphized dur<strong>in</strong>g implantation, extended defects did not form <strong>and</strong> the diffusivity of<br />

the ion was very close to that <strong>in</strong> material which had been diffused from a solid source.<br />

When amorphization did not occur, extended defects formed after implantation, <strong>and</strong><br />

329


Si <strong>GaAs</strong> Si<br />

diffusion was <strong>in</strong>hibited; especially after low doses, <strong>in</strong> the short term, or at low<br />

temperatures. The higher the density of extended defects, the greater was the suppression<br />

of diffusion. The diffusion was time-dependent. It was concluded that the results (table<br />

22) were consistent with a diffusion mechanism <strong>in</strong> which the mobile species was the<br />

donor that was coupled with a charged Ga vacancy. The equilibrium vacancy<br />

concentration was thought to be suppressed by the presence of extended defects <strong>and</strong>/or<br />

excess Ga <strong>in</strong>terstitials which resulted from implantation.<br />

E.L.Allen, J.J.Murray, M.D.Deal, J.D.Plummer, K.S.Jones, W.S.Rubart: Journal of the<br />

Electrochemical Society, 1991, 138[11], 3440-9<br />

[446-84/85-016]<br />

Table 22<br />

Diffusivity of Implanted Si <strong>in</strong> <strong>GaAs</strong><br />

Dose (/cm 2 ) Temperature (C) D (cm 2 /s)<br />

5 x 10 15 950 2.5 x 10 -15<br />

1 x 10 14 900 2.6 x 10 -15<br />

1 x 10 15 900 2.0 x 10 -15<br />

1 x 10 15 900 1.2 x 10 -15<br />

1 x 10 15 900 8.3 x 10 -16<br />

1 x 10 15 900 6.2 x 10 -16<br />

1 x 10 14 900 4.4 x 10 -16<br />

1 x 10 14 900 3.0 x 10 -16<br />

1 x 10 14 900 2.2 x 10 -16<br />

1 x 10 15 900 1.3 x 10 -16<br />

5 x 10 15 850 1.4 x 10 -16<br />

1 x 10 15 850 4.2 x 10 -17<br />

5 x 10 15 750 7.5 x 10 -18<br />

5 x 10 14 750 3.7 x 10 -18<br />

5 x 10 14 750 1.8 x 10 -18<br />

1 x 10 15 750 1.0 x 10 -18<br />

1 x 10 14 850 1.3 x 10 -17<br />

1 x 10 14 900 1.4 x 10 -17<br />

1 x 10 14 900 2.9 x 10 -17<br />

1 x 10 15 900 4.0 x 10 -17<br />

1 x 10 14 900 5.4 x 10 -17<br />

1 x 10 14 950 5.0 x 10 -16<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The behavior of Si, after be<strong>in</strong>g <strong>in</strong>itially d-doped <strong>in</strong>to very pure gas-source molecular beam<br />

epitaxial <strong>GaAs</strong> layers, was studied by us<strong>in</strong>g capacitance-voltage profil<strong>in</strong>g. A non-l<strong>in</strong>ear<br />

330


Si <strong>GaAs</strong> Si<br />

behavior of the diffusivity of Si as a function of the reciprocal temperature was observed.<br />

This was expla<strong>in</strong>ed <strong>in</strong> terms of a 2-component Arrhenius dependence <strong>in</strong> which the<br />

activation energies changed by 1.5eV. When Si diffusion was governed by the lower<br />

activation energy, the impurity profile grew <strong>in</strong> width as a l<strong>in</strong>ear function of the anneal<strong>in</strong>g<br />

time. Deviations of the measured Si diffusivity from classical impurity diffusion behavior<br />

were attributed to the existence of a non-equilibrium concentration of vacancies which<br />

was generated at the d-source position dur<strong>in</strong>g anneal<strong>in</strong>g.<br />

J.E.Cunn<strong>in</strong>gham, T.H.Chui, W.Jan, T.Y.Kuo: Applied Physics Letters, 1991, 59[12],<br />

1452-4<br />

[446-84/85-016]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Samples which had been implanted with 220keV Si ions to doses rang<strong>in</strong>g from 3 x 10 13 to<br />

10 15 /cm 2 , <strong>and</strong> annealed at 850C, were studied. By us<strong>in</strong>g transmission electron<br />

microscopy, voids were observed <strong>in</strong> samples with implanted doses of more than 3 x<br />

10 14 /cm 2 ; after anneal<strong>in</strong>g times which were as short as 5s. In the same region where voids<br />

were found, capacitance-voltage measurements revealed abnormally low electron<br />

concentrations. Also <strong>in</strong> the same region, secondary ion mass spectrometry measurements<br />

detected anomalies <strong>in</strong> the Si concentration profiles. It was therefore deduced that Si<br />

redistribution had occurred. At high Si doses, the onset of void formation, the abnormally<br />

low electron concentration, <strong>and</strong> the Si accumulation anomaly were concurrent. On the<br />

basis of the results, it was concluded that voids <strong>in</strong>hibited electrical activity <strong>and</strong> led to the<br />

Si diffusion anomaly.<br />

S.Chen, S.T.Lee, G.Braunste<strong>in</strong>, K.Y.Ko, L.R.Zheng, T.Y.Tan: Japanese Journal of<br />

Applied Physics, 1990, 29[11], L1950-3<br />

[446-81/82-010]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The growth of Si-doped liquid-encapsulated Czochralski material exhibited a significant<br />

deviation, <strong>in</strong> Si <strong>in</strong>corporation, from that which was predicted by a classical segregation<br />

model. It was usually expected that, for a given impurity segregation coefficient, dopant<br />

<strong>in</strong>corporation throughout the crystal could be calculated with fairly good accuracy.<br />

Profiles for Te-doped liquid-encapsulated Czochralski material gave a close<br />

approximation to this classical model. However, <strong>in</strong> the case of Si-doped material, the<br />

dopant distribution <strong>in</strong> the crystals deviated significantly from the segregation model. In<br />

some cases, this deviation amounted to several orders of magnitude. The degree of<br />

deviation was found to depend upon the growth conditions. The present work was<br />

performed <strong>in</strong> order to underst<strong>and</strong> the source of the deviation from the model <strong>and</strong> to permit<br />

accurate account to be taken of Si <strong>in</strong>corporation. Therefore, Si-doped crystals were grown<br />

by us<strong>in</strong>g a low-pressure liquid-encapsulated Czochralski technique <strong>and</strong> were <strong>in</strong>tended to<br />

be doped with between 10 16 <strong>and</strong> 10 18 /cm 3 . The actual dopant <strong>in</strong>corporation was<br />

significantly lower than that predicted by the segregation model, <strong>and</strong> the axial dopant<br />

variation was much too flat. Also, <strong>in</strong> spite of the use of <strong>in</strong>tentional Si dop<strong>in</strong>g, a significant<br />

number of the crystals was semi-<strong>in</strong>sulat<strong>in</strong>g. An analysis of the experimental data showed<br />

331


Si <strong>GaAs</strong> Si<br />

that this dop<strong>in</strong>g behavior could not be expla<strong>in</strong>ed by assum<strong>in</strong>g the existence of an<br />

unknown compensat<strong>in</strong>g impurity. A different model was developed which was consistent<br />

with all of the observations <strong>and</strong> which provided an accurate account of Si <strong>in</strong>corporation.<br />

An underst<strong>and</strong><strong>in</strong>g of the Si <strong>in</strong>corporation anomaly permitted successful process changes<br />

<strong>and</strong> improvements to be made. This model expla<strong>in</strong>ed the observed anomaly by tak<strong>in</strong>g<br />

account of Si diffusion between the <strong>GaAs</strong> melt <strong>and</strong> the B 2 O 3 encapsulant, as well as of<br />

permanent Si trapp<strong>in</strong>g <strong>in</strong> the B 2 O 3 . When the <strong>in</strong>teraction of Si with B 2 O 3 was taken <strong>in</strong>to<br />

account, the Si <strong>in</strong>corporation obeyed the segregation model <strong>and</strong> the anomaly vanished.<br />

A.Flat: Journal of Crystal Growth, 1991, 109[1-4], 224-7<br />

[446-81/82-011]<br />

Table 23<br />

Diffusivity of Si <strong>in</strong> <strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

1000 1.5 x 10 -12<br />

950 4.1 x 10 -13<br />

900 3.1 x 10 -14<br />

800 5.7 x 10 -15<br />

705 1.0 x 10 -16<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The lateral variation <strong>in</strong> the emission energy of <strong>GaAs</strong> which had been masklessly grown<br />

on V-grooved Si was studied by us<strong>in</strong>g cathodolum<strong>in</strong>escence wavelength imag<strong>in</strong>g. This<br />

new experimental approach permitted, for the first time, the direct visualization <strong>and</strong><br />

quantification of the extreme homogeneity of this novel growth mode <strong>and</strong> of the lateral<br />

variations of Si impurity <strong>in</strong>corporation <strong>in</strong>to such semiconductor microstructures. It was<br />

thus an important method for characteriz<strong>in</strong>g micro-patterned opto-electronic monolithic<br />

<strong>in</strong>tegrated circuits.<br />

M.Grundmann, J.Christen, D.Bimberg, A.Hashimoto, T.Fukunaga, N.Watanabe: Applied<br />

Physics Letters, 1991, 58[19], 2090-2<br />

[446-81/82-012]<br />

323 <strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Capacitance-voltage methods were used to profile d-doped <strong>GaAs</strong> layers which had been<br />

grown on Si substrates via metalorganic chemical vapor deposition. It was found that<br />

there was a close correlation between dislocation densities, <strong>in</strong> the epitaxial layers, <strong>and</strong> the<br />

associated diffusion coefficients. After rapid thermal anneal<strong>in</strong>g (800-1000C, 7s), the<br />

diffusion data (table 23) could be described by:<br />

D(cm 2 /s) = 30 exp[-3.4(eV)/kT]<br />

332


Si <strong>GaAs</strong> Si<br />

for a relatively thick buffer layer of 0.0033mm. It was concluded that the dislocationenhanced<br />

diffusion of Si impurities was appreciable, <strong>and</strong> that the <strong>in</strong>clusion of an<br />

0.003mm buffer layer was <strong>in</strong>sufficient to prevent the diffusion of impurities.<br />

Y.Kim, M.S.Kim, S.K.M<strong>in</strong>, C.Lee: Journal of Applied Physics, 1991, 69[3], 1355-8<br />

[446-78/79-014]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

High depth-resolution secondary ion mass spectrometry profil<strong>in</strong>g was used to <strong>in</strong>vestigate<br />

the broaden<strong>in</strong>g of d-doped planes of Si plus Be <strong>in</strong> material which had been prepared by<br />

us<strong>in</strong>g molecular beam epitaxial methods. It was confirmed that concentration-dependent<br />

diffusion was the predom<strong>in</strong>ant broaden<strong>in</strong>g process for Be at growth temperatures of less<br />

than 600C. By <strong>in</strong>corporat<strong>in</strong>g Si atoms <strong>in</strong>to the same plane, it was shown that the<br />

broaden<strong>in</strong>g could be completely <strong>in</strong>hibited. This suggested that the rapid diffusion process<br />

resulted from mutual repulsion between the Be Ga - ions, <strong>and</strong> was prevented by the reverse<br />

field which arose from Si Ga + ions or by the formation of low-mobility Si Ga + -Be Ga<br />

-<br />

complexes. The rapid diffusion of Si as Si Ga -Si As pairs was also reduced. The latter was<br />

attributed to a Fermi-level effect, with compensation by Be tend<strong>in</strong>g to reduce the<br />

probability of Si As formation. The surface segregation of Si was unaffected, whereas that<br />

of Be was reduced. This <strong>in</strong>dicated that the surface fields which existed dur<strong>in</strong>g growth<br />

contributed to the behavior of Be, but not to that of Si.<br />

J.J.Harris, J.B.Clegg, R.B.Beall, J.Castagné: Semiconductor Science <strong>and</strong> Technology,<br />

1990, 5[7], 785-8<br />

[446-76/77-007]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

A quantitative model was proposed for the behavior of Si. The model took account of the<br />

fact that Si was an amphoteric impurity which acted as a shallow donor when it occupied<br />

the Ga site (Si Ga + ) <strong>and</strong> as a shallow acceptor when it occupied the As site (Si As - ). Both<br />

Si Ga + <strong>and</strong> Si As - existed at high concentrations. The amphoteric behavior of Si could be<br />

viewed as be<strong>in</strong>g an effect of the Fermi level. It was assumed that Si Ga + <strong>and</strong> Si As - diffused<br />

on the Ga <strong>and</strong> As sub-lattices, respectively. Thus, the diffusion of Si Ga + <strong>and</strong> Si As - was<br />

governed by the variously charged Ga vacancies <strong>and</strong> As vacancies (or self-<strong>in</strong>terstitials)<br />

respectively. The experimentally observed Si diffusivity was concentration dependent,<br />

<strong>and</strong> this was attributed to the amphoteric nature of Si as well as to an<strong>other</strong> effect of the<br />

Fermi level. The latter <strong>in</strong>volved its <strong>in</strong>fluence upon the concentrations of charged po<strong>in</strong>tdefect<br />

species. Satisfactory quantitative descriptions of available experimental data were<br />

obta<strong>in</strong>ed. An analysis of results on the diffusion of Si <strong>in</strong>to a Sn-doped <strong>GaAs</strong> substrate<br />

suggested that a previously proposed Si-pair diffusion model was unfavorable.<br />

S.Yu, U.M.Gösele, T.Y.Tan: Journal of Applied Physics, 1989, 66[7], 2952-61<br />

[446-74-012]<br />

333


Si <strong>GaAs</strong> Si<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

A model for Si diffusion was developed which was based upon the formation of Si Ga + -<br />

V Ga - pairs. By us<strong>in</strong>g recent data on the diffusivity of Ga vacancies, it was shown that the<br />

pair diffusion coefficient was approximately equal to 75% of the former value. The<br />

predictions of the model were found to be <strong>in</strong> good agreement with Si diffusion data.<br />

K.B.Kahen: Journal of Applied Physics, 1989, 66[12], 6176-8<br />

[446-74-012]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The diffusion of Si was studied after implant<strong>in</strong>g Si <strong>in</strong>to un-doped, Se-, Si- or Zn-doped<br />

material. The diffusion of grown-<strong>in</strong> Si <strong>in</strong> epitaxial layer structures was also studied. No<br />

diffusion was detected <strong>in</strong> un-doped or Zn-doped material, while a moderate amount of<br />

diffusion was detected <strong>in</strong> Si-doped samples. A significant amount of diffusion occurred <strong>in</strong><br />

Se-doped material <strong>and</strong> <strong>in</strong> non-implanted Si-doped epitaxial samples. The results <strong>in</strong>dicated<br />

that diffusion was controlled by a Fermi-level mechanism which probably <strong>in</strong>volved<br />

ionized Ga vacancies, <strong>and</strong> that implantation damage <strong>in</strong>hibited diffusion by keep<strong>in</strong>g the<br />

electron concentration <strong>and</strong>/or the ionized Ga vacancy concentration at a low level.<br />

J.J.Murray, M.D.Deal, D.A.Stevenson: Applied Physics Letters, 1990, 56[5], 472-4<br />

[446-74-012]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The effect of the substrate temperature, dur<strong>in</strong>g molecular beam epitaxial growth, upon the<br />

migration of Si atoms <strong>in</strong> d-doped or planar-doped <strong>GaAs</strong> was <strong>in</strong>vestigated by us<strong>in</strong>g<br />

secondary ion mass spectrometry. The results for d-doped <strong>GaAs</strong> revealed a measurable<br />

spread of Si, which <strong>in</strong>creased by about 8nm when the substrate temperature was <strong>in</strong>creased<br />

from 580 to 640C. For substrate temperatures below 580C, the measured width of the Si<br />

profile was limited by the resolution of the secondary ion mass spectrometer.<br />

Magnetotransport measurements were also performed <strong>in</strong> order to determ<strong>in</strong>e dopant<br />

spread<strong>in</strong>g. The Si migration which was measured by means of secondary ion mass<br />

spectrometry was <strong>in</strong> qualitative agreement with the transport results. However, the<br />

secondary ion mass spectrometry data <strong>in</strong>dicated larger Si areal densities. Two<br />

mechanisms, auto-compensation <strong>and</strong> electron localization by a DX center, were believed<br />

to be responsible for the latter observations.<br />

A.M.Lanzillotto, M.Santos, M.Shayegan: Applied Physics Letters, 1989, 55[14], 1445-7<br />

[446-72/73-002]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Electron-beam deposited films of phosphosilicate were used for the encapsulation of Siimplanted<br />

material dur<strong>in</strong>g post-implantation anneal<strong>in</strong>g. Depth profiles which were <strong>in</strong><br />

excellent agreement with the L<strong>in</strong>dhard-Scharff-Schiott curves were obta<strong>in</strong>ed by us<strong>in</strong>g<br />

100nm thick films <strong>and</strong> by anneal<strong>in</strong>g for 0.5h at 850C. The diffusion coefficient of<br />

334


Si <strong>GaAs</strong> Si<br />

implanted Si was found to be an order of magnitude smaller for phosphosilicate films<br />

than for conventional plasma-assisted chemical vapor deposited SiO 2 films.<br />

S.S<strong>in</strong>gh, F.Baiocchi, A.D.Butherus, W.H.Grodkiewicz, B.Schwartz, L.G.Van Uitert,<br />

L.Yesis, G.J.Zydzik: Journal of Applied Physics, 1988, 64[8], 4194-8<br />

[446-72/73-011]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Layers were grown, by molecular beam epitaxy, at a substrate temperature of 520C. The<br />

layers conta<strong>in</strong>ed three d-doped planes, with Si concentrations of 4 x 10 12 , 10 13 or 4 x<br />

10 13 /cm 3 , <strong>and</strong> were annealed at temperatures of up to 648C. Secondary ion mass<br />

spectroscopy <strong>and</strong> capacitance-voltage methods were then used to monitor broaden<strong>in</strong>g of<br />

the profiles. It was found that the most lightly doped plane gave a near-Gaussian profile,<br />

<strong>and</strong> the diffusion coefficient was comparable with published data on the simple diffusion<br />

of isolated Si Ga atoms. The more highly doped planes exhibited complex profiles with 2<br />

components; <strong>in</strong> which some of the atoms were conf<strong>in</strong>ed to the orig<strong>in</strong>al plane, while there<br />

was an essentially square-shaped profile of fast-diffus<strong>in</strong>g atoms. A comparison of the 2<br />

types of experimental data suggested that the formation of Si isl<strong>and</strong>s took place dur<strong>in</strong>g<br />

deposition of the d-doped plane. This gave rise to electrically <strong>in</strong>active atoms which could<br />

then diffuse <strong>in</strong>to the surround<strong>in</strong>g material dur<strong>in</strong>g heat treatment.<br />

R.B.Beall, J.B.Clegg, J.Castagné, J.J.Harris, R.Murray, R.C.Newman: Semiconductor<br />

Science <strong>and</strong> Technology, 1989, 4[12], 1171-5<br />

[446-72/73-011]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Quantum oscillations <strong>in</strong> the magnetoresistance of Si-doped material were analyzed <strong>in</strong><br />

order to obta<strong>in</strong> the electron densities of the electrical sub-b<strong>and</strong>s. These densities were<br />

compared with the results for self-consistently calculated sub-b<strong>and</strong> structures of d-doped<br />

material <strong>in</strong> which the spread of Si dopant atoms <strong>in</strong> the growth direction was a fitt<strong>in</strong>g<br />

parameter. The results <strong>in</strong>dicated that there was a negligible spread <strong>in</strong> structures which<br />

were grown at substrate temperatures of up to 530C. In structures which were grown at<br />

higher substrate temperatures, there was a measurable spread. This <strong>in</strong>creased with<br />

<strong>in</strong>creas<strong>in</strong>g substrate temperature. At a substrate temperature of 640C, the Si spread was<br />

about 22nm. An exam<strong>in</strong>ation of the three-dimensional Si densities <strong>in</strong> these layers<br />

<strong>in</strong>dicated that the predom<strong>in</strong>ant mechanism of Si spread<strong>in</strong>g at substrate temperatures<br />

greater than 600C was the Si migration which was necessary <strong>in</strong> order to satisfy the solid<br />

solubility limit.<br />

M.Santos, T.Sajoto, A.Zrenner, M.Shayegan: Applied Physics Letters, 1988, 53[25],<br />

2504-6<br />

[446-64/65-163]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

A review was presented of self-diffusion mechanisms <strong>and</strong> dop<strong>in</strong>g-enhanced superlattice<br />

disorder<strong>in</strong>g. The mechanism of enhanced superlattice disorder<strong>in</strong>g due to Si dop<strong>in</strong>g was a<br />

Fermi-level effect which <strong>in</strong>creased the concentrations of charged po<strong>in</strong>t defects. The<br />

335


Si <strong>GaAs</strong> Si<br />

diffusion of Si appeared to be governed by Ga vacancies, <strong>and</strong> was well described by<br />

current Si pair diffusion models. It was concluded that dislocations <strong>in</strong> this material <strong>and</strong> <strong>in</strong><br />

<strong>other</strong> <strong>III</strong>-V compounds were only moderately efficient s<strong>in</strong>ks or sources for po<strong>in</strong>t defects.<br />

T.Y.Tan, U.Gösele: Materials Science <strong>and</strong> Eng<strong>in</strong>eer<strong>in</strong>g, 1988, B1, 47-65<br />

[446-62/63-208]<br />

Table 24<br />

Diffusivity of Si <strong>in</strong> <strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

995 7.8 x 10 -14<br />

940 3.4 x 10 -14<br />

890 1.2 x 10 -14<br />

800 1.3 x 10 -15<br />

685 9.5 x 10 -17<br />

590 1.2 x 10 -17<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The migration of atomic Si <strong>in</strong> d-doped material was studied by means of capacitancevoltage<br />

measurements <strong>and</strong> rapid thermal anneal<strong>in</strong>g. It was shown that these methods<br />

could detect diffusion which occurred at a length scale as small as 1nm. The capacitancevoltage<br />

profile widths broadened from less than 4nm, to 13.7nm, upon anneal<strong>in</strong>g (1000C,<br />

5s). It was found that the results could be described by:<br />

D(cm 2 /s) = 0.0004 exp[-2.45(eV)/kT]<br />

E.F.Schubert, T.H.Chiu, J.E.Cunn<strong>in</strong>gham, B.Tell, J.B.Stark: Journal of Electronic<br />

Materials, 1988, 17[6], 527-31<br />

[446-62/63-210]<br />

324 <strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Impurities, with an <strong>in</strong>itially Dirac d-like distribution profile, were diffused <strong>in</strong>to <strong>GaAs</strong> by<br />

us<strong>in</strong>g rapid thermal anneal<strong>in</strong>g. The diffusion of atomic Si was monitored via a novel<br />

method which <strong>in</strong>volved compar<strong>in</strong>g the experimental capacitance-voltage profiles with<br />

predicted self-consistent profiles. The capacitance-voltage profiles broadened dur<strong>in</strong>g<br />

rapid thermal anneal<strong>in</strong>g (1000C, 5s). It was found that the diffusion data (table 24) could<br />

be described by:<br />

D(cm 2 /s) = 0.0004 exp[-2.45(eV)/kT]<br />

The diffusivity values were 2 orders of magnitude smaller than those for Si-pair diffusion<br />

<strong>in</strong> <strong>GaAs</strong>.<br />

E.F.Schubert, J.B.Stark, T.H.Chiu, B.Tell: Applied Physics Letters, 1988, 53[4], 293-5<br />

[446-62/63-210]<br />

336


Si <strong>GaAs</strong> Si<br />

Table 25<br />

Diffusivity of Si <strong>in</strong> <strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

960 7.5 x 10 -15<br />

955 4.7 x 10 -15<br />

895 4.4 x 10 -15<br />

900 2.0 x 10 -15<br />

795 2.0 x 10 -15<br />

790 1.1 x 10 -15<br />

695 1.0 x 10 -15<br />

695 5.1 x 10 -16<br />

645 1.9 x 10 -16<br />

590 7.8 x 10 -17<br />

535 4.9 x 10 -17<br />

605 3.7 x 10 -17<br />

485 2.0 x 10 -17<br />

590 1.9 x 10 -17<br />

555 8.2 x 10 -18<br />

515 1.0 x 10 -18<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

High-quality samples were prepared by means of chemical beam epitaxy, at a substrate<br />

temperature of 500C, by us<strong>in</strong>g triethylgallium <strong>and</strong> ars<strong>in</strong>e. Capacitance-voltage<br />

measurements of Si-doped material revealed profile widths of 2.2nm at 300K <strong>and</strong> 1.8nm<br />

at 77K. This <strong>in</strong>dicated that a high degree of Si spatial localization had been achieved.<br />

Subsequent anneal<strong>in</strong>g treatments showed that appreciable Si segregation <strong>and</strong> diffusion<br />

occurred at a growth temperature of about 600C. The capacitance-voltage widths of<br />

annealed doped structures provided a good estimate of the diffusion coefficient of Si <strong>in</strong><br />

<strong>GaAs</strong>.<br />

T.H.Chiu, J.E.Cunn<strong>in</strong>gham, B.Tell, E.F.Schubert: Journal of Applied Physics, 1988,<br />

64[3], 1578-80<br />

[446-61-067]<br />

325 <strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Planar conf<strong>in</strong>ement, diffusion, <strong>and</strong> surface segregation results were reported for Si which<br />

had been d-doped <strong>in</strong>to <strong>GaAs</strong>. In the case of gas-source molecular beam epitaxy, the Si<br />

diffusion (table 25) as a function of the reciprocal anneal<strong>in</strong>g temperature exhibited an<br />

337


Si <strong>GaAs</strong> Si<br />

unique 2-component Arrhenius form <strong>in</strong> which the activation energies differed by the<br />

fundamental <strong>GaAs</strong> b<strong>and</strong>-gap energy of 1.5eV.<br />

J.E.Cunn<strong>in</strong>gham, T.H.Chiu, B.Tell, W.Jan: Journal of Vacuum Science <strong>and</strong> Technology<br />

B, 1990, 8[2], 157-9<br />

[446-74-012]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Secondary ion mass spectroscopy <strong>and</strong> carrier concentration measurements were used to<br />

characterize Si diffusion <strong>in</strong>to <strong>GaAs</strong> wafers which conta<strong>in</strong>ed 2 fundamentally different<br />

types of donor. These were column-IV donors (Si, Sn) <strong>and</strong> column-VI donors (Se, Te). A<br />

decrease <strong>in</strong> the Si diffusion rate was found <strong>in</strong> material which conta<strong>in</strong>ed column-VI donors<br />

as compared with that which conta<strong>in</strong>ed column-IV donors. This trend was consistent with<br />

a model <strong>in</strong> which Si diffused as donor Ga-vacancy complexes. The decrease <strong>in</strong> the Si<br />

diffusion coefficient was attributed to the greater b<strong>in</strong>d<strong>in</strong>g energy of column-VI donor Gavacancy<br />

nearest-neighbor complexes. This reduced the numbers of free Ga vacancies<br />

which were available to complex with the Si.<br />

D.G.Deppe, N.Holonyak, J.E.Baker: Applied Physics Letters, 1988, 52[2], 129-31<br />

[446-60-003]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Data were presented which demonstrated that the surface encapsulant <strong>and</strong> As 4 overpressure<br />

strongly affected Si. An <strong>in</strong>crease <strong>in</strong> the As 4 over-pressure resulted <strong>in</strong> an <strong>in</strong>crease<br />

<strong>in</strong> the diffusion depth. No b<strong>and</strong>-edge exciton was observed dur<strong>in</strong>g absorption on material<br />

that was diffused with Si, <strong>in</strong> spite of the high degree of compensation.<br />

L.J.Guido, W.E.Plano, D.W.Nam, N.Holonyak, J.E.Baker, R.D.Burnham, P.Gavrilovic:<br />

Journal of Electronic Materials, 1988, 17[1], 53-6<br />

[446-60-004]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The diffusion profiles of buried Si dopant which had been implanted by us<strong>in</strong>g a focussed<br />

ion beam were determ<strong>in</strong>ed after anneal<strong>in</strong>g. The diffusion coefficient of the Si was<br />

determ<strong>in</strong>ed by fitt<strong>in</strong>g the results of computer calculations. Concentration-dependent<br />

diffusion of Si which had been <strong>in</strong>troduced by us<strong>in</strong>g a molecular beam was observed.<br />

However, the diffusion coefficient of Si which had been <strong>in</strong>troduced by us<strong>in</strong>g a focussed<br />

ion beam was undetectably small when compared with that of Be at 850C.<br />

T.Morita, J.Kobayashi, T.Takamori, A.Takamori, E.Miyauchi, H.Hashimoto: Japanese<br />

Journal of Applied Physics, 1987, 26[8], 1324-7<br />

[446-55/56-005]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Junction-depth measurements, performed us<strong>in</strong>g scann<strong>in</strong>g electron microscopy <strong>and</strong><br />

secondary ion mass spectroscopy, were used to characterize Si diffusion <strong>in</strong> <strong>GaAs</strong> crystals<br />

which conta<strong>in</strong>ed various amounts of Zn background dop<strong>in</strong>g. The Zn concentration was<br />

338


Si <strong>GaAs</strong> Si<br />

found to control Si diffusion. This behavior was attributed to a shift <strong>in</strong> the Fermi level<br />

with <strong>in</strong>creas<strong>in</strong>g n-type dop<strong>in</strong>g. Also, the electric field which was due to the p-n junction<br />

that formed at the Si diffusion front had a large effect upon the Zn background dop<strong>in</strong>g<br />

profile.<br />

D.G.Deppe, N.Holonyak, F.A.Kish, J.E.Baker: Applied Physics Letters, 1987, 50[15],<br />

998-1000<br />

[446-51/52-116]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

A new method for self-aligned Si-Zn diffusion was described. In this method, closed-tube<br />

Si diffusion was carried out by us<strong>in</strong>g a sputtered SiN x film. Then, Zn diffusion which was<br />

self-aligned to the Si diffusion w<strong>in</strong>dow was carried out by re-us<strong>in</strong>g the SiN x film as a<br />

mask. The key factor was that the SiN x film should have the correct refractive <strong>in</strong>dex<br />

profile.<br />

W.X.Zou, R.Boudreau, H.T.Han, T.Bowen, S.S.Shi, D.S.L.Mui, J.L.Merz: Journal of<br />

Applied Physics, 1995, 77[12], 6244-6<br />

[446-121/122-045]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The reported anomalous so-called up-hill diffusion behavior of implanted Si <strong>in</strong> this<br />

material was simulated. The up-hill diffusion had been found to be implantation-dose<br />

dependent. No up-hill diffusion was observed below a threshold dose of 3 x 10 14 /cm 2 . It<br />

was suggested here that the anomalous behavior was due to the formation of vacancy<br />

clusters when the implantation dose was sufficiently high. Atoms of Si were assumed to<br />

migrate through the Ga vacancies, which were comparatively mobile. On the <strong>other</strong> h<strong>and</strong>,<br />

the immobile clusters collected Si atoms around the vacancy peak; thus result<strong>in</strong>g <strong>in</strong> uphill<br />

diffusion. A dip <strong>in</strong> the carrier concentration profile also occurred at high implantation<br />

doses. The occurrence of this dip was attributed to the effect of the trapp<strong>in</strong>g of Si atoms<br />

<strong>in</strong>to clusters where the former were not electrically activated.<br />

V.C.Lo, J.Z.Sun: Modell<strong>in</strong>g <strong>and</strong> Simulation <strong>in</strong> Materials Science <strong>and</strong> Eng<strong>in</strong>eer<strong>in</strong>g, 1996,<br />

4[6], 613-21<br />

[446-157/159-339]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The diffusion of Si was carried out (900C, 5h, As pressure), from a 50nm sputtered film<br />

<strong>and</strong> <strong>in</strong>to undoped semi-<strong>in</strong>sulat<strong>in</strong>g material or Te-doped or Zn-doped liquid-encapsulated<br />

Czochralski material. Secondary ion mass spectroscopy <strong>and</strong> spread<strong>in</strong>g resistance<br />

techniques were used to characterize the Si <strong>in</strong>-diffusion profiles. Lattice defects <strong>in</strong> highly<br />

Si-doped diffusion regions were studied as a function of the post-diffusion heat treatment<br />

(700C, 0.25h; 1000C, 0.5h) via the transmission electron microscopy of plan-view <strong>and</strong><br />

cross-sectional samples. Two types of defect were observed <strong>in</strong> the diffused region. These<br />

were perfect prismatic loops of <strong>in</strong>terstitial type on {110} planes, <strong>and</strong> Frank faulted loops<br />

on {111}; aga<strong>in</strong> of <strong>in</strong>terstitial type. Defect formation, <strong>and</strong> the role of Si <strong>in</strong> defect<br />

generation, were expla<strong>in</strong>ed <strong>in</strong> terms of a negative temperature dependence of the thermal<br />

339


Si <strong>GaAs</strong> Si<br />

equilibrium concentrations of V Ga 3- ; which were assumed to mediate Si diffusion under<br />

highly n-doped conditions. Cathodolum<strong>in</strong>escence spectra at 4 <strong>and</strong> 77K were obta<strong>in</strong>ed<br />

from the diffusion layer. It was found that the Si diffusion affected the b<strong>and</strong>-gap<br />

lum<strong>in</strong>escence <strong>and</strong> generated 2 deep-level emission b<strong>and</strong>s <strong>in</strong> the 0.9 to 1.3eV spectral<br />

region. It was suggested that these deep levels were associated with diffusion-<strong>in</strong>duced<br />

defects <strong>and</strong> defect complexes.<br />

L.Herzog, U.Egger, O.Breitenste<strong>in</strong>, H.G.Hettwer: Materials Science <strong>and</strong> Eng<strong>in</strong>eer<strong>in</strong>g B,<br />

1995, 30[1], 43-53<br />

[446-134/135-125]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The migration of Si acceptors (Si on As sites) <strong>in</strong> δ-doped <strong>GaAs</strong> which had been grown<br />

onto <strong>GaAs</strong>(111)A, was <strong>in</strong>vestigated by means of secondary ion mass spectrometry. It was<br />

found that the diffusion parameters for <strong>GaAs</strong>(111)A differed from those for <strong>GaAs</strong>(001).<br />

The diffusion coefficient <strong>in</strong> <strong>GaAs</strong>(111)A was smaller than that <strong>in</strong> <strong>GaAs</strong>(001), <strong>and</strong> the<br />

activation energy <strong>in</strong> <strong>GaAs</strong>(111)A was higher than that <strong>in</strong> <strong>GaAs</strong>(001). The diffusion<br />

mechanism of Si <strong>in</strong> <strong>GaAs</strong>(111)A was <strong>in</strong>vestigated by means of photolum<strong>in</strong>escence <strong>and</strong> it<br />

was found that, <strong>in</strong> p-type layers, Si-donors (Si on Ga sites) diffused easily to As sites. The<br />

data on Si acceptor diffusion could be described by:<br />

D(cm 2 /s) = 0.0114 exp[-2.74(eV)/kT]<br />

The results <strong>in</strong>dicated that Si-acceptors were more stable than Si donors.<br />

M.Hirai, H.Ohnishi, K.Fujita, P.Vaccaro, T.Watanabe: Journal of Crystal Growth, 1995,<br />

150[1-4], 209-13<br />

[446-127/128-120]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

An <strong>in</strong>vestigation was made of the fast diffusion of Si from deposited surface layers when<br />

oxidized <strong>in</strong> an Ar/H 2 O ambient. This revealed the presence of excess As which was<br />

formed by the oxidation of Ga which orig<strong>in</strong>ated from the substrate, <strong>and</strong> was used to<br />

expla<strong>in</strong> the enhanced diffusion of Si <strong>in</strong>to the substrate. The formation of SiO 2 on the<br />

surface dur<strong>in</strong>g oxidation prevented the loss of the excess As, which then accumulated <strong>in</strong><br />

the rema<strong>in</strong><strong>in</strong>g Si film. The use of higher H 2 O partial pressures dur<strong>in</strong>g oxidation produced<br />

higher As/Si ratios; thus result<strong>in</strong>g <strong>in</strong> an <strong>in</strong>crease <strong>in</strong> Si diffusion depth <strong>and</strong> concentration.<br />

However, the n-type carrier concentration decreased with <strong>in</strong>creas<strong>in</strong>g As/Si ratios <strong>in</strong> the<br />

rema<strong>in</strong><strong>in</strong>g Si layer. A second non-oxidiz<strong>in</strong>g anneal<strong>in</strong>g treatment, of samples from which<br />

the SiO 2 <strong>and</strong> Si layer had been removed, had differ<strong>in</strong>g effects upon the carrier<br />

concentration; depend<strong>in</strong>g upon whether the As was free to escape from the substrate. The<br />

results <strong>in</strong>dicated that excess As-related defects, such as Ga vacancies, were probably<br />

responsible for n-type compensation <strong>in</strong> fast diffused samples.<br />

R.C.Keller, C.R.Helms: Applied Physics Letters, 1995, 67[3], 398-400<br />

[446-123/124-162]<br />

340


Si <strong>GaAs</strong> Si<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The diffusion of Si was studied by us<strong>in</strong>g secondary ion mass spectroscopic <strong>and</strong><br />

transmission electron microscopic methods after implant<strong>in</strong>g it us<strong>in</strong>g energies rang<strong>in</strong>g<br />

from 20 to 200keV <strong>and</strong> doses rang<strong>in</strong>g from 10 13 to 10 14 /cm 2 , followed by furnace<br />

anneal<strong>in</strong>g. It was found that little or no diffusion occurred after implantation at energies<br />

greater than 100keV. At energies of less than 100keV, there was usually appreciable<br />

dopant redistribution; regardless of the peak implant concentration. Both concentrationdependent<br />

<strong>and</strong> concentration-<strong>in</strong>dependent diffusion was observed. The dislocation loop<br />

density varied <strong>in</strong>versely with the amount of diffusion as a function of implantation<br />

energy. A st<strong>and</strong>ard Monte Carlo computer program was able to predict the trends <strong>in</strong> the<br />

implant energy dependence of diffusion by consider<strong>in</strong>g the excess po<strong>in</strong>t defect content<br />

which was produced by implantation. The effect of this excess defect dose <strong>and</strong> of surface<br />

effects upon Si diffusion was consistent with vacancy-assisted diffusion.<br />

C.C.Lee, M.D.Deal, K.S.Jones, H.G.Rob<strong>in</strong>son, J.C.Bravman: Journal of the<br />

Electrochemical Society, 1994, 141[8], 2245-9<br />

[446-119/120-192]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Implantation of Si <strong>in</strong>to (x11)A-oriented semi-<strong>in</strong>sulat<strong>in</strong>g <strong>GaAs</strong> substrates (where x took<br />

values of up to 4) was carried out. For comparison, (110)- <strong>and</strong> (100)-oriented substrates<br />

were also implanted. No <strong>in</strong>-diffusion of Si was observed after anneal<strong>in</strong>g substrates with<br />

any orientation. A similar behavior was observed for Si implants <strong>in</strong> <strong>GaAs</strong> <strong>and</strong> for Si/B coimplants.<br />

M.V.Rao, H.B.Dietrich, P.B.Kle<strong>in</strong>, A.Fathimulla, D.S.Simons, P.H.Chi: Journal of<br />

Applied Physics, 1994, 75[12], 7774-8<br />

[446-117/118-166]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Dopant diffusion <strong>and</strong> defect formation were studied as a function of implantation<br />

temperature <strong>in</strong> Si-implanted material. It was found that the diffusion of Si dur<strong>in</strong>g postimplantation<br />

anneal<strong>in</strong>g decreased by a factor of 2.5 as the implantation temperature was<br />

<strong>in</strong>creased from -2 to 40C. With<strong>in</strong> the same temperature range, the maximum depth <strong>and</strong><br />

density of extr<strong>in</strong>sic dislocation loops <strong>in</strong>creased by factors of 3 <strong>and</strong> 4, respectively.<br />

Rutherford back-scatter<strong>in</strong>g channell<strong>in</strong>g measurements <strong>in</strong>dicated that Si-implanted <strong>GaAs</strong><br />

underwent an amorphous to crystall<strong>in</strong>e transition at Si implantation temperatures of<br />

between -51 <strong>and</strong> 40C. A unified explanation was proposed, for the effects of implantation<br />

temperature upon diffusion <strong>and</strong> dislocation formation, which was based upon known<br />

differences <strong>in</strong> sputter yield between crystall<strong>in</strong>e <strong>and</strong> amorphous semiconductors. The<br />

model assumed that the sputter yield was enhanced by amorphization at lower<br />

temperatures; which <strong>in</strong>creased the excess vacancy concentration. Estimates of the latter<br />

341


Si <strong>GaAs</strong> Si<br />

were obta<strong>in</strong>ed by simulat<strong>in</strong>g the diffusion profiles, <strong>and</strong> were found to be quantitatively<br />

consistent with sputter yield enhancement.<br />

H.G.Rob<strong>in</strong>son, T.E.Haynes, E.L.Allen, C.C.Lee, M.D.Deal, K.S.Jones: Journal of<br />

Applied Physics, 1994, 76[8], 4571-5<br />

[446-117/118-166]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

A study was made of the diffusion of Si <strong>in</strong> δ-doped layers of (111)A- or (100)-type, <strong>and</strong><br />

of the evaporation of As atoms from the surfaces. It was found that the diffusion of<br />

dopants <strong>in</strong> (111)A layers was slower than <strong>in</strong> (100), regardless of the presence of As<br />

vacancies. On the <strong>other</strong> h<strong>and</strong>, diffusion <strong>in</strong> (100) layers was enhanced by the presence of<br />

As vacancies. It was noted that As atoms on the (111)A surface did not evaporate easily,<br />

as compared with those on the (100) surface.<br />

A.Sh<strong>in</strong>oda, T.Yamamoto, M.Inai, T.Takebe, T.Watanabe: Japanese Journal of Applied<br />

Physics, 1993, 32[2-10A], L1374-6<br />

[446-115/116-116]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

First-pr<strong>in</strong>ciples calculations were presented for the vacancy-mediated diffusion of Si. It<br />

was shown that a DX-like mechanism facilitated the migration of lattice-site atoms <strong>in</strong>to<br />

the <strong>in</strong>terstitial region, <strong>and</strong> that the dangl<strong>in</strong>g bonds of a second-nearest neighbor vacancy<br />

assisted migration through the <strong>in</strong>terstitial region. Due to these 2 mechanisms, vacancyassisted<br />

diffusion of Si occurred with a low-energy barrier.<br />

J.Dabrowski, J.E.Northrup: Physical Review B, 1994, 49[20], 14286-9<br />

[446-115/116-117]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

A study was made of molecular beam epitaxially grown samples which were δ-doped<br />

with Si <strong>and</strong> Al layers. Long-term diffusion anneal<strong>in</strong>g was carried out at temperatures<br />

rang<strong>in</strong>g from 550 to 800C, <strong>and</strong> the concentration profiles were determ<strong>in</strong>ed by means of<br />

secondary ion mass spectrometry. It was found that the results could be described by the<br />

expression,<br />

D(cm 2 /s) = 7.9 exp[-2.25(eV)/kT]<br />

The Si diffusion coefficients which were obta<strong>in</strong>ed were <strong>in</strong> good agreement with data<br />

which had been obta<strong>in</strong>ed by us<strong>in</strong>g rapid thermal anneal<strong>in</strong>g, capacitance-voltage profil<strong>in</strong>g,<br />

<strong>and</strong> s<strong>and</strong>wiched diffusion sources. They differed from earlier measurements which had<br />

been based upon the diffusion of implanted dopants that were much more widely spread.<br />

It was concluded that the more accurate data which resulted from δ-dop<strong>in</strong>g showed that<br />

the diffusion coefficient was an <strong>in</strong>tr<strong>in</strong>sic parameter, provided that the amount of dopant<br />

<strong>and</strong> the dislocation density were kept sufficiently small.<br />

F.Bénière, R.Chapla<strong>in</strong>, M.Gauneau, V.Reddy, A.Régrény: Journal de Physique <strong>III</strong>, 1993,<br />

3[12], 2165-71<br />

[446-115/116-119]<br />

342


Si <strong>GaAs</strong> Si<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Various mechanisms of Si diffusion <strong>in</strong> <strong>GaAs</strong> were <strong>in</strong>vestigated by us<strong>in</strong>g first-pr<strong>in</strong>ciples<br />

molecular dynamics methods. It was found that the predom<strong>in</strong>ant mechanism <strong>in</strong>volved the<br />

motion of negatively charged Si <strong>III</strong> -V <strong>III</strong> pairs via second-nearest neighbor jumps. This<br />

mechanism expla<strong>in</strong>ed the ability of Si to disorder superlattices (regardless of whether it<br />

was <strong>in</strong>troduced dur<strong>in</strong>g growth or was <strong>in</strong>-diffused later), <strong>and</strong> the suppression of<br />

<strong>in</strong>terdiffusion by compensation dop<strong>in</strong>g. The calculated activation energies were <strong>in</strong> very<br />

good agreement with experimental data.<br />

B.Chen, Q.M.Zhang, J.Bernholc: Physical Review B, 1994, 49[4], 2985-8<br />

[446-115/116-119]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

It was recalled that implantation damage was believed to play a significant role <strong>in</strong> dopant<br />

diffusion. An attempt was made here to modify the po<strong>in</strong>t defect damage profile of 40keV<br />

29 Si implants by chemically etch<strong>in</strong>g away the top 10nm of the sample after implantation.<br />

No Si diffusion was observed <strong>in</strong> these samples after anneal<strong>in</strong>g, whereas significant Si<br />

redistribution occurred <strong>in</strong> a similar sample which had received no post-implantation<br />

etch<strong>in</strong>g. The results of TRIM simulations predicted an excess Ga vacancy surface layer,<br />

<strong>and</strong> the presence of excess Ga <strong>in</strong>terstitials deeper with<strong>in</strong> the sample. It was thought that,<br />

by remov<strong>in</strong>g the vacancy-rich surface layer, the etch altered the implant damage profile<br />

<strong>and</strong> thus the diffusion behavior of Si. The surface effects of etch<strong>in</strong>g which were related to<br />

Si diffusion were shown to be consistent with a vacancy-assisted diffusion mechanism.<br />

There was some evidence that this model might be applicable to B implants <strong>in</strong> Si.<br />

C.C.Lee, M.D.Deal, J.C.Bravman: Applied Physics Letters, 1994, 64[24], 3302-4<br />

[446-115/116-119]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

A study was made of the transport properties of the 2-dimensional electron gas <strong>in</strong><br />

annealed Si δ-doped structures. The diffusion rate of Si atoms was deduced from the<br />

temperature dependence of the sub-b<strong>and</strong> population. In samples with large selfcompensation,<br />

it was found that the electron density <strong>in</strong>creased after anneal<strong>in</strong>g at<br />

temperatures below 800C. After anneal<strong>in</strong>g at temperatures above 800C, there was a<br />

reduction <strong>in</strong> the electron concentration of the 2-dimensional electron gas. The results<br />

showed that, after anneal<strong>in</strong>g, the quantum mobility <strong>in</strong> the lowest sub-b<strong>and</strong> <strong>in</strong>creased<br />

slightly, while the quantum mobility <strong>in</strong> the higher sub-b<strong>and</strong>s markedly decreased.<br />

P.M.Koenraad, I.Bársony, A.F.W.Van der Stadt, J.A.A.J.Perenboom, J.H.Wolter:<br />

Materials Science Forum, 1994, 143-147, 663-8<br />

[446-113/114-012]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The diffusion of Si out of δ-planes <strong>in</strong> <strong>GaAs</strong> was <strong>in</strong>vestigated by means of high-resolution<br />

X-ray diffractometry, <strong>in</strong>fra-red absorption localized vibrational mode spectroscopy <strong>and</strong><br />

secondary ion mass spectroscopy. In the case of a Si δ-doped sample which had been<br />

343


Si <strong>GaAs</strong> Si<br />

grown at 400C with a Si areal concentration of 3.4 x 10 14 /cm 2 , the Si was conf<strong>in</strong>ed to a<br />

layer which was no more than 0.5nm <strong>in</strong> thickness. After post-growth anneal<strong>in</strong>g (600C,<br />

3h), 22.6% of the Si rema<strong>in</strong>ed on the δ-planes, while the rema<strong>in</strong>der had diffused<br />

homogeneously throughout the epilayer to give a Si concentration of 2.1 x 10 19 /cm 3 .<br />

Localized vibrational mode spectroscopy <strong>in</strong>dicated that these Si atoms were located<br />

ma<strong>in</strong>ly on Ga sites (Si Ga ). The Si atoms were also found to occupy As sites, or were<br />

present as Si Ga -Si As pairs, Si-X <strong>and</strong> Si Ga -V Ga complexes. At 950C, all of the Si had<br />

diffused away from the δ-planes to form precipitates <strong>and</strong> dislocation loops near to the<br />

surface.<br />

L.Hart, P.F.Fewster, M.J.Ashw<strong>in</strong>, M.R.Fahy, R.C.Newman: Materials Science Forum,<br />

1994, 143-147, 647-52<br />

[446-113/114-013]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

It was recalled that, <strong>in</strong> n-type material, the diffusion of atoms which resided on the Ga<br />

sub-lattice occurred ma<strong>in</strong>ly via Ga vacancies. In order to elucidate the microscopic details<br />

<strong>and</strong> energetics of these processes, local density approximation estimates were made of the<br />

total energies, electronic structures, <strong>and</strong> relaxed positions of atoms <strong>in</strong> microscopic<br />

configurations which were relevant to the diffusion of Si donors. It was found that a DXlike<br />

mechanism facilitated the migration of lattice site atoms <strong>in</strong>to the <strong>in</strong>terstitial region. A<br />

bond-pass<strong>in</strong>g mechanism was also identified, <strong>in</strong> which a second-nearest neighbor vacancy<br />

assisted migration through the <strong>in</strong>terstitial region. Due to these 2 mechanisms, vacancyassisted<br />

Si diffusion <strong>in</strong> n-type material was characterized by an energy barrier of about<br />

1.0eV.<br />

J.Dabrowski, J.Northrup: Materials Science Forum, 1994, 143-147, 1263-8<br />

[446-113/114-013]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The thermal equilibrium concentrations of the various negatively charged Ga vacancy<br />

species were calculated. The triply negatively charged Ga vacancy, V Ga 3- , was studied <strong>in</strong><br />

particular because it dom<strong>in</strong>ated Ga self-diffusion <strong>and</strong> Ga/Al <strong>in</strong>terdiffusion under <strong>in</strong>tr<strong>in</strong>sic<br />

<strong>and</strong> n-dop<strong>in</strong>g conditions, as well as the diffusion of Si donor atoms which occupied Ga<br />

sites.<br />

T.Y.Tan, H.M.You, U.M.Gösele: Applied Physics A, 1993, 56[3], 249-58<br />

[446-111/112-050]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Samples were doped with Si to a concentration of about 2.7 x 10 18 /cm 3 <strong>and</strong> were annealed<br />

(800 to 1000C, 3 to 20h) under As-rich or As-poor conditions. The Si out-diffusion was<br />

measured by us<strong>in</strong>g the capacitance-voltage method <strong>and</strong> an electrochemical profiler. It was<br />

found that the Si diffusivity exhibited a marked dependence upon the As 4 vapor phase<br />

pressure <strong>and</strong> upon the electron concentration. When reduced to <strong>in</strong>tr<strong>in</strong>sic conditions,<br />

activation enthalpies of 3.91 <strong>and</strong> 4.19eV were obta<strong>in</strong>ed for As-rich <strong>and</strong> As-poor anneal<strong>in</strong>g<br />

344


Si <strong>GaAs</strong> Si<br />

conditions, respectively. On the basis of these results, it was concluded that Si outdiffusion<br />

was governed by the triply negatively charged vacancies, V Ga 3- .<br />

H.M.You, U.M.Gösele, T.Y.Tan: Materials Science Forum, 1993, 117-118, 399-404<br />

[446-111/112-051]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Samples of n-type <strong>and</strong> p-type d-doped material, grown by molecular beam epitaxy <strong>and</strong><br />

with quite high doses of Si, were <strong>in</strong>vestigated by means of transmission electron<br />

microscopy. The magnitude of the doses ranged from half a monolayer to 2 monolayers.<br />

The microscopic structures of the d-doped regions <strong>and</strong> of the adjacent epilayers were<br />

observed directly. The effect of impurity spread<strong>in</strong>g upon the hetero-<strong>in</strong>terfaces <strong>and</strong><br />

superlattices was studied, <strong>and</strong> it was found that the Si atoms <strong>in</strong> Si d-doped samples were<br />

conf<strong>in</strong>ed to with<strong>in</strong> a few atomic layers. Stack<strong>in</strong>g faults were found <strong>in</strong> d-doped samples<br />

when they were grown at low temperatures. Their presence was attributed to local stra<strong>in</strong>s<br />

that were caused by heavy dop<strong>in</strong>g.<br />

D.G.Liu, J.C.Fan, C.P.Lee, K.H.Chang, D.C.Liou: Journal of Applied Physics, 1993,<br />

73[2], 608-14<br />

[446-106/107-034]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The lateral diffusion of sources dur<strong>in</strong>g the selective growth of metalorganic vapor-phase<br />

epitaxial Si-doped layers was analyzed. The diffusion lengths of Si species were deduced<br />

from the carrier concentration profiles which were measured by us<strong>in</strong>g Raman<br />

spectroscopy <strong>and</strong> thickness profil<strong>in</strong>g. On the basis of these diffusion lengths, it was<br />

speculated that the effective diffusion material was silyl ars<strong>in</strong>e. It was suggested that there<br />

was no difference between ars<strong>in</strong>e <strong>and</strong> tertiary butyl ars<strong>in</strong>e, as diffusion sources.<br />

N.Hara, K.Shi<strong>in</strong>a, T.Ohori, K.Kasai, J.Komeno: Journal of Applied Physics, 1993, 74[2],<br />

1327-30<br />

[446-106/107-036]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The formation energy of Si donors, acceptors, <strong>and</strong> defect complexes were calculated. The<br />

equilibrium concentrations of native defects <strong>and</strong> Si-defect complexes were deduced from<br />

these energies, as was the total solubility of Si. The calculated equilibrium solubility limit<br />

for Si was <strong>in</strong> good agreement with experimental data. The (Si Ga -V Ga ) 2- complex occurred<br />

at relatively high concentrations under As-rich conditions, <strong>and</strong> could therefore mediate Si<br />

<strong>and</strong> Ga diffusion. It was concluded that the donor-vacancy complex was an important<br />

compensation mechanism <strong>in</strong> heavily doped <strong>GaAs</strong>.<br />

J.E.Northrup, S.B.Zhang: Physical Review B, 1993, 47[11], 6791-4<br />

[446-106/107-036]<br />

345


Si <strong>GaAs</strong> Si<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

An experimental study was made of Si out-diffusion from <strong>GaAs</strong>, by us<strong>in</strong>g pre-doped<br />

samples. The results showed that the Si diffusivity depended upon the As 4 vapor-phase<br />

pressure <strong>in</strong> the ambient, <strong>and</strong> upon the electron concentration <strong>in</strong> the crystal. It was<br />

concluded that, <strong>in</strong> <strong>GaAs</strong>, diffusion of the Si donor species which occupied Ga sites, Si + Ga ,<br />

was governed by the triply negatively charged Ga vacancies, V 3- Ga . However, the present<br />

V 3- Ga -dom<strong>in</strong>ated Si + Ga out-diffusivities were larger, by many orders of magnitude, than<br />

those which were obta<strong>in</strong>ed under Si <strong>in</strong>-diffusion conditions. A tentative explanation of<br />

this large difference was given <strong>in</strong> terms of an undersaturation of V 3- Ga <strong>in</strong> <strong>in</strong>tr<strong>in</strong>sic material<br />

3-<br />

dur<strong>in</strong>g <strong>in</strong>-diffusion experiments, <strong>and</strong> of a supersaturation of V Ga which developed<br />

dur<strong>in</strong>g the out-diffusion of Si from n-type Si-doped material.<br />

H.M.You, U.M.Gösele, T.Y.Tan: Journal of Applied Physics, 1993, 73[11], 7207-16<br />

[446-106/107-037]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Heavily Si-doped (5 x 10 19 /cm 3 ) low-temperature <strong>GaAs</strong>, s<strong>and</strong>wiched between undoped<br />

low-temperature <strong>GaAs</strong> layers, was grown by us<strong>in</strong>g molecular beam epitaxy <strong>and</strong> was<br />

annealed at up to 900C. Transmission electron microscopy showed that, with<strong>in</strong> the first<br />

few m<strong>in</strong>utes of anneal<strong>in</strong>g, an accumulation of As precipitates formed near to each<br />

doped/undoped low-temperature <strong>in</strong>terface. Dur<strong>in</strong>g further anneal<strong>in</strong>g, Si segregation to As<br />

precipitates was detected, us<strong>in</strong>g secondary ion mass spectroscopy, <strong>in</strong> the form of deltalike<br />

peaks at the As precipitate accumulations. It was found that the Si diffusion<br />

coefficient was <strong>in</strong>itially <strong>in</strong>dependent of concentration, at a value of 2.5 x 10 -13 cm 2 /s, <strong>and</strong><br />

was comparable to diffusion under <strong>in</strong>tr<strong>in</strong>sic conditions <strong>in</strong> As-rich material when grown at<br />

normal temperatures. Dur<strong>in</strong>g anneal<strong>in</strong>g for 1h, the Si concentration <strong>in</strong> the As precipitates<br />

reached 2.5 x 10 20 /cm 3 .<br />

K.L.Kavanagh, J.C.P.Chang, P.D.Kirchner, A.C.Warren, J.M.Woodall: Applied Physics<br />

Letters, 1993, 62[3], 286-8<br />

[446-106/107-038]<br />

<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The diffusion of Si from a novel diffusion source, consist<strong>in</strong>g of an undoped SiO x /SiN<br />

double-layered film, was studied by rapid thermal anneal<strong>in</strong>g at 860 to 940C. The<br />

characteristics of the Si-diffused layers were <strong>in</strong>vestigated by us<strong>in</strong>g secondary ion mass<br />

spectrometric, capacitance-voltage, <strong>and</strong> Hall methods. The carrier profiles exceeded 2 x<br />

10 18 /cm 3 , <strong>and</strong> featured an abrupt diffusion front. A maximum electron concentration of 6<br />

x 10 18 /cm 3 was obta<strong>in</strong>ed at 940C. The diffused Si profiles were consistent with the<br />

operation of Si Ga + -V Ga - pair diffusion.<br />

S.Matsushita, S.Terada, E.Fujii, Y.Harada: Applied Physics Letters, 1993, 63[2], 225-7<br />

[446-106/107-038]<br />

346


Si <strong>GaAs</strong> Si<br />

<strong>GaAs</strong>/AlAs: Si <strong>Diffusion</strong><br />

Various mechanisms of Si-<strong>in</strong>duced <strong>in</strong>terdiffusion <strong>in</strong> <strong>GaAs</strong>/AlAs superlattices were<br />

<strong>in</strong>vestigated by us<strong>in</strong>g first-pr<strong>in</strong>ciples molecular dynamics methods. It was found that the<br />

predom<strong>in</strong>ant mechanism <strong>in</strong>volved the motion of negatively charged Si <strong>III</strong> -V <strong>III</strong> pairs via<br />

second-nearest neighbor jumps. This mechanism expla<strong>in</strong>ed the ability of Si to disorder<br />

superlattices (regardless of whether it was <strong>in</strong>troduced dur<strong>in</strong>g growth or was <strong>in</strong>-diffused<br />

later), <strong>and</strong> the suppression of <strong>in</strong>terdiffusion by compensation dop<strong>in</strong>g. The calculated<br />

activation energies were <strong>in</strong> very good agreement with experimental data.<br />

B.Chen, Q.M.Zhang, J.Bernholc: Physical Review B, 1994, 49[4], 2985-8<br />

[446-115/116-119]<br />

<strong>GaAs</strong>/AlAs: Si <strong>Diffusion</strong><br />

Hall <strong>and</strong> photo-Hall measurements were performed on <strong>GaAs</strong>/AlAs short-period<br />

superlattices which were selectively doped with Si. The dopant was <strong>in</strong>troduced<br />

selectively <strong>in</strong>to the <strong>GaAs</strong> or AlAs layers, or at the <strong>in</strong>terface. A superlattice which was<br />

doped uniformly <strong>in</strong> both layers was <strong>in</strong>vestigated for comparison. It was found that the<br />

electrical properties were controlled by DX deep donors which lay <strong>in</strong> the gap of the<br />

superlattice. The Hall data were expla<strong>in</strong>ed <strong>in</strong> terms of a model which took account of the<br />

existence of two DX deep donors <strong>and</strong> a shallow donor which were both related to the Si<br />

impurity. It was found that the Si donor state <strong>in</strong> AlAs lay 0.06eV below the Si donor state<br />

<strong>in</strong> <strong>GaAs</strong>. The ionization energies of the DX states <strong>in</strong> <strong>GaAs</strong> <strong>and</strong> AlAs were calculated <strong>in</strong><br />

order to account for the experimental results. The <strong>in</strong>terpretation of Hall data <strong>in</strong> selectively<br />

doped samples required the assumption of Si segregation dur<strong>in</strong>g epitaxy.<br />

P.Sellitto, P.Jeanjean, J.Sicart, J.L.Robert, R.Planel: Journal of Applied Physics, 1993,<br />

74[12], 7166-72<br />

[446-109/110-033]<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

Laser-assisted disorder<strong>in</strong>g was studied by us<strong>in</strong>g scann<strong>in</strong>g electron microscopy <strong>and</strong><br />

secondary ion mass spectrometry. This permitted the extent of the layer-disordered region<br />

to be correlated with the presence of laser-<strong>in</strong>corporated Si <strong>and</strong> O. Transmission electron<br />

microscopic studies permitted the determ<strong>in</strong>ation of the distribution of Al <strong>and</strong> Ga at the<br />

<strong>in</strong>terface between the impurity-disordered alloy <strong>and</strong> the as-grown crystal. The data<br />

revealed the occurrence of more rapid Si diffusion <strong>in</strong> the <strong>GaAs</strong> layers as compared with<br />

the Al-rich layers.<br />

J.E.Epler, F.A.Ponce, F.J.Endicott, T.L.Paoli: Journal of Applied Physics, 1988, 64[7],<br />

3439-44<br />

[446-72/73-026]<br />

<strong>GaAs</strong>/Si: Si <strong>Diffusion</strong><br />

Samples of <strong>GaAs</strong>, which were encapsulated with th<strong>in</strong> films of amorphous Si at 450C,<br />

were annealed at temperatures of up to 1050C. The resultant poly-Si/<strong>GaAs</strong> <strong>in</strong>terfaces<br />

were <strong>in</strong>vestigated by us<strong>in</strong>g secondary ion mass spectroscopy, Rutherford back-scatter<strong>in</strong>g<br />

347


Si <strong>GaAs</strong> Sn<br />

spectrometry, <strong>and</strong> transmission electron microscopy. Little or no <strong>in</strong>terdiffusion was<br />

detected at undoped Si/<strong>GaAs</strong> <strong>in</strong>terfaces, whereas Si diffused <strong>in</strong>to the <strong>GaAs</strong> (from P-doped<br />

or As-doped Si) to depths as great as 550nm after only 10s of anneal<strong>in</strong>g at 1050C. The<br />

flux of Si <strong>in</strong>to <strong>GaAs</strong> was related to the fluxes of Ga <strong>and</strong> As <strong>in</strong>to Si. Both fluxes <strong>in</strong>creased<br />

with <strong>in</strong>creas<strong>in</strong>g dopant concentration <strong>in</strong> the Si. An enhanced diffusivity of Si <strong>in</strong>to <strong>GaAs</strong><br />

was attributed to the diffusion of po<strong>in</strong>t defects which were created by the diffusion of As<br />

<strong>and</strong> Ga <strong>in</strong>to the encapsulant. It was deduced that the Si diffusivities <strong>in</strong> <strong>GaAs</strong> at doped<br />

polycrystall<strong>in</strong>e Si <strong>in</strong>terfaces were enhanced by factors of about 10000.<br />

K.L.Kavanagh, C.W.Magee, J.Sheets, J.W.Meyer: Journal of Applied Physics, 1988,<br />

64[4], 1845-54<br />

[446-72/73-027]<br />

<strong>GaAs</strong>/Si: Si <strong>Diffusion</strong><br />

A new mechanism was proposed for the <strong>in</strong>corporation of Si <strong>in</strong>to <strong>GaAs</strong>/Si hetero-epitaxial<br />

layers which were grown by metalorganic chemical vapor deposition. The mechanism<br />

<strong>in</strong>volved gas-phase transport of Si to hetero-epitaxial layers dur<strong>in</strong>g growth. This mode of<br />

Si uptake could operate as well as the previously proposed mechanism. The latter<br />

<strong>in</strong>volved <strong>in</strong>corporation by enhanced diffusion, from the hetero-<strong>in</strong>terface, via defects <strong>in</strong> the<br />

<strong>GaAs</strong> layer.<br />

S.Nozaki, J.J.Murray, A.T.Wu, T.George, E.R.Weber, M.Umeno: Applied Physics<br />

Letters, 1989, 55[16], 1674-6<br />

[446-72/73-027]<br />

<strong>GaAs</strong>/Si: Si Pipe <strong>Diffusion</strong><br />

Preferential diffusion channels for Si were found <strong>in</strong> <strong>GaAs</strong> which had been grown, by<br />

means of metal-organic vapor phase epitaxy, onto Si(100). The density of these diffusion<br />

channels was consistent with the measured dislocation density. Also, by comb<strong>in</strong><strong>in</strong>g<br />

scann<strong>in</strong>g electron microscopy <strong>and</strong> X-ray fluorescence, it was shown that a large amount<br />

of Si emerged at the surface with<strong>in</strong> small [011]-type overgrowth-oriented defects that<br />

were present at the surface.<br />

A.Freundlich, A.Leycuras, J.C.Grenet, C.Grattepa<strong>in</strong>: Applied Physics Letters, 1988,<br />

53[26], 2635-7<br />

[446-64/65-168]<br />

Sn<br />

<strong>GaAs</strong>: Sn <strong>Diffusion</strong><br />

A study was made of the contributions of segregation, diffusion <strong>and</strong> aggregation to the<br />

broaden<strong>in</strong>g of d-doped planes of Sn. It was found that the Sn planes were severely<br />

broadened by all 3 processes. At higher temperatures or densities, segregation or<br />

348


Sn <strong>GaAs</strong> Sn<br />

concentration-dependent rapid diffusion could occur; thus caus<strong>in</strong>g significant spread<strong>in</strong>g<br />

even dur<strong>in</strong>g growth.<br />

J.J.Harris, J.B.Clegg, R.B.Beall, J.Castagné, K.Woodbridge, C.Roberts: Journal of<br />

Crystal Growth, 1991, 111[1-4], 239-45<br />

[446-91/92-001]<br />

Table 26<br />

Diffusivity of Implanted Sn <strong>in</strong> <strong>GaAs</strong><br />

Dose (/cm 2 ) Temperature (C) D (cm 2 /s)<br />

1 x 10 14 1000 1.7 x 10 -13<br />

1 x 10 14 1000 1.1 x 10 -13<br />

1 x 10 15 1000 7.0 x 10 -14<br />

1 x 10 13 1000 3.0 x 10 -14<br />

1 x 10 14 1000 2.3 x 10 -14<br />

1 x 10 15 900 6.0 x 10 -14<br />

1 x 10 13 900 2.1 x 10 -14<br />

1 x 10 14 900 5.2 x 10 -15<br />

1 x 10 15 900 3.8 x 10 -15<br />

1 x 10 14 870 1.3 x 10 -15<br />

1 x 10 15 900 8.2 x 10 -16<br />

1 x 10 14 870 8.4 x 10 -16<br />

1 x 10 13 700 7.8 x 10 -19<br />

1 x 10 15 700 4.8 x 10 -19<br />

5 x 10 15 750 9.8 x 10 -18<br />

1 x 10 15 800 5.2 x 10 -17<br />

5 x 10 15 850 1.5 x 10 -16<br />

1 x 10 15 850 2.1 x 10 -16<br />

5 x 10 15 950 9.8 x 10 -16<br />

5 x 10 14 950 5.1 x 10 -15<br />

1 x 10 14 1000 1.5 x 10 -14<br />

326 <strong>GaAs</strong>: Sn <strong>Diffusion</strong><br />

This dopant diffused extensively after implantation <strong>and</strong> long-term anneal<strong>in</strong>g. The results<br />

could be expla<strong>in</strong>ed by assum<strong>in</strong>g that the diffusivity depended upon the square of the<br />

electron concentration. The dopant diffusion was affected by the presence of implantation<br />

damage; the higher the concentration of extended defects, the slower was the diffusivity<br />

as compared with the values for conventional diffusion from a solid source. If the sample<br />

was amorphized dur<strong>in</strong>g implantation, extended defects did not form <strong>and</strong> the diffusivity of<br />

the ion was very close to that <strong>in</strong> material which had been diffused from a solid source.<br />

When amorphization did not occur, extended defects formed after implantation, <strong>and</strong><br />

349


Sn <strong>GaAs</strong> Sn<br />

diffusion was <strong>in</strong>hibited; especially after low doses, <strong>in</strong> the short term, or at low<br />

temperatures. The higher the density of extended defects, the greater was the suppression<br />

of diffusion. No time-dependence was observed. It was concluded that the results (table<br />

26) were consistent with a diffusion mechanism <strong>in</strong> which the mobile species was the<br />

donor that was coupled with a charged Ga vacancy. The equilibrium vacancy<br />

concentration was thought to be suppressed by the presence of extended defects <strong>and</strong>/or<br />

excess Ga <strong>in</strong>terstitials which resulted from implantation.<br />

E.L.Allen, J.J.Murray, M.D.Deal, J.D.Plummer, K.S.Jones, W.S.Rubart: Journal of the<br />

Electrochemical Society, 1991, 138[11], 3440-9<br />

[446-84/85-016]<br />

<strong>GaAs</strong>: Sn <strong>Diffusion</strong><br />

The migration of implanted Sn was studied, as a function of dose <strong>and</strong> background dop<strong>in</strong>g,<br />

by means of secondary ion mass spectrometry. It was concluded that the diffusivity of the<br />

Sn depended upon the electron concentration, <strong>and</strong> that the Sn diffused via negatively<br />

charged Ga vacancies. The diffusivity of Sn outside of the implanted region was<br />

<strong>in</strong>dependent of the dose, <strong>and</strong> was associated with an activation energy of 1.98eV.<br />

E.L.Allen, M.D.Deal, J.D.Plummer: Journal of Applied Physics, 1990, 67[7], 3311-4<br />

[446-78/79-014]<br />

<strong>GaAs</strong>: Sn <strong>Diffusion</strong><br />

A scanned electron beam was used to diffuse Sn <strong>in</strong>to <strong>GaAs</strong> from doped emulsions, <strong>and</strong><br />

Rutherford back-scatter<strong>in</strong>g was used to <strong>in</strong>vestigate the results. It was found that diffusion<br />

was greatly enhanced by capp<strong>in</strong>g the emulsion with evaporated SiO 2 .<br />

Z.Meglicki, D.D.Cohen, A.G.Nassibian: Journal of Applied Physics, 1987, 62[5], 1778-<br />

81<br />

[446-55/56-006]<br />

<strong>GaAs</strong>: Sn <strong>Diffusion</strong><br />

The effects of stra<strong>in</strong> upon the diffusion of Sn were studied by us<strong>in</strong>g laser Raman <strong>and</strong><br />

photolum<strong>in</strong>escence spectroscopy. It was found that an <strong>in</strong>crease <strong>in</strong> compressive stra<strong>in</strong><br />

produced an <strong>in</strong>crease <strong>in</strong> the carrier concentration, while a decrease <strong>in</strong> compressive stra<strong>in</strong><br />

or an <strong>in</strong>crease <strong>in</strong> tensile stra<strong>in</strong> led to a decrease <strong>in</strong> the carrier concentration at the surface.<br />

This behavior was attributed to a decrease <strong>in</strong> the diffusion coefficient of Sn with<br />

compressive stra<strong>in</strong>, <strong>and</strong> an <strong>in</strong>crease with tensile stra<strong>in</strong>. The data showed that the peak<br />

which was due to Ga antisite defects <strong>in</strong>creased with <strong>in</strong>creas<strong>in</strong>g compressive stress. This<br />

<strong>in</strong>dicated a decrease, <strong>in</strong> the Ga vacancy concentration, from the equilibrium concentration<br />

<strong>in</strong> an unstressed sample. Photolum<strong>in</strong>escence data for tensile-stressed samples revealed an<br />

<strong>in</strong>crease <strong>in</strong> Ga vacancy concentration with respect to the equilibrium concentration <strong>in</strong> an<br />

unstressed sample. It was concluded that the change <strong>in</strong> diffusion coefficient with stra<strong>in</strong><br />

was related to Ga vacancies. It was found that the diffusion coefficient decreased<br />

350


Sn <strong>GaAs</strong> Ti<br />

exponentially with the compressive stra<strong>in</strong>, <strong>and</strong> <strong>in</strong>creased exponentially with tensile stra<strong>in</strong>.<br />

The activation energy for Sn diffusion therefore varied l<strong>in</strong>early as a function of stra<strong>in</strong>.<br />

A.B.M.Harun-ur Rashid, T.Katoda: Journal of Applied Physics, 1997, 81[4], 1661-9<br />

[446-148/149-173]<br />

<strong>GaAs</strong>: Sn <strong>Diffusion</strong><br />

The stresses which were generated at the <strong>GaAs</strong>/SiO 2 <strong>in</strong>terface dur<strong>in</strong>g anneal<strong>in</strong>g were<br />

<strong>in</strong>vestigated by means of laser Raman spectroscopy. It was found that compressive<br />

stresses existed at the surface of the <strong>GaAs</strong> after anneal<strong>in</strong>g, <strong>and</strong> that these <strong>in</strong>creased with<br />

<strong>in</strong>creas<strong>in</strong>g SiO 2 cap layer thickness. The compressive stress on the <strong>GaAs</strong> was generated<br />

dur<strong>in</strong>g anneal<strong>in</strong>g <strong>and</strong> was attributed to the difference <strong>in</strong> the thermal expansion<br />

coefficients of <strong>GaAs</strong> <strong>and</strong> SiO 2 . The <strong>in</strong>crease <strong>in</strong> compressive stress on the surface of <strong>GaAs</strong><br />

decreased the diffusion coefficient of Sn <strong>in</strong> the <strong>GaAs</strong>. This was due to a reduction <strong>in</strong> the<br />

number of Ga vacancies <strong>in</strong> the compressively stressed sample, as compared with the<br />

equilibrium Ga vacancy content of an unstressed sample.<br />

A.B.M.Harun-ur Rashid, M.Kishi, T.Katoda: Journal of Applied Physics, 1996, 80[6],<br />

3540-5<br />

[446-138/139-078]<br />

<strong>GaAs</strong>: Sn <strong>Diffusion</strong><br />

Films of SiO 2 spun-on glass, that were doped with Sn <strong>and</strong>/or Ga, were used as diffusion<br />

sources. The diffusion was studied dur<strong>in</strong>g rapid thermal anneal<strong>in</strong>g, with or without any<br />

As over-pressure. It was found that the diffusivity of Sn decreased as the As over-pressure<br />

was <strong>in</strong>creased. Modify<strong>in</strong>g the Sn-doped spun-on glass, so as to conta<strong>in</strong> 4mol%Ga,<br />

slightly reduced the Sn diffusivity. These results were expla<strong>in</strong>ed <strong>in</strong> terms of chemical<br />

reactions between the spun-on glass <strong>and</strong> the <strong>GaAs</strong>. Highly doped layers (10 18 to 3 x<br />

10 18 /cm 3 ) were obta<strong>in</strong>ed.<br />

C.S.Hern<strong>and</strong>es, J.W.Swart, M.A.A.Pudenzi, G.T.Kraus, Y.Shacham-Diam<strong>and</strong>,<br />

E.P.Giannelis: Journal of the Electrochemical Society, 1995, 142[8], 2829-32<br />

[446-134/135-125]<br />

Ti<br />

<strong>GaAs</strong>: Ti <strong>Diffusion</strong><br />

High-resolution X-ray photo-emission spectroscopy <strong>and</strong> Ar ion bombardment were used<br />

to study temperature-dependent chemical reactions <strong>and</strong> species redistribution <strong>in</strong> the<br />

Ti/<strong>GaAs</strong>(100) system. The results showed that Ti, deposited at room temperature,<br />

disrupted the <strong>GaAs</strong> substrate (by react<strong>in</strong>g with As) <strong>and</strong> released Ga <strong>in</strong>to the over-layer.<br />

The As was found to accumulate, near to the buried <strong>in</strong>terface, <strong>in</strong> the form of a Ti-As<br />

compound. The Ga was depleted, but accumulated beyond the reaction region. Sputter<br />

depth profiles <strong>in</strong>dicated that high-temperature anneal<strong>in</strong>g caused Ti diffusion <strong>in</strong>to the<br />

<strong>GaAs</strong> substrate <strong>and</strong> enhanced reaction with As. The rejection of Ga from the form<strong>in</strong>g Ti-<br />

As compound became more severe when the amount of Ti-As <strong>in</strong>creased. Heat<strong>in</strong>g<br />

351


Ti <strong>GaAs</strong> Ti<br />

promoted the segregation of rejected Ga atoms to the vacuum surface, but had little effect<br />

upon As segregation.<br />

F.Xu, D.M.Hill, Z.L<strong>in</strong>, S.G.Anderson, Y.Shapira, J.H.Weaver: Physical Review B, 1988,<br />

37[17], 10295-300<br />

[446-62/63-211]<br />

1.0E-10<br />

1.0E-11<br />

1.0E-12<br />

1.0E-13<br />

1.0E-14<br />

D (cm 2 /s)<br />

1.0E-15<br />

1.0E-16<br />

1.0E-17<br />

1.0E-18<br />

1.0E-19<br />

table 27<br />

table 28<br />

table 29<br />

table 30<br />

table 31<br />

1.0E-20<br />

7 8 9 10 11 12 13<br />

10 4 /T(K)<br />

Figure 7: Diffusivity of Zn <strong>in</strong> <strong>GaAs</strong><br />

<strong>GaAs</strong>: Ti <strong>Diffusion</strong><br />

The structural properties of samples which had been implanted with 150 or 400keV Ti, to<br />

doses of between 10 12 <strong>and</strong> 10 15 /cm 2 , were studied. The depth distributions of the implants<br />

were compared before <strong>and</strong> after anneal<strong>in</strong>g with, or without, a Si 3 N 4 cap. Rutherford backscatter<strong>in</strong>g,<br />

X-ray double-crystal diffractometry, <strong>and</strong> secondary ion mass spectroscopy<br />

results <strong>in</strong>dicated that the Ti did not redistribute at all.<br />

H.Ullrich, A.Knecht, D.Bimberg, H.Kräutle, W.Schlaak: Journal of Applied Physics,<br />

1992, 72[8], 3514-21<br />

[446-106/107-034]<br />

352


Zn <strong>GaAs</strong> Zn<br />

Zn<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The migration of th<strong>in</strong> highly p-doped layers <strong>in</strong> s<strong>in</strong>gle <strong>and</strong> double heterostructures, grown<br />

us<strong>in</strong>g metalorganic vapor-phase epitaxy, was studied us<strong>in</strong>g capacitance-voltage etch<br />

profil<strong>in</strong>g <strong>and</strong> secondary ion mass spectrometry. It was deduced that the diffusivity of Zn<br />

<strong>in</strong> <strong>GaAs</strong> could be described by:<br />

D (cm 2 /s) = 4.6 x 10 -4 exp[-2.1(eV)/kT]<br />

for rapid thermal anneal<strong>in</strong>g, while the diffusivity could be described by:<br />

D (cm 2 /s) = 1.2 x 10 -6 exp[-1.8(eV)/kT]<br />

for furnace anneal<strong>in</strong>g. A model which was based upon an <strong>in</strong>terstitial cum substitutional<br />

diffusion mechanism, with certa<strong>in</strong> k<strong>in</strong>etic limitations, was successfully used to simulate<br />

the observed dopant concentration profiles. Markedly anomalous diffusion of Zn, from<br />

<strong>GaAs</strong> <strong>and</strong> <strong>in</strong>to highly n-doped GaAlAs, was found.<br />

N.Nordell, P.Ojala, W.H.Van Berlo, G.L<strong>and</strong>gren, M.K.L<strong>in</strong>narsson: Journal of Applied<br />

Physics, 1990, 67[2], 778-86<br />

[446-74-004]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

It was noted that the formation of dislocations <strong>and</strong> precipitates <strong>in</strong> monocrystall<strong>in</strong>e<br />

substrates occurred when elemental gas sources with a sufficiently high Zn partial<br />

pressure were used to diffuse Zn <strong>in</strong>to the surface regions of the substrate. A study of the<br />

nature of these defects at various depths <strong>in</strong> the Zn concentration profiles permitted<br />

conclusions to be drawn concern<strong>in</strong>g defect formation <strong>and</strong> evolution dur<strong>in</strong>g diffusion, <strong>and</strong><br />

revealed valuable <strong>in</strong>formation concern<strong>in</strong>g the po<strong>in</strong>t defects which were <strong>in</strong>volved <strong>in</strong> Zn<br />

diffusion.<br />

A.Rucki, W.Jäger: Defect <strong>and</strong> <strong>Diffusion</strong> Forum, 1997, 143-147, 1095-100<br />

[446-143/147-1095]<br />

327 <strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

It was recalled that Zn <strong>in</strong> <strong>GaAs</strong> exhibited a complicated diffusion behavior, as well as a<br />

high solid-solubility. The diffusion of Zn, at temperatures of between 700 <strong>and</strong> 1100C,<br />

was studied here by us<strong>in</strong>g 3 different Zn diffusion sources. In order to compare the<br />

penetration curves for the various sources, reduced penetration depths <strong>and</strong> reduced<br />

concentrations were calculated. Numerical simulation of Zn transport furnished a good<br />

description for particular sources. Effective diffusivities (table 27) were deduced from the<br />

experimental data <strong>and</strong> were normalized to st<strong>and</strong>ard vapor pressures <strong>and</strong> electronically<br />

<strong>in</strong>tr<strong>in</strong>sic conditions. The data for normalized Zn diffusion could be described by:<br />

D (cm 2 /s) = 82.3 exp[-4.03(eV)/kT]<br />

H.G.Hettwer, N.A.Stolwijk, H.Mehrer: Defect <strong>and</strong> <strong>Diffusion</strong> Forum, 1997, 143-147,<br />

1117-24<br />

[446-143/147-1117]<br />

353


Zn <strong>GaAs</strong> Zn<br />

Table 27<br />

Diffusivity of Zn <strong>in</strong> <strong>GaAs</strong> as a Function of <strong>Diffusion</strong>-Source Composition<br />

Temperature (C) As Zn Ga D (cm 2 /s)<br />

700 0.054 0.698 0.248 1.08 x 10 -19<br />

900 0.547 0.414 0.039 2.52 x 10 -16<br />

900 0.535 0.429 0.036 1.88 x 10 -16<br />

900 0.246 0.508 0.246 1.20 x 10 -15<br />

900 0.075 0.130 0.795 1.66 x 10 -15<br />

1050 0.542 0.323 0.135 2.24 x 10 -14<br />

1050 0.474 0.366 0.160 2.94 x 10 -14<br />

1050 0.330 0.333 0.337 5.26 x 10 -14<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

It was po<strong>in</strong>ted out that Zn was one of the ma<strong>in</strong> p-type dopants which were used for the<br />

fabrication of devices that were based upon <strong>GaAs</strong> or related <strong>III</strong>-V materials. The element<br />

dissolved substitutionally on the group-<strong>III</strong> sub-lattice, <strong>and</strong> diffused via a kick-out<br />

mechanism which <strong>in</strong>volved group-<strong>III</strong> self-<strong>in</strong>terstitials. Non-equilibrium concentrations of<br />

these self-<strong>in</strong>terstitials had a marked effect upon the diffusivity of Zn. Various situations<br />

were considered <strong>in</strong> which non-equilibrium po<strong>in</strong>t defects played a role <strong>in</strong> Zn diffusion.<br />

These <strong>in</strong>cluded the <strong>in</strong>-diffusion of such dopants from an external source, the diffusion of<br />

grown-<strong>in</strong> dopants, <strong>and</strong> self-<strong>in</strong>terstitial generation by Fermi-level surface p<strong>in</strong>n<strong>in</strong>g. It was<br />

noted that the diffusion behavior of C, which was found on the group-V sub-lattice of<br />

<strong>GaAs</strong>, was much less sensitive to non-equilibrium po<strong>in</strong>t defects. It was therefore used to<br />

replace Zn as a p-type dopant.<br />

M.Uematsu, K.Wada, U.Gösele: Applied Physics A, 1992, 55[4], 301-12<br />

[446-93/94-008]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The chemical processes which occurred dur<strong>in</strong>g the diffusion of Zn from spun-on SiO 2<br />

films <strong>and</strong> <strong>in</strong>to wafer samples were <strong>in</strong>vestigated. It was found that, at 600C, only about<br />

45% of the film was changed to SiO 2 glass. At 700C, Zn silicates began to form. When<br />

the Zn was <strong>in</strong> this form, it was much less able to diffuse <strong>in</strong>to the <strong>GaAs</strong>.<br />

E.Nowak, G.Kühn, B.Schumann, R.Hesse, H.Sobotta: Crystal Research <strong>and</strong> Technology,<br />

1992, 27[4], 503-8<br />

[446-91/92-008]<br />

354


Zn <strong>GaAs</strong> Zn<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The rapid thermal diffusion of Zn <strong>in</strong>to semi-<strong>in</strong>sulat<strong>in</strong>g material, from spun-on silica films,<br />

was <strong>in</strong>vestigated. Depend<strong>in</strong>g upon the heat<strong>in</strong>g rate, 2 types of secondary ion mass<br />

spectrometry profile were observed.<br />

E.Nowak, G.Kühn, T.Morgenstern, B.Schumann: Crystal Research <strong>and</strong> Technology,<br />

1991, 26[8], 981-6<br />

[446-88/89-014]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The migration of Zn <strong>in</strong>to material which had been grown by means of molecular beam<br />

epitaxy at 200C was studied. The diffusion coefficient was found, us<strong>in</strong>g the sealed<br />

ampoule technique, to be an order of magnitude higher <strong>in</strong> so-called low-temperature<br />

material (1.3 x 10 -11 cm 2 /s) than it was <strong>in</strong> normal material (1.5 x 10 -12 cm 2 /s). This<br />

difference was attributed to the large number of defects (<strong>in</strong>clud<strong>in</strong>g As antisites) which<br />

were present <strong>in</strong> the low-temperature material. It was noted that the effectiveness of a<br />

Si 3 N 4 diffusion mask depended upon whether the mask was deposited directly onto the<br />

low-temperature material. A failure of such masks to stop Zn diffusion was attributed to<br />

the effect of As atoms which out-diffused from the As-rich low-temperature material <strong>and</strong><br />

<strong>in</strong>to the Si 3 N 4 . The presence of a 10nm <strong>GaAs</strong> layer on the low-temperature material was<br />

effective <strong>in</strong> conserv<strong>in</strong>g the mask<strong>in</strong>g properties of the nitride.<br />

Y.K.S<strong>in</strong>, Y.Hwang, T.Zhang, R.M.Kolbas: Journal of Electronic Materials, 1991, 20[6],<br />

465-70<br />

[446-88/89-015]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The diffusion of Zn was studied by us<strong>in</strong>g liquid-phase epitaxy methods, <strong>and</strong> Si-doped n-<br />

type substrate material. The measurements were carried out at 850C, <strong>and</strong> dopant<br />

concentrations which ranged from 10 18 to 10 19 /cm 3 were <strong>in</strong>troduced. It was found that the<br />

Zn concentration <strong>in</strong> the solid depended upon the square root of the atomic fraction of Zn<br />

<strong>in</strong> the liquid. The diffusivity was dom<strong>in</strong>ated by the <strong>in</strong>terstitial-substitutional process, <strong>and</strong><br />

exhibited a cubic dependence upon the Zn content. The Zn <strong>in</strong>terstitial was ma<strong>in</strong>ly doublyionized<br />

Zn i 2+ .<br />

C.Algora, G.L.Araujo, A.Marti: Journal of Applied Physics, 1990, 68[6], 2723-30<br />

[446-86/87-002]<br />

328 <strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The migration of Zn <strong>in</strong> Al<strong>GaAs</strong> at 650C was studied by us<strong>in</strong>g the sealed-ampoule method<br />

<strong>and</strong> a ZnAs 2 source. It was found that the results for zero Al content (table 28) could be<br />

described by the expression:<br />

D (cm 2 /s) = 26 exp[-2.47(eV)/kT]<br />

V.Qu<strong>in</strong>tana, J.J.Clemencon, A.K.Ch<strong>in</strong>: Journal of Applied Physics, 1988, 63[7], 2454-5<br />

[446-72/73-003]<br />

355


Zn <strong>GaAs</strong> Zn<br />

Table 28<br />

Bulk Diffusivity of Zn <strong>in</strong> <strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

700 4.08 x 10 -12<br />

650 1.00 x 10 -12<br />

600 1.84 x 10 -13<br />

550 1.83 x 10 -14<br />

329 <strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

In order to study its diffusion mechanism, Zn was diffused from a ZnAs 2 source <strong>in</strong>to Sidoped<br />

samples, at temperatures rang<strong>in</strong>g from 575 to 700C, <strong>in</strong> sealed evacuated quartz<br />

tubes. The samples were characterized by means of the depth profile of the<br />

photolum<strong>in</strong>escence at various temperatures. The photolum<strong>in</strong>escence spectra conta<strong>in</strong>ed<br />

characteristic emissions which were associated with deep levels of Ga <strong>and</strong> As vacancies.<br />

A detailed analysis of the spectra revealed the role that was played by vacancies <strong>in</strong> the Zn<br />

diffusion process. A spatial correlation between the lum<strong>in</strong>escence spectra, <strong>and</strong> the Zn<br />

concentrations deduced from secondary ion mass spectrometric data, was demonstrated. It<br />

was found that the data (table 29) could be described by the expression:<br />

D(cm 2 /s) = 2.05 exp[-2.28(eV)/kT]<br />

N.H.Ky, L.Pavesi, D.Araujo, J.D.Ganière, F.K.Re<strong>in</strong>hart: Journal of Applied Physics,<br />

1991, 69[11], 7585-93<br />

[446-86/87-012]<br />

Table 29<br />

Diffusivity of Zn <strong>in</strong> <strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

700 3.33 x 10 -12<br />

650 7.12 x 10 -13<br />

600 1.72 x 10 -13<br />

575 5.43 x 10 -14<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

A marked difference between the Be <strong>and</strong> C distributions <strong>in</strong> this material was noted after<br />

Zn diffusion. It was shown that grown-<strong>in</strong> C was stable, <strong>and</strong> rema<strong>in</strong>ed localized even after<br />

Zn diffusion. On the <strong>other</strong> h<strong>and</strong>, Be diffused very rapidly <strong>in</strong> the presence of diffus<strong>in</strong>g Zn.<br />

356


Zn <strong>GaAs</strong> Zn<br />

This effect was attributed to differences <strong>in</strong> the crystal lattice sites which the dopant<br />

occupied.<br />

E.Tokumitsu, T.H.Chiu, H.S.Luftman, N.T.Ha: Journal of Applied Physics, 1991, 69[12],<br />

8426-8<br />

[446-86/87-013]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The Zn diffusion dop<strong>in</strong>g of <strong>GaAs</strong>, by us<strong>in</strong>g metalorganic vapor-phase epitaxy <strong>and</strong><br />

diethylz<strong>in</strong>c as a dopant source, was exam<strong>in</strong>ed. Typical Zn concentrations <strong>and</strong> depths<br />

which were obta<strong>in</strong>ed were 10 19 to 10 21 /cm 3 , <strong>and</strong> 40 to 200nm. The highest concentration<br />

gradient which was obta<strong>in</strong>ed <strong>in</strong> this way was 4 orders of magnitude per 50nm, <strong>and</strong> the<br />

highest Zn concentration was 2 x 10 21 /cm 3 at the sample surface.<br />

Z.F.Paska, D.Haga, B.Willén, M.K.L<strong>in</strong>narsson: Applied Physics Letters, 1992, 60[13],<br />

1594-6<br />

[446-86/87-013]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

A simple method for the open-tube diffusion of Zn from (ZnO) x (SiO 2 ) 1-x film sources,<br />

<strong>and</strong> <strong>in</strong>to <strong>GaAs</strong> was described. The oxide films were deposited by us<strong>in</strong>g metal-organic<br />

chemical vapor deposition. A capp<strong>in</strong>g layer of SiO 2 was deposited on top of the source<br />

films, <strong>and</strong> diffusion was carried out <strong>in</strong> flow<strong>in</strong>g N at 650C. <strong>Diffusion</strong> depths of between<br />

200nm <strong>and</strong> several microns could be easily obta<strong>in</strong>ed. The diffusion front <strong>in</strong> n-type<br />

substrates was sharp. The dependence of the diffusion depth upon the source film<br />

composition (for x-values of 0.04 to 1) was determ<strong>in</strong>ed by us<strong>in</strong>g section<strong>in</strong>g methods.<br />

D.J.Lawrence, F.T.Smith, S.T.Lee: Journal of Applied Physics, 1991, 69[5], 3011-5<br />

[446-78/79-002]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

Rapid thermal anneal<strong>in</strong>g was used to diffuse Zn <strong>in</strong>to <strong>GaAs</strong> from a th<strong>in</strong>-film z<strong>in</strong>c silicate<br />

source that was prepared by chemical vapor deposition at atmospheric pressure.<br />

Comparisons were made with conventional open-tube furnace anneal<strong>in</strong>g, for a diffusion<br />

temperature of 650C. The diffusivities were found to be similar; <strong>in</strong> contrast to previous<br />

results. At temperatures rang<strong>in</strong>g from 650 to 750C, sharp Zn diffusion profiles were<br />

observed. At temperatures above 750C, k<strong>in</strong>ks <strong>in</strong> the diffusion profiles were found. Such<br />

k<strong>in</strong>ks were also observed when semi-<strong>in</strong>sulat<strong>in</strong>g substrates were used <strong>in</strong>stead of Si-doped<br />

n + -type substrates. A model for Zn diffusion had already been developed, <strong>and</strong> this was<br />

based upon the pair<strong>in</strong>g of <strong>in</strong>terstitial Zn with all of the acceptor species which were<br />

present dur<strong>in</strong>g diffusion. The predom<strong>in</strong>ant species were found to be substitutional Zn, <strong>and</strong><br />

Ga vacancies. The concentration of the latter was a function of the background dopant<br />

concentration. The results of the model were shown to agree with all of the present<br />

experimental evidence, <strong>and</strong> were also consistent with the experimental observation of 2<br />

dist<strong>in</strong>ct activation energies for Zn diffusion <strong>in</strong>to n + -doped substrates. These energies were<br />

357


Zn <strong>GaAs</strong> Zn<br />

equal to 1.1 <strong>and</strong> 2.6eV for temperatures which were above or below about 790C,<br />

respectively.<br />

G.Rajeswaran, K.B.Kahen, D.J.Lawrence: Journal of Applied Physics, 1991, 69[3],<br />

1359-65<br />

[446-78/79-014]<br />

330 <strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

Samples were diffused with Zn, via a 200 to 300nm protective ZrO 2 layer. The diffusion<br />

depth exhibited a square-root time dependence. The absolute diffusivity values depended<br />

slightly upon the diffusion conditions (table 30). The layer had essentially no effect upon<br />

the carrier concentration profile or the activation energy.<br />

J.E.Bisberg, A.K.Ch<strong>in</strong>, F.P.Dabkowski: Journal of Applied Physics, 1990, 67[3], 1347-51<br />

[446-74-003]<br />

Table 30<br />

Diffusivity of Zn <strong>in</strong> <strong>GaAs</strong><br />

Protection Source Temperature (C) D (cm 2 /s)<br />

ZrO 2 Zn 755 3.8 x 10 -11<br />

ZrO 2 Zn 700 1.2 x 10 -11<br />

ZrO 2 Zn 650 6.9 x 10 -12<br />

ZrO 2 Zn 600 1.3 x 10 -12<br />

ZrO 2 <strong>GaAs</strong>/Zn 2 As 3 650 2.0 x 10 -12<br />

- <strong>GaAs</strong>/Zn 2 As 3 650 1.6 x 10 -12<br />

ZrO 2 <strong>GaAs</strong>/Zn 2 As 3 600 4.5 x 10 -13<br />

- <strong>GaAs</strong>/Zn 2 As 3 600 3.8 x 10 -13<br />

- <strong>GaAs</strong>/Zn 2 As 3 755 1.2 x 10 -11<br />

- <strong>GaAs</strong>/Zn 2 As 3 700 4.0 x 10 -12<br />

ZrO 2 <strong>GaAs</strong>/Zn 2 As 3 700 3.1 x 10 -12<br />

ZrO 2 <strong>GaAs</strong>/Zn 2 As 3 755 9.7 x 10 -12<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The saturation behavior of the free carrier concentrations <strong>in</strong> p-type <strong>GaAs</strong> monocrystals<br />

which had been doped by Zn diffusion was <strong>in</strong>vestigated. Free-hole saturation occurred at<br />

10 20 /cm 3 . The difference <strong>in</strong> saturation hole concentrations of materials was <strong>in</strong>vestigated<br />

by study<strong>in</strong>g the <strong>in</strong>corporation <strong>and</strong> lattice location of Zn. The latter was an acceptor when<br />

located on a group-<strong>III</strong> atom site. Z<strong>in</strong>c was diffused <strong>in</strong>to <strong>III</strong>-V wafers <strong>in</strong> a sealed quartz<br />

ampoule. Particle-<strong>in</strong>duced X-ray emission <strong>and</strong> ion-channell<strong>in</strong>g techniques were then used<br />

to determ<strong>in</strong>e the exact lattice location of Zn atoms. It was found that more than 90% of<br />

the Zn atoms occupied Ga sites <strong>in</strong> diffused <strong>GaAs</strong> samples. The results were analyzed <strong>in</strong><br />

terms of the amphoteric native defect model. It was shown that differences <strong>in</strong> the<br />

358


Zn <strong>GaAs</strong> Zn<br />

electrical activities of Zn atoms <strong>in</strong> various materials were a consequence of the differ<strong>in</strong>g<br />

locations of the Fermi-level stabilization energy.<br />

L.Y.Chan, K.M.Yu, M.Ben-Tzur, E.E.Haller, J.M.Jaklevic, W.Walukiewicz,<br />

C.M.Hanson: Journal of Applied Physics, 1991, 69[5], 2998-3006<br />

[446-78/79-015]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The use of th<strong>in</strong> Si films for the selective-area diffusion of Si <strong>and</strong> Zn was described. It was<br />

found that Si films behaved as ideal masks for Zn diffusion at temperatures below 750C.<br />

Ideal lateral Zn diffusion profiles were also observed when us<strong>in</strong>g these films; regardless<br />

of the stress at the <strong>in</strong>terface.<br />

G.A.Vawter, E.Omura, X.S.Wu, J.L.Merz, L.Coldren, E.Hu: Journal of Applied Physics,<br />

1988, 63[11], 5541-7<br />

[446-72/73-003]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The growth <strong>and</strong> diffusion of abrupt Zn profiles <strong>in</strong> un-doped, or Si-doped, material was<br />

monitored by means of secondary ion mass spectrometry. The sharp diffusion fronts<br />

which resulted from anneal<strong>in</strong>g treatments <strong>in</strong>dicated that the Zn diffusion coefficient was<br />

concentration-dependent. The data were encompassed by the curves:<br />

D(cm 2 /s) = 5.5 x 10 -10 exp[-0.8(eV)/kT]<br />

<strong>and</strong><br />

D(cm 2 /s) = 1.8 x 10 -9 exp[-1.0(eV)/kT]<br />

However, the diffusion of Zn at high concentrations appeared to be <strong>in</strong>hibited by crystal<br />

defect k<strong>in</strong>etics <strong>and</strong> resulted <strong>in</strong> a relatively concentration-<strong>in</strong>dependent Zn diffusion<br />

coefficient. The V/<strong>III</strong> growth ratio did not have any effect upon Zn diffusion <strong>in</strong> un-doped<br />

or Si-doped material. The diffusion of Zn <strong>in</strong> heterojunction bipolar transistor structures<br />

was different; <strong>in</strong> that the diffusion of Zn <strong>in</strong>to a <strong>GaAs</strong> collector was larger by an order of<br />

magnitude, <strong>and</strong> decreased with an <strong>in</strong>crease <strong>in</strong> the V/<strong>III</strong> growth ratio. In addition, the<br />

diffusion of Zn <strong>in</strong>to an Al<strong>GaAs</strong> emitter was markedly lower <strong>and</strong> was <strong>in</strong>hibited by an<br />

<strong>in</strong>crease <strong>in</strong> the V/<strong>III</strong> ratio. These data could be summarized by the expressions:<br />

<strong>GaAs</strong> (V/<strong>III</strong> = 60): D(cm 2 /s) = 2.0 x 10 3 exp[-3.0(eV)/kT]<br />

<strong>GaAs</strong> (V/<strong>III</strong> = 120): D(cm 2 /s) = 1.0 x 10 6 exp[-3.6(eV)/kT]<br />

Ga 0.78 Al 0.22 As (V/<strong>III</strong> = 60): D(cm 2 /s) = 6.8 x 10 4 exp[-3.4(eV)/kT]<br />

Ga 0.78 Al 0.22 As (V/<strong>III</strong> = 120): D(cm 2 /s) = 1.1 x 10 -5 exp[-1.7(eV)/kT]<br />

P.Enquist, J.A.Hutchby, T.J.De Lyon. Journal of Applied Physics, 1988, 63[9], 4485-93<br />

[446-72/73-012]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The anomalous shape of Zn diffusion profiles <strong>in</strong> <strong>GaAs</strong> was quantitatively expla<strong>in</strong>ed. The<br />

Frank-Turnbull mechanism was suggested to govern <strong>in</strong>terchanges, between <strong>in</strong>terstitial <strong>and</strong><br />

substitutional Zn, via Ga vacancies. It was proposed that these vacancies were either<br />

neutral or were s<strong>in</strong>gly ionized; depend<strong>in</strong>g upon the position of the Fermi level. In<br />

359


Zn <strong>GaAs</strong> Zn<br />

addition, 2 physical phenomena were proposed. Substitutional Zn thermally generated<br />

<strong>in</strong>terstitial Zn-Ga vacancy pairs, <strong>and</strong> there was pair<strong>in</strong>g between the donor (<strong>in</strong>terstitial Zn)<br />

<strong>and</strong> the acceptor (substitutional Zn). The model furnished good agreement with<br />

experimental data.<br />

K.B.Kahen: Applied Physics Letters, 1989, 55[20], 2117-9<br />

[446-72/73-012]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

A review was presented of self-diffusion mechanisms <strong>and</strong> dop<strong>in</strong>g-enhanced superlattice<br />

disorder<strong>in</strong>g. With regard to the <strong>in</strong>fluence of Zn p-type dopants, the Fermi level effect had<br />

to be considered. In accord with its effect upon superlattice disorder<strong>in</strong>g, Zn diffusion<br />

appeared to be governed by the kick-out mechanism. It was concluded that dislocations <strong>in</strong><br />

this material <strong>and</strong> <strong>in</strong> <strong>other</strong> <strong>III</strong>-V compounds were only moderately efficient s<strong>in</strong>ks or<br />

sources for po<strong>in</strong>t defects.<br />

T.Y.Tan, U.Gösele: Materials Science <strong>and</strong> Eng<strong>in</strong>eer<strong>in</strong>g, 1988, B1, 47-65<br />

[446-62/63-208]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The diffusion of Zn <strong>in</strong>to <strong>GaAs</strong>, from diethylz<strong>in</strong>c <strong>and</strong> trimethylarsenic, was studied. The<br />

process produced surface hole concentrations which were greater than 10 20 /cm 3 . Wellcontrolled<br />

junction depths which could be as shallow as 0.000lmm were obta<strong>in</strong>ed, <strong>and</strong> a<br />

smooth surface morphology was reta<strong>in</strong>ed. The profile shape was much more complex<br />

than those predicted by accepted <strong>in</strong>terstitial cum substitutional Zn diffusion models. In<br />

order to expla<strong>in</strong> the observed profiles, a new model for Zn diffusion was proposed, <strong>and</strong><br />

used <strong>in</strong> a computer simulation.<br />

S.Reynolds, D.W.Vook, J.F.Gibbons: Journal of Applied Physics, 1988, 63[4], 1052-9<br />

[446-62/63-211]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

Data were presented which showed that low-temperature (680C) Zn diffusion was<br />

effective <strong>in</strong> reduc<strong>in</strong>g the dislocation density of epitaxial <strong>GaAs</strong> which was grown onto Si.<br />

The <strong>GaAs</strong>/Si system was analyzed by us<strong>in</strong>g both cross-sectional <strong>and</strong> plan-view<br />

transmission electron microscopy. Simple thermal anneal<strong>in</strong>g of <strong>GaAs</strong>/Si at a higher<br />

temperature (850C) was also studied. The reduction <strong>in</strong> dislocation density which occurred<br />

due to Zn diffusion was suggested to arise from the <strong>in</strong>creased concentrations of po<strong>in</strong>t<br />

defects which were generated dur<strong>in</strong>g Zn diffusion. This resulted <strong>in</strong> enhanced dislocation<br />

climb. The mechanism was consistent with impurity-<strong>in</strong>duced layer disorder<strong>in</strong>g, via Zn<br />

diffusion, <strong>in</strong> Al<strong>GaAs</strong>/<strong>GaAs</strong> heterostructures.<br />

D.G.Deppe, N.Holonyak, K.C.Hsieh, D.W.Nam, W.E.Plano, R.J.Matyi, H.Shichijo:<br />

Applied Physics Letters, 1988, 52[21], 1812-4<br />

[446-62/63-211]<br />

360


Zn <strong>GaAs</strong> Zn<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

Second diffusion of Zn was observed us<strong>in</strong>g low concentrations. The behavior was similar<br />

to that of double diffusion <strong>in</strong> InP. The effect of the Zn activity <strong>in</strong> the vapor phase was<br />

studied by us<strong>in</strong>g a semi-closed box system. The observed Zn profiles were expla<strong>in</strong>ed <strong>in</strong><br />

terms of a model which <strong>in</strong>volved vary<strong>in</strong>g charge transfer by vacancy centers dur<strong>in</strong>g<br />

<strong>in</strong>terstitial-substitutional <strong>in</strong>terchanges.<br />

K.Kazmierski, F.Launay, B.De Cremoux: Japanese Journal of Applied Physics, 1987,<br />

26[10], 1630-3<br />

[446-55/56-007]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

A ternary Zn 3 As 2 -ZnAs 2 -<strong>GaAs</strong> source for the diffusion of Zn <strong>in</strong>to <strong>GaAs</strong> was developed<br />

by us<strong>in</strong>g a low-temperature s<strong>in</strong>ter<strong>in</strong>g technique. It was noted that wafers which were<br />

diffused by us<strong>in</strong>g this source rema<strong>in</strong>ed free from damage. The dopant concentrations <strong>and</strong><br />

diffusion depths agreed with the results which were obta<strong>in</strong>ed by us<strong>in</strong>g high-temperature<br />

Ga/As/Zn diffusion sources.<br />

J.Werner, H.Melchior: Japanese Journal of Applied Physics, 1987, 26[4], 641-2<br />

[446-51/52-117]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

A new method for self-aligned Si-Zn diffusion was described. In this method, closed-tube<br />

Si diffusion was carried out by us<strong>in</strong>g a sputtered SiN x film. Then, Zn diffusion which was<br />

self-aligned to the Si diffusion w<strong>in</strong>dow was carried out by re-us<strong>in</strong>g the SiN x film as a<br />

mask. The key factor was that the SiN x film should have the correct refractive <strong>in</strong>dex<br />

profile.<br />

W.X.Zou, R.Boudreau, H.T.Han, T.Bowen, S.S.Shi, D.S.L.Mui, J.L.Merz: Journal of<br />

Applied Physics, 1995, 77[12], 6244-6<br />

[446-121/122-045]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The is<strong>other</strong>mal diffusion of Zn was described <strong>in</strong> terms of the Long<strong>in</strong>i reaction <strong>and</strong> a<br />

recomb<strong>in</strong>ation process. It was assumed that, dur<strong>in</strong>g diffusion, highly mobile Zn<br />

<strong>in</strong>terstitials recomb<strong>in</strong>ed with a Ga vacancy <strong>and</strong> became a relatively immobile site defect.<br />

It was noted that the long-term profile of the total Zn concentration was governed ma<strong>in</strong>ly<br />

by the vacancy concentration profile. All of the known concentration profiles could be<br />

obta<strong>in</strong>ed for a constant diffusivity, without us<strong>in</strong>g fitt<strong>in</strong>g parameters. A new method was<br />

proposed for determ<strong>in</strong><strong>in</strong>g the diffusivities of <strong>in</strong>terstitial Zn, of Zn <strong>in</strong> Ga lattice sites, <strong>and</strong><br />

of vacancies. These coefficients were deduced from exist<strong>in</strong>g experimental data. It was<br />

shown that the apparent dependence of the Zn diffusivity upon its background<br />

concentration was due to its recomb<strong>in</strong>ation with Ga vacancies.<br />

N.N.Grigorev, T.A.Kudyk<strong>in</strong>a: Fizika i Tekhnika Poluprovodnikov, 1997, 31[6], 697-702<br />

(<strong>Semiconductors</strong>, 1997, 31[6], 595-9)<br />

[446-157/159-361]<br />

361


Zn <strong>GaAs</strong> Zn<br />

331 <strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The diffusion of implanted Zn was studied, at anneal<strong>in</strong>g temperatures of between 625 <strong>and</strong><br />

850C, by means of secondary ion mass spectrometry. A substitutional-<strong>in</strong>terstitial<br />

diffusion mechanism was proposed <strong>in</strong> order to expla<strong>in</strong> how deviations of the local Ga<br />

<strong>in</strong>terstitial concentration, from its equilibrium value, regulated Zn diffusion. It was found<br />

that it was possible to simulate both box-shaped profiles, that resulted from hightemperature<br />

anneal<strong>in</strong>g, <strong>and</strong> k<strong>in</strong>k-<strong>and</strong>-tail profiles which resulted from lower-temperature<br />

anneal<strong>in</strong>g. The simulation data permitted the determ<strong>in</strong>ation of Arrhenius relationships for<br />

the <strong>in</strong>tr<strong>in</strong>sic diffusion coefficient of implanted Zn. This (table 31) could be described by:<br />

D (cm 2 /s) = 0.6075 exp[-3.21(eV)/kT]<br />

M.P.Chase, M.D.Deal, J.D.Plummer: Journal of Applied Physics, 1997, 81[4], 1670-6<br />

[446-148/149-172]<br />

Table 31<br />

Diffusivity of Zn <strong>in</strong> <strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

850 1.9 x 10 -15<br />

800 1.2 x 10 -15<br />

800 6.0 x 10 -16<br />

750 1.2 x 10 -16<br />

750 8.4 x 10 -17<br />

675 8.4 x 10 -18<br />

675 3.1 x 10 -18<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The diffusion behavior was <strong>in</strong>vestigated by us<strong>in</strong>g a spun-on doped silica film. <strong>Diffusion</strong><br />

anneal<strong>in</strong>g was carried out, at temperatures which ranged from 650 to 800C, us<strong>in</strong>g a SiO 2<br />

or Si 3 N 4 cap. At higher temperatures, the sample surface after diffusion was found to be<br />

damaged. The profiles which were obta<strong>in</strong>ed at temperatures rang<strong>in</strong>g from 700 to 780C<br />

were of abrupt or k<strong>in</strong>k-<strong>and</strong>-tail type. <strong>Diffusion</strong> anneal<strong>in</strong>g which was carried out us<strong>in</strong>g a<br />

nitride cap were termed Ga-rich, <strong>and</strong> produced abrupt box-like profiles. The use of an<br />

oxide cap was expected to produce a superposition of shallow As-rich <strong>and</strong> deeper Ga-rich<br />

profiles; lead<strong>in</strong>g to a characteristic k<strong>in</strong>k-<strong>and</strong>-tail profile. This behavior was observed at<br />

higher temperatures, but the junctions were too shallow to exhibit a tail at lower<br />

temperatures. The junction depths were attributed to the existence of differ<strong>in</strong>g activation<br />

energies for the two types of cap material. The energy that was required for oxide-cap<br />

diffusion was close to published values for oxide caps. The lower values which were<br />

found for nitride-cap diffusion were closer to published values for a phosphosilicate glass<br />

362


Zn <strong>GaAs</strong> Zn<br />

cap. These results clearly showed that the nature of the cap material determ<strong>in</strong>ed the type<br />

of diffusion profile.<br />

S.Chatterjee, K.N.Bhat, P.R.S.Rao: Solid-State Electronics, 1997, 41[3], 496-500<br />

[446-148/149-174]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The <strong>in</strong>-diffusion of Zn at high concentrations, under As-deficient conditions, caused the<br />

generation of dislocation loops, elongated dislocations <strong>and</strong> Ga precipitates decorated with<br />

voids with<strong>in</strong> the diffusion zone. Similar treatment under As vapor led to the recovery of<br />

diffusion-<strong>in</strong>duced damage <strong>in</strong> the sub-surface region. This was accompanied by the<br />

appearance of 2 dist<strong>in</strong>ct steps <strong>in</strong> the Zn concentration profile. Previous work had<br />

suggested that these phenomena were connected with the out-diffusion of Ga from<br />

precipitates <strong>and</strong> towards the surface. The present results showed that the replacement, by<br />

P or Sb, of As <strong>in</strong> the diffusion ambient produced similar recovery effects.<br />

G.Boesker, H.G.Hettwer, A.Rucki, N.A.Stolwijk, H.Mehrer, W.Jaeger, K.Urban:<br />

Materials Chemistry <strong>and</strong> Physics, 1995, 42[1], 68-71<br />

[446-136/137-110]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The diffusion of Zn, at high concentrations, <strong>in</strong>to samples under As-deficient ambient<br />

conditions led to the generation of defects such as dislocation loops, elongated<br />

dislocations <strong>and</strong> Ga precipitates - decorated with voids- throughout the diffusion zone.<br />

Similar treatments under As vapor led to the recovery of diffusion-<strong>in</strong>duced damage <strong>in</strong><br />

regions beneath the surface. This was accompanied by the appearance of 2 dist<strong>in</strong>ct steps<br />

<strong>in</strong> the Zn concentration profile. Previous work had suggested that these phenomena were<br />

associated with the out-diffusion of Ga from the precipitates <strong>and</strong> towards the surface. It<br />

was shown here that the replacement of As by P or Sb produced similar recovery effects.<br />

This <strong>in</strong>dicated that ambient group-V elements made the near-surface of <strong>GaAs</strong> <strong>in</strong>to an<br />

effective s<strong>in</strong>k for diffusion-<strong>in</strong>duced excess Ga.<br />

G.Bösker, H.G.Hettwer, A.Rucki, N.A.Stolwijk, H.Mehrer, W.Jäger, K.Urban: Materials<br />

Chemistry <strong>and</strong> Physics, 1995, 42[1], 68-71<br />

[446-134/135-126]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The diffusivity of Zn <strong>in</strong> heavily-doped pnpn structures was measured after growth <strong>and</strong><br />

anneal<strong>in</strong>g. Dur<strong>in</strong>g growth at 650C, the Zn diffusivity of about 10 -12 cm 2 /s <strong>in</strong> the buried p-<br />

type layer was found to be more than 10000 times the Zn diffusivity <strong>in</strong> the top p-type<br />

layer. Dur<strong>in</strong>g anneal<strong>in</strong>g at 800C, the Zn diffusivity of about 5 x 10 -14 cm 2 /s <strong>in</strong> the buried<br />

layer rema<strong>in</strong>ed orders of magnitude greater than the Zn diffusivity <strong>in</strong> the top layer. The<br />

measurements provided clear experimental evidence that a large flux of Ga <strong>in</strong>terstitials<br />

was <strong>in</strong>jected from the surface dur<strong>in</strong>g the growth of n-type layers, <strong>and</strong> that the Ga<br />

<strong>in</strong>terstitials were trapped <strong>in</strong> the buried p-type layer by the electric field of the pn junctions<br />

(<strong>and</strong> were therefore positively charged). It was suggested that the resultant large<br />

concentration of Ga <strong>in</strong>terstitials <strong>in</strong> the buried layer accounted for the <strong>in</strong>creased Zn<br />

363


Zn <strong>GaAs</strong> Zn<br />

diffusivity via a kick-out mechanism. F<strong>in</strong>ally, it was deduced that the mobile Zn<br />

<strong>in</strong>terstitial was positively charged.<br />

C.Y.Chen, R.M.Cohen, D.S.Simons, P.H.Chi: Applied Physics Letters, 1995, 67[10],<br />

1402-4<br />

[446-125/126-121]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

It was recalled that the rapid diffusion of Zn <strong>in</strong>to <strong>GaAs</strong> had recently been attributed to the<br />

fact that a small fraction of the Zn <strong>in</strong>terstitials changed to Ga sites, thereby produc<strong>in</strong>g<br />

<strong>in</strong>terstitial Ga (I Ga ). This kick-out reaction led to the possibility of determ<strong>in</strong><strong>in</strong>g the I Ga<br />

transport properties from Zn diffusion experiments on essentially perfect <strong>GaAs</strong>. However,<br />

previous attempts had been impeded by the diffusion-<strong>in</strong>duced generation of<br />

microstructural defects which acted as I Ga s<strong>in</strong>ks. In the present study, this was prevented<br />

by us<strong>in</strong>g Zn-doped <strong>GaAs</strong> powder as a diffusion source. The measured 2-stage profiles<br />

showed that, under these conditions, Zn diffusion at 906C was controlled by I Ga 3+ <strong>in</strong><br />

addition to I Ga 2+ . Analysis of the profiles yielded quantitative data on the Ga- <strong>and</strong> Znrelated<br />

diffusivities, <strong>and</strong> on the concentration of I Ga .<br />

G.Bösker, N.A.Stolwijk, H.G.Hettwer, A.Rucki, W.Jäger, U.Södervall: Physical Review<br />

B, 1995, 52[16], 11927-31<br />

[446-125/126-121]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The Zn was diffused <strong>in</strong>to <strong>GaAs</strong> through anodic oxide layers with various thicknesses <strong>and</strong><br />

densities. Electrochemical profil<strong>in</strong>g was used to determ<strong>in</strong>e the electrically active Zn<br />

concentration <strong>and</strong> the diffusion depth. It was found that the depth of the junction varied<br />

<strong>in</strong>versely with the thickness <strong>and</strong> density of the oxide. However, the surface concentration<br />

appeared to be <strong>in</strong>dependent of the oxide thickness or density <strong>and</strong> atta<strong>in</strong>ed a value which<br />

was identical to that which was found for diffusion <strong>in</strong>to a bare <strong>GaAs</strong> sample. The results<br />

demonstrated that the most important effect of the oxide was to delay the <strong>in</strong>troduction of<br />

Zn <strong>in</strong>to the lattice. Thus, the anodic oxide could not be used as a mask or a Zn<br />

concentration attenuator.<br />

H.Cutlerywala, R.J.Roedel: Journal of the Electrochemical Society, 1994, 141[6], 1639-<br />

43<br />

[446-119/120-192]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The Zn was <strong>in</strong>troduced by us<strong>in</strong>g the electron beam dop<strong>in</strong>g method. That is, a Zn sheet<br />

was s<strong>and</strong>wiched between <strong>GaAs</strong> wafers <strong>and</strong> the surface of the <strong>GaAs</strong> was irradiated with<br />

7MeV electrons. The use of secondary ion mass spectroscopy revealed U-shaped<br />

diffusion profiles for impurities <strong>in</strong> the substrates. The results could be expla<strong>in</strong>ed <strong>in</strong> terms<br />

of the kick-out mechanism, <strong>and</strong> surface diffusion processes.<br />

A.Takeda, T.Wada: Materials Science Forum, 1994, 143-147, 1421-6<br />

[446-113/114-013]<br />

364


Zn <strong>GaAs</strong> Zn<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

Secondary ion mass spectroscopy <strong>and</strong> photolum<strong>in</strong>escence studies were made of V Ga<br />

dur<strong>in</strong>g Zn diffusion <strong>in</strong>to Si-doped material. Photolum<strong>in</strong>escence spectra were obta<strong>in</strong>ed at<br />

various etch<strong>in</strong>g depths below the sample surface. After anneal<strong>in</strong>g the samples <strong>in</strong> excess<br />

As 4 vapor at 650C, the conversion of n-type material <strong>in</strong>to p-type material was observed<br />

by mak<strong>in</strong>g electrical measurements <strong>in</strong> the region near to the sample surface. The<br />

importance of the Si Ga -V Ga emission b<strong>and</strong> <strong>in</strong> photolum<strong>in</strong>escence spectra from thermally<br />

converted regions <strong>in</strong>dicated that V Ga which were generated at the surface dur<strong>in</strong>g<br />

anneal<strong>in</strong>g were responsible for the thermal conversion. The results also showed that the<br />

ma<strong>in</strong> po<strong>in</strong>t defects which were generated dur<strong>in</strong>g anneal<strong>in</strong>g under Ga-rich conditions were<br />

V As <strong>and</strong> Ga As . In the case of Zn-diffused Si-doped substrates at 600C, the disappearance<br />

of the V Ga -related b<strong>and</strong> from the photolum<strong>in</strong>escence spectra of diffused regions furnished<br />

evidence for the <strong>in</strong>corporation of Zn <strong>in</strong>terstitials <strong>in</strong>to Ga sites dur<strong>in</strong>g diffusion. An<br />

accumulation of V Ga was found ahead of the Zn diffusion front. The Zn-diffused samples<br />

were also annealed at 800C for 2h <strong>in</strong> vacuum, or <strong>in</strong> As 4 vapor with or without a Si 3 N 4<br />

cap. In the case of samples which were annealed <strong>in</strong> vacuum, an abrupt diffusion front<br />

advanced slightly <strong>in</strong>to the bulk; with a supersaturation of V Ga ahead of the front. On the<br />

<strong>other</strong> h<strong>and</strong>, samples which were annealed <strong>in</strong> As vapor, with or without a cap on the<br />

surface, exhibited double Zn concentration profiles with an undersaturation of V Ga around<br />

the tail region. These results revealed the important role which was played by nonequilibrium<br />

po<strong>in</strong>t defects, <strong>and</strong> were expla<strong>in</strong>ed <strong>in</strong> terms of a kick-out mechanism for Zn<br />

diffusion.<br />

N.H.Ky, J.D.Ganière, F.K.Re<strong>in</strong>hart, B.Blanchard: Materials Science Forum, 1994, 143-<br />

147, 1397-402<br />

[446-113/114-014]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

Experimental studies were made of Zn diffusion, us<strong>in</strong>g high temperatures, open tubes,<br />

<strong>and</strong> SiO 2 protective layers. Precisely controlled diffusion depths of less than 0.2 or 0.4µ<br />

could be obta<strong>in</strong>ed by us<strong>in</strong>g SiO 2 , doped with 0.1%Zn, as a diffusion source. Under these<br />

conditions, the diffusion coefficients were equal to 1.31 x 10 -13 <strong>and</strong> 1.76 x 10 -12 cm 2 /s at<br />

temperatures of 700 <strong>and</strong> 1025C, respectively.<br />

D.K.Gautam, Y.Shimogaki, Y.Nakano, K.Tada: Materials Science Forum, 1993, 117-<br />

118, 417-22<br />

[446-111/112-051]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

After Zn diffusion <strong>in</strong>to Si-doped material, the diffused samples were annealed <strong>in</strong> vacuum,<br />

<strong>in</strong> As vapor, or with a Si 3 N 4 mask capp<strong>in</strong>g the surface. The Zn concentration profiles<br />

which were obta<strong>in</strong>ed by secondary ion mass spectroscopy, <strong>and</strong> photolum<strong>in</strong>escence spectra<br />

for various depths below the sample surface, were studied <strong>in</strong> detail. After anneal<strong>in</strong>g <strong>in</strong><br />

vacuum, the steep (p + -n) Zn diffusion front advanced <strong>in</strong>to the bulk. It was observed that<br />

365


Zn <strong>GaAs</strong> Zn<br />

the <strong>in</strong>tensity ratio between the Si donor-Ga vacancy complex (Si Ga -V Ga ) related emission<br />

b<strong>and</strong> <strong>and</strong> the b<strong>and</strong>-to-b<strong>and</strong> (e-h) transition was enhanced <strong>in</strong> the region ahead of the Zn<br />

diffusion front. On the <strong>other</strong> h<strong>and</strong>, Zn atoms diffused deeper <strong>in</strong>to the bulk of samples<br />

which were annealed <strong>in</strong> As vapor, with or without a capp<strong>in</strong>g layer. These samples<br />

exhibited k<strong>in</strong>k-<strong>and</strong>-tail (p + -p-n ) Zn concentration profiles with a decrease <strong>in</strong> the <strong>in</strong>tensity<br />

ratio around the tail region. Analysis of the photolum<strong>in</strong>escence data suggested that there<br />

was a supersaturation of Ga vacancies ahead of the diffusion front of samples which were<br />

annealed <strong>in</strong> vacuum, <strong>and</strong> an under-saturation of this defect around the tail region of<br />

samples which were annealed <strong>in</strong> As vapor. The results emphasized the important role<br />

which was played by non-equilibrium of the defect concentration dur<strong>in</strong>g post-diffusion<br />

anneal<strong>in</strong>g. This permitted an anomalous Zn double profile to be expla<strong>in</strong>ed <strong>in</strong> terms of the<br />

<strong>in</strong>terstitial-substitutional mechanism.<br />

N.H.Ky, J.D.Ganière, F.K.Re<strong>in</strong>hart, B.Blanchard, J.C.Pfister: Journal of Applied Physics,<br />

1993, 74[9], 5493-500<br />

[446-111/112-051]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The microscopic mechanisms of Zn diffusion <strong>in</strong> <strong>GaAs</strong> were <strong>in</strong>vestigated by us<strong>in</strong>g ab<br />

<strong>in</strong>itio molecular dynamics techniques. It was found that, among the various proposed<br />

mechanisms for Zn diffusion, kick-out by Ga <strong>in</strong>terstitials had the lowest activation<br />

energy. The occurrence of Zn <strong>in</strong>-diffusion generated non-equilibrium group-<strong>III</strong><br />

<strong>in</strong>terstitials which were bound to Zn by Coulomb forces. The <strong>in</strong>terstitials followed the Zn<br />

diffusion front <strong>and</strong> disordered the superlattice. The calculated activation energies for<br />

these processes were <strong>in</strong> good agreement with experimental data.<br />

C.Wang, Q.M.Zhang, J.Bernholc: Physical Review Letters, 1992, 69[26], 3789-92<br />

[446-106/107-039]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The electrical activity <strong>and</strong> lattice-site locations of Zn atoms which had been diffused <strong>in</strong>to<br />

<strong>GaAs</strong> were studied by us<strong>in</strong>g various characterization techniques. Particle-<strong>in</strong>duced X-ray<br />

emission channell<strong>in</strong>g showed that all of the Zn atoms were substitutional <strong>and</strong> were<br />

electrically active acceptors. A difference between the behaviors of Zn <strong>in</strong> <strong>GaAs</strong> <strong>and</strong> InP<br />

could be understood <strong>in</strong> terms of the amphoteric native defect model. It was also shown<br />

that the Fermi level stabilization energy provided a convenient energy reference for the<br />

treatment of dopant diffusion at semiconductor hetero-<strong>in</strong>terfaces.<br />

W.Walukiewicz, K.M.Yu, L.Y.Chan, J.Jaklevic, E.E.Haller: Materials Science Forum,<br />

1992, 83-87, 941-6<br />

[446-99/100-065]<br />

<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

A model was developed which could expla<strong>in</strong> the nature of Zn diffusion profiles <strong>in</strong> n + -type<br />

material. The model was based upon the effect of Coulomb pair<strong>in</strong>g between <strong>in</strong>terstitial<br />

<strong>and</strong> substitutional Zn. By extend<strong>in</strong>g the model so as to <strong>in</strong>clude the Coulomb pair<strong>in</strong>g of<br />

<strong>in</strong>terstitial Zn with all of the acceptors which were present dur<strong>in</strong>g diffusion, the<br />

366


Zn <strong>GaAs</strong> Zn<br />

predictions were caused to be <strong>in</strong> good agreement with experimental data. Only one<br />

adjustable parameter was <strong>in</strong>volved.<br />

K.B.Kahen, J.P.Spence, G.Rajeswaran: Journal of Applied Physics, 1991, 70[4], 2464-6<br />

[446-91/92-008]<br />

<strong>GaAs</strong>/AlAs: Zn <strong>Diffusion</strong><br />

The microscopic mechanisms of Zn-<strong>in</strong>duced <strong>in</strong>terdiffusion <strong>in</strong> <strong>GaAs</strong>/AlAs superlattices,<br />

were <strong>in</strong>vestigated by us<strong>in</strong>g ab <strong>in</strong>itio molecular dynamics techniques. It was found that,<br />

among the various proposed mechanisms for Zn diffusion, kick-out by Ga <strong>in</strong>terstitials had<br />

the lowest activation energy. The occurrence of Zn <strong>in</strong>-diffusion generated nonequilibrium<br />

group-<strong>III</strong> <strong>in</strong>terstitials which were bound to Zn by Coulomb forces. The<br />

<strong>in</strong>terstitials followed the Zn diffusion front <strong>and</strong> disordered the superlattice. The calculated<br />

activation energies for these processes were <strong>in</strong> good agreement with experimental data.<br />

C.Wang, Q.M.Zhang, J.Bernholc: Physical Review Letters, 1992, 69[26], 3789-92<br />

[446-106/107-039]<br />

<strong>GaAs</strong>/AlAs: Zn <strong>Diffusion</strong><br />

A model was proposed for the effect of Zn <strong>in</strong>-diffusion <strong>in</strong> enhanc<strong>in</strong>g superlattice<br />

disorder<strong>in</strong>g. It comb<strong>in</strong>ed recently proposed models for Ga self-diffusion <strong>and</strong> Zn diffusion<br />

<strong>in</strong> <strong>GaAs</strong>. Four coupled partial differential equations, which described the process, were<br />

solved numerically. Satisfactory agreement was obta<strong>in</strong>ed between the simulated results<br />

<strong>and</strong> published experimental data. At a given temperature, <strong>and</strong> for the values which were<br />

assumed for the diffusion coefficient <strong>and</strong> thermal equilibrium concentration of po<strong>in</strong>t<br />

defects, doubly positively charged Ga self-<strong>in</strong>terstitials, I Ga 2+ , were deduced to be a<br />

consistent splitt<strong>in</strong>g of the known Ga self-diffusion coefficient which was dom<strong>in</strong>ated by<br />

I Ga 2+ . The superlattice disorder<strong>in</strong>g enhancement was due ma<strong>in</strong>ly to the Fermi-level effect,<br />

but I Ga 2+ supersaturation also made a small contribution. Because of p-dop<strong>in</strong>g by Zn<br />

acceptor atoms, the I Ga 2+ concentration was greatly <strong>in</strong>creased via the Fermi-level effect.<br />

An I Ga 2+ supersaturation also developed because the I Ga 2+ generation rate was higher than<br />

its removal rate. Enhanced superlattice disorder<strong>in</strong>g occurred ma<strong>in</strong>ly under Ga-rich<br />

superlattice conditions. The Zn <strong>in</strong>-diffusion enhanced Al-Ga <strong>in</strong>terdiffusion coefficient<br />

exhibited an apparent dependence, upon the Zn s - concentration, which differed slightly<br />

from a quadratic relationship.<br />

H.Zimmermann, U.Gösele, T.Y.Tan: Journal of Applied Physics, 1993, 73[1], 150-7<br />

[446-106/107-078]<br />

<strong>GaAs</strong>/AlAs: Zn <strong>Diffusion</strong><br />

The Car-Parr<strong>in</strong>ello method was used to study Zn-enhanced <strong>in</strong>terdiffusion <strong>in</strong> superlattices.<br />

The energetics of several mechanisms for the diffusion of Zn were exam<strong>in</strong>ed. It was<br />

found that a pair which consisted of a substitutional Zn acceptor <strong>and</strong> an <strong>in</strong>terstitial group-<br />

<strong>III</strong> atom had a substantially lower formation energy than did an isolated <strong>in</strong>terstitial. The<br />

low formation energy of this pair resulted <strong>in</strong> the <strong>in</strong>terstitial kick-out mechanism hav<strong>in</strong>g a<br />

much lower activation energy than those which <strong>in</strong>volved vacancies or the dissociative<br />

367


Zn <strong>GaAs</strong> Zn<br />

(Frank-Turnbull or Long<strong>in</strong>i) mechanism. The lowest-energy path for the <strong>in</strong>terchange of<br />

group-<strong>III</strong> atoms <strong>in</strong>volved the kick-out of Zn by a group-<strong>III</strong> <strong>in</strong>terstitial, followed by fast Zn<br />

<strong>in</strong>terstitial diffusion <strong>and</strong> the subsequent ejection of an<strong>other</strong> group-<strong>III</strong> atom <strong>in</strong>to the<br />

<strong>in</strong>terstitial channel. The activation energies for these processes, as determ<strong>in</strong>ed by<br />

follow<strong>in</strong>g the kick-out trajectories <strong>and</strong> <strong>in</strong>clud<strong>in</strong>g full relaxation of all of the atoms, were<br />

<strong>in</strong> good agreement with experimental data.<br />

Q.M.Zhang, C.Wang, J.Bernholc: Materials Science Forum, 1992, 83-87, 1351-6<br />

[446-99/100-083]<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

Multiple quantum well structures, with the same well thickness but differ<strong>in</strong>g Al x Ga 1-x As<br />

compositions (with x = 0.1, 0.2, 0.45, or 1), were grown by us<strong>in</strong>g molecular beam<br />

epitaxy. After Zn diffusion at 575C (1 to 16h), the structures were studied by means of<br />

the transmission electron microscopy of cleaved wedges, secondary electron imag<strong>in</strong>g <strong>in</strong> a<br />

scann<strong>in</strong>g electron microscope, <strong>and</strong> by means of secondary ion mass spectroscopy. The<br />

results showed that totally <strong>and</strong> partially disordered regions always lay beyond the Zn<br />

diffusion front. The extent of partial disorder<strong>in</strong>g depended upon the value of x. As x<br />

<strong>in</strong>creased, the disorder<strong>in</strong>g rate <strong>in</strong>creased due to an <strong>in</strong>crease <strong>in</strong> Zn diffusivity. The effect of<br />

a high Zn concentration was <strong>in</strong>vestigated by monitor<strong>in</strong>g the photolum<strong>in</strong>escence <strong>and</strong><br />

Raman scatter<strong>in</strong>g. Analysis of the photolum<strong>in</strong>escence spectra of structures which had<br />

been diffused for various times, <strong>and</strong> of the photolum<strong>in</strong>escence spectra from various<br />

depths below the sample surface, made it possible to describe the physical processes<br />

which occurred dur<strong>in</strong>g Zn diffusion. Column-<strong>III</strong> vacancies were created at the sample<br />

surface <strong>and</strong> diffused <strong>in</strong>to the bulk of the sample, where they were filled by <strong>other</strong> defects.<br />

By us<strong>in</strong>g X-ray diffraction techniques, an expansion of the lattice constant <strong>in</strong> the region<br />

beyond the Zn diffusion front was detected. This was attributed to a supersaturation of<br />

column-<strong>III</strong> <strong>in</strong>terstitials. Dur<strong>in</strong>g the <strong>in</strong>corporation of Zn <strong>in</strong>to the crystal lattice, column-<strong>III</strong><br />

<strong>in</strong>terstitials were generated. These were suggested to be responsible for the enhancement<br />

of Al-Ga <strong>in</strong>terdiffusion. An important role was played by the electric field at the p-n<br />

junction that was formed by Zn diffusion. That is, the negatively charged column-<strong>III</strong><br />

vacancies <strong>and</strong> the positively charged column-<strong>III</strong> <strong>in</strong>terstitials were conf<strong>in</strong>ed to the n <strong>and</strong> p<br />

sides of the p-n junction, respectively. These results provided evidence for a self<strong>in</strong>terstitial<br />

mechanism of Zn diffusion-<strong>in</strong>duced disorder<strong>in</strong>g <strong>in</strong> these multiple quantum well<br />

structures.<br />

N.H.Ky, J.D.Ganière, M.Gailhanou, B.Blanchard, L.Pavesi, G.Burri, D.Araújo,<br />

F.K.Re<strong>in</strong>hart: Journal of Applied Physics, 1993, 73[8], 3769-81<br />

[446-106/107-079]<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

An <strong>in</strong>vestigation was made of the diffusion of Zn <strong>in</strong> n-<strong>GaAs</strong>/Zn-Al<strong>GaAs</strong>/Se-Al<strong>GaAs</strong><br />

structures dur<strong>in</strong>g the growth of n-<strong>GaAs</strong> layers which were doped with Se <strong>and</strong> Si. It was<br />

found that the diffusion of Zn <strong>in</strong> these structures depended strongly upon the type of<br />

dopant as well as upon the carrier concentration <strong>in</strong> the n-<strong>GaAs</strong> layer. The amount of Zn<br />

which diffused <strong>in</strong>to both n-<strong>GaAs</strong> <strong>and</strong> Se-Al<strong>GaAs</strong> layers was much smaller for a Si-doped<br />

368


Zn <strong>GaAs</strong> Zn<br />

<strong>GaAs</strong> layer than for a Se-doped layer. The slower diffusion which occurred dur<strong>in</strong>g the<br />

growth of Si-doped <strong>GaAs</strong> layers <strong>in</strong> these structures could be reasonably well expla<strong>in</strong>ed by<br />

modify<strong>in</strong>g a model <strong>in</strong> which <strong>in</strong>terstitial Ga which diffused from the n-<strong>GaAs</strong> layer <strong>and</strong> <strong>in</strong>to<br />

the Zn-Al<strong>GaAs</strong> layer was supposed to kick out substitutional Zn. The density of<br />

<strong>in</strong>terstitial Ga <strong>in</strong> the Si-doped <strong>GaAs</strong> layer could be lower than <strong>in</strong> the Se-doped <strong>GaAs</strong><br />

because the <strong>in</strong>terstitial Ga atoms replaced Si which occupied the column-<strong>III</strong> site. This was<br />

not the case for Se-doped <strong>GaAs</strong>, where Se occupied the column-V site.<br />

N.Fujii, T.Kimura, M.Tsugami, T.Sonoda, S.Takamiya, S.Mitsui: Journal of Crystal<br />

Growth, 1994, 145[1-4], 808-12<br />

[446-119/120-201]<br />

<strong>GaAs</strong>/GaAlAs: Zn <strong>Diffusion</strong><br />

Heterostructures of <strong>GaAs</strong>/Ga 0.7 Al 0.3 As, which conta<strong>in</strong>ed Zn <strong>and</strong> Se as <strong>in</strong>tr<strong>in</strong>sic p <strong>and</strong> n<br />

dopants, were subjected to comb<strong>in</strong>ed Be <strong>and</strong> O implantation. Rapid thermal anneal<strong>in</strong>g<br />

then resulted <strong>in</strong> the enhanced out-diffusion of Zn. The Se dopant profile rema<strong>in</strong>ed<br />

essentially unchanged. The atomic profile of Zn could be related to the microscopic<br />

defect distributions. A change <strong>in</strong> the photolum<strong>in</strong>escence spectrum, due to overcompensation<br />

of the n-doped <strong>GaAs</strong> <strong>and</strong> GaAlAs layers, was observed. Anneal<strong>in</strong>g without<br />

preced<strong>in</strong>g implantation had no effect upon the Zn atomic profile.<br />

T.Humer-Hager, R.Treichler, P.Wurz<strong>in</strong>ger, H.Tews, P.Zwicknagl: Journal of Applied<br />

Physics, 1989, 66[1], 181-6<br />

[446-74-026]<br />

<strong>GaAs</strong>/GaAlAs: Zn <strong>Diffusion</strong><br />

Multiple quantum-well <strong>GaAs</strong>/Ga 0.8 Al 0.2 As structures which were uniformly Si-doped, to<br />

concentrations rang<strong>in</strong>g from 10 17 to 10 19 /cm 3 , were grown by means of molecular-beam<br />

epitaxy <strong>in</strong> order to study the effects of the background Si dopant level upon the Zn<br />

diffusion-<strong>in</strong>duced disorder<strong>in</strong>g. After Zn diffusion (575C, 4 or 16h), cleaved wedges of the<br />

samples were <strong>in</strong>vestigated by means of secondary-ion mass spectrometry <strong>and</strong><br />

transmission electron microscopy. The results showed that completely <strong>and</strong> partially<br />

disordered regions were always beh<strong>in</strong>d the Zn diffusion front. A dependence of the<br />

effective Zn diffusivity <strong>and</strong> of the disorder<strong>in</strong>g rate of the structure upon the background<br />

Si dopant level was observed. The effective Zn diffusivity <strong>and</strong> the disorder<strong>in</strong>g rate<br />

significantly decreased with <strong>in</strong>creas<strong>in</strong>g background Si concentration. Before Zn diffusion,<br />

the photolum<strong>in</strong>escence spectra of Si-doped structures exhibited an <strong>in</strong>crease <strong>in</strong> <strong>in</strong>tensity of<br />

the Si donor column-<strong>III</strong> vacancy complex emission b<strong>and</strong> with <strong>in</strong>creas<strong>in</strong>g Si dopant level.<br />

This <strong>in</strong>dicated that the concentration of column-<strong>III</strong> vacancies <strong>in</strong> the structures <strong>in</strong>creased<br />

as the background Si concentration was <strong>in</strong>creased. After Zn diffusion, a large decrease <strong>in</strong><br />

<strong>in</strong>tensity of the column-<strong>III</strong> vacancy-related emission b<strong>and</strong> was observed <strong>in</strong> the<br />

photolum<strong>in</strong>escence spectra from Zn-diffused regions. A model that was based upon the<br />

so-called kick-out mechanism of Zn diffusion was proposed <strong>in</strong> order to expla<strong>in</strong> the effect<br />

of the background Si dop<strong>in</strong>g level upon the effective Zn diffusivity. The model showed<br />

that the effective Zn diffusivity was controlled by the concentration of column-<strong>III</strong><br />

<strong>in</strong>terstitials beh<strong>in</strong>d the Zn diffusion front, <strong>and</strong> by the donor concentration <strong>in</strong> the sample.<br />

369


Zn <strong>GaAs</strong> Muons<br />

Column-<strong>III</strong> <strong>in</strong>terstitials were generated dur<strong>in</strong>g the <strong>in</strong>corporation of Zn <strong>in</strong>to the crystal<br />

lattice. The supersaturation of these <strong>in</strong>terstitials beh<strong>in</strong>d the Zn diffusion front was<br />

responsible for the enhancement of Al-Ga <strong>in</strong>terdiffusion. Because column-<strong>III</strong> <strong>in</strong>terstitials<br />

<strong>and</strong> column-<strong>III</strong> vacancies could mutually annihilate, the concentration of column-<strong>III</strong><br />

<strong>in</strong>terstitial <strong>and</strong> column-<strong>III</strong> vacancies <strong>in</strong> Zn-diffused regions decreased with <strong>in</strong>creas<strong>in</strong>g Si<br />

dop<strong>in</strong>g level; thus lead<strong>in</strong>g to a retardation of Zn diffusion <strong>in</strong>to the structure. A decrease <strong>in</strong><br />

effective Zn diffusivity, caused by an <strong>in</strong>crease <strong>in</strong> the donor concentration of the samples,<br />

was also demonstrated. The results revealed effects of the Fermi level <strong>and</strong> of <strong>in</strong>teractions<br />

between po<strong>in</strong>t defects dur<strong>in</strong>g the Zn diffusion-<strong>in</strong>duced disorder<strong>in</strong>g of <strong>GaAs</strong>/Al<strong>GaAs</strong><br />

multi-layered structures.<br />

N.H.Ky, J.D.Ganière, F.K.Re<strong>in</strong>hart, B.Blanchard: Journal of Applied Physics, 1996,<br />

79[8], 4009-16<br />

[446-134/135-136]<br />

<strong>GaAs</strong>/GaAlAs: Zn <strong>Diffusion</strong><br />

Secondary-ion mass spectrometry <strong>and</strong> photolum<strong>in</strong>escence methods were used to study Zn<br />

diffusion <strong>in</strong>to <strong>GaAs</strong>/Ga 0.8 Al 0.2 As multi quantum-well structures which were uniformly<br />

doped with Si to concentrations of between 10 17 <strong>and</strong> 10 19 /cm 3 . The secondary-ion mass<br />

spectrometry profiles which were measured after Zn diffusion at 575C revealed a large<br />

effect of the background Si dop<strong>in</strong>g level upon the Zn diffusion process <strong>and</strong> upon Zn<br />

diffusion-<strong>in</strong>duced disorder<strong>in</strong>g of the multi quantum-well structures. It was found that the<br />

Zn diffusivity <strong>and</strong> the disorder<strong>in</strong>g rate significantly decreased with <strong>in</strong>creas<strong>in</strong>g Si<br />

background concentration. Before Zn diffusion, the photolum<strong>in</strong>escence spectra of the<br />

multi quantum-well samples revealed an <strong>in</strong>creased <strong>in</strong>tensity of the Si donor-V <strong>III</strong> complex<br />

emission b<strong>and</strong> with <strong>in</strong>creased Si dop<strong>in</strong>g level. This <strong>in</strong>dicated that the V <strong>III</strong> concentration <strong>in</strong><br />

the multi quantum-well structures <strong>in</strong>creased as the Si background concentration <strong>in</strong>creased.<br />

After Zn diffusion, a large decrease <strong>in</strong> the <strong>in</strong>tensity of the V <strong>III</strong> -related emission b<strong>and</strong> was<br />

detected <strong>in</strong> the photolum<strong>in</strong>escence spectra which were obta<strong>in</strong>ed from Zn-diffused regions.<br />

The results were expla<strong>in</strong>ed <strong>in</strong> terms of the kick-out mechanism of Zn diffusion. Dur<strong>in</strong>g<br />

the <strong>in</strong>corporation of Zn <strong>in</strong>to the crystal lattice, column-<strong>III</strong> <strong>in</strong>terstitials were generated. The<br />

supersaturation of I <strong>III</strong> beh<strong>in</strong>d the Zn diffusion front resulted <strong>in</strong> an enhancement of Al-Ga<br />

<strong>in</strong>terdiffusion <strong>in</strong> the Zn-diffused region. S<strong>in</strong>ce I <strong>III</strong> <strong>and</strong> V <strong>III</strong> could mutually annihilate, a<br />

reduction <strong>in</strong> I <strong>III</strong> concentration occurred <strong>in</strong> the Zn-diffused region of Si-doped samples<br />

which conta<strong>in</strong>ed a high V <strong>III</strong> concentration. As a result, Zn diffusion <strong>and</strong> disorder<strong>in</strong>g of the<br />

multi quantum-well structures were retarded with <strong>in</strong>creas<strong>in</strong>g Si-dop<strong>in</strong>g level.<br />

N.H.Ky: Materials Science Forum, 1995, 196-201, 1643-8<br />

Muons<br />

[446-127/128-139]<br />

<strong>GaAs</strong>: Muon <strong>Diffusion</strong><br />

Measurements were made of the dynamics of negatively charged muonium <strong>in</strong> heavily Sidoped<br />

material at temperatures rang<strong>in</strong>g from 295 to 1000K. The muonium began to<br />

370


Muons <strong>GaAs</strong> Muons<br />

diffuse, at temperatures above 500K, at a hopp<strong>in</strong>g rate which was described by an attempt<br />

frequency of 5.6 x 10 12 /s <strong>and</strong> an activation energy of 0.73eV. At temperatures above<br />

700K, relaxation from charge-state fluctuations was observed. The data implied that Mu -<br />

to Mu o conversion occurred, via the alternat<strong>in</strong>g capture of holes <strong>and</strong> electrons. This<br />

established that Mu was a deep recomb<strong>in</strong>ation center. Similar dynamics were expected<br />

for the isolated H- center <strong>in</strong> n-type material.<br />

K.H.Chow, B.Hitti, R.F.Kiefl, S.R.Dunsiger, R.L.Lichti, T.L.Estle: Physical Review<br />

Letters, 1996, 76[20], 3790-3<br />

[446-136/137-111]<br />

<strong>GaAs</strong>: Muon <strong>Diffusion</strong><br />

The results of <strong>in</strong>vestigations of transitions among the sites <strong>and</strong> charge states of muonium<br />

were summarized. Energy parameters were determ<strong>in</strong>ed for the full description of<br />

muonium dynamics <strong>and</strong> were found to correlate well with available data for H. A model<br />

was developed which accounted for all of the major features which were observed. Its<br />

validity for a wide range of dopant concentrations, us<strong>in</strong>g a s<strong>in</strong>gle set of parameters,<br />

reflected its predictive strength. The near equality of the energy parameters for muonium<br />

as compared with those which were available for H, strongly implied that the results for<br />

muonium dynamic behavior should be applicable to H, with very little change. The model<br />

could be applied to all tetrahedrally coord<strong>in</strong>ated semiconductors, with few modifications,<br />

<strong>and</strong> served as a basis for the underst<strong>and</strong><strong>in</strong>g of muonium dynamics <strong>in</strong> <strong>GaAs</strong> <strong>and</strong> <strong>other</strong><br />

materials. Differences <strong>in</strong> H properties could be understood by exam<strong>in</strong><strong>in</strong>g materialspecific<br />

deviations from the basic model.<br />

R.L.Lichti, C.Schwab, T.L.Estle: Materials Science Forum, 1995, 196-201, 831-6<br />

[446-127/128-119]<br />

<strong>GaAs</strong>: Muon <strong>Diffusion</strong><br />

It was noted that the presence of Si donors had a marked effect upon the charge state <strong>and</strong><br />

diffusion of muonium at the tetrahedral <strong>in</strong>terstitial site (Mu o T ), while it had a relatively<br />

weak effect upon bond-center muonium (Mu o BC ). In metallic Si-doped material, the<br />

highly mobile Mu o T center which was observed <strong>in</strong> non-metallic material was replaced by<br />

a charged species (probably Mu - ) which had a diffusion rate that was smaller than that of<br />

the Mu o T center. The Mu o BC center was metastable, <strong>and</strong> its transition to Mu o T or Mu -<br />

centers (depend<strong>in</strong>g upon dopant concentration) was observed at 50 to 150K. The results<br />

o<br />

also suggested that the Mu T state underwent fast sp<strong>in</strong>-exchange <strong>in</strong>teraction with<br />

conductive carriers before the transition: Mu o T → Mu - .<br />

R.Kadono, A.Matsushita, K.Nagam<strong>in</strong>e, K.Nishiyama, K.H.Chow, R.F.Kiefl,<br />

A.MacFarlane, D.Schumann, S.Fujii, S.Tanigawa: Physical Review B, 1994, 50[3], 1999-<br />

2002<br />

[446-115/116-118]<br />

371


Muons <strong>GaAs</strong> General<br />

<strong>GaAs</strong>: Muon <strong>Diffusion</strong><br />

By measur<strong>in</strong>g the muon sp<strong>in</strong> relaxation rate of muonium as it moved from site to site <strong>in</strong> a<br />

host lattice with nuclear sp<strong>in</strong>s, it was possible to derive an average correlation time, t, for<br />

fluctuations <strong>in</strong> the nuclear hyperf<strong>in</strong>e field which acted upon the unpaired electron.<br />

Provided that the motion was <strong>in</strong>coherent (diffusive), t was <strong>in</strong>versely proportional to the<br />

diffusion constant. Such measurements were reported here for isotropic muonium at<br />

temperatures rang<strong>in</strong>g from 20mK to 300K. Large differences were found <strong>in</strong> the muonium<br />

motion, <strong>and</strong> were most marked at low temperature; where l/t became temperature<strong>in</strong>dependent<br />

below about 10K <strong>in</strong> <strong>GaAs</strong>. A study was also made of the diffusive behavior<br />

of muonium <strong>in</strong> various <strong>GaAs</strong> samples <strong>in</strong> order to detect the possible effect of slight<br />

crystall<strong>in</strong>e imperfections.<br />

W.Schneider, R.F.Kiefl, E.J.Ansaldo, J.H.Brewer, K.Chow, S.F.J.Cox, S.A.Dodds,<br />

R.C.Duvarney, T.L.Estle, E.E.Haller, R.Kadono, S.R.Kreitzman, R.L.Lichti,<br />

C.Niedermayer, T.Pfiz, T.M.Riseman, C.Schwab: Materials Science Forum, 1992, 83-87,<br />

569-74<br />

[446-99/100-065]<br />

General<br />

<strong>GaAs</strong>: <strong>Diffusion</strong><br />

A review was presented of developments <strong>in</strong> the underst<strong>and</strong><strong>in</strong>g of self-diffusion <strong>and</strong><br />

impurity diffusion processes <strong>in</strong> this material; with particular emphasis be<strong>in</strong>g placed on the<br />

<strong>in</strong>clusion of recent Ga-isotope diffusion data. Specific diffusion mechanisms were<br />

suggested for C, P, Sb <strong>and</strong> S; which were all substitutionally dissolved on the As sublattice.<br />

U.Gösele, T.Y.Tan, M.Schultz, U.Egger, P.Werner, R.Scholz, O.Breitenste<strong>in</strong>: Defect <strong>and</strong><br />

<strong>Diffusion</strong> Forum, 1997, 143-147, 1079-94<br />

[446-143/147-1079]<br />

<strong>GaAs</strong>[l]: Electromigration<br />

A new experimental technique was proposed which permitted the measurement of the<br />

effective mobility of solute elements <strong>in</strong> <strong>III</strong>-V solutions. It was based upon the concept of<br />

an electro-epitaxial growth experiment <strong>in</strong> which the Peltier effect contribution to growth<br />

was elim<strong>in</strong>ated by adjust<strong>in</strong>g the current flow direction so that it was parallel to the<br />

substrate surface. Thus, only diffusion <strong>and</strong> electrotransport contributed to growth.<br />

Z.R.Zytkiewicz: Journal of Crystal Growth, 1987, 82[4], 647-51<br />

[446-55/56-168]<br />

<strong>GaAs</strong>: <strong>Diffusion</strong><br />

Dur<strong>in</strong>g the fabrication of refractory gate MESFET devices, sputter deposition of a WSi x<br />

gate <strong>and</strong> reactive ion etch<strong>in</strong>g of the gate pattern could lead to surface damage <strong>and</strong><br />

contam<strong>in</strong>ation. In order to study these effects, material with a shallow Si implant was<br />

372


General <strong>GaAs</strong> General<br />

subjected to reactive ion etch<strong>in</strong>g alone, or to both WSi x sputter deposition <strong>and</strong> reactive<br />

ion etch<strong>in</strong>g, before anneal<strong>in</strong>g. The surface damage, due to WSi x sputter deposition <strong>and</strong><br />

reactive ion etch<strong>in</strong>g at self-bias under 200V, was healed by capped (SiN x ) furnace<br />

anneal<strong>in</strong>g at 800C. It was found that sheet resistance <strong>and</strong> Hall mobility measurements<br />

could be correlated with the diffusion of compensat<strong>in</strong>g impurities <strong>in</strong>to the bulk.<br />

Secondary ion mass spectrometry profiles <strong>in</strong>dicated that the major contam<strong>in</strong>ants (Fe, Cr,<br />

Ni, Cu, V) were already present <strong>in</strong> the W targets <strong>and</strong> were thus present <strong>in</strong> the WSi x<br />

layers. These contam<strong>in</strong>ants were left on the surface of the <strong>GaAs</strong> after gate reactive ion<br />

etch<strong>in</strong>g <strong>and</strong> were driven <strong>in</strong>to the bulk dur<strong>in</strong>g capped anneal<strong>in</strong>g. An HCl etch was found to<br />

remove the contam<strong>in</strong>ants; thus result<strong>in</strong>g <strong>in</strong> lower sheet resistances for implanted <strong>and</strong><br />

processed <strong>GaAs</strong>. Refractory gate sub-micron MESFET devices which were fabricated by<br />

us<strong>in</strong>g an HCl etch after gate reactive ion etch<strong>in</strong>g exhibited a reduced access resistance.<br />

H.Baratte, A.J.Fleischman, G.J.Scilla, T.N.Jackson, H.J.Hovel, F.Cardone: Journal of the<br />

Electrochemical Society, 1991, 138[1], 219-22<br />

[446-78/79-010]<br />

<strong>GaAs</strong>: <strong>Diffusion</strong><br />

A long-term reliable spun-on source was developed for open-tube p + diffusion <strong>in</strong>to <strong>III</strong>/V<br />

compound semiconductors. The source consisted of an aqueous/alcoholic solution of<br />

zircon alcoholate, doped with z<strong>in</strong>c chloride. After sp<strong>in</strong>n<strong>in</strong>g <strong>and</strong> dry<strong>in</strong>g <strong>in</strong> air at 300C, a<br />

glassy film of ZrO 2 :ZnO on the semiconductor surface acted as a solid-state source for<br />

subsequent diffusion. This solution exhibited an excellent long-term durability.<br />

Furthermore, the thermal expansion coefficient of zirconia was well matched to that of<br />

most <strong>III</strong>/V compound semiconductors <strong>and</strong> yielded very stable films, even at high<br />

temperatures. This permitted essentially stress-free diffusion. <strong>Diffusion</strong> <strong>in</strong>to <strong>GaAs</strong> was<br />

carried out at temperatures rang<strong>in</strong>g from 600 <strong>and</strong> 700C. The hole concentrations <strong>and</strong><br />

diffusion coefficients which were obta<strong>in</strong>ed by us<strong>in</strong>g this source were rather close to those<br />

obta<strong>in</strong>ed by us<strong>in</strong>g <strong>other</strong> diffusion techniques. The simpler h<strong>and</strong>l<strong>in</strong>g <strong>and</strong> long-term<br />

durability of zirconia-based solutions offered significant advantages over <strong>other</strong> p + -<br />

diffusion techniques which were used <strong>in</strong> the preparation of <strong>III</strong>/V compound<br />

semiconductors.<br />

G.Franz, M.C.Amann: Journal of the Electrochemical Society, 1989, 136[8], 2410-3<br />

[446-70/71-105]<br />

<strong>GaAs</strong>: <strong>Diffusion</strong><br />

Deep-level transient Fourier spectroscopic, Hall-effect <strong>and</strong> DSL etch<strong>in</strong>g techniques were<br />

used to analyze <strong>in</strong>itially semi-<strong>in</strong>sulat<strong>in</strong>g bulk samples after anneal<strong>in</strong>g at temperatures of<br />

between 800 <strong>and</strong> 1100C under As pressures of between 0 <strong>and</strong> 3bar. It was found that the<br />

electrical resistivity, electronic mobility <strong>and</strong> EL2 concentration <strong>in</strong>creased with <strong>in</strong>creas<strong>in</strong>g<br />

As pressure at all temperatures. Initially n-type samples were converted to p-type at<br />

pressures below 0.5bar <strong>and</strong> temperatures above 1000C. It was noted that As precipitates<br />

disappeared at high temperatures, <strong>and</strong> reappeared dur<strong>in</strong>g low temperature anneal<strong>in</strong>g. In<br />

order to expla<strong>in</strong> the observations, it was proposed that long-range rapid As <strong>in</strong>terstitial<br />

transport occurred between the surface <strong>and</strong> the bulk, together with short-range Ga<br />

373


General <strong>GaAs</strong> Surface<br />

vacancy migration <strong>in</strong>volv<strong>in</strong>g dislocations as sources <strong>and</strong> s<strong>in</strong>ks. It was suggested that a<br />

discrepancy, with regard to very small reported tracer diffusion coefficients, could be<br />

resolved by assum<strong>in</strong>g that rapidly diffus<strong>in</strong>g marked <strong>in</strong>terstitials which entered the lattice<br />

at the crystal surface tended to exchange sites with unmarked lattice species. They then<br />

became immobile.<br />

M.Noack, K.W.Kehr, H.Wenzl: Journal of Crystal Growth, 1997, 178, 438-44<br />

[446-152-0374]<br />

Surface <strong>Diffusion</strong><br />

Al<br />

<strong>GaAs</strong>: Al Surface <strong>Diffusion</strong><br />

Dur<strong>in</strong>g the molecular beam epitaxial growth of <strong>GaAs</strong> on the vic<strong>in</strong>al (100) surface of<br />

<strong>GaAs</strong>, reflection high-energy electron diffraction was used to measure the transition<br />

temperature between 2-dimensional nucleation <strong>and</strong> pure step propagation which occurred<br />

when sub-monolayer amounts of Sn were present on the surface. In the case of samples<br />

which were misoriented by 0.5º with respect to the [011] or [01¯1] direction, the transition<br />

temperature decreased by approximately 100C after the deposition of 0.6 of a monolayer<br />

of Sn. The presence of Sn <strong>in</strong>creased the surface mobility of Al adatoms on (100) AlAs<br />

surfaces; as <strong>in</strong>dicated by the anneal<strong>in</strong>g behavior of the AlAs surface at 600C.<br />

G.S.Petrich, A.M.Dabiran, P.I.Cohen: Applied Physics Letters, 1992, 61[2], 162-4<br />

[446-93/94-001]<br />

<strong>GaAs</strong>: Al Surface <strong>Diffusion</strong><br />

Misoriented samples which were tilted by 1 or 2º away from the (001) plane <strong>and</strong> towards<br />

[110] or (111), or towards [¯110] or (111), were prepared. The critical temperature at<br />

which a transition <strong>in</strong> growth mode occurred was studied by us<strong>in</strong>g reflection high-energy<br />

electron diffraction methods dur<strong>in</strong>g the growth of AlAs on <strong>GaAs</strong> vic<strong>in</strong>al surfaces. The<br />

critical temperature of AlAs was higher than that of <strong>GaAs</strong>, thus <strong>in</strong>dicat<strong>in</strong>g that the stepflow<br />

growth of AlAs occurred at a higher temperature. By comb<strong>in</strong><strong>in</strong>g these data with a<br />

theory that took account of 2-dimensional nucleation <strong>and</strong> surface diffusion, the surface<br />

diffusion length of Al was deduced. It was found to be greater than that of Ga for both<br />

types of substrate surface. By tak<strong>in</strong>g account of chemical equilibrium at the step edge of<br />

each surface, it was predicted that the surface diffusion length of Al should be<br />

anisotropic.<br />

M.Tanaka, T.Suzuki, T.Nish<strong>in</strong>aga: Japanese Journal of Applied Physics, 1990, 29[5],<br />

L706-8<br />

[446-76/77-006]<br />

<strong>GaAs</strong>: Al Surface <strong>Diffusion</strong><br />

The adsorption <strong>and</strong> migration characteristics of Al atoms on (100) surfaces were<br />

<strong>in</strong>vestigated by us<strong>in</strong>g reflection high-energy electron diffraction <strong>and</strong> the alternate<br />

374


Surface <strong>GaAs</strong> Surface<br />

deposition of Al <strong>and</strong> As 4 onto the grow<strong>in</strong>g surface. All of the results were attributed to<br />

the existence of differ<strong>in</strong>g migration velocities for various atoms.<br />

Y.Horikoshi, H.Yamaguchi, M.Kawashima. Japanese Journal of Applied Physics, 1989,<br />

28[8], 1307-11<br />

[446-72/73-001]<br />

As<br />

<strong>GaAs</strong>: As Surface <strong>Diffusion</strong><br />

The <strong>in</strong>corporation diffusion length of surface-migrat<strong>in</strong>g adatoms on molecular beam<br />

epitaxial material was studied theoretically <strong>and</strong> experimentally. By us<strong>in</strong>g microprobe<br />

reflection high-energy electron diffraction <strong>and</strong> scann<strong>in</strong>g electron microscopic techniques,<br />

it was demonstrated that the <strong>in</strong>corporation diffusion length depended strongly upon the<br />

As partial pressure, <strong>and</strong> was of the order of 1µ. This behavior was expla<strong>in</strong>ed by assum<strong>in</strong>g<br />

that the Ga flux which entered the step exceeded the As flux. When the Ga flux which<br />

entered the step was lower than that of As, the <strong>in</strong>corporation diffusion length decreased to<br />

that of the step separation; <strong>and</strong> was then typically of the order of tens of nm. It was<br />

concluded that the balance of Ga <strong>and</strong> As fluxes which entered the step edge determ<strong>in</strong>ed<br />

the <strong>in</strong>corporation diffusion length of Ga.<br />

T.Nish<strong>in</strong>aga, X.Q.Shen: Applied Surface Science, 1994, 82-83, 141-8<br />

[446-123/124-161]<br />

Ga<br />

<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

Dur<strong>in</strong>g the molecular beam epitaxial growth of <strong>GaAs</strong> on the vic<strong>in</strong>al (100) surface of<br />

<strong>GaAs</strong>, reflection high-energy electron diffraction was used to measure the transition<br />

temperature between 2-dimensional nucleation <strong>and</strong> pure step propagation which occurred<br />

when sub-monolayer amounts of Sn were present on the surface. In the case of samples<br />

which were misoriented by 0.5º with respect to the [011] or [01¯1] direction, the transition<br />

temperature decreased by approximately 100C after the deposition of 0.6 of a monolayer<br />

of Sn. This <strong>in</strong>dicated that the Ga mobility had <strong>in</strong>creased.<br />

G.S.Petrich, A.M.Dabiran, P.I.Cohen: Applied Physics Letters, 1992, 61[2], 162-4<br />

[446-93/94-001]<br />

<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

The dependence of Ga adatom surface diffusion upon the As flux dur<strong>in</strong>g molecular beam<br />

epitaxial growth was <strong>in</strong>vestigated. Variations of the growth rate of <strong>GaAs</strong> layers, grown<br />

onto (001) surfaces adjacent to (111) surfaces, were measured by means of scann<strong>in</strong>g<br />

microprobe reflection high-energy electron diffraction. The surface diffusion length was<br />

375


Surface <strong>GaAs</strong> Surface<br />

deduced from the variations <strong>in</strong> the growth rate. It was found that the surface diffusion<br />

length of the Ga adatoms became larger under a lower As flux.<br />

M.Hata, A.Watanabe, T.Isu: Journal of Crystal Growth, 1991, 111[1-4], 83-7<br />

[446-91/92-007]<br />

<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

The microscopic details of Ga adatom diffusion upon an As-stabilized (001) surface were<br />

<strong>in</strong>vestigated by us<strong>in</strong>g an ab <strong>in</strong>itio pseudopotential method. The results showed that Ga<br />

adatoms diffused on the surface by pass<strong>in</strong>g through the miss<strong>in</strong>g As dimer rows. A<br />

comparison with scann<strong>in</strong>g tunnell<strong>in</strong>g microscopic experiments dur<strong>in</strong>g molecular beam<br />

epitaxial growth suggested that a low pressure of As <strong>in</strong>creased surface Ga adatom<br />

diffusion due to the creation of a cont<strong>in</strong>uous Ga adatom diffusion path. This conclusion<br />

was consistent with the observation that low-temperature growth was possible via<br />

migration-enhanced epitaxy <strong>in</strong> which As <strong>and</strong> Ga sources were supplied alternately.<br />

K.Shiraishi: Applied Physics Letters, 1992, 60[11], 1363-5<br />

[446-86/87-010]<br />

<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

Misoriented samples which were tilted by 1 or 2º away from the (001) plane <strong>and</strong> towards<br />

[110] or (111), or towards [¯110] or (111), were prepared. The critical temperature at<br />

which a transition <strong>in</strong> growth mode occurred was studied by us<strong>in</strong>g reflection high-energy<br />

electron diffraction methods dur<strong>in</strong>g the growth of AlAs on <strong>GaAs</strong> vic<strong>in</strong>al surfaces. The<br />

critical temperature of AlAs was higher than that of <strong>GaAs</strong>, thus <strong>in</strong>dicat<strong>in</strong>g that the stepflow<br />

growth of AlAs occurred at a higher temperature. The surface diffusion length of Ga<br />

was found to be shorter than that of Ga for both types of substrate surface.<br />

M.Tanaka, T.Suzuki, T.Nish<strong>in</strong>aga: Japanese Journal of Applied Physics, 1990, 29[5],<br />

L706-8<br />

[446-76/77-006]<br />

<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

An experimental study of RHEED <strong>in</strong>tensity oscillations was performed on (111)B<br />

substrates which were misoriented by 1 or 2º towards [110], [2¯1¯1], or [¯211] orientations.<br />

The behavior of the RHEED oscillations on (111)B was similar to that on (001).<br />

However, the temperature at which the RHEED <strong>in</strong>tensity oscillations began to appear on<br />

(111)B was lower than that on (001). The surface diffusion length of Ga on (111)B was<br />

evaluated by tak<strong>in</strong>g account of the supersaturation ratio of adatoms on the terrace.<br />

T.Shitara, E.Kondo, T.Nish<strong>in</strong>aga: Journal of Crystal Growth, 1990, 99, 530-4<br />

[446-76/77-008]<br />

<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

Surface diffusion dur<strong>in</strong>g molecular beam epitaxy was studied. Firstly, the mode transition<br />

between 2-dimensional nucleation <strong>and</strong> step flow dur<strong>in</strong>g molecular beam epitaxial growth<br />

on vic<strong>in</strong>al surfaces was studied theoretically <strong>and</strong> experimentally. The basis of the theory<br />

was to assume that the transition occurred when the surface supersaturation on the step<br />

376


Surface <strong>GaAs</strong> Surface<br />

terrace became identical to the critical supersaturation for 2-dimensional nucleation. This<br />

permitted the diffusion length of Ga to be calculated at the experimentally determ<strong>in</strong>ed<br />

critical temperature for the mode transition. It was found that the diffusion length<br />

<strong>in</strong>creased, as the temperature decreased, due to an <strong>in</strong>creased residence time. Also, the<br />

diffusion length on (111)B was longer than that on (001) when the same formation energy<br />

for 2-dimensional nuclei was assumed for both surfaces. The theory gave good agreement<br />

with the experimental data, <strong>and</strong> it was concluded that surface diffusion was one of the<br />

most important processes which controlled molecular beam epitaxial growth <strong>and</strong> impurity<br />

<strong>in</strong>corporation.<br />

T.Nish<strong>in</strong>aga, T.Shitara, K.Mochizuki, K.I.Cho: Journal of Crystal Growth, 1990, 99, 482-<br />

90<br />

[446-76/77-009]<br />

<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

The adsorption <strong>and</strong> migration characteristics of Ga atoms on (100) surfaces were<br />

<strong>in</strong>vestigated by us<strong>in</strong>g reflection high-energy electron diffraction <strong>and</strong> the alternate<br />

deposition of Ga <strong>and</strong> As 4 onto the grow<strong>in</strong>g surface. Excess Ga deposition onto the surface<br />

produced Ga clusters or droplets on the first Ga layer. These dissolved very quickly after<br />

As 4 deposition <strong>and</strong> formed flat <strong>GaAs</strong> layers when the number of Ga atoms was near to 2<br />

or 3 times the surface site number. All of the results were attributed to the existence of<br />

differ<strong>in</strong>g migration velocities for various atoms.<br />

Y.Horikoshi, H.Yamaguchi, M.Kawashima. Japanese Journal of Applied Physics, 1989,<br />

28[8], 1307-11<br />

[446-72/73-001]<br />

<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

The generation of extra facets on ridge-type triangles, with (001)-, (110)- <strong>and</strong> (201)-<br />

related equivalent slopes on <strong>GaAs</strong> (111) A substrates, <strong>and</strong> stripes runn<strong>in</strong>g <strong>in</strong> [¯110], [110]<br />

<strong>and</strong> [100] directions on (001) substrates, was <strong>in</strong>vestigated dur<strong>in</strong>g the molecular beam<br />

epitaxy of <strong>GaAs</strong>/Al<strong>GaAs</strong> multi-layers. By <strong>in</strong>vestigat<strong>in</strong>g local variations <strong>in</strong> the layer<br />

thickness <strong>in</strong> regions adjacent to extra (114)A, (110) <strong>and</strong> (¯1¯1¯1)B facets which were<br />

common to the (111)A <strong>and</strong> (001) patterned substrates, <strong>and</strong> extra facets which were related<br />

to the respective substrates <strong>and</strong> growth rates of the facets relative to the growth rate on<br />

the substrate plane, the orientation-dependent Ga surface diffusion lengths were<br />

determ<strong>in</strong>ed. They <strong>in</strong>creased <strong>in</strong> the order: (001), (¯1¯1¯1)B-related, (111)A-related <strong>and</strong> (110).<br />

Or, diagrammatically,<br />

(001) < (¯1¯1¯1)B < (159) < (110)<br />

(¯1¯1¯3)B (¯3¯3¯1)B (114)A<br />

(013)B (111)A<br />

(113)B<br />

T.Takebe, M.Fujii, T.Yamamoto, K.Fujita, T.Watanabe: Journal of Applied Physics,<br />

1997, 81[11], 7273-8<br />

[446-152-0377]<br />

377


Surface <strong>GaAs</strong> Surface<br />

<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

The migration potentials of Ga adatoms near to step edges on the c(4 x 4) surface were<br />

<strong>in</strong>vestigated by us<strong>in</strong>g an empirical <strong>in</strong>teratomic potential <strong>and</strong> an energy term that<br />

accounted for charge redistribution on the surface. The energy term, as a function of the<br />

number of electrons which rema<strong>in</strong>ed <strong>in</strong> the Ga dangl<strong>in</strong>g bonds, was deduced from firstpr<strong>in</strong>ciples<br />

calculations. The latter results implied that lattice sites along the A-type step<br />

edges were stable for Ga adatoms, whereas no preferential adsorption site was found near<br />

to B-type step edges. This was because the number of electrons which rema<strong>in</strong>ed <strong>in</strong> the Ga<br />

dangl<strong>in</strong>g bond was reduced by Ga adatoms that occupied lattice sites along A-type step<br />

edges, rather than be<strong>in</strong>g unchanged by those which occupied lattice sites near to B-type<br />

step edges.<br />

T.Ito, K.Shiraishi: Japanese Journal of Applied Physics, 1996, 35[2-8B], L1016-8<br />

[446-138/139-076]<br />

<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

The migration potentials of Ga adatoms near to k<strong>in</strong>k <strong>and</strong> step edges were qualitatively<br />

<strong>in</strong>vestigated by us<strong>in</strong>g empirical <strong>in</strong>ter-atomic potentials <strong>and</strong> an energy term. The latter<br />

term, as a function of the number of electrons that rema<strong>in</strong>ed <strong>in</strong> the Ga dangl<strong>in</strong>g bond, was<br />

deduced from first-pr<strong>in</strong>ciples pseudopotential calculations. The calculated results implied<br />

that the lattice sites <strong>in</strong> the miss<strong>in</strong>g dimer row were favorable for Ga adatoms on the<br />

<strong>GaAs</strong>(001)-(2 x 4)β2 surface. This was because the formation of Ga dimers reduced the<br />

number of electrons that rema<strong>in</strong>ed <strong>in</strong> Ga dangl<strong>in</strong>g bonds. Lattice sites <strong>in</strong> the miss<strong>in</strong>g<br />

dimer row, near to a k<strong>in</strong>k <strong>and</strong> a B-type step edge, were stable locations for a Ga adatom.<br />

On the <strong>other</strong> h<strong>and</strong>, no preferential adsorption site was found near to an A-type step edge.<br />

This was because a Ga adatom <strong>in</strong> the miss<strong>in</strong>g dimer row near to a k<strong>in</strong>k <strong>and</strong> a B-type step<br />

edge was slightly stretched by an As atom <strong>and</strong> As-dimer on the plane that was 1 layer<br />

below, rather than be<strong>in</strong>g strongly stretched by two As-dimers near to an A-type step edge.<br />

T.Ito, K.Shiraishi: Japanese Journal of Applied Physics, 1996, 35[2-8A], L949-52<br />

[446-138/139-077]<br />

<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

A valence force field was optimized <strong>in</strong> order to reproduce the phonon dispersion curves<br />

of crystall<strong>in</strong>e <strong>GaAs</strong> <strong>and</strong> derived <strong>in</strong>teraction energies of Ga adatoms on the (001) surface.<br />

Calculations of the diffusion constant of isolated Ga atoms on the <strong>GaAs</strong> surface were<br />

performed by means of molecular dynamics simulations. All of the bulk Ga <strong>and</strong> As<br />

atoms, <strong>and</strong> the adsorbed Ga atoms, were completely free to move <strong>and</strong> no normalization of<br />

the velocity was performed after the trajectory had begun. Averages were taken of the<br />

results of hundreds of such trajectories for each temperature. Surface diffusion constants<br />

were then obta<strong>in</strong>ed from the (001) <strong>in</strong>-plane components. It was found that the data could<br />

be described by:<br />

378


Surface <strong>GaAs</strong> Surface<br />

D(cm 2 /s) = 2.41 x 10 -5 exp[-0.0971(eV)/kT]<br />

A.Palma, E.Sempr<strong>in</strong>i, A.Talamo, N.Tomass<strong>in</strong>i: Journal of Crystal Growth, 1995, 150[1-<br />

4], 180-4<br />

[446-127/128-118]<br />

<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

Facets of (110)-type were formed, on mesa edges which def<strong>in</strong>ed (100)-(110) facet<br />

structures, by the molecular beam epitaxial growth of <strong>GaAs</strong> onto [001]-mesa stripes on<br />

(100) <strong>GaAs</strong> substrates. The surface diffusion length of Ga adatoms along the [010]<br />

direction on the mesa stripes was estimated for various growth conditions by means of <strong>in</strong><br />

situ scann<strong>in</strong>g microprobe reflection high-energy electron diffraction. By us<strong>in</strong>g these<br />

values, <strong>and</strong> the correspond<strong>in</strong>g growth rate on the (110) <strong>GaAs</strong> facets, the diffusion length<br />

on the (110) plane was deduced. It was found that the Ga diffusion length on the (110)<br />

plane was greater than that on the (100) <strong>and</strong> (111)B planes. The long diffusion length on<br />

the (110) plane was expla<strong>in</strong>ed <strong>in</strong> terms of the particular surface reconstruction on this<br />

plane.<br />

M.López, Y.Nomura: Journal of Crystal Growth, 1995, 150[1-4], 68-72<br />

[446-127/128-118]<br />

<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

Systematic measurements were made of the surface diffusion lengths of Ga adatoms<br />

dur<strong>in</strong>g the molecular beam epitaxy of this material <strong>in</strong> the presence of H or H 2 . The spatial<br />

variation <strong>in</strong> the growth rate on the (100) surface adjacent to the (111)A surface was<br />

deduced from the period of reflection high-energy electron diffraction <strong>in</strong>tensity<br />

oscillations. The surface diffusion length of Ga adatoms, as estimated from the spatial<br />

variation <strong>in</strong> the growth rate, <strong>in</strong>creased with <strong>in</strong>creas<strong>in</strong>g H or H 2 pressure. It also <strong>in</strong>creased<br />

as the substrate temperature was <strong>in</strong>creased at a given H or H 2 pressure. The diffusion<br />

length <strong>in</strong> the case of H was greater than that <strong>in</strong> the case of H 2 .<br />

Y.Morishita, Y.Nomura, S.Goto, Y.Katayama: Applied Physics Letters, 1995, 67[17],<br />

2500-2<br />

[446-125/126-121]<br />

<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

The <strong>in</strong>corporation diffusion length of surface-migrat<strong>in</strong>g adatoms on molecular beam<br />

epitaxial material was studied theoretically <strong>and</strong> experimentally. By us<strong>in</strong>g microprobe<br />

reflection high-energy electron diffraction <strong>and</strong> scann<strong>in</strong>g electron microscopic techniques,<br />

it was demonstrated that the <strong>in</strong>corporation diffusion length depended strongly upon the<br />

As partial pressure, <strong>and</strong> was of the order of 1µ. This behavior was expla<strong>in</strong>ed by assum<strong>in</strong>g<br />

that the Ga flux which entered the step exceeded the As flux. When the Ga flux which<br />

entered the step was lower than that of As, the <strong>in</strong>corporation diffusion length decreased to<br />

that of the step separation; <strong>and</strong> was then typically of the order of tens of nm. It was<br />

379


Surface <strong>GaAs</strong> Surface<br />

concluded that the balance of Ga <strong>and</strong> As fluxes which entered the step edge determ<strong>in</strong>ed<br />

the <strong>in</strong>corporation diffusion length of Ga.<br />

T.Nish<strong>in</strong>aga, X.Q.Shen: Applied Surface Science, 1994, 82-83, 141-8<br />

[446-123/124-161]<br />

<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

Two-dimensional nuclei of <strong>GaAs</strong> which were grown on a s<strong>in</strong>gular (001) surface by<br />

metalorganic chemical vapor deposition were observed by means of high-vacuum<br />

scann<strong>in</strong>g tunnell<strong>in</strong>g microscopy. The nuclei extended <strong>in</strong> the [110] direction, which was<br />

opposite to that of molecular beam epitaxial growth. The nucleus number density was<br />

obta<strong>in</strong>ed from scann<strong>in</strong>g tunnell<strong>in</strong>g microscopic images, <strong>and</strong> the relationship between the<br />

density <strong>and</strong> the surface diffusion coefficient of Ga species was estimated by simulat<strong>in</strong>g<br />

growth at the surface. The surface diffusion coefficient was estimated to be equal to about<br />

10 -7 cm 2 /s at 530C.<br />

M.Kasu, N.Kobayashi: Journal of Crystal Growth, 1994, 145[1-4], 120-5<br />

[446-119/120-191]<br />

<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

Quantitative features of Ga atom surface migration (migration length, migration time,<br />

hopp<strong>in</strong>g frequency) were obta<strong>in</strong>ed by the Monte Carlo simulation of molecular beam<br />

epitaxial growth of <strong>GaAs</strong> on flat <strong>and</strong> stepped (001) surfaces under various growth<br />

conditions. Changes <strong>in</strong> these parameters dur<strong>in</strong>g growth, <strong>and</strong> the relationship between<br />

migration, surface roughness <strong>and</strong> bulk defect concentration, were considered.<br />

N.V.Peskov: Surface Science, 1994, 306[1-2], 227-32<br />

[446-119/120-191]<br />

<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

Spatial variations <strong>in</strong> the growth rate on mesa-etched <strong>GaAs</strong> (¯1¯1¯1)B substrates dur<strong>in</strong>g the<br />

molecular beam epitaxy of <strong>GaAs</strong> were deduced from the period of reflection high-energy<br />

electron diffraction <strong>in</strong>tensity oscillations by us<strong>in</strong>g <strong>in</strong> situ scann<strong>in</strong>g microprobe methods.<br />

The surface diffusion length of Ga adatoms on the (¯1¯1¯1)B surface was deduced from the<br />

spatial variation <strong>in</strong> the growth rate. The surface diffusion length on the (¯1¯1¯1)B surface<br />

<strong>in</strong>creased as the substrate temperature was <strong>in</strong>creased or the As pressure was decreased. A<br />

typical value of the diffusion length was about 10µ, at a substrate temperature of 580C<br />

<strong>and</strong> an As pressure of 5.7 x 10 -4 Pa. This was an order of magnitude larger than that on the<br />

(100) surface, <strong>in</strong> the [011] direction. The activation energy of the surface diffusion length<br />

changed with surface reconstruction. Anisotropic diffusion, which had been reported for<br />

the (100) surface, was not observed on the (¯1¯1¯1)B surface.<br />

Y.Nomura, Y.Morishita, S.Goto, Y.Katayama, T.Isu: Applied Physics Letters, 1994,<br />

64[9], 1123-5<br />

[446-115/116-117]<br />

380


Surface <strong>GaAs</strong> Surface<br />

<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

The faceted surface morphologies of homo-epitaxial films which had been grown onto<br />

exactly (¯1¯1¯1)-oriented <strong>GaAs</strong> substrates <strong>in</strong> the √19 x √19 regime were studied with the aid<br />

of an atomic force microscope. The facets were composed of 3 vic<strong>in</strong>al surfaces which<br />

were tilted by about 2° towards the [2¯1¯1], [¯12¯1], <strong>and</strong> [¯1¯12] directions. The diffusion<br />

length was deduced from the surface morphologies, <strong>and</strong> was found to be equal to some<br />

hundreds of nm. It was comparable to the diffusion length on (100) surfaces which were<br />

grown under the same conditions. It was concluded that facet formation on <strong>GaAs</strong> (¯1¯1¯1)<br />

films was unlikely to be caused by a lower surface mobility.<br />

K.Yang, L.J.Schowalter, T.Thundat: Applied Physics Letters, 1994, 64[13], 1641-3<br />

[446-115/116-117]<br />

<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

The As pressure dependence of Ga adatom surface diffusion dur<strong>in</strong>g molecular beam<br />

epitaxy onto non-planar substrates was <strong>in</strong>vestigated. By us<strong>in</strong>g <strong>in</strong> situ scann<strong>in</strong>g<br />

microprobe reflection high-energy electron diffraction techniques, the distribution of the<br />

growth rate of <strong>GaAs</strong> on the (001) surface near to the edge of the (111)A or (111)B sidewall<br />

was measured under various As pressures. It was found that the surface diffusion<br />

length of Ga adatom <strong>in</strong>corporation on the (001) surface, as deduced from the growth rate<br />

distribution, was of the order of µ <strong>and</strong> exhibited a marked dependence upon the As<br />

pressure. A simple model which was based upon 1-dimensional surface diffusion was<br />

used to estimate the lifetime for Ga adatom <strong>in</strong>corporation on <strong>other</strong> surfaces.<br />

X.Q.Shen, D.Kishimoto, T.Nish<strong>in</strong>aga: Japanese Journal of Applied Physics, 1994, 33[1-<br />

1A], 11-7<br />

[446-113/114-011]<br />

<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

The As pressure dependence of the surface diffusion of Ga adatoms on molecular beam<br />

epitaxial (111) B-(001) mesa-etched substrates was <strong>in</strong>vestigated by us<strong>in</strong>g <strong>in</strong> situ scann<strong>in</strong>g<br />

microprobe reflection high-energy electron diffraction techniques. It was observed, for<br />

the first time, that the direction of Ga adatom migration from, or to, the (111)B side-wall<br />

changed; depend<strong>in</strong>g upon the As pressure. Moreover, the diffusion length of Ga adatoms<br />

on the (001) surface, <strong>in</strong> the [¯110] direction, was found to depend exponentially upon the<br />

As pressure. However, it was <strong>in</strong>dependent of the direction of lateral migration. The<br />

diffusion length of Ga adatoms on the (001) surface, <strong>in</strong> the [¯110] direction, varied from<br />

about 0.25 to 1.2µ at 600C, with<strong>in</strong> the present range of As pressures. It was suggested<br />

that the lifetime of Ga adatoms, before <strong>in</strong>corporation <strong>in</strong>to the crystal, depended strongly<br />

upon the As pressure.<br />

X.Q.Shen, T.Nish<strong>in</strong>aga: Japanese Journal of Applied Physics, 1993, 32[2-8B], L1117-9<br />

[446-109/110-029]<br />

381


Surface <strong>GaAs</strong> Surface<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

Thickness variations <strong>in</strong> quantum wells which had been grown on patterned substrates, by<br />

means of molecular beam epitaxy, were analyzed by us<strong>in</strong>g spatially <strong>and</strong> spectrally<br />

resolved low-temperature cathodolum<strong>in</strong>escence methods. In the case of lower <strong>and</strong> upper<br />

(100) facets which were jo<strong>in</strong>ed by an angled (311)A facet, relative <strong>in</strong>creases <strong>in</strong> the<br />

quantum well thickness, of up to about 6% <strong>and</strong> 20% respectively, were observed <strong>in</strong> the<br />

vic<strong>in</strong>ity of the facet <strong>in</strong>tersection. The Ga adatom migration length obeyed an exponential<br />

behavior, <strong>and</strong> ranged from 0.001 to 0.002mm on both the lower <strong>and</strong> upper (100) facets. It<br />

was <strong>in</strong>dependent of the quantum well thickness. The present migration length was some<br />

orders of magnitude greater than that which had previously been reported for Ga adatoms<br />

dur<strong>in</strong>g molecular beam epitaxial growth.<br />

S.Nilsson, E.Van Gieson, D.J.Arent, H.P.Meier, W.Walter, T.Forster: Applied Physics<br />

Letters, 1989, 55[10], 972-4<br />

[446-72/73-026]<br />

K<br />

<strong>GaAs</strong>: K Surface <strong>Diffusion</strong><br />

A self-consistent semi-empirical molecular orbital method was used to determ<strong>in</strong>e whether<br />

the adsorption properties of K atoms, <strong>and</strong> the formation of K adsorbate cha<strong>in</strong>s or clusters<br />

<strong>in</strong> the low-coverage regime, could be affected by the nature of the semiconductor surface<br />

(that is, perfect or stepped). It was found to be possible to determ<strong>in</strong>e the microscopic<br />

structures of monatomic <strong>and</strong> diatomic K molecules on perfect <strong>and</strong> stepped (110) <strong>GaAs</strong><br />

surfaces. The results for K adsorption on the perfect <strong>GaAs</strong>(110) surface were consistent<br />

with scann<strong>in</strong>g tunnell<strong>in</strong>g microscopic observations of Na on (110) <strong>GaAs</strong>; with the stable<br />

site for K be<strong>in</strong>g the bridge site which encompassed one Ga <strong>and</strong> two As surface atoms.<br />

The equilibrium geometry for diatomic K <strong>in</strong>volved the second K atom occupy<strong>in</strong>g the<br />

next-nearest neighbor bridge site; thus strongly support<strong>in</strong>g the formation of an open l<strong>in</strong>ear<br />

structure parallel to the zig-zag surface atomic cha<strong>in</strong>s. The calculated K-K distance <strong>in</strong> this<br />

equilibrium configuration was 0.802nm. This was similar to the Na-Na distance (0.8nm)<br />

which was deduced from scann<strong>in</strong>g tunnell<strong>in</strong>g microscopy experiments. The results for the<br />

stepped (110) <strong>GaAs</strong> surface suggested that a step was unlikely to assist the cluster<strong>in</strong>g of<br />

K atoms. However, the formation of the l<strong>in</strong>ear adsorbate cha<strong>in</strong> appeared to be <strong>in</strong>fluenced<br />

more by the orientation of the steps. On the perfect surface, the K adsorbates were bound<br />

more strongly at steps than at bridge sites.<br />

G.S.Khoo, C.K.Ong: Journal of Physics - Condensed Matter, 1993, 5[36], 6507-14<br />

[446-106/107-037]<br />

382


Surface <strong>GaAs</strong> Surface<br />

Ti<br />

<strong>GaAs</strong>: Ti Surface <strong>Diffusion</strong><br />

Scann<strong>in</strong>g tunnell<strong>in</strong>g microscopy was used to study the morphology of the Ti/(110)<strong>GaAs</strong><br />

<strong>in</strong>terface after atom deposition at 300K. The results <strong>in</strong>dicated that thermally activated<br />

surface diffusion was m<strong>in</strong>imal dur<strong>in</strong>g the growth process. This was because of the high<br />

activation energy which was imposed by chemical bond<strong>in</strong>g. On the <strong>other</strong> h<strong>and</strong>, surface<br />

hopp<strong>in</strong>g was observed because of the dynamics which were associated with the cool<strong>in</strong>g of<br />

imp<strong>in</strong>g<strong>in</strong>g atoms. It was noted that s<strong>in</strong>gle Ti atoms formed stable bonds with the<br />

substrate, but unique bond<strong>in</strong>g sites were not dist<strong>in</strong>guishable <strong>and</strong> the areas around these<br />

sites were not disrupted.<br />

Y.N.Yang, B.M.Trafas, Y.S.Luo, R.L.Siefert, J.H.Weaver: Physical Review B, 1991,<br />

44[11], 5720-5<br />

[446-84/85-017]<br />

- miscellaneous<br />

<strong>GaAs</strong>: Surface <strong>Diffusion</strong><br />

The bombardment of (110) samples with Ar + ions of normal <strong>in</strong>cidence, at temperatures of<br />

300 to 775K, created surface layer defects that usually spanned 1 or 2 unit cells (as<br />

revealed by scann<strong>in</strong>g tunnell<strong>in</strong>g microscopy). Vacancies which were produced <strong>in</strong> this way<br />

diffused via thermal activation to form s<strong>in</strong>gle-layer vacancy isl<strong>and</strong>s. The diffusion of divacancies<br />

favored [1¯10], <strong>and</strong> accommodation at isl<strong>and</strong>s produced roughly isotropic<br />

isl<strong>and</strong>s. Modell<strong>in</strong>g of this growth process revealed an overall Arrhenius behavior of the<br />

diffusion, with an activation energy of 1.3eV. Investigations of the surface morphology<br />

dur<strong>in</strong>g multi-layer erosion revealed deviations from layer-by-layer removal, with scal<strong>in</strong>g<br />

exponents of between 0.4 <strong>and</strong> 0.5 at temperatures of between 626 <strong>and</strong> 775K.<br />

R.J.Pechman, X.S.Wang, J.H.Weaver: Physical Review B, 1995, 51[16], 10929-36<br />

[446-121/122-053]<br />

<strong>GaAs</strong>: Surface <strong>Diffusion</strong><br />

It was found that surface migration was effectively enhanced by evaporat<strong>in</strong>g Ga or Al<br />

atoms onto a clean <strong>GaAs</strong> surface under an As-free atmosphere or low As pressure. This<br />

characteristic was exploited by alternately supply<strong>in</strong>g Ga <strong>and</strong>/or Al <strong>and</strong> As to the substrate<br />

surface <strong>in</strong> order to grow atomically-flat <strong>GaAs</strong>-Al<strong>GaAs</strong> hetero-<strong>in</strong>terfaces, <strong>and</strong> also to grow<br />

high-quality <strong>GaAs</strong> layers at very low temperatures. The migration characteristics of<br />

surface adatoms were <strong>in</strong>vestigated by us<strong>in</strong>g reflection high-energy electron diffraction<br />

measurements. It was found that differ<strong>in</strong>g growth mechanisms operated at high <strong>and</strong> low<br />

temperatures. Both mechanisms were expected to yield flat heterojunction <strong>in</strong>terfaces. By<br />

383


Surface <strong>GaAs</strong> Surface<br />

apply<strong>in</strong>g this method, <strong>GaAs</strong> layers could be grown at substrate temperatures of 200 <strong>and</strong><br />

300C, respectively.<br />

Y.Horikoshi, M.Kawashima, H.Yamaguchi: Japanese Journal of Applied. Physics, 1988,<br />

27[2], 169-79<br />

[446-60-001]<br />

<strong>GaAs</strong>: Surface <strong>Diffusion</strong><br />

A 90° double-reflection high-energy electron diffraction method was used to carry out a<br />

study of the morphology of vic<strong>in</strong>al (001) surfaces dur<strong>in</strong>g molecular beam epitaxy. The<br />

technique permitted the simultaneous record<strong>in</strong>g of reflection high-energy electron<br />

diffraction <strong>in</strong>tensifies <strong>in</strong> the [¯110] <strong>and</strong> [110] azimuths. Comparative measurements of<br />

surfaces with 2° misorientations towards (111)Ga (A surface) or (1¯11)As (B surface)<br />

showed that, regardless of the step-type <strong>and</strong> reconstruction anisotropy, record<strong>in</strong>gs of the<br />

specular beam <strong>in</strong>tensity <strong>in</strong> the azimuth perpendicular to the steps were dom<strong>in</strong>ated by<br />

changes <strong>in</strong> the staircase order whereas <strong>in</strong>tensity record<strong>in</strong>gs <strong>in</strong> the azimuth parallel to the<br />

steps revealed changes <strong>in</strong> the step-edge roughness. Measurements which were performed<br />

over a wide range of substrate temperatures clarified the competition between k<strong>in</strong>etic<br />

processes <strong>and</strong> thermodynamic equilibrium at the length scale which was accessible to<br />

reflection high-energy electron diffraction techniques. In the case of the A surface, the<br />

transition between 2-dimensional growth <strong>and</strong> step-flow growth not only occurred at a<br />

higher temperature than it did on the B surface, but the disappearance of <strong>in</strong>tensity<br />

oscillations also occurred at differ<strong>in</strong>g substrate temperatures for different azimuths. An<br />

approximately 20C higher disappearance temperature for the [¯110] azimuth was<br />

expla<strong>in</strong>ed <strong>in</strong> terms of a model that was based upon previous scann<strong>in</strong>g tunnell<strong>in</strong>g<br />

microscopy results which had revealed an <strong>in</strong>creas<strong>in</strong>g elongation of isl<strong>and</strong>s, <strong>in</strong> the [¯110]<br />

direction, with <strong>in</strong>creas<strong>in</strong>g substrate temperature. The B surface was more isotropic, <strong>and</strong><br />

therefore no difference <strong>in</strong> the transition temperature for the 2 azimuths could be detected.<br />

Dur<strong>in</strong>g growth <strong>in</strong> the transition range between 2-dimensional <strong>and</strong> step-flow growth,<br />

<strong>in</strong>creased terrace-width fluctuations were observed on the B surface whereas the A<br />

surface became more uniformly stepped. It was concluded that, <strong>in</strong> the k<strong>in</strong>etically<br />

controlled regime, the anisotropic barrier height for downward diffusion of adatoms over<br />

step edges played an important role <strong>in</strong> the evolution of surface morphology. At high<br />

temperatures, the barrier height permitted downward jumps of the adatoms over B-type<br />

steps but not over A-type steps. Under conditions that were close to thermodynamic<br />

equilibrium, k<strong>in</strong>etic smooth<strong>in</strong>g was observed on the A surface as well as on the B surface.<br />

This <strong>in</strong>dicated that an<strong>other</strong> mechanism became operative upon a change <strong>in</strong> the energetics<br />

due to order<strong>in</strong>g of the steps <strong>and</strong> disorder<strong>in</strong>g of the reconstruction on the terraces. This<br />

surface was metastable <strong>and</strong> rapidly recovered (with<strong>in</strong> less than 1s) to give the equilibrium<br />

bunched surface after <strong>in</strong>terruptions <strong>in</strong> growth at substrate temperatures above 580C.<br />

H.Nörenberg, L.Däweritz, P.Schützendübe, K.Ploog: Journal of Applied Physics, 1997,<br />

81[6], 2611-20<br />

[446-148/149-175]<br />

384


Surface <strong>GaAs</strong> Surface<br />

<strong>GaAs</strong>: Surface <strong>Diffusion</strong><br />

An <strong>in</strong>vestigation was made of surface k<strong>in</strong>etics, dur<strong>in</strong>g metalorganic vapor-phase epitaxial<br />

growth, by means of high-vacuum scann<strong>in</strong>g tunnell<strong>in</strong>g microscopic observations of 2-<br />

dimensional nuclei <strong>and</strong> denuded zones. Monte Carlo simulations were carried out which<br />

were based upon the solid-on-solid model. Two-dimensional nucleus densities were used<br />

to deduce that the surface diffusion coefficient of <strong>GaAs</strong> was equal to 2 x 10 -6 cm 2 /s at<br />

530C. The activation energy for migration was estimated to be 0.62eV. The 2-<br />

dimensional nucleus size <strong>in</strong> the [110] direction was about twice that <strong>in</strong> the [¯110]<br />

direction. This anisotropy was attributed ma<strong>in</strong>ly to a difference <strong>in</strong> the lateral stick<strong>in</strong>g<br />

probabilities between steps along [¯110] <strong>and</strong> those along [110]. The ratio of the stick<strong>in</strong>g<br />

probabilities was estimated to be greater than 3:1. The denuded zone widths on the upper<br />

terraces were some 2 times wider than those on the lower terraces. This suggested that the<br />

stick<strong>in</strong>g probability at descend<strong>in</strong>g steps was 10 to 300 times larger than the probability at<br />

ascend<strong>in</strong>g steps.<br />

M.Kasu, N.Kobayashi: Journal of Crystal Growth, 1997, 170, 246-50<br />

[446-141/142-093]<br />

<strong>GaAs</strong>: Surface <strong>Diffusion</strong><br />

The surface diffusion of group-<strong>III</strong> atom <strong>in</strong>corporation dur<strong>in</strong>g molecular beam epitaxial<br />

growth was considered. Firstly, the diffusion length for <strong>in</strong>corporation on the (001) top<br />

surface, with (111)A or (411)A side surfaces on V grooves, was studied. It was shown<br />

that the diffusion length took the same value for both cases <strong>and</strong> was <strong>in</strong>versely<br />

proportional to the As pressure. However, the diffusion length of Ga on (111)B exhibited<br />

an <strong>in</strong>verse parabolic dependence of the As pressure. It was suggested that, on the (001)<br />

surface, two As 4 molecules met to furnish active As atoms for growth. On the <strong>other</strong> h<strong>and</strong>,<br />

the behavior of the As 4 molecule on the (111)B surface rema<strong>in</strong>ed unclear. The ratio of the<br />

surface diffusion coefficients on (111)B <strong>and</strong> (001) was calculated. It was found that the<br />

ratio took a value of about 140. Us<strong>in</strong>g this ratio, the <strong>in</strong>corporation lifetimes on (111)B<br />

<strong>and</strong> (001) surfaces were calculated as functions of the As pressure. It was found that the<br />

curves of <strong>in</strong>corporation lifetime <strong>in</strong>tersected at the As pressure where flow <strong>in</strong>version<br />

occurred.<br />

T.Nish<strong>in</strong>aga, X.Q.Shen, D.Kishimoto: Journal of Crystal Growth, 1996, 163[1], 60-6<br />

[446-138/139-078]<br />

<strong>GaAs</strong>: Surface <strong>Diffusion</strong><br />

The migration of anion <strong>and</strong> cation vacancies on the (110) surface was studied by means<br />

of scann<strong>in</strong>g tunnell<strong>in</strong>g microscopy. Marked asymmetries were found <strong>in</strong> the direction <strong>and</strong><br />

bias-polarity dependence of the migration probability. This <strong>in</strong>dicated the importance of<br />

the bond<strong>in</strong>g topology at the surface. The asymmetry showed that vacancy motion was<br />

driven by the recomb<strong>in</strong>ation of carriers, <strong>in</strong>jected by the scann<strong>in</strong>g tunnell<strong>in</strong>g microscope<br />

tip, with carriers from the bulk. The impact-parameter dependence of the reaction cross-<br />

385


Surface <strong>GaAs</strong> Surface<br />

section showed that this <strong>in</strong>jection occurred via resonant tunnell<strong>in</strong>g <strong>in</strong>to dangl<strong>in</strong>g-bond<br />

defect states.<br />

G.Lengel, J.Harper, M.Weimer: Physical Review Letters, 1996, 76[25], 4725-8<br />

[446-136/137-111]<br />

<strong>GaAs</strong>: Surface <strong>Diffusion</strong><br />

Reconstruction dur<strong>in</strong>g chemical beam etch<strong>in</strong>g with AsCl 3 was studied. The detection of<br />

reflection high-energy electron diffraction <strong>in</strong>tensity oscillations <strong>in</strong>dicated the occurrence<br />

of a planar etch<strong>in</strong>g mode <strong>in</strong> the <strong>in</strong>itial stages. Its change to 3-dimensional etch<strong>in</strong>g could<br />

be understood <strong>in</strong> terms of suppressed cation diffusion dur<strong>in</strong>g etch<strong>in</strong>g. It was concluded<br />

that a suitable choice of etch<strong>in</strong>g parameters, which would enhance cation diffusion,<br />

would lead to a smooth etch<strong>in</strong>g morphology. The effectiveness of etch clean<strong>in</strong>g depended<br />

upon the planarity of the surface dur<strong>in</strong>g etch<strong>in</strong>g, <strong>and</strong> upon the reactivity of the<br />

contam<strong>in</strong>ants with the etch<strong>in</strong>g gas. This was illustrated by etch<strong>in</strong>g Be <strong>and</strong> Si δ-doped<br />

structures <strong>in</strong> <strong>GaAs</strong>.<br />

T.H.Chiu, W.T.Tsang, M.D.Williams, C.A.C.Mendonça, K.Dreyer, F.G.Storz: Journal of<br />

Crystal Growth, 1995, 150[1-4], 546-50<br />

[446-127/128-120]<br />

<strong>GaAs</strong>: Surface <strong>Diffusion</strong><br />

The atomic structures of Ga <strong>and</strong> As atoms on (110) planes were studied by us<strong>in</strong>g a firstpr<strong>in</strong>ciples<br />

pseudopotential method. It was found that both Ga <strong>and</strong> As atoms resided <strong>in</strong> the<br />

center of a triangle that consisted of a surface Ga atom <strong>and</strong> 2 surface As atoms <strong>in</strong> the<br />

s<strong>in</strong>gle-atom chemisorbed state. The adsorption energies for Ga <strong>and</strong> As atoms were 3.1<br />

<strong>and</strong> 3.5eV, respectively. The energy barrier heights for Ga <strong>and</strong> As atoms which migrated<br />

along the path through the <strong>in</strong>terstitial channel were found to be 0.6 <strong>and</strong> 1.0eV,<br />

respectively. Simulation of the deposition of 2 atoms revealed that pair formation was<br />

stable with respect to separate s<strong>in</strong>gle-atom chemisorption.<br />

J.Y.Yi, J.Y.Koo, S.Lee, J.S.Ha, E.Lee: Physical Review B, 1995, 52[15], 10733-6<br />

[446-125/126-122]<br />

<strong>GaAs</strong>: Surface <strong>Diffusion</strong><br />

A 1/6 monolayer of <strong>GaAs</strong> was grown onto a very flat <strong>GaAs</strong> surface by us<strong>in</strong>g metalorganic<br />

vapor-phase epitaxial techniques, <strong>and</strong> 2-dimensional nuclei were studied by us<strong>in</strong>g highvacuum<br />

scann<strong>in</strong>g tunnell<strong>in</strong>g microscopy. On the basis of the 2-dimensional nuclei densities,<br />

the surface diffusion coefficient at 530C was estimated to be equal to 2 x 10 -6 cm 2 /s. It was<br />

found that, dur<strong>in</strong>g growth, the bunched-step (multi-step) separation saturated <strong>and</strong> was<br />

<strong>in</strong>dependent of the substrate misorientation angle. The results could be expla<strong>in</strong>ed <strong>in</strong> terms of<br />

a mechanism that took account of 2-dimensional nucleus formation on the wider terraces,<br />

<strong>and</strong> their coalescence on ascend<strong>in</strong>g steps. A step-bunch<strong>in</strong>g simulation which was based<br />

upon this model revealed that the saturated multi-step separation was proportional to the 2-<br />

dimensional nucleus separation (that is, to the reciprocal of the square root of the density).<br />

M.Kasu, N.Kobayashi: Journal of Applied Physics, 1995, 78[5], 3026-35<br />

[446-123/124-162]<br />

386


Surface <strong>GaAs</strong> Surface<br />

<strong>GaAs</strong>: Surface <strong>Diffusion</strong><br />

The mechanisms of molecular beam epitaxy were <strong>in</strong>vestigated by grow<strong>in</strong>g <strong>and</strong> analyz<strong>in</strong>g<br />

the shapes of facet structures which consisted of an (001) top surface <strong>and</strong> two (111)B side<br />

surfaces. It was found that all of the Ga flux on the 3 facet planes was <strong>in</strong>corporated <strong>in</strong>to<br />

the film, but the growth rates on (111)B <strong>and</strong> (001) depended strongly upon the As flux<br />

<strong>and</strong> were governed ma<strong>in</strong>ly by the diffusion of Ga adatoms between the 2 planes. By<br />

analyz<strong>in</strong>g the shape of the facet, the diffusion length of Ga on a (001) surface was<br />

estimated to be about 1µ at 580C. On (111)B, the diffusion length of Ga was found to be<br />

equal to several µ. The reflectivity of diffus<strong>in</strong>g Ga atoms was found to be far less than<br />

unity for the (001)/(111)B boundary, <strong>and</strong> was almost equal to unity at facet boundaries<br />

where the (111)B side surfaces were bounded by the (1¯10) side walls.<br />

S.Koshiba, Y.Nakamura, M.Tsuchiya, H.Noge, H.Kano, Y.Nagamune, T.Noda,<br />

H.Sakaki: Journal of Applied Physics, 1994, 76[7], 4138-44<br />

[446-117/118-159]<br />

<strong>GaAs</strong>: Surface <strong>Diffusion</strong><br />

It was shown that Te <strong>and</strong> Pb which segregated at the surface dur<strong>in</strong>g epitaxial growth<br />

decreased <strong>and</strong> <strong>in</strong>creased, respectively, the surface diffusion length. This <strong>in</strong>dicated that,<br />

under the generic term of surfactant, there were 2 types of surface-segregat<strong>in</strong>g species<br />

which had opposite effects upon surface diffusion. It was suggested that the key<br />

parameter which governed the surfactant-<strong>in</strong>duced modification of epitaxial growth<br />

k<strong>in</strong>etics was the reactivity of a given pair of surfactant <strong>and</strong> grow<strong>in</strong>g materials. Accord<strong>in</strong>g<br />

to this theory, it was predicted that surfactants which occupied <strong>in</strong>terstitial surface sites<br />

(non-reactive surfactants) <strong>in</strong>creased the surface diffusion length, whereas surfactants<br />

which were <strong>in</strong> substitutional sites (reactive surfactants) decreased it.<br />

J.Massies, N.Gr<strong>and</strong>jean: Physical Review B, 1993, 48[11], 8502-5<br />

[446-106/107-039]<br />

<strong>GaAs</strong>: Surface <strong>Diffusion</strong><br />

The growth rates of layers which were grown on a mesa-etched (001) surface were<br />

measured by us<strong>in</strong>g <strong>in</strong> situ scann<strong>in</strong>g microprobe reflection high-energy electron diffraction<br />

methods. The diffusion lengths of the surface adatoms of column-<strong>III</strong> elements were<br />

deduced from the gradient of the variation of the growth rate. The diffusion lengths were<br />

of the order of one micron for every source/material comb<strong>in</strong>ation. When metalorganic<br />

materials were used as a Ga source, it was found that the diffusion length was larger than<br />

that of Ga atoms from a metallic Ga source. Because the substrate temperatures which<br />

were used <strong>in</strong> the present experiments were high enough to decompose trimethylgallium<br />

<strong>and</strong> triethylgallium on the surface, Ga adatoms were considered to be responsible for the<br />

surface diffusion. It was concluded that derivatives of the metalorganic molecules, such<br />

as methyl radicals, affected the diffusion of Ga adatoms.<br />

T.Isu, M.Hata, Y.Morishita, Y.Nomura, S.Goto, Y.Katayama: Journal of Crystal Growth,<br />

1992, 120[1-4], 45-9<br />

[446-106/107-039]<br />

387


Interdiffusion <strong>GaAs</strong> Interdiffusion<br />

Interdiffusion<br />

<strong>GaAs</strong>: Interdiffusion<br />

<strong>Diffusion</strong> <strong>and</strong> <strong>in</strong>terdiffusion <strong>in</strong> <strong>GaAs</strong> was shown to be consistent with a charged po<strong>in</strong>t<br />

defect model. Charged Ga vacancies, V Ga 3- , <strong>and</strong> <strong>in</strong>terstitials, I Ga 2+ , appeared to control the<br />

diffusion of group-II, group-<strong>III</strong> <strong>and</strong> (probably) group-V elements. After adjust<strong>in</strong>g for<br />

carrier concentration <strong>and</strong> As pressure, these elements were found to have an almost<br />

identical <strong>in</strong>tr<strong>in</strong>sic diffusivity <strong>and</strong> activation energy over a wide range of temperatures. A<br />

natural consequence, of Ga diffusion via negative or positive po<strong>in</strong>t defects, was that<br />

enhanced group-<strong>III</strong> <strong>in</strong>terdiffusion was expected under either n-type or p-type dop<strong>in</strong>g.<br />

Anomalous enhancements of group-II dopant diffusivity were related to the<br />

supersaturation of Ga <strong>in</strong>terstitials.<br />

R.M.Cohen: Journal of Applied Physics, 1990, 67[12], 7268-73<br />

[446-78/79-011]<br />

<strong>GaAs</strong>: Interdiffusion<br />

The <strong>in</strong>terdiffusion between silicides (W, Ta, Mo, Ti) <strong>and</strong> <strong>GaAs</strong>, at temperatures of 825 to<br />

975C, was studied by us<strong>in</strong>g Rutherford back-scatter<strong>in</strong>g spectrometry, particle-<strong>in</strong>duced X-<br />

ray emission, <strong>and</strong> X-ray diffraction techniques. The specimens were furnace-annealed at<br />

825 or 875C, or rapid thermally annealed at 975C. Almost no <strong>in</strong>terdiffusion was observed<br />

<strong>in</strong> the cases of WSi 2 , TaSi 2 <strong>and</strong> TiSi 2 at 825C. It was not observed at all at 975C. In the<br />

case of MoSi 2 , there was marked atomic migration at temperatures as low as 825C. It was<br />

concluded that rapid thermal anneal<strong>in</strong>g was more beneficial from the po<strong>in</strong>t of view of<br />

<strong>in</strong>terface stability.<br />

J.Osvald, R.S<strong>and</strong>rik: Th<strong>in</strong> Solid Films, 1989, 169[2], 223-8<br />

[446-64/65-163]<br />

<strong>GaAs</strong>/Al/<strong>GaAs</strong>: Interdiffusion<br />

It was shown that the presence of low-temperature-grown <strong>GaAs</strong>, <strong>in</strong> <strong>GaAs</strong>/AlAs on Sidoped<br />

<strong>GaAs</strong> heterostructures, <strong>in</strong>creased Al/Ga <strong>in</strong>terdiffusion at the heterostructure<br />

<strong>in</strong>terfaces. The <strong>in</strong>terdiffusivity enhancement was attributed to the presence of Ga<br />

vacancies, V Ga , <strong>in</strong> the As-rich low-temperature <strong>GaAs</strong>, which diffused from a<br />

supersaturation of V Ga which was frozen <strong>in</strong> dur<strong>in</strong>g growth. Chemical mapp<strong>in</strong>g, which<br />

dist<strong>in</strong>guished between the AlAs <strong>and</strong> <strong>GaAs</strong> lattices at the atomic scale, was used to<br />

measure the Al concentration gradient <strong>in</strong> adjacent Si-doped <strong>GaAs</strong> layers.<br />

C.Kisielowski, A.R.Calawa, Z.Liliental-Weber: Journal of Applied Physics, 1996, 80[1],<br />

156-60<br />

[446-136/137-115]<br />

<strong>GaAs</strong>/AlAs: Interdiffusion<br />

The results of ab <strong>in</strong>itio molecular dynamics simulations of impurity-<strong>in</strong>duced disorder<strong>in</strong>g<br />

<strong>in</strong> superlattices were presented. Two typical impurities, the Zn acceptor <strong>and</strong> the Si donor,<br />

388


Interdiffusion <strong>GaAs</strong> Interdiffusion<br />

were studied. It was found that Zn-<strong>in</strong>duced <strong>in</strong>terdiffusion was due to the formation of<br />

non-equilibrium group-<strong>III</strong> <strong>in</strong>terstitials dur<strong>in</strong>g Zn <strong>in</strong>-diffusion. The <strong>in</strong>terstitials, which<br />

were bound to Zn acceptors via Coulomb forces, disordered the superlattice via kick-out<br />

processes on the group-<strong>III</strong> sub-lattice. On the <strong>other</strong> h<strong>and</strong>, Si-<strong>in</strong>duced <strong>in</strong>terdiffusion<br />

occurred via Si Ga -V Ga pairs. Their motion, via second-nearest neighbor jumps, disordered<br />

the group-<strong>III</strong> sub-lattice. The calculated activation energies for these processes were <strong>in</strong><br />

good agreement with experiment.<br />

J.Bernholc, B.Chen, Q.Zhang, C.Wang, B.Yakobson: Materials Science Forum, 1994,<br />

143-147, 593-8<br />

[446-113/114-028]<br />

<strong>GaAs</strong>/AlAs: Interdiffusion<br />

Photolum<strong>in</strong>escence measurements were used to <strong>in</strong>vestigate C impurity effects upon the<br />

<strong>in</strong>termix<strong>in</strong>g behavior of multiple quantum wells which had been grown by means of<br />

molecular beam epitaxy. The wells were furnace annealed, with a C source. The<br />

photolum<strong>in</strong>escence spectra revealed that the degree of <strong>in</strong>termix<strong>in</strong>g of Al <strong>and</strong> Ga, which<br />

was <strong>in</strong>duced by thermal anneal<strong>in</strong>g, <strong>in</strong>creased with depth. This behavior did not agree with<br />

<strong>in</strong>termix<strong>in</strong>g mechanisms which considered vacancy <strong>in</strong>jection of the surface. The nonuniformity<br />

of <strong>in</strong>termix<strong>in</strong>g as a function of depth was attributed to the effect of C<br />

impurities which were <strong>in</strong>jected dur<strong>in</strong>g heat treatment.<br />

Y.T.Oh, S.K.Kim, Y.H.Kim, T.W.Kang, C.Y.Hong, T.W.Kim: Journal of Applied<br />

Physics, 1995, 77[6], 2415-8<br />

[446-121/122-062]<br />

<strong>GaAs</strong>/AlAs: Interdiffusion<br />

The relative stability of ideal-geometry versus <strong>in</strong>termixed buried semiconductor <strong>in</strong>terfaces<br />

was studied by us<strong>in</strong>g first-pr<strong>in</strong>ciples density functional methods. It was found that,<br />

although <strong>in</strong>termix<strong>in</strong>g of an ideal lattice-matched <strong>GaAs</strong>/AlAs(001) <strong>in</strong>terface required<br />

energy, <strong>in</strong>termix<strong>in</strong>g was an energy-lower<strong>in</strong>g process at the coherent stra<strong>in</strong>ed-layer (001)<br />

<strong>in</strong>terfaces of lattice-mismatched materials. Intermix<strong>in</strong>g of an ideal stra<strong>in</strong>ed-layer (001)<br />

<strong>in</strong>terface lowered the energy ma<strong>in</strong>ly because it partially relieved the stra<strong>in</strong> <strong>in</strong> a local<br />

region near to the <strong>in</strong>terface. It was predicted that lattice-mismatched <strong>in</strong>terfaces would<br />

have a greater degree of atomic-scale micro-roughness than would the analogous<br />

<strong>in</strong>terfaces of lattice-matched materials.<br />

R.G.D<strong>and</strong>rea, C.B.Duke: Physical Review B, 1992, 45[24), 14065-8<br />

[446-93/94-027]<br />

<strong>GaAs</strong>/AlAs: Interdiffusion<br />

Anneal<strong>in</strong>g experiments were performed on undoped s<strong>in</strong>gle-well heterostructures which<br />

had been grown by means of molecular beam epitaxy. The anneal<strong>in</strong>g was carried out <strong>in</strong><br />

evacuated <strong>and</strong> sealed silica ampoules. The anneal<strong>in</strong>g temperature <strong>and</strong> time were 1000C<br />

<strong>and</strong> 4h, respectively, while the As vapor pressure <strong>in</strong> the ampoules was varied from the<br />

dissociation pressure to about 108kPa. Compositional profiles were obta<strong>in</strong>ed by us<strong>in</strong>g<br />

dynamic secondary ion mass spectrometry. It was found that the amount of <strong>in</strong>termix<strong>in</strong>g <strong>in</strong><br />

389


Interdiffusion <strong>GaAs</strong> Interdiffusion<br />

the layers depended upon both the As pressure <strong>and</strong> the distance from the sample surface.<br />

In contrast with previous studies, the complex variation <strong>in</strong> <strong>in</strong>terdiffusion as a function of<br />

As pressure which was observed here could not be expla<strong>in</strong>ed <strong>in</strong> terms of <strong>in</strong>terdiffusion<br />

via group-<strong>III</strong> vacancies <strong>and</strong> <strong>in</strong>terstitials alone.<br />

N.Baba-Ali, I.Harrison, B.Tuck: Journal of Materials Science - Materials <strong>in</strong> Electronics,<br />

1995, 6[3], 127-34<br />

[446-127/128-138]<br />

<strong>GaAs</strong>/AlAs: Interdiffusion<br />

Comb<strong>in</strong>ed Raman <strong>and</strong> lattice dynamics studies of ultra-th<strong>in</strong> (<strong>GaAs</strong>) 4 (AlAs) 4 superlattices<br />

showed that a marked degree of <strong>in</strong>terdiffusion occurred <strong>in</strong> these samples when they were<br />

grown by us<strong>in</strong>g conventional molecular beam epitaxial temperatures of between 580 <strong>and</strong><br />

640C. At these temperatures, an <strong>in</strong>terruption of growth had little effect upon the structural<br />

quality of the superlattices.<br />

J.Grant, J.Menéndez, L.N.Pfeiffer, K.W.West, E.Mol<strong>in</strong>ari, S.Baroni: Applied Physics -<br />

Letters, 1991, 59[22], 2859-61<br />

[446-88/89-030]<br />

<strong>GaAs</strong>/AlAs: Interdiffusion<br />

A new method was proposed for the estimation of <strong>in</strong>terdiffusion coefficients <strong>in</strong><br />

superlattices. This was based upon measurements of the thickness of layers which<br />

rema<strong>in</strong>ed, without form<strong>in</strong>g an alloy, after an anneal<strong>in</strong>g process which caused<br />

<strong>in</strong>terdiffusion. The measurements were based upon the frequency of phonons from the<br />

Raman spectra. The <strong>in</strong>terdiffusion coefficient values which were found <strong>in</strong> this way were<br />

almost the same as those previously published. It was noted that Ga atoms <strong>in</strong> the present<br />

superlattices diffused more rapidly <strong>in</strong>to AlAs layers than Al atoms diffused <strong>in</strong>to <strong>GaAs</strong><br />

layers. The <strong>in</strong>terdiffusion coefficients first decreased with anneal<strong>in</strong>g time <strong>and</strong> <strong>in</strong>creased<br />

slightly when anneal<strong>in</strong>g was performed for more than 1.5h at 860C.<br />

N.Hara, T.Katoda: Journal of Applied Physics, 1991, 69[4], 2112-6<br />

[446-78/79-030]<br />

<strong>GaAs</strong>/AlAs/Al<strong>GaAs</strong>: Interdiffusion<br />

The <strong>in</strong>termix<strong>in</strong>g of double-barrier quantum wells by 50keV Ga + implantation was studied<br />

experimentally <strong>and</strong> theoretically. It was found that, even at low doses (less than<br />

10 12 /cm 2 ), a considerably broadened emission peak with an appreciable lum<strong>in</strong>escence<br />

blue-shift could be obta<strong>in</strong>ed. At medium doses (about 10 13 /cm 2 ), very large blue-shifts of<br />

the order of 0.2eV were observed which reta<strong>in</strong>ed a reasonable emission <strong>in</strong>tensity. At high<br />

doses (above 3 x 10 14 /cm 2 ), total <strong>in</strong>termix<strong>in</strong>g occurred <strong>and</strong> no photolum<strong>in</strong>escence could<br />

be recovered. The photolum<strong>in</strong>escence blue-shifts after low-dose implantation were not<br />

affected by the anneal<strong>in</strong>g temperature, whereas the blue-shift at high doses depended<br />

greatly upon the anneal<strong>in</strong>g conditions. The data <strong>in</strong>dicated heterogeneously enhanced<br />

390


Interdiffusion <strong>GaAs</strong> Interdiffusion<br />

<strong>in</strong>terdiffusion that was based upon a defect cluster model. It was noted that close control<br />

of the anneal<strong>in</strong>g ambient <strong>and</strong> sample surface was important.<br />

R.K.Kupka, Y.Chen: Journal of Applied Physics, 1995, 78[4], 2355-61<br />

[446-123/124-172]<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Interdiffusion<br />

The occurrence of <strong>in</strong>terdiffusion dur<strong>in</strong>g anneal<strong>in</strong>g was characterized by Raman scatter<strong>in</strong>g<br />

studies of a <strong>GaAs</strong>/Al x Ga 1-x As superlattice <strong>in</strong>to which various degrees of damage had been<br />

<strong>in</strong>troduced by the implantation of electrically <strong>in</strong>active iso-electronic 31 P + . This process<br />

elim<strong>in</strong>ated impurity charge associated effects. Auger <strong>and</strong> secondary ion mass<br />

spectrometry methods were used to determ<strong>in</strong>e the amount of mix<strong>in</strong>g beyond the damage<br />

zone <strong>in</strong> the superlattice. As a result, it was possible to dist<strong>in</strong>guish between the<br />

contributions of impurities <strong>and</strong> implantation-<strong>in</strong>duced defects to Ga/Al <strong>in</strong>termix<strong>in</strong>g.<br />

J.Sapriel, E.V.K.Rao, F.Brillouet, J.Chavignon, P.Ossart, Y.Gao, P.Krauz.: Superlattices<br />

<strong>and</strong> Microstructures, 1988, 4[1], 115-20<br />

[446-61-076]<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Interdiffusion<br />

The <strong>in</strong>termix<strong>in</strong>g of hetero-<strong>in</strong>terfaces, by Ga + implantation <strong>and</strong> anneal<strong>in</strong>g, was<br />

<strong>in</strong>vestigated. Damage accumulation <strong>in</strong> a <strong>GaAs</strong>/AlAs superlattice was found to be less<br />

rapid than that <strong>in</strong> a <strong>GaAs</strong>/GaAlAs quantum-well structure. Low-temperature<br />

photolum<strong>in</strong>escence spectroscopy of a <strong>GaAs</strong>/AlAs superlattice was performed for doses as<br />

high as 10 16 /cm 2 . The photolum<strong>in</strong>escence spectra exhibited several emission b<strong>and</strong>s on the<br />

high-energy side. The number <strong>and</strong> energy of these blue-shifted peaks were found to<br />

depend upon the implanted dose. Secondary ion mass spectrometric data suggested that<br />

they could be <strong>in</strong>terpreted as be<strong>in</strong>g due to emission from several quantum wells, of the<br />

superlattice, which disordered at differ<strong>in</strong>g mix<strong>in</strong>g rates. Two regimes were revealed.<br />

Thus, while the depth extension of the disorder<strong>in</strong>g was directly related to the postimplantation<br />

defect distribution <strong>in</strong> the high-dose regime, some diffusion of these defects<br />

dur<strong>in</strong>g anneal<strong>in</strong>g occurred <strong>in</strong> the low-dose regime. Cross-sectional transmission electron<br />

microscopy confirmed that there was an effect of the structure, of the implanted sample,<br />

upon damage accumulation. A decrease <strong>in</strong> the photolum<strong>in</strong>escence <strong>in</strong>tensity after<br />

anneal<strong>in</strong>g was attributed to the presence of extended residual defects <strong>in</strong> the implanted<br />

layers. A study of the effect of anneal<strong>in</strong>g time at 760C showed that the<br />

photolum<strong>in</strong>escence <strong>in</strong>tensity progressively recovered, whereas the <strong>in</strong>termix<strong>in</strong>g rapidly<br />

saturated.<br />

C.Vieu, M.Schneider, R.Planel, H.Launois, B.Descouts, Y.Gao: Journal of Applied<br />

Physics, 1991, 70[3], 1433-43<br />

[446-93/94-028]<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Interdiffusion<br />

The shape of the conf<strong>in</strong>ement potential which resulted from <strong>in</strong>terdiffusion of a <strong>GaAs</strong><br />

quantum well, locally enhanced by defects due to Ga implantation, was computed. The<br />

simplest model which could take account of lateral diffusion of the defects was used. A<br />

391


Interdiffusion <strong>GaAs</strong> Interdiffusion<br />

variational calculation of the first 2 electronic levels with<strong>in</strong> this two-dimensional potential<br />

supported the assignment of recently observed new cathodolum<strong>in</strong>escence l<strong>in</strong>es to<br />

electrons which were laterally conf<strong>in</strong>ed <strong>in</strong> a graded potential.<br />

J.Cibert, P.M.Petroff: Physical Review B, 1987, 36[6], 3243-6<br />

[446-55/56-020]<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Interdiffusion<br />

Low-energy As + implantation, followed by rapid thermal anneal<strong>in</strong>g, was used to modify<br />

the exciton transition energies of quantum wells. Various structures were irradiated us<strong>in</strong>g<br />

an energy which was sufficiently low that the disordered region was spatially separated<br />

from the quantum wells. After rapid thermal anneal<strong>in</strong>g, the exciton energies exhibited<br />

large <strong>in</strong>creases which depended upon the quantum-well width <strong>and</strong> the implantation<br />

fluence. There was no appreciable <strong>in</strong>crease <strong>in</strong> the peak l<strong>in</strong>e-widths. The observed energy<br />

shifts were attributed to modifications of the shapes of the quantum wells, due to<br />

enhanced Ga <strong>and</strong> Al <strong>in</strong>terdiffusion at hetero-<strong>in</strong>terfaces <strong>in</strong> irradiated areas. The results<br />

were consistent with a model which was based upon an enhanced <strong>in</strong>termix<strong>in</strong>g of Al <strong>and</strong><br />

Ga atoms, <strong>in</strong> the depth of the material, due to the diffusion of vacancies which were<br />

generated near to the surface.<br />

B.Elman, E.S.Koteles, P.Melman, C.A.Armiento: Journal of Applied Physics, 1989,<br />

66[5], 2104-7<br />

[446-74-027]<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Interdiffusion<br />

Implantation with Se + ions, <strong>and</strong> the anneal<strong>in</strong>g of quantum-well samples, were studied by<br />

us<strong>in</strong>g transmission electron microscopy, photolum<strong>in</strong>escence spectroscopy <strong>and</strong> Monte<br />

Carlo simulation. It was concluded that enhanced layer <strong>in</strong>terdiffusion occurred at depths<br />

which were several times greater than the projected range of the implanted Se. There<br />

were signs of residual stress at similar depths.<br />

E.G.Bithell, W.M.Stobbs, C.Phillips, R.Eccleston, R.Gwilliam: Journal of Applied<br />

Physics, 1990, 67[3], 1279-87<br />

[446-74-027]<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Interdiffusion<br />

The transient-enhanced <strong>in</strong>terdiffusion of <strong>in</strong>terfaces dur<strong>in</strong>g the rapid thermal anneal<strong>in</strong>g of<br />

ion-implanted heterostructures was studied. It was shown that the factors which most<br />

<strong>in</strong>fluenced the degree of <strong>in</strong>terdiffusion were the temperature, the concentration of excess<br />

vacancies, <strong>and</strong> the ability of the vacancies to diffuse freely. A model was developed <strong>in</strong><br />

order to expla<strong>in</strong> these observations. It was based upon the solution of coupled diffusion<br />

equations which <strong>in</strong>volved excess vacancies <strong>and</strong> the distribution of Al after ion<br />

implantation. Both <strong>in</strong>itial distributions were deduced from a 3-dimensional Monte Carlo<br />

simulation of ion implantation <strong>in</strong>to a heterostructure sample. The model was found to give<br />

392


Interdiffusion <strong>GaAs</strong> Interdiffusion<br />

excellent agreement with experiment. In particular, it was valid for as-implanted vacancy<br />

concentrations of less than 6 x 10 19 /cm 3 .<br />

K.B.Kahen, G.Rajeswaran: Journal of Applied Physics, 1989, 66[2], 545-51<br />

[446-74-026]<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Interdiffusion<br />

A high-energy (up to 150keV) Ga + focussed ion beam was used to implant quantum-well<br />

structures, <strong>and</strong> to <strong>in</strong>terdiffuse heterojunctions so as to create quantum wires <strong>and</strong> boxes.<br />

Wires as wide as 80nm were found 200nm below the surface. Optical damage <strong>and</strong><br />

<strong>in</strong>terdiffusion processes were studied as a function of the implantation parameters <strong>and</strong> the<br />

(rapid) thermal anneal<strong>in</strong>g time. A universal correlation was found between the optical<br />

damage <strong>and</strong> the <strong>in</strong>terdiffusion length.<br />

F.Laruelle, P.Hu, R.Simes, R.Kubena, W.Rob<strong>in</strong>son, J.Merz, P.M.Petroff: Journal of<br />

Vacuum Science <strong>and</strong> Technology B, 1989, 7[6], 2034-8<br />

[446-74-026]<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Interdiffusion<br />

The occurrence of enhanced <strong>in</strong>terdiffusion at the <strong>in</strong>terface was studied. Ions of Ar were<br />

implanted to doses which ranged from 2 x 10 13 to 5 x 10 14 /cm 2 , <strong>and</strong> the samples were<br />

then rapidly thermally annealed (950C, 30s). It was found that the degree of <strong>in</strong>termix<strong>in</strong>g<br />

decreased from the surface towards the projected ion range, <strong>and</strong> was a function of the<br />

implantation dose. It was suggested that this variation arose from the coalescence of some<br />

of the excess vacancies <strong>in</strong>to extended defects, so that they were then unavailable to the<br />

enhanced diffusion mechanism. By assum<strong>in</strong>g that the concentration of mobile vacancies<br />

at a given depth was proportional to the electronic energy of the ion, <strong>and</strong> <strong>in</strong>versely<br />

proportional to the nuclear energy loss of the ion, predictions were obta<strong>in</strong>ed which were<br />

<strong>in</strong> good agreement with the experimental results.<br />

K.B.Kahen, D.L.Peterson, G.Rajeswaran: Journal of Applied Physics, 1990, 68[5], 2087-<br />

90<br />

[446-86/87-031]<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Interdiffusion<br />

<strong>Diffusion</strong> <strong>and</strong> <strong>in</strong>terdiffusion <strong>in</strong> <strong>GaAs</strong>/Al<strong>GaAs</strong> superlattices was shown to be consistent<br />

with a charged po<strong>in</strong>t defect model. Charged Ga vacancies, V Ga 3- , <strong>and</strong> <strong>in</strong>terstitials, I Ga 2+ ,<br />

appeared to control the diffusion of group-II, group-<strong>III</strong> <strong>and</strong> (probably) group-V elements.<br />

After adjust<strong>in</strong>g for carrier concentration <strong>and</strong> As pressure, these elements were found to<br />

have an almost identical <strong>in</strong>tr<strong>in</strong>sic diffusivity <strong>and</strong> activation energy over a wide range of<br />

temperatures. A natural consequence, of Ga diffusion via negative or positive po<strong>in</strong>t<br />

defects, was that enhanced group-<strong>III</strong> <strong>in</strong>terdiffusion was expected under either n-type or p-<br />

type dop<strong>in</strong>g. Anomalous enhancements of group-II dopant diffusivity were related to the<br />

supersaturation of Ga <strong>in</strong>terstitials.<br />

R.M.Cohen: Journal of Applied Physics, 1990, 67[12], 7268-73<br />

[446-78/79-011]<br />

393


Interdiffusion <strong>GaAs</strong> Interdiffusion<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Interdiffusion<br />

The superlattice <strong>and</strong> multi-quantum well structures of various <strong>III</strong>-V semiconductors were<br />

remarkably stable to thermal anneal<strong>in</strong>g. This was very different to the <strong>in</strong>stability of these<br />

structures which occurred when an impurity such as Zn or Si was diffused <strong>in</strong>to them. It<br />

was also recalled that selective area impurity-<strong>in</strong>duced disorder<strong>in</strong>g of multi-quantum well<br />

structures had become a potentially powerful technique for the fabrication of certa<strong>in</strong><br />

devices. The current state of underst<strong>and</strong><strong>in</strong>g of the fundamental mechanisms was reviewed<br />

here. It was po<strong>in</strong>ted out that, <strong>in</strong> n-type material, the <strong>in</strong>volvement of group-<strong>III</strong> vacancies<br />

was generally accepted. However, <strong>in</strong> the case of p-type dopants such as Zn, an oversaturation<br />

of self-<strong>in</strong>terstitials was required for the enhancement of <strong>in</strong>terdiffusion.<br />

I.Harrison, H.P.Ho, N.Baba-Ali: Journal of Electronic Materials, 1991, 20[6], 449-56<br />

[446-91/92-016]<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Interdiffusion<br />

The compositional disorder<strong>in</strong>g of <strong>GaAs</strong>/Al<strong>GaAs</strong> quantum wells, due to the presence of<br />

low-temperature molecular beam epitaxially grown <strong>GaAs</strong>, was studied. It was found that<br />

Ga vacancy-enhanced <strong>in</strong>terdiffusion was the mechanism which was responsible for the<br />

observed <strong>in</strong>termix<strong>in</strong>g. The diffusion equations were solved numerically <strong>in</strong> order to obta<strong>in</strong><br />

the b<strong>and</strong> profile after <strong>in</strong>termix<strong>in</strong>g. The transition energies <strong>in</strong> the quantum wells under<br />

various anneal<strong>in</strong>g conditions were predicted, <strong>and</strong> were found to agree very well with<br />

observed photolum<strong>in</strong>escence emission peaks. The vacancy-<strong>in</strong>duced <strong>in</strong>terdiffusivity was<br />

found to require an activation energy of 4.08eV. This was smaller than the activation<br />

energy for the <strong>in</strong>terdiffusion of <strong>GaAs</strong>/Al<strong>GaAs</strong> heterostructures which were grown at<br />

normal temperatures. It was concluded that the present results clearly <strong>in</strong>dicated an<br />

enhanced <strong>in</strong>terdiffusion that was due to the presence of <strong>GaAs</strong> which had been grown at<br />

low temperatures.<br />

J.S.Tsang, C.P.Lee, S.H.Lee, K.L.Tsai, H.R.Chen: Journal of Applied Physics, 1995,<br />

77[9], 4302-6<br />

[446-121/122-062]<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Interdiffusion<br />

The implantation of Ga + , followed by rapid thermal anneal<strong>in</strong>g, was used to enhance<br />

<strong>in</strong>terdiffusion <strong>in</strong> s<strong>in</strong>gle quantum wells. The extent of <strong>in</strong>termix<strong>in</strong>g was found to depend<br />

upon the well depth, the number of implanted ions <strong>and</strong> the anneal<strong>in</strong>g time. Very rapid<br />

<strong>in</strong>terdiffusion occurred <strong>in</strong> the <strong>in</strong>itial anneal<strong>in</strong>g stage. The enhanced diffusion coefficient<br />

subsequently returned to the non-implanted value. A 2-step model was proposed <strong>in</strong> order<br />

to expla<strong>in</strong> the diffusion process as a function of anneal<strong>in</strong>g time. This <strong>in</strong>volved a rapid<br />

diffusion process <strong>and</strong> a saturated diffusion process. The <strong>in</strong>terdiffusion coefficient for the<br />

rapid diffusion was found to be well depth-dependent <strong>and</strong> was estimated to be between<br />

5.4 x 10 -16 <strong>and</strong> 1. 5 x 10 -15 cm 2 /s.<br />

N.Sai, B.Zheng, J.Xu, P.Zhang, X.Yang, Z.Xu: Solid State Communications, 1996,<br />

98[12], 1039-42<br />

[446-136/137-116]<br />

394


Interdiffusion <strong>GaAs</strong> Interdiffusion<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Interdiffusion<br />

A new impurity-free <strong>in</strong>terdiffusion technique was described which <strong>in</strong>volved pulsed<br />

anodization followed by rapid thermal anneal<strong>in</strong>g at temperatures near to 900C. Enhanced<br />

<strong>in</strong>terdiffusion was observed, <strong>in</strong> the presence of an anodized <strong>GaAs</strong> capp<strong>in</strong>g layer, <strong>in</strong><br />

<strong>GaAs</strong>/Al<strong>GaAs</strong> quantum-well structures. The use of transmission electron microscopy<br />

revealed evidence of <strong>in</strong>terdiffusion. The photolum<strong>in</strong>escence spectra from <strong>in</strong>terdiffused<br />

samples exhibited a large blue-shift, with no appreciable l<strong>in</strong>e-width broaden<strong>in</strong>g.<br />

S.Yuan, Y.Kim, C.Jagadish, P.T.Burke, M.Gal, J.Zou, D.Q.Cai, D.J.H.Cockayne,<br />

R.M.Cohen: Applied Physics Letters, 1997, 70[10], 1269-71<br />

[446-148/149-178]<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Interdiffusion<br />

A formula was derived which described the <strong>in</strong>terdiffusion profiles of quantum wells. It<br />

was shown that it accurately modelled <strong>in</strong>terdiffusion <strong>in</strong> quantum wells of lattice-matched<br />

Al<strong>GaAs</strong>. The formula took account of the differ<strong>in</strong>g <strong>in</strong>terdiffusion coefficients between<br />

layers, <strong>and</strong> of the <strong>in</strong>terfacial discont<strong>in</strong>uity of <strong>in</strong>terdiffused species. The formula expla<strong>in</strong>ed<br />

how quantum energy shifts due to <strong>in</strong>terdiffusion varied with anneal<strong>in</strong>g time <strong>and</strong> anneal<strong>in</strong>g<br />

temperature <strong>in</strong> various wide-well layers of both In<strong>GaAs</strong>P/InP <strong>and</strong> <strong>GaAs</strong>/Al<strong>GaAs</strong> quantum<br />

wells. The quantitative difference between the <strong>in</strong>terdiffusion profiles <strong>in</strong> these two<br />

materials was also demonstrated.<br />

K.Mukai, M.Sugawara, S.Yamazaki: Physical Review B, 1994, 50[4], 2273-82<br />

[446-115/116-130]<br />

<strong>GaAs</strong>/Al<strong>GaAs</strong>: Interdiffusion<br />

The dependence of impurity-free <strong>in</strong>terdiffusion upon the properties of a dielectric cap<br />

layer was studied <strong>in</strong> unstra<strong>in</strong>ed multi-quantum well structures that had been grown by<br />

means of molecular beam epitaxy. Electron-beam evaporated SiO 2 films, chemical vapor<br />

deposited SiO x N y films, <strong>and</strong> spun-on SiO 2 films were used as cap layers dur<strong>in</strong>g rapid<br />

thermal anneal<strong>in</strong>g at temperatures of between 850 <strong>and</strong> 950C. The photolum<strong>in</strong>escence at<br />

10K was used to monitor <strong>in</strong>terdiffusion-<strong>in</strong>duced b<strong>and</strong>-gap shifts, <strong>and</strong> to calculate the<br />

correspond<strong>in</strong>g Al-Ga <strong>in</strong>terdiffusion coefficients. The latter were found to <strong>in</strong>crease with<br />

cap layer thickness (electron-beam SiO 2 ) up to a limit which was governed by saturation<br />

of the out-diffused Ga concentration <strong>in</strong> the SiO 2 caps. A maximum concentration of<br />

between 4 x 10 19 <strong>and</strong> 7 x 10 19 /cm 3 <strong>in</strong> the SiO 2 caps was found by us<strong>in</strong>g secondary ion<br />

mass spectroscopic profil<strong>in</strong>g. Larger b<strong>and</strong>-edge shifts were also obta<strong>in</strong>ed when the O<br />

content of SiO x N y cap layers was <strong>in</strong>creased, but the differences were <strong>in</strong>sufficient to<br />

suggest a laterally selective <strong>in</strong>terdiffusion process that was based upon variations <strong>in</strong> cap<br />

layer composition alone. Much larger differences were obta<strong>in</strong>ed by us<strong>in</strong>g various<br />

deposition techniques for the cap layers. This <strong>in</strong>dicated that the porosity of the cap layer<br />

395


Interdiffusion <strong>GaAs</strong> Interdiffusion<br />

was a much more important factor than was the film composition <strong>in</strong> obta<strong>in</strong><strong>in</strong>g a laterally<br />

selective <strong>in</strong>terdiffusion process.<br />

S.Bürkner, M.Maier, E.C.Lark<strong>in</strong>s, W.Rothemund, E.P.O’Reilly, J.D.Ralston: Journal of<br />

Electronic Materials, 1995, 24[7], 805-12<br />

[446-125/126-131]<br />

<strong>GaAs</strong>/AlInP: Interdiffusion<br />

An analysis was made of the <strong>in</strong>terdiffusion of a discrete <strong>GaAs</strong> layer, <strong>in</strong>to an Al 0.5 In 0.5 P<br />

half-space, by us<strong>in</strong>g Si dop<strong>in</strong>g as an agent for enhanced layer <strong>in</strong>terdiffusion. Enhanced<br />

<strong>in</strong>terdiffusion was observed on both column-<strong>III</strong> <strong>and</strong> column-V sites; but the column-<strong>III</strong><br />

<strong>in</strong>terdiffusion coefficient exceeded the column-V <strong>in</strong>terdiffusion coefficient by 2 orders of<br />

magnitude. Due to this disparity between the diffusion coefficients, large defectproduc<strong>in</strong>g<br />

stra<strong>in</strong>s were <strong>in</strong>troduced by the <strong>in</strong>terdiffusion. It was shown that, by modell<strong>in</strong>g<br />

the resultant stra<strong>in</strong> profiles <strong>and</strong> by apply<strong>in</strong>g a critical thickness analysis, the <strong>in</strong>stability of<br />

such <strong>in</strong>terdiffused structures could be understood.<br />

R.L.Thornton, F.A.Ponce, G.B.Anderson, F.J.Endicott: Applied Physics Letters, 1993,<br />

62[17], 2060-2<br />

[446-106/107-082]<br />

<strong>GaAs</strong>/GaAlAs: Interdiffusion<br />

Simple analytical expressions were derived for the approximate estimation of the<br />

<strong>in</strong>terdiffusion coefficient, of partially disordered quantum-well heterostructures, directly<br />

from measurements of the photolum<strong>in</strong>escence phase shift which was associated with<br />

layer <strong>in</strong>terdiffusion. The phase shift was calculated as a function of the <strong>in</strong>terdiffusion<br />

length, (Dt) ½ , <strong>in</strong> the lattice-matched system, <strong>GaAs</strong>/Ga 0.7 Al 0.3 As. The calculations were<br />

performed with<strong>in</strong> the framework of the envelope function approximation <strong>and</strong> Fick's law.<br />

A simple relationship was derived for the variation <strong>in</strong> phase shift as a function of the<br />

dimensionless parameter, (Dt) ½ /L, where L was the quantum-well thickness. This<br />

satisfactorily accounted for most of the published <strong>in</strong>terdiffusivity values, to with<strong>in</strong> a<br />

factor of 2.<br />

M.T.Furtado, M.S.S.Loural: Superlattices <strong>and</strong> Microstructures, 1993, 14[1], 21-5<br />

[446-113/114-029]<br />

332 <strong>GaAs</strong>/GaAlAs: Interdiffusion<br />

The effects of Si <strong>and</strong> Be, at dop<strong>in</strong>g levels of up to 10 19 /cm 3 , upon the <strong>in</strong>terdiffusion of<br />

quantum wells after anneal<strong>in</strong>g were studied by us<strong>in</strong>g photolum<strong>in</strong>escence techniques (table<br />

32). It was found that, for Be concentrations of up to 2.5 x 10 19 /cm 3 , <strong>and</strong> for Si<br />

concentrations of up to 10 18 /cm 3 , no change <strong>in</strong> the <strong>in</strong>terdiffusion coefficients could be<br />

measured. At a Si dopant concentration of 6 x 10 18 /cm 3 , there was a dramatic degradation<br />

of the material quality after anneal<strong>in</strong>g (750C, 15s). This caused the lum<strong>in</strong>escence from<br />

the well to disappear, while a deep-level lum<strong>in</strong>escence that was related to donor-Ga<br />

vacancy complexes <strong>and</strong> As antisite defects appeared. On the basis of these results, it was<br />

suggested that the position of the Fermi level played no role <strong>in</strong> the <strong>in</strong>termix<strong>in</strong>g of <strong>III</strong>-V<br />

396


Interdiffusion <strong>GaAs</strong> Interdiffusion<br />

heterostructures. It was also concluded that most of the enhanced <strong>in</strong>termix<strong>in</strong>g which was<br />

observed <strong>in</strong> Si-doped <strong>GaAs</strong>/Al<strong>GaAs</strong> structures was related to Si relocation at very high<br />

dop<strong>in</strong>g levels.<br />

W.P.Gill<strong>in</strong>, I.V.Bradley, L.K.Howard, R.Gwilliam, K.P.Homewood: Journal of Applied<br />

Physics, 1993, 73[11], 7715-9<br />

[446-106/107-082]<br />

Table 32<br />

Interdiffusion Data for Ga 0.8 Al 0.2 As/<strong>GaAs</strong><br />

Dopant Amount (/cm 3 ) Temperature (C) Coefficient (cm 2 /s)<br />

- - 1000 1.54 x 10 -16<br />

- - 1050 6.95 x 10 -16<br />

- - 1100 3.76 x 10 -15<br />

Si 5 x 10 17 1000 4.50 x 10 -16<br />

Si 5 x 10 17 1050 5.60 x 10 -16<br />

Si 5 x 10 17 1100 1.19 x 10 -15<br />

Si 10 18 1000 8.60 x 10 -17<br />

Si 10 18 1050 4.50 x 10 -16<br />

Si 10 18 1100 1.79 x 10 -15<br />

<strong>GaAs</strong>/GaAlAs: Interdiffusion<br />

The Al-Ga <strong>in</strong>terdiffusion which was produced by focussed Si ion-implantation <strong>and</strong> rapid<br />

thermal anneal<strong>in</strong>g was <strong>in</strong>vestigated <strong>in</strong> a Ga 0.7 Al 0.3 As/<strong>GaAs</strong> superlattice structure with<br />

equal (3.5nm) barrier <strong>and</strong> well widths. Ions of Si 2+ were accelerated to 50 or 100kV <strong>and</strong><br />

were implanted, parallel to the sample normal, to doses which ranged from 10 13 to<br />

10 15 /cm 2 . The effect of rapid thermal anneal<strong>in</strong>g (950C, 10s) was characterized by means of<br />

secondary ion mass spectrometry. It was found that, <strong>in</strong> the implanted region, the<br />

<strong>in</strong>terdiffusion was significantly enhanced by Si implantation. Ion doses which were as low<br />

as 10 14 /cm 2 led to a 2 orders of magnitude <strong>in</strong>crease, <strong>in</strong> the <strong>in</strong>terdiffusion coefficient, to a<br />

value of 4.5 x 10 -14 cm 2 /s. This led to a mix<strong>in</strong>g effectiveness of about 90%. On the <strong>other</strong><br />

h<strong>and</strong>, the use of rapid thermal anneal<strong>in</strong>g alone produced an <strong>in</strong>terdiffusion coefficient of 1.3<br />

x 10 -16 cm 2 /s; with very little mix<strong>in</strong>g. A marked depth dependence of the mix<strong>in</strong>g process<br />

was observed at an implantation energy of 100keV, with a more heavily mixed, so-called<br />

p<strong>in</strong>ch-off, region be<strong>in</strong>g formed at a certa<strong>in</strong> depth. It was noted that the depth at which this<br />

enhancement occurred was not associated with a maximum concentration of Si ions or of<br />

vacancies. It <strong>in</strong>stead co<strong>in</strong>cided with a positive maximum <strong>in</strong> the second derivative of the<br />

vacancy profile. This, <strong>in</strong> turn, represented a maximum <strong>in</strong> the vacancy<br />

397


Interdiffusion <strong>GaAs</strong> Interdiffusion<br />

<strong>in</strong>jection process that was caused by the presence of a transient vacancy concentration<br />

gradient.<br />

P.Chen, A.J.Steckl: Journal of Applied Physics, 1995, 77[11], 5616-24<br />

[446-121/122-065398]<br />

<strong>GaAs</strong>/GaAlAs: Interdiffusion<br />

Electrically <strong>in</strong>active iso-electronic 31 P + <strong>and</strong> 27 Al + were implanted <strong>in</strong>to molecular beam<br />

epitaxially grown <strong>GaAs</strong>/Ga 0.7 Al 0.3 As superlattices at 25 or 250C. Evidence was found for<br />

implantation damage <strong>and</strong> for Al/Ga <strong>in</strong>terdiffusion which depended upon the anneal<strong>in</strong>g<br />

time.<br />

E.V.K.Rao, F.Brillouet, P.Ossart, Y.Gao, J.Sapriel, P.Krauz: Journal de Physique -<br />

Colloque C5, 1987, 48[11], 113-6<br />

[446-61-076]<br />

<strong>GaAs</strong>/GaAlAs: Interdiffusion<br />

By carry<strong>in</strong>g out various anneal<strong>in</strong>g treatments on Sn-doped molecular beam epitaxially<br />

grown <strong>GaAs</strong>-Ga 0.72 Al 0.28 As quantum well structures it was shown that Sn, like <strong>other</strong><br />

donor atoms (Si, S), <strong>in</strong>duced disorder<strong>in</strong>g by enhanc<strong>in</strong>g <strong>in</strong>terdiffusion. Also, the voluntary<br />

<strong>in</strong>troduction of B atoms <strong>in</strong>to a Sn-doped structure before anneal<strong>in</strong>g led to a retardation of<br />

Sn-enhanced <strong>in</strong>terdiffusion.<br />

E.V.K.Rao, P.Ossart, F.Alex<strong>and</strong>re, H.Thibierge: Applied Physics Letters, 1987, 50[10],<br />

588-91<br />

[446-51/52-125]<br />

<strong>GaAs</strong>/<strong>GaAs</strong>P, <strong>GaAs</strong>/<strong>GaAs</strong>Sb: Interdiffusion<br />

Interdiffusion <strong>in</strong> superlattices was studied at various temperatures <strong>and</strong> under various As<br />

partial pressures. An analysis of the As pressure-dependence of the effective diffusion<br />

coefficient revealed that a substitutional-<strong>in</strong>terstitial diffusion mechanism governed the<br />

<strong>in</strong>terdiffusion process. Computer simulations were used to study the profile shapes of<br />

annealed samples, <strong>and</strong> the As pressure dependence of the effective diffusion coefficient.<br />

It was found that the Frank-Turnbull diffusion mechanism governed the <strong>in</strong>terdiffusion of<br />

these superlattices. The As pressure-dependence of the effective diffusion coefficients, as<br />

measured <strong>in</strong> <strong>in</strong>terdiffusion experiments, was opposite to the reported pressure<br />

dependences which had been measured <strong>in</strong> As <strong>and</strong> P <strong>in</strong>-diffusion experiments. The<br />

apparently contradictory <strong>in</strong>-diffusion <strong>and</strong> out-diffusion behaviors could be reconciled by<br />

a diffusion model which <strong>in</strong>volved As vacancies, fast-diffus<strong>in</strong>g As-vacancy plus P-<br />

<strong>in</strong>terstitial complexes, <strong>and</strong> fast-diffus<strong>in</strong>g P <strong>in</strong>terstitials (or the analogous Sb-related<br />

defects).<br />

M.Schultz, U.Egger, R.Scholz, O.Breitenste<strong>in</strong>, P.Werner, U.Gösele, R.Franzheld,<br />

M.Uematsu, H.Ito: Defect <strong>and</strong> <strong>Diffusion</strong> Forum, 1997, 143-147, 1101-8<br />

[446-143/147-1101]<br />

398


Interdiffusion <strong>GaAs</strong> Interdiffusion<br />

<strong>GaAs</strong>/GaInP: Interdiffusion<br />

An analytical <strong>and</strong> experimental <strong>in</strong>vestigation was made of the mechanisms of defect<br />

formation dur<strong>in</strong>g the <strong>in</strong>terdiffusion of these materials. It was found that the analytical<br />

model predicted a critical thickness, below which defects were not produced dur<strong>in</strong>g this<br />

highly stra<strong>in</strong>ed <strong>in</strong>terdiffusion process. Transmission electron microscopic analysis of<br />

diffused buried layers of various thickness revealed a very good qualitative agreement<br />

with the present model.<br />

R.L.Thornton, D.P.Bour, D.Treat, F.A.Ponce, J.C.Tramontana, F.J.Endicott: Applied<br />

Physics Letters, 1994, 65[21], 2696-8<br />

[446-119/120-203]<br />

<strong>GaAs</strong>/In: Interdiffusion<br />

The effects of <strong>in</strong>teractions between a thick In layer <strong>and</strong> heat-treated <strong>GaAs</strong> at 570C were<br />

<strong>in</strong>vestigated by means of scann<strong>in</strong>g electron microscopy, secondary ion mass<br />

spectrometry, Rutherford back-scatter<strong>in</strong>g spectrometry, X-ray diffraction <strong>and</strong> Nomarski<br />

techniques. It was shown that, as well as the usual In<strong>GaAs</strong> crystallites which grew<br />

epitaxially upon dissolution of the substrate, an array of In-rich dendrites was present<br />

whose numbers were related to the dislocation density. The driv<strong>in</strong>g force for In to migrate<br />

along the dislocations <strong>and</strong> eventually form In<strong>GaAs</strong> spikes was assumed to be an excess of<br />

As which had been reported to be present <strong>in</strong> the vic<strong>in</strong>ity of <strong>in</strong>dividual dislocations. It was<br />

suggested that exist<strong>in</strong>g data on the coefficient for the conventional diffusion of In <strong>in</strong><br />

<strong>GaAs</strong> had been overestimated by a factor of 10 6 .<br />

A.J.Barcz, J.M.Baranowski, S.Kwiatkowski: Applied Physics A, 1995, 60[3], 321-4<br />

[446-134/135-136]<br />

<strong>GaAs</strong>/In<strong>GaAs</strong>: Interdiffusion<br />

The <strong>in</strong>terdiffusion of a multiple quantum-well sample, due to a th<strong>in</strong> source of vacancies,<br />

was used as a probe for the simultaneous measurement of the <strong>in</strong>terdiffusion coefficient,<br />

the diffusivity of group-<strong>III</strong> vacancies <strong>and</strong> the background concentration of these<br />

vacancies. It was shown that <strong>in</strong>terdiffusion at all temperatures was governed by a constant<br />

background concentration of vacancies, <strong>and</strong> that this concentration was equal to that of<br />

the vacancies <strong>in</strong> the substrate. The measured vacancy concentration was about 2 x<br />

10 17 /cm 3 . This showed that the vacancy concentrations were not <strong>in</strong> thermal equilibrium,<br />

contrary to the usual assumption. It <strong>in</strong>stead had a value which was frozen <strong>in</strong>; probably at<br />

the growth temperature. It was shown that the activation energy for the <strong>in</strong>termix<strong>in</strong>g of<br />

In<strong>GaAs</strong> <strong>and</strong> <strong>GaAs</strong> was governed only by the activation energy for vacancy diffusion,<br />

which was estimated to have a value of 3.4eV.<br />

O.M.Khreis, W.P.Gill<strong>in</strong>, K.P.Homewood: Physical Review B, 1997, 55[23], 15813-8<br />

[446-152-0399]<br />

<strong>GaAs</strong>/Ir: Interdiffusion<br />

Interfacial reactions, <strong>in</strong> both th<strong>in</strong>-film <strong>and</strong> bulk samples, were studied at temperatures of<br />

between 400 <strong>and</strong> 1000C. The diffusion path for Ir/<strong>GaAs</strong> was found to be:<br />

399


Interdiffusion <strong>GaAs</strong>|Ga(As,P) Zn<br />

Ir/lrGa/IrAs 2 /<strong>GaAs</strong>. In the th<strong>in</strong>-film case, where the Ir supply was limited, the f<strong>in</strong>al<br />

configuration was: Ga 5 Ir 3 /IrAs 2 /<strong>GaAs</strong>. The activation energies for diffusion <strong>in</strong> the th<strong>in</strong>film<br />

<strong>and</strong> bulk cases were 3.15 <strong>and</strong> 2.96eV, respectively.<br />

K.J.Schulz, O.A.Musbah, Y.A.Chang: Journal of Applied Physics, 1990, 67[11], 6798-<br />

806<br />

[446-78/79-032]<br />

Ga(As,P)<br />

Zn<br />

<strong>GaAs</strong>P: Zn <strong>Diffusion</strong><br />

A simple method for the open-tube diffusion of Zn from (ZnO) x (SiO 2 ) 1-x film sources,<br />

<strong>and</strong> <strong>in</strong>to <strong>GaAs</strong> 0.6 P 0.4 , was described. The oxide films were deposited by us<strong>in</strong>g metalorganic<br />

chemical vapor deposition. A capp<strong>in</strong>g layer of SiO 2 was deposited on top of the<br />

source films, <strong>and</strong> diffusion was carried out <strong>in</strong> flow<strong>in</strong>g N at 650C. <strong>Diffusion</strong> depths of<br />

between 200nm <strong>and</strong> several microns could be easily obta<strong>in</strong>ed. The diffusion front <strong>in</strong> n-<br />

type substrates was sharp. The dependence of the diffusion depth upon the source film<br />

composition (for x-values of 0.04 to 1) was determ<strong>in</strong>ed by us<strong>in</strong>g section<strong>in</strong>g methods.<br />

D.J.Lawrence, F.T.Smith, S.T.Lee: Journal of Applied Physics, 1991, 69[5], 3011-5<br />

[446-78/79-002]<br />

<strong>GaAs</strong>P: Zn <strong>Diffusion</strong><br />

Samples of <strong>GaAs</strong> 0.6 P 0.4 were diffused with Zn, via a 200 to 300nm protective ZrO 2 layer.<br />

The diffusion depth exhibited a square-root time dependence. The absolute diffusivity<br />

values depended slightly upon the diffusion conditions. The layer had essentially no<br />

effect upon the carrier concentration profile or the activation energy.<br />

J.E.Bisberg, A.K.Ch<strong>in</strong>, F.P.Dabkowski: Journal of Applied Physics, 1990, 67[3], 1347-51<br />

[446-74-003]<br />

400


Zn Ga(As,P)|Ga(As,Sb) Interdiffusion<br />

<strong>GaAs</strong>P: Zn <strong>Diffusion</strong><br />

The open-tube diffusion of Zn <strong>in</strong>to <strong>GaAs</strong> 0.8 P 0.2 , from a Zn-doped silica film, was<br />

<strong>in</strong>vestigated by us<strong>in</strong>g films of AlN or SiN x as anneal<strong>in</strong>g caps. The dependence of the<br />

diffusion depth upon the thickness of an AlN cap was found to differ from the<br />

dependence upon SiN x cap thickness. Selective masked diffusion of Zn, us<strong>in</strong>g an AlN<br />

mask, was also studied. It was found that the diffusion depth dur<strong>in</strong>g selective masked<br />

diffusion depended upon both the AlN cap thickness <strong>and</strong> the AlN diffusion-mask<br />

thickness. The results suggested that the diffusion depth was not necessarily governed by<br />

either the cap thickness or the diffusion-mask thickness. It was concluded that the total<br />

film stress was the ma<strong>in</strong> factor which determ<strong>in</strong>ed the diffusion depth under the present<br />

conditions.<br />

M.Ogihara, M.Tan<strong>in</strong>aka, Y.Nakamura: Journal of Applied Physics, 1996, 79[6], 2995-<br />

3002<br />

[446-131/132-175]<br />

Interdiffusion<br />

<strong>GaAs</strong>P/<strong>GaAs</strong>: Interdiffusion<br />

Interdiffusion on the group-V sub-lattice of <strong>GaAs</strong> was studied. Stra<strong>in</strong>ed<br />

<strong>GaAs</strong> 0.86 P 0.14 /<strong>GaAs</strong> <strong>and</strong> <strong>GaAs</strong> 0.8 P 0.2 /<strong>GaAs</strong> 0.975 P 0.025 superlattices were annealed at<br />

temperatures of between 850 <strong>and</strong> 1100C, under various As vapor pressures. The diffusion<br />

coefficients were measured by means of secondary ion mass spectroscopy <strong>and</strong><br />

cathodolum<strong>in</strong>escence spectroscopy. It was found that the <strong>in</strong>terdiffusion coefficients were<br />

higher under As-rich conditions than under Ga-rich conditions; thus <strong>in</strong>dicat<strong>in</strong>g an<br />

<strong>in</strong>terstitial-substitutional diffusion mechanism.<br />

U.Egger, M.Schultz, P.Werner, O.Breitenste<strong>in</strong>, T.Y.Tan, U.Gösele, R.Franzheld,<br />

M.Uematsu, H.Ito: Journal of Applied Physics, 1997, 81[9], 6056-61<br />

[446-150/151-144]<br />

Ga(As,Sb)<br />

<strong>GaAs</strong>Sb/<strong>GaAs</strong>: Interdiffusion<br />

Interdiffusion on the group-V sub-lattice of <strong>GaAs</strong> was studied. Stra<strong>in</strong>ed<br />

<strong>GaAs</strong> 0.98 Sb 0.02 /<strong>GaAs</strong> superlattices were annealed at temperatures of between 850 <strong>and</strong><br />

1100C, under various As vapor pressures. The diffusion coefficients were measured by<br />

means of secondary ion mass spectroscopy <strong>and</strong> cathodolum<strong>in</strong>escence spectroscopy. It was<br />

found that the <strong>in</strong>terdiffusion coefficients were higher under As-rich conditions than under<br />

Ga-rich conditions; thus <strong>in</strong>dicat<strong>in</strong>g an <strong>in</strong>terstitial-substitutional diffusion mechanism.<br />

U.Egger, M.Schultz, P.Werner, O.Breitenste<strong>in</strong>, T.Y.Tan, U.Gösele, R.Franzheld,<br />

M.Uematsu, H.Ito: Journal of Applied Physics, 1997, 81[9], 6056-61<br />

[446-150/151-144]<br />

401


(Ga,In)As<br />

Al<br />

In<strong>GaAs</strong>/InAlAs: Al <strong>Diffusion</strong><br />

It was shown that implantation of O (a basically non-dopant impurity), after adequate<br />

high-temperature furnace anneal<strong>in</strong>g (750C, 1h), led to significant <strong>in</strong>terdiffusion of group-<br />

<strong>III</strong> atoms <strong>in</strong> molecular beam epitaxially grown In 0.53 Ga 0.47 As/In 0.52 Al 0.48 As multiquantum<br />

wells. Photolum<strong>in</strong>escence <strong>and</strong> Auger electron spectroscopic measurements<br />

(coupled with Ar + ion etch<strong>in</strong>g) were used to monitor the disorder<strong>in</strong>g of multi-quantum<br />

wells which were implanted with O (5 x 10 13 to 5 x 10 14 /cm 2 ) <strong>and</strong> then annealed by us<strong>in</strong>g<br />

either rapid thermal anneal<strong>in</strong>g or long-term furnace anneal<strong>in</strong>g. The role which was played<br />

by O <strong>in</strong> enhanc<strong>in</strong>g Al <strong>in</strong>terdiffusion was unambiguously established, <strong>and</strong> a tentative<br />

explanation for this was based upon the possible migration of O <strong>in</strong> these multi-quantum<br />

wells.<br />

E.V.K.Rao, P.Ossart, H.Thibierge, M.Quillec, P.Krauz: Applied Physics Letters, 1990,<br />

57[21], 2190-2<br />

[446-78/79-043]<br />

As<br />

In<strong>GaAs</strong>: As <strong>Diffusion</strong><br />

A theoretical model was proposed for the calculation of compositional variations <strong>in</strong> <strong>III</strong>-V<br />

ternary crystals dur<strong>in</strong>g growth. In this novel model, phase equilibrium between the crystal<br />

<strong>and</strong> the growth solution was ma<strong>in</strong>ta<strong>in</strong>ed; together with a simultaneous constancy of the<br />

transported <strong>and</strong> <strong>in</strong>corporated mass of solute atoms at the crystal/solution <strong>in</strong>terface. This<br />

model could be applied to the calculation of diffusion-limited growth <strong>in</strong> a temperaturegradient<br />

solution; as <strong>in</strong> the case of the source current-controlled growth method. The<br />

compositional variations of In<strong>GaAs</strong> crystals which were grown via diffusion were<br />

calculated by us<strong>in</strong>g this model. The <strong>in</strong>corporation of As via the crystal/solution <strong>in</strong>terface<br />

was considered on the basis of phase equilibrium laws <strong>and</strong> mass-constancy. Upon<br />

compar<strong>in</strong>g the experimental results with the calculated ones it was found that, <strong>in</strong> the In-Ga-<br />

As solution, the diffusion coefficient of Ga was about twice as great as that of As. The<br />

402


As (Ga,In)As Be<br />

calculated compositional variations showed that the compositional uniformity of In<strong>GaAs</strong><br />

crystals could be markedly improved by controll<strong>in</strong>g growth parameters such as the<br />

temperature gradient <strong>in</strong> the solution.<br />

K.Nakajima: Journal of Crystal Growth, 1991, 110[4], 781-94<br />

[446-81/82-039]<br />

In<strong>GaAs</strong>/In<strong>GaAs</strong>P: As <strong>Diffusion</strong><br />

A photolum<strong>in</strong>escence study was made of the <strong>in</strong>terdiffusion of As <strong>in</strong> the<br />

In 0.66 Ga 0.33 As/In 0.66 Ga 0.33 As 0.7 P 0.3 system at temperatures rang<strong>in</strong>g from 950 to 600C. It<br />

was shown that the diffusion was Fickian, with no dependence of the diffusion coefficient<br />

upon the substrate dopant-type or etch-pit density. In the case of Sn- or S-doped<br />

substrates, the diffusion could be described by:<br />

D(cm 2 /s) = 23 exp[-3.7(eV)/kT]<br />

at temperatures greater than 675C. This activation energy was the same as that deduced<br />

for group-<strong>III</strong> <strong>in</strong>terdiffusion <strong>in</strong> Ga 0.8 In 0.2 As/<strong>GaAs</strong>. At lower temperatures, diffusivity <strong>in</strong> the<br />

present system could be described by:<br />

D(cm 2 /s) = 5 x 10 -10 exp[-1.7(eV)/kT]<br />

W.P.Gill<strong>in</strong>, S.S.Rao, I.V.Bradley, K.P.Homewood, A.D.Smith, A.T.R.Briggs: Applied<br />

Physics Letters, 1993, 63[6], 797-9<br />

[446-106/107-116]<br />

Be<br />

333 In<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

The diffusion mechanisms which operated dur<strong>in</strong>g post-growth anneal<strong>in</strong>g were<br />

<strong>in</strong>vestigated <strong>in</strong> p-type Be-doped epitaxial layers which had been grown between 2<br />

undoped In<strong>GaAs</strong> layers. Anneal<strong>in</strong>g (30 to 180s, 500 to 800C) was applied to samples<br />

with dopant concentrations of 5 x 10 18 , 10 19 or 3 x 10 19 /cm 3 . It was found that no<br />

appreciable Be diffusion occurred dur<strong>in</strong>g post-growth rapid thermal anneal<strong>in</strong>g, for all of<br />

the dopant levels, when anneal<strong>in</strong>g was carried out at 500 or 600C for times of less than<br />

60s. The same was true, for a Be concentration of 5 x 10 18 /cm 3 , dur<strong>in</strong>g anneal<strong>in</strong>g at 700C<br />

for 30s. At a dop<strong>in</strong>g level of 3 x 10 19 /cm 3 , significant Be diffusion occurred dur<strong>in</strong>g<br />

anneal<strong>in</strong>g (>700C, 60s). The resultant curves exhibited a concave k<strong>in</strong>k region. It was<br />

deduced that the effective Be diffusion coefficient (table 33) was approximately constant<br />

<strong>in</strong> one part of the concentration profile, <strong>and</strong> was proportional to the square root of the<br />

concentration <strong>in</strong> an<strong>other</strong> part. By assum<strong>in</strong>g the latter dependence, an effective diffusivity<br />

could be substituted <strong>in</strong>to Fick’s second law. The resultant differential equation was<br />

solved numerically by us<strong>in</strong>g an explicit f<strong>in</strong>ite difference method. Good agreement was<br />

obta<strong>in</strong>ed between the measured depth profiles <strong>and</strong> the simulated distributions.<br />

S.Koumetz, J.Marcon, K.Ketata, M.Ketata, P.Launay: Journal of Physics D, 1997, 30[5],<br />

757-62<br />

[446-148/149-181]<br />

403


Be (Ga,In)As Be<br />

Table 33<br />

Diffusivity of Be <strong>in</strong> In<strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

800 1.7 x 10 -12<br />

700 3.4 x 10 -13<br />

In<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

The diffusion of Be dur<strong>in</strong>g post-growth anneal<strong>in</strong>g was studied <strong>in</strong> epitaxial layers. Two<br />

models were proposed <strong>in</strong> order to expla<strong>in</strong> the observed concentration profiles. In the first<br />

model, the Boltzmann-Matano method was used while tak<strong>in</strong>g account of the diffusivity<br />

time-<strong>in</strong>dependence. An observed double profile was expla<strong>in</strong>ed <strong>in</strong> terms of a change <strong>in</strong><br />

diffusivity. In a second model, vacancy equilibrium was not assumed, <strong>and</strong> k<strong>in</strong>etic terms<br />

which were related to vacancy production had to be <strong>in</strong>cluded <strong>in</strong> the diffusion equations.<br />

The observed double profile was then expla<strong>in</strong>ed <strong>in</strong> terms of a reduction <strong>in</strong> the vacancy<br />

concentration <strong>in</strong> the crystal bulk. By us<strong>in</strong>g these two approaches, it was shown that good<br />

agreement between the experimental <strong>and</strong> simulated profiles could be obta<strong>in</strong>ed. In the case<br />

of the non-equilibrium model, the vacancy concentration was close to its equilibrium<br />

value. It was therefore possible to assume that Be diffusion <strong>in</strong> In<strong>GaAs</strong> epitaxial layers<br />

occurred with quasi-equilibrium po<strong>in</strong>t defect concentrations. Under these conditions, the<br />

diffusion depth varied as the square root of the diffusion time. There was no significant<br />

under-saturation of vacancies or super-saturation of self-<strong>in</strong>terstitials at the diffusion front<br />

which could change the normalized Be diffusion depth <strong>in</strong> In<strong>GaAs</strong> epitaxial layers.<br />

J.Marcon, S.Gautier, S.Koumetz, K.Ketata, M.Ketata, P.Launay: Solid State<br />

Communications, 1997, 101[3], 159-62<br />

[446-141/142-110]<br />

In<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

The diffusion of Be from In<strong>GaAs</strong> epitaxial layers that had been grown between undoped<br />

In<strong>GaAs</strong> layers was <strong>in</strong>vestigated dur<strong>in</strong>g post-growth anneal<strong>in</strong>g. A general substitutional<strong>in</strong>terstitial<br />

diffusion mechanism was developed <strong>in</strong> order to expla<strong>in</strong> the concentration<br />

profiles which were observed. The possibility of a concentration-dependent diffusivity<br />

was also considered <strong>in</strong> order to improve the fit to Be diffusion profiles.<br />

S.Koumetz, J.Marcon, K.Ketata, M.Ketata, F.Lefebvre, P.Mart<strong>in</strong>, P.Launay: Materials<br />

Science <strong>and</strong> Eng<strong>in</strong>eer<strong>in</strong>g B, 1996, 37[1-3], 208-11<br />

[446-140-007]<br />

In<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

The behavior of implanted Be + ions was <strong>in</strong>vestigated dur<strong>in</strong>g rapid thermal anneal<strong>in</strong>g at<br />

temperatures of between 600 <strong>and</strong> 900C. It was found that the apparent activation energy<br />

for Be was equal to 0.38eV. Higher activation efficiencies were found for the dopant <strong>in</strong><br />

In<strong>GaAs</strong>, as compared with InAlAs. Anomalously low activation was detected for low-<br />

404


Be (Ga,In)As Be<br />

dose Be implants. The latter effect was attributed to a lack of vacant sites for the Be<br />

atoms to occupy. Extensive redistribution of the Be was observed after anneal<strong>in</strong>g (750C,<br />

10s).<br />

E.Hailemariam, S.J.<strong>Pearton</strong>, W.S.Hobson, H.S.Luftman, A.P.Perley: Journal of Applied<br />

Physics, 1992, 71[1], 215-20<br />

[446-86/87-038]<br />

In<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

The occurrence of Be diffusion dur<strong>in</strong>g post-growth anneal<strong>in</strong>g was studied <strong>in</strong> epitaxial<br />

In<strong>GaAs</strong> layers which were grown between 2 undoped In<strong>GaAs</strong> layers. In order to expla<strong>in</strong><br />

the observed concentration profiles <strong>and</strong> related diffusion mechanisms, a general<br />

substitutional-<strong>in</strong>terstitial model was proposed. On one h<strong>and</strong>, simultaneous diffusion by<br />

dissociative <strong>and</strong> kick-out mechanisms was suggested <strong>and</strong>, on the <strong>other</strong> h<strong>and</strong>, the Fermilevel<br />

effect was used to expla<strong>in</strong> changes <strong>in</strong> the effective diffusion coefficient of Be<br />

species as a function of concentration. A concentration-dependent diffusivity was also<br />

used to obta<strong>in</strong> an improved fit to Be diffusion profiles.<br />

S.Koumetz, J.Marcon, K.Ketata, M.Ketata, C.Dubon-Chevallier, P.Launay,<br />

J.L.Benchimol: Applied Physics Letters, 1995, 67[15], 2161-3<br />

[446-125/126-140]<br />

GaInAs: Be <strong>Diffusion</strong><br />

The diffusion of Be from buried Be-doped layers was studied at temperatures of between<br />

600 <strong>and</strong> 700C. Four types of Be diffusion profile were identified. An <strong>in</strong>terstitial cum<br />

substitutional model was proposed to be the diffusion mechanism, which depended upon<br />

the growth conditions. The values of the effective Be diffusion coefficient, for dopant<br />

levels which were above <strong>and</strong> below about 5 x 10 16 /cm 3 , were found to be 1.9 x 10 -15 <strong>and</strong><br />

9 x 10 -16 cm 2 /s, respectively.<br />

E.G.Scott, D.Wake, G.D.T.Spiller, G.J.Davies: Journal of Applied Physics, 1989, 66[11],<br />

5344-8<br />

[446-74-030]<br />

In<strong>GaAs</strong>/<strong>GaAs</strong>: Be <strong>Diffusion</strong><br />

The suppression, of Be out-diffusion from a Be-doped <strong>GaAs</strong> layer, by stra<strong>in</strong>ed In<strong>GaAs</strong><br />

layers was studied by us<strong>in</strong>g secondary ion mass spectroscopy. The experimental<br />

structures consisted of an 80nm Be-doped (about 10 19 /cm 3 ) <strong>GaAs</strong> layer that was<br />

s<strong>and</strong>wiched between 8nm In x Ga 1-x As layers, where x was equal to 0, 0.1, or 0.25. The<br />

samples were subjected to rapid thermal anneal<strong>in</strong>g (750C, 360s), <strong>and</strong> it was clearly<br />

observed that Be diffusion beyond the In<strong>GaAs</strong> layers was most rapid for a structure with<br />

x = 0, <strong>and</strong> was slowest for a structure with x = 0.25.<br />

K.Zhang, Y.C.Chen, J.S<strong>in</strong>gh, P.Bhattacharya: Applied Physics Letters, 1994, 65[7], 872-<br />

4<br />

[446-119/120-214]<br />

405


Be (Ga,In)As Cd<br />

In<strong>GaAs</strong>/InP: Be <strong>Diffusion</strong><br />

Heterojunctions which were grown by us<strong>in</strong>g liquid-phase epitaxy were <strong>in</strong>vestigated by<br />

us<strong>in</strong>g I-V <strong>and</strong> I o (T) measurements. It was shown that the <strong>in</strong>jection of electrons across the<br />

hetero-<strong>in</strong>terface was described well by a thermionic emission model which <strong>in</strong>volved a<br />

barrier height that was strongly affected by p-type dopant diffusion. Heterojunction<br />

bipolar transistors with Schottky collectors were used as tools to measure the <strong>in</strong>jection<br />

current <strong>in</strong> as-grown p + -In 0.53 Ga 0.47 As/n-InP diodes, <strong>and</strong> the results were compared with<br />

the predictions of the thermionic emission model. In this way, the relative diffusivities of<br />

the three p-type dopants <strong>in</strong> InP at 600C were evaluated. The results could be summarized<br />

by:<br />

D Be /D Mn ˜ 20, D Mg /D Be ˜ 1<br />

P.S.Whitney, J.C.Vicek, C.G.Fonstad: Journal of Applied Physics, 1987, 62[5], 1920-4<br />

[446-55/56-028]<br />

Cd<br />

In<strong>GaAs</strong>: Cd <strong>Diffusion</strong><br />

The diffusion of Cd <strong>in</strong> In 0.53 Ga 0.47 As at 600C was studied by us<strong>in</strong>g a closed-ampoule<br />

technique, secondary ion mass spectrometry, <strong>and</strong> Hall effect measurements. It was found<br />

that the diffusivity could be described by an activation energy of 2.6eV. In some cases, a<br />

slower second diffusion front was found as well as a first steep junction.<br />

P.Ambreé, B.Gruska: Crystal Research <strong>and</strong> Technology, 1989, 24[3], 299-305<br />

[446-70/71-118]<br />

In<strong>GaAs</strong>: Cd <strong>Diffusion</strong><br />

The Cd was diffused <strong>in</strong>to In<strong>GaAs</strong> by us<strong>in</strong>g Cd 3 P 2 plus P or Cd 3 P 2 plus Cd 3 As 2 as<br />

diffusion sources. Two diffusion fronts were observed. The diffusion characteristics of<br />

Cd 3 P 2 plus P sources were expla<strong>in</strong>ed <strong>in</strong> terms of the <strong>in</strong>terstitial-substitutional model or<br />

the vacancy complex model. The charge state of the diffus<strong>in</strong>g <strong>in</strong>terstitial Cd atom was a<br />

s<strong>in</strong>gly ionized donor. Gaseous Cd orig<strong>in</strong>ated from solid-phase CdP 2 . In the case of Cd 3 P 2<br />

plus Cd 3 As 2 diffusion sources, the effective diffusion coefficient <strong>and</strong> the surface acceptor<br />

concentration decreased with <strong>in</strong>creas<strong>in</strong>g weight fraction of Cd 3 As 2 . The relative depth of<br />

the deeper diffusion front <strong>in</strong>creased when the supply of vacancies was suppressed.<br />

K.I.Ohtsuka, T.Matsui, H.Ogata: Japanese Journal of Applied Physics, 1988, 27[2], 253-9<br />

[446-60-009]<br />

In<strong>GaAs</strong>: Cd <strong>Diffusion</strong><br />

A new source for Cd diffusion <strong>in</strong>to In 0.53 Ga 0.47 As was developed <strong>in</strong> which Langmuir-<br />

Blodgett deposited monolayers of cadmium arachidate were used as the Cd source.<br />

D.M.Shah, W.K.Chan, R.Bhat, H.M.Cox, N.E.Schlotter, C.C.Chang: Applied Physics<br />

Letters, 1990, 56[21], 2132-4<br />

[446-76/77-023]<br />

406


Cd (Ga,In)As Ga<br />

In<strong>GaAs</strong>/InP: Cd <strong>Diffusion</strong><br />

A closed-ampoule technique was used to study the simultaneous diffusion of Cd <strong>and</strong> Zn<br />

<strong>in</strong>to material which was lattice-matched to InP. It was found that the Cd atoms, whose<br />

<strong>in</strong>terstitial diffusion coefficient was small when compared with that of Zn <strong>in</strong>terstitials,<br />

penetrated <strong>in</strong>to the crystal as deeply as the Zn atoms. These data agreed with the results<br />

of numerical simulations of simultaneous diffusion, <strong>in</strong> <strong>III</strong>-V compounds, which were<br />

based upon an <strong>in</strong>terstitial-substitutional diffusion mechanism.<br />

U.Wielsch, P.Ambrée, B.Gruska: Semiconductor Science <strong>and</strong> Technology, 1990, 5[9],<br />

923-7<br />

[446-76/77-024]<br />

In<strong>GaAs</strong>/InP: Cd <strong>Diffusion</strong><br />

Results on diffusion across In<strong>GaAs</strong>/InP <strong>and</strong> InP/In<strong>GaAs</strong> hetero-<strong>in</strong>terfaces were described.<br />

<strong>Diffusion</strong> from an InP top layer, Cd diffusion, or simple anneal<strong>in</strong>g of the samples, had no<br />

measurable effect upon the stability of the <strong>in</strong>terfaces. The marked <strong>in</strong>terdiffusion of In <strong>and</strong><br />

Ga host atoms, as well as Zn getter<strong>in</strong>g at the <strong>in</strong>terface, were analyzed <strong>in</strong> terms of kick-out<br />

<strong>and</strong> vacancy mechanisms.<br />

P.Ambrée, A.Hangleiter, M.H.Pilkuhn, K.W<strong>and</strong>el: Applied Physics Letters, 1990, 56[10],<br />

931-3<br />

[446-74-039]<br />

D<br />

In<strong>GaAs</strong>/Al<strong>GaAs</strong>: D <strong>Diffusion</strong><br />

The effects of monatomic D diffusion <strong>in</strong> quantum wells were studied by us<strong>in</strong>g<br />

photolum<strong>in</strong>escence <strong>and</strong> secondary ion mass spectroscopic methods. The multiple<br />

quantum well structures were grown by means of molecular beam epitaxy <strong>and</strong> were<br />

hydrogenated with a remote plasma. A significant <strong>in</strong>crease <strong>in</strong> the 77K photolum<strong>in</strong>escence<br />

<strong>in</strong>tegrated <strong>in</strong>tensity of bound excitons was observed after hydrogenation. This was<br />

attributed to the passivation of non-radiative recomb<strong>in</strong>ation centers with<strong>in</strong> the quantum<br />

wells. The studies demonstrated that there was an <strong>in</strong>crease <strong>in</strong> passivation efficiency with<br />

<strong>in</strong>creas<strong>in</strong>g Al concentration <strong>in</strong> the barriers, <strong>and</strong> that the hydrogenation was stable to<br />

temperatures above 450C. Overall, the results all strongly suggested that the passivated<br />

non-radiative recomb<strong>in</strong>ation centers were <strong>in</strong>terface defects.<br />

S.M.Lord, G.Roos, J.S.Harris, N.M.Johnson: Journal of Applied Physics, 1993, 73[2],<br />

740-8<br />

[446-106/107-113]<br />

Ga<br />

In<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

A theoretical model was proposed for the calculation of compositional variations <strong>in</strong> <strong>III</strong>-V<br />

ternary crystals dur<strong>in</strong>g growth. In this novel model, phase equilibrium between the crystal<br />

407


Ga (Ga,In)As Ga<br />

<strong>and</strong> the growth solution was ma<strong>in</strong>ta<strong>in</strong>ed; together with a simultaneous constancy of the<br />

transported <strong>and</strong> <strong>in</strong>corporated mass of solute atoms at the crystal/solution <strong>in</strong>terface. This<br />

model could be applied to the calculation of diffusion-limited growth <strong>in</strong> a temperaturegradient<br />

solution; as <strong>in</strong> the case of the source current-controlled growth method. The<br />

compositional variations of In<strong>GaAs</strong> crystals which were grown via diffusion were<br />

calculated by us<strong>in</strong>g this model. The <strong>in</strong>corporation of Ga via the crystal/solution <strong>in</strong>terface<br />

was considered on the basis of phase equilibrium laws <strong>and</strong> mass constancy. Upon<br />

compar<strong>in</strong>g the experimental results with the calculated ones it was found that, <strong>in</strong> the In-<br />

Ga-As solution, the diffusion coefficient of Ga was about twice as great as that of As.<br />

The calculated compositional variations showed that the compositional uniformity of<br />

In<strong>GaAs</strong> crystals could be markedly improved by controll<strong>in</strong>g growth parameters such as<br />

the temperature gradient <strong>in</strong> the solution.<br />

K.Nakajima: Journal of Crystal Growth, 1991, 110[4], 781-94<br />

[446-81/82-039]<br />

GaInAs/<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

A stra<strong>in</strong>ed s<strong>in</strong>gle quantum well of <strong>GaAs</strong>/Ga 0.77 In 0.23 Ga/<strong>GaAs</strong> was grown, via lowpressure<br />

metal-organic vapor phase epitaxy, onto a (100)<strong>GaAs</strong> substrate at 625C. Samples<br />

were annealed under AsH 3 /H 2 at temperatures rang<strong>in</strong>g from 750 to 900C. S<strong>in</strong>ce the<br />

quantum well thickness of about 8nm was below the critical value for this latticemismatched<br />

system, it was assumed that the GaInAs layer was commensurate with the<br />

<strong>GaAs</strong> substrate. The low-temperature (2K) photolum<strong>in</strong>escence of the electron to heavy<br />

hole transition <strong>in</strong> the quantum well of these samples was analyzed <strong>in</strong> order to study In/Ga<br />

<strong>in</strong>terdiffusion at the GaInAs/<strong>GaAs</strong> <strong>in</strong>terfaces. The energies of the photolum<strong>in</strong>escence<br />

peaks shifted to higher values dur<strong>in</strong>g anneal<strong>in</strong>g. These shifts were quantitatively<br />

<strong>in</strong>terpreted <strong>in</strong> terms of changes <strong>in</strong> the quantum well profile, due to In <strong>and</strong> Ga<br />

<strong>in</strong>terdiffusion. The <strong>in</strong>terdiffusion coefficient at 850C was deduced to be 3 x 10 -17 cm 2 /s;<br />

with an activation energy of 2.07eV. The values which were obta<strong>in</strong>ed for the In/Ga<br />

<strong>in</strong>terdiffusion coefficient were larger than the published values for Al <strong>and</strong> Ga<br />

<strong>in</strong>terdiffusion <strong>in</strong> Al<strong>GaAs</strong>/<strong>GaAs</strong> heterojunctions.<br />

F.Iikawa, P.Motisuke, J.A.Brum, M.A.Sacilotti, A.P.Roth, R.A.Masut: Journal of Crystal<br />

Growth, 1988, 93, 336-41<br />

[446-64/65-173]<br />

334 GaInAs/<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

The enhancement of Ga <strong>in</strong>terdiffusion, due to Zn diffusion, was studied <strong>in</strong> stra<strong>in</strong>ed<br />

In x Ga 1-x As/<strong>GaAs</strong> s<strong>in</strong>gle quantum well structures, where x was between 0.20 <strong>and</strong> 0.24.<br />

The structures were grown by means of metalorganic vapor base epitaxy, <strong>and</strong> consisted of<br />

a 1000 to 2000nm-thick <strong>GaAs</strong> buffer layer plus an 8 to 10nm GaInAs layer, <strong>and</strong> a 50 to<br />

100nm <strong>GaAs</strong> cap layer. Shallow Zn <strong>in</strong>-diffusion was carried out at temperatures of<br />

between 585 <strong>and</strong> 620C, <strong>in</strong> order to vary the Zn content of the quantum wells. The<br />

samples were then annealed at temperatures of between 650 <strong>and</strong> 785C (<strong>in</strong> AsH 3 -H 2<br />

mixtures) <strong>in</strong> order to determ<strong>in</strong>e the enhanced In-Ga <strong>in</strong>terdiffusion coefficient. The results<br />

for a typical case (table 34) could be described by:<br />

408


Ga (Ga,In)As Ga<br />

D(cm 2 /s) = 3 x 10 -6 exp[-2.33(eV)/kT]<br />

The <strong>in</strong>terdiffusion enhancement was expla<strong>in</strong>ed <strong>in</strong> terms of an <strong>in</strong>terstitial migration<br />

process.<br />

M.T.Furtado, E.A.Sato, M.A.Sacilotti: Superlattices <strong>and</strong> Microstructures, 1991, 10[2],<br />

225-30<br />

[446-88/89-042]<br />

Table 34<br />

Interdiffusivity (In-Ga) <strong>in</strong> In<strong>GaAs</strong>/<strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

785 2.2 x 10 -17<br />

750 1.1 x 10 -17<br />

700 2.5 x 10 -18<br />

650 6.1 x 10 -19<br />

In<strong>GaAs</strong>/<strong>GaAs</strong>/Al<strong>GaAs</strong>: Ga <strong>Diffusion</strong><br />

The <strong>in</strong>terdiffusion of In <strong>and</strong> Ga at an In<strong>GaAs</strong>/<strong>GaAs</strong> <strong>in</strong>terface was studied. The<br />

<strong>in</strong>terdiffusion coefficients <strong>and</strong> activation energies were determ<strong>in</strong>ed by correlat<strong>in</strong>g shifts <strong>in</strong><br />

photolum<strong>in</strong>escence peaks with calculated quantum well transition energies that were<br />

based upon an erf concentration profile. The results <strong>in</strong>dicated that a higher x-value, <strong>in</strong><br />

In x Ga 1-x As s<strong>in</strong>gle quantum wells, led to a higher <strong>in</strong>terdiffusion coefficient for Ga under<br />

As over-pressure anneal<strong>in</strong>g conditions. Moreover, an <strong>in</strong>crease <strong>in</strong> the As over-pressure<br />

<strong>in</strong>creased the tendency to <strong>in</strong>terdiffusion, whereas a Ga over-pressure reduced<br />

<strong>in</strong>terdiffusion. The thermal activation energies for x-values of 0.057, 0.1 or 0.15 ranged<br />

from 3.3 to 2.6eV under an As over-pressure, <strong>and</strong> from 3 to 2.23eV under a Ga overpressure.<br />

K.Y.Hsieh, Y.L.Hwang, J.H.Lee, R.M.Kolbas: Journal of Electronic Materials, 1990,<br />

19[12], 1417-23<br />

[446-84/85-052]<br />

In<strong>GaAs</strong>/InAlAs: Ga <strong>Diffusion</strong><br />

It was shown that implantation of O (a basically non-dopant impurity), after adequate hightemperature<br />

furnace anneal<strong>in</strong>g (750C, 1h), led to significant <strong>in</strong>terdiffusion of group-<strong>III</strong><br />

atoms <strong>in</strong> molecular beam epitaxially grown In 0.53 Ga 0.47 As-In 0.52 Al 0.48 As multi-quantum<br />

wells. Photolum<strong>in</strong>escence <strong>and</strong> Auger electron spectroscopic measurements (coupled with<br />

Ar + ion etch<strong>in</strong>g) were used to monitor the disorder<strong>in</strong>g of multi-quantum wells which were<br />

implanted with O (5 x 10 13 to 5 x 10 14 /cm 2 ) <strong>and</strong> then annealed by us<strong>in</strong>g either rapid<br />

thermal anneal<strong>in</strong>g or long-term furnace anneal<strong>in</strong>g. The role which was played by O <strong>in</strong><br />

409


Ga (Ga,In)As H<br />

enhanc<strong>in</strong>g Ga <strong>in</strong>terdiffusion was unambiguously established, <strong>and</strong> a tentative explanation<br />

for this was based upon the possible migration of O <strong>in</strong> these multi-quantum wells.<br />

E.V.K.Rao, P.Ossart, H.Thibierge, M.Quillec, P.Krauz: Applied Physics Letters, 1990,<br />

57[21], 2190-2<br />

[446-78/79-043]<br />

Table 35<br />

Diffusivity of H <strong>in</strong> In<strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

605 2.1 x 10 -9<br />

550 7.9 x 10 -10<br />

500 2.7 x 10 -10<br />

In<strong>GaAs</strong>/InP: Ga <strong>Diffusion</strong><br />

Results on diffusion across In<strong>GaAs</strong>/InP <strong>and</strong> InP/In<strong>GaAs</strong> hetero-<strong>in</strong>terfaces were described.<br />

Marked <strong>in</strong>terdiffusion of In <strong>and</strong> Ga host atoms, as well as Zn getter<strong>in</strong>g at the <strong>in</strong>terface,<br />

were analyzed <strong>in</strong> terms of kick-out <strong>and</strong> vacancy mechanisms. The activation energy for<br />

Zn-stimulated Ga <strong>in</strong>terdiffusion across the In<strong>GaAs</strong>/InP heterojunction was estimated to be<br />

3.8eV.<br />

P.Ambrée, A.Hangleiter, M.H.Pilkuhn, K.W<strong>and</strong>el: Applied Physics Letters, 1990, 56[10],<br />

931-3<br />

[446-74-039]<br />

H<br />

In<strong>GaAs</strong>: H <strong>Diffusion</strong><br />

The migration of H was studied <strong>in</strong> In 0.53 Ga 0.47 As. The H diffusivity was much higher <strong>in</strong><br />

p-type than <strong>in</strong> n-type samples. It was concluded that H was a deep donor <strong>in</strong> this material.<br />

E.M.Omeljanovsky, A.V.Pakhomov, A.Y.Polyakov, O.M.Borod<strong>in</strong>a, E.A.Kozhukhova,<br />

A.Y.Nashelskii, S.V.Yakobson, V.V.Novikova: Solid State Communications, 1989,<br />

72[5], 409-11<br />

[446-72/73-035]<br />

335 In<strong>GaAs</strong>: H <strong>Diffusion</strong><br />

The diffusion length of H <strong>in</strong> C-doped material was estimated from resistance variations<br />

which occurred <strong>in</strong> samples with n-type cap layers dur<strong>in</strong>g anneal<strong>in</strong>g <strong>in</strong> an N 2 ambient. It<br />

was found that the resultant diffusion lengths were proportional to the square root of the<br />

anneal<strong>in</strong>g time; as <strong>in</strong> a normal diffusion process. The activation energy for H diffusion<br />

was estimated to be 1.2eV (table 35). The effective diffusion coefficients were smaller<br />

410


H (Ga,In)As In<br />

than those <strong>in</strong> <strong>GaAs</strong>, thus imply<strong>in</strong>g that group-<strong>III</strong> atoms strongly affected H diffusion <strong>in</strong> C-<br />

doped compound semiconductors.<br />

H.Ito: Japanese Journal of Applied Physics, 1996, 35[2-9B], L1155-7<br />

[446-141/142-110]<br />

In<strong>GaAs</strong>/Al<strong>GaAs</strong>: H <strong>Diffusion</strong><br />

The effects of monatomic H diffusion <strong>in</strong> quantum wells were studied by us<strong>in</strong>g<br />

photolum<strong>in</strong>escence <strong>and</strong> secondary ion mass spectroscopic methods. The multiple<br />

quantum well structures were grown by means of molecular beam epitaxy <strong>and</strong> were<br />

hydrogenated with a remote plasma. A significant <strong>in</strong>crease <strong>in</strong> the 77K photolum<strong>in</strong>escence<br />

<strong>in</strong>tegrated <strong>in</strong>tensity of bound excitons was observed after hydrogenation. This was<br />

attributed to the passivation of non-radiative recomb<strong>in</strong>ation centers with<strong>in</strong> the quantum<br />

wells. The studies demonstrated that there was an <strong>in</strong>crease <strong>in</strong> passivation efficiency with<br />

<strong>in</strong>creas<strong>in</strong>g Al concentration <strong>in</strong> the barriers, <strong>and</strong> that the hydrogenation was stable to<br />

temperatures above 450C. Overall, the results all strongly suggested that the passivated<br />

non-radiative recomb<strong>in</strong>ation centers were <strong>in</strong>terface defects.<br />

S.M.Lord, G.Roos, J.S.Harris, N.M.Johnson: Journal of Applied Physics, 1993, 73[2],<br />

740-8<br />

[446-106/107-113]<br />

In<br />

GaInAs/<strong>GaAs</strong>: In <strong>Diffusion</strong><br />

The enhancement of In <strong>in</strong>terdiffusion, due to Zn diffusion, was studied <strong>in</strong> stra<strong>in</strong>ed In x<br />

Ga 1-x As/<strong>GaAs</strong> s<strong>in</strong>gle quantum well structures, where x was between 0.20 <strong>and</strong> 0.24. The<br />

structures were grown by means of metalorganic vapor base epitaxy, <strong>and</strong> consisted of a<br />

1000 to 2000nm-thick <strong>GaAs</strong> buffer layer plus an 8 to 10nm GaInAs layer, <strong>and</strong> a 50 to<br />

100nm <strong>GaAs</strong> cap layer. Shallow Zn <strong>in</strong>-diffusion was carried out at temperatures of<br />

between 585 <strong>and</strong> 620C, <strong>in</strong> order to vary the Zn content of the quantum wells. The<br />

samples were then annealed at temperatures of between 650 <strong>and</strong> 785C (<strong>in</strong> AsH 3 -H 2<br />

mixtures) <strong>in</strong> order to determ<strong>in</strong>e the enhanced In-Ga <strong>in</strong>terdiffusion coefficient. The results<br />

for a typical case could be described by:<br />

D(cm 2 /s) = 3 x 10 -6 exp[-2.33(eV)/kT]<br />

The <strong>in</strong>terdiffusion enhancement was expla<strong>in</strong>ed <strong>in</strong> terms of an <strong>in</strong>terstitial migration<br />

process.<br />

M.T.Furtado, E.A.Sato, M.A.Sacilotti: Superlattices <strong>and</strong> Microstructures, 1991, 10[2],<br />

225-30<br />

[446-88/89-042]<br />

GaInAs/<strong>GaAs</strong>: In <strong>Diffusion</strong><br />

A stra<strong>in</strong>ed s<strong>in</strong>gle quantum well of <strong>GaAs</strong>/Ga 0.77 In 0.23 Ga/<strong>GaAs</strong> was grown, via low-pressure<br />

metal-organic vapor phase epitaxy, onto a (100)<strong>GaAs</strong> substrate at 625C. Samples were<br />

annealed under AsH 3 /H 2 at temperatures rang<strong>in</strong>g from 750 to 900C. S<strong>in</strong>ce the quantum<br />

411


In (Ga,In)As In<br />

well thickness of about 8nm was below the critical value for this lattice-mismatched<br />

system, it was assumed that the GaInAs layer was commensurate with the <strong>GaAs</strong> substrate.<br />

The low-temperature (2K) photolum<strong>in</strong>escence of the electron to heavy hole transition <strong>in</strong><br />

the quantum well of these samples was analyzed <strong>in</strong> order to study In/Ga <strong>in</strong>terdiffusion at<br />

the GaInAs/<strong>GaAs</strong> <strong>in</strong>terfaces. The energies of the photolum<strong>in</strong>escence peaks shifted to<br />

higher values dur<strong>in</strong>g anneal<strong>in</strong>g. These shifts were quantitatively <strong>in</strong>terpreted <strong>in</strong> terms of<br />

changes <strong>in</strong> the quantum well profile, due to In <strong>and</strong> Ga <strong>in</strong>terdiffusion. The <strong>in</strong>terdiffusion<br />

coefficient at 850C was deduced to be 3 x 10 -17 cm 2 /s; with an activation energy of<br />

2.07eV. The values which were obta<strong>in</strong>ed for the In/Ga <strong>in</strong>terdiffusion coefficient were<br />

larger than the published values for Al <strong>and</strong> Ga <strong>in</strong>terdiffusion <strong>in</strong> Al<strong>GaAs</strong>/<strong>GaAs</strong><br />

heterojunctions.<br />

F.Iikawa, P.Motisuke, J.A.Brum, M.A.Sacilotti, A.P.Roth, R.A.Masut: Journal of Crystal<br />

Growth, 1988, 93, 336-41<br />

[446-64/65-173]<br />

In<strong>GaAs</strong>/<strong>GaAs</strong>/Al<strong>GaAs</strong>: In <strong>Diffusion</strong><br />

The <strong>in</strong>terdiffusion of In <strong>and</strong> Ga at an In<strong>GaAs</strong>/<strong>GaAs</strong> <strong>in</strong>terface was studied. The<br />

<strong>in</strong>terdiffusion coefficients <strong>and</strong> activation energies were determ<strong>in</strong>ed by correlat<strong>in</strong>g shifts <strong>in</strong><br />

photolum<strong>in</strong>escence peaks with calculated quantum well transition energies that were<br />

based upon an erf concentration profile. The results <strong>in</strong>dicated that a higher x-value, <strong>in</strong><br />

In x Ga 1-x As s<strong>in</strong>gle quantum wells, led to a higher <strong>in</strong>terdiffusion coefficient for In under As<br />

over-pressure anneal<strong>in</strong>g conditions. Moreover, an <strong>in</strong>crease <strong>in</strong> the As over-pressure<br />

<strong>in</strong>creased the tendency to <strong>in</strong>terdiffusion, whereas a Ga over-pressure reduced<br />

<strong>in</strong>terdiffusion. The thermal activation energies for x-values of 0.057, 0.1 or 0.15 ranged<br />

from 3.3 to 2.6eV under an As over-pressure, <strong>and</strong> from 3 to 2.23eV under a Ga overpressure.<br />

K.Y.Hsieh, Y.L.Hwang, J.H.Lee, R.M.Kolbas: Journal of Electronic Materials, 1990,<br />

19[12], 1417-23<br />

[446-84/85-052]<br />

In<strong>GaAs</strong>/InAlAs: In <strong>Diffusion</strong><br />

The distribution of group-<strong>III</strong> metals at In 0.53 Ga 0.47 As/In 0.52 Al 0.48 As <strong>in</strong>terfaces, before <strong>and</strong><br />

after anneal<strong>in</strong>g at 1085K, were measured. Little evidence for Al <strong>in</strong>terdiffusion was found,<br />

but the Ga concentration profiles exhibited some broaden<strong>in</strong>g after anneal<strong>in</strong>g. The almost<br />

constant orig<strong>in</strong>al In profiles developed strong modulations; with near-discont<strong>in</strong>uities at<br />

the <strong>in</strong>itial <strong>in</strong>terfaces. This behavior was expla<strong>in</strong>ed <strong>in</strong> terms of In diffusion <strong>in</strong> the chemical<br />

potential gradient which was established by the disparity <strong>in</strong> Al <strong>and</strong> Ga mobilities <strong>and</strong> by<br />

the requirement for <strong>III</strong>-V stoichiometry <strong>in</strong> the alloys.<br />

R.J.Baird, T.J.Potter, G.P.Kothiyal, P.K.Bhattacharya: Applied Physics Letters, 1988,<br />

52[24], 2055-7<br />

[446-62/63-223]<br />

412


In (Ga,In)As Mn<br />

In<strong>GaAs</strong>/InAlAs: In <strong>Diffusion</strong><br />

It was shown that implantation of O (a basically non-dopant impurity), after adequate<br />

high-temperature furnace anneal<strong>in</strong>g (750C, 1h), led to significant <strong>in</strong>terdiffusion of group-<br />

<strong>III</strong> atoms <strong>in</strong> molecular beam epitaxially grown In 0.53 Ga 0.47 As-In 0.52 Al 0.48 As multiquantum<br />

wells. Photolum<strong>in</strong>escence <strong>and</strong> Auger electron spectroscopic measurements<br />

(coupled with Ar + ion etch<strong>in</strong>g) were used to monitor the disorder<strong>in</strong>g of multi-quantum<br />

wells which were implanted with O (5 x 10 13 to 5 x 10 14 /cm 2 ) <strong>and</strong> then annealed by us<strong>in</strong>g<br />

either rapid thermal anneal<strong>in</strong>g or long-term furnace anneal<strong>in</strong>g. The role which was played<br />

by O <strong>in</strong> enhanc<strong>in</strong>g In <strong>in</strong>terdiffusion was unambiguously established, <strong>and</strong> a tentative<br />

explanation for this was based upon the possible migration of O <strong>in</strong> these multi-quantum<br />

wells.<br />

E.V.K.Rao, P.Ossart, H.Thibierge, M.Quillec, P.Krauz: Applied Physics Letters, 1990,<br />

57[21], 2190-2<br />

[446-78/79-043]<br />

Mg<br />

In<strong>GaAs</strong>/InP: Mg <strong>Diffusion</strong><br />

Heterojunctions which were grown by us<strong>in</strong>g liquid-phase epitaxy were <strong>in</strong>vestigated by<br />

us<strong>in</strong>g I-V <strong>and</strong> I o (T) measurements. It was shown that the <strong>in</strong>jection of electrons across the<br />

hetero-<strong>in</strong>terface was described well by a thermionic emission model which <strong>in</strong>volved a<br />

barrier height that was strongly affected by p-type dopant diffusion. Heterojunction<br />

bipolar transistors with Schottky collectors were used as tools to measure the <strong>in</strong>jection<br />

current <strong>in</strong> as-grown p + -In 0.53 Ga 0.47 As/n-InP diodes, <strong>and</strong> the results were compared with<br />

the predictions of the thermionic emission model. In this way, the relative diffusivities of<br />

the three p-type dopants <strong>in</strong> InP at 600C were evaluated. The results could be summarized<br />

by:<br />

D Be /D Mn ˜ 20, D Mg /D Be ˜ 1<br />

P.S.Whitney, J.C.Vicek, C.G.Fonstad: Journal of Applied Physics, 1987, 62[5], 1920-4<br />

[446-55/56-028]<br />

Mn<br />

In<strong>GaAs</strong>/InP: Mn <strong>Diffusion</strong><br />

Heterojunctions which were grown by us<strong>in</strong>g liquid-phase epitaxy were <strong>in</strong>vestigated by<br />

us<strong>in</strong>g I-V <strong>and</strong> I o (T) measurements. It was shown that the <strong>in</strong>jection of electrons across the<br />

hetero-<strong>in</strong>terface was described well by a thermionic emission model which <strong>in</strong>volved a<br />

barrier height that was strongly affected by p-type dopant diffusion. Heterojunction<br />

bipolar transistors with Schottky collectors were used as tools to measure the <strong>in</strong>jection<br />

current <strong>in</strong> as-grown p + -In 0.53 Ga 0.47 As/n-InP diodes, <strong>and</strong> the results were compared with<br />

the predictions of the thermionic emission model. In this way, the relative diffusivities of<br />

413


Mn (Ga,In)As Si<br />

the three p-type dopants <strong>in</strong> InP at 600C were evaluated. The results could be summarized<br />

by:<br />

D Be /D Mn ˜ 20, D Mg /D Be ˜ 1<br />

P.S.Whitney, J.C.Vicek, C.G.Fonstad: Journal of Applied Physics, 1987, 62[5], 1920-4<br />

[446-55/56-028]<br />

P<br />

In<strong>GaAs</strong>/In<strong>GaAs</strong>P: P <strong>Diffusion</strong><br />

A photolum<strong>in</strong>escence study was made of the <strong>in</strong>terdiffusion of P <strong>in</strong> the<br />

In 0.66 Ga 0.33 As/In 0.66 Ga 0.33 As 0.7 P 0.3 system at temperatures rang<strong>in</strong>g from 950 to 600C. It<br />

was shown that the diffusion was Fickian, with no dependence of the diffusion coefficient<br />

upon the substrate dopant-type or etch-pit density. In the case of Sn- or S-doped<br />

substrates, the diffusion could be described by:<br />

D(cm 2 /s) = 23 exp[-3.7(eV)/kT]<br />

at temperatures greater than 675C. This activation energy was the same as that deduced<br />

for group-<strong>III</strong> <strong>in</strong>terdiffusion <strong>in</strong> Ga 0.8 In 0.2 As/<strong>GaAs</strong>. At lower temperatures, diffusivity <strong>in</strong> the<br />

present system could be described by:<br />

D(cm 2 /s) = 5 x 10 -10 exp[-1.7(eV)/kT]<br />

W.P.Gill<strong>in</strong>, S.S.Rao, I.V.Bradley, K.P.Homewood, A.D.Smith, A.T.R.Briggs: Applied<br />

Physics Letters, 1993, 63[6], 797-9<br />

[446-106/107-116]<br />

Si<br />

In<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The behavior of implanted Si + ions was <strong>in</strong>vestigated dur<strong>in</strong>g rapid thermal anneal<strong>in</strong>g at<br />

temperatures of between 600 <strong>and</strong> 900C. The apparent activation energy for Si was equal<br />

to 0.64eV. Higher activation efficiencies were found for the dopant <strong>in</strong> In<strong>GaAs</strong>, as<br />

compared with InAlAs. The Si underwent no migration, even after anneal<strong>in</strong>g at 850C.<br />

E.Hailemariam, S.J.<strong>Pearton</strong>, W.S.Hobson, H.S.Luftman, A.P.Perley: Journal of Applied<br />

Physics, 1992, 71[1], 215-20<br />

[446-86/87-038]<br />

In<strong>GaAs</strong>/InP: Si <strong>Diffusion</strong><br />

It was found that Si-<strong>in</strong>duced mix<strong>in</strong>g produced comparable anion <strong>and</strong> cation <strong>in</strong>terdiffusion.<br />

This was consistent with a di-vacancy mechanism. The mix<strong>in</strong>g depended markedly upon<br />

the Si content <strong>and</strong> occurred above the diffusion shoulder, where the vacancy <strong>and</strong> defect<br />

pair concentrations were greatly enhanced. The absence of stra<strong>in</strong>-related growth defects <strong>in</strong><br />

414


Si (Ga,In)As Zn<br />

the un-stra<strong>in</strong>ed start<strong>in</strong>g material permitted the formation of a high-quality stra<strong>in</strong>ed-layer<br />

superlattice by mix<strong>in</strong>g.<br />

S.A.Schwarz, P.Mei, T.Venkatesan, R.Bhat, D.M.Hwang, C.L.Schwartz, M.Koza,<br />

L.Nazar, B.J.Skromme: Applied Physics Letters, 1988, 53[12], 1051-3<br />

[446-62/63-223]<br />

Ti<br />

In<strong>GaAs</strong>: Ti <strong>Diffusion</strong><br />

Samples of n-type In 0.53 Ga 0.47 As were implanted with Co <strong>and</strong> Fe, <strong>and</strong> p-type samples of<br />

the same material were implanted with Ti. In the case of high-temperature s<strong>in</strong>gle-energy<br />

Co <strong>and</strong> Fe implantation, no satellite peaks were observed at locations such as 0.8R, R +<br />

dR, or 2R, where R was the projected range <strong>and</strong> dR was the straggle of the implant.<br />

Dur<strong>in</strong>g high-temperature anneal<strong>in</strong>g, out-diffusion of the implant was as severe as that for<br />

room-temperature implants. In-diffusion of the implant also occurred, but it was not as<br />

severe as the out-diffusion. High-temperature anneal<strong>in</strong>g of Ti-implanted material resulted<br />

<strong>in</strong> slight Ti <strong>in</strong>-diffusion, with m<strong>in</strong>imal redistribution or out-diffusion. In the case of hightemperature<br />

implants, the lattice quality of the annealed material was close to that of<br />

virg<strong>in</strong> material. Regardless of the ion type, resistivities that were close to the <strong>in</strong>tr<strong>in</strong>sic<br />

limit were measured <strong>in</strong> implanted <strong>and</strong> annealed materials.<br />

M.V.Rao, S.M.Gulwadi, S.Mulpuri, D.S.Simons, P.H.Chi, C.Caneau, W.P.Hong,<br />

O.W.Holl<strong>and</strong>, H.B.Dietrich: Journal of Electronic Materials, 1992, 21[9], 923-8<br />

[446-93/94-038]<br />

Zn<br />

In<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The diffusion of Zn <strong>in</strong>to In 0.57 Ga 0.43 As was studied by us<strong>in</strong>g boat diffusion, diffusion<br />

from As- or P-doped spun-on films, or diffusion from In-doped spun-on films. The depth<br />

profiles were deduced from junction positions. It was found that the p + /p - junction<br />

position depended upon the diffusion method which was used, but not upon the sample<br />

growth technique. The p - /n junction position depended upon both factors. Because the<br />

amounts of As, Ga <strong>and</strong> In (or of the respective vacancies) differed, it was possible to<br />

identify diffusion mechanisms. It was proposed that <strong>in</strong>terstitially diffus<strong>in</strong>g Zn was<br />

<strong>in</strong>dependently trapped by 2 immobile vacancy centers. These consisted of Zn on V In or<br />

V Ga <strong>in</strong> the p + region, <strong>and</strong> Zn on V As ZnV As <strong>in</strong> the p - region.<br />

U.König, H.Haspeklo, P.Marschall, M.Kuisl: Journal of Applied Physics, 1989, 65[2],<br />

548-52<br />

[446-72/73-035]<br />

415


Zn (Ga,In)As Zn<br />

In<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

A secondary ion mass spectroscopic study demonstrated that, dur<strong>in</strong>g the growth of Zndoped<br />

In 0.53 Ga 0.47 As layers by atmospheric-pressure organometallic vapor-phase epitaxy,<br />

Zn atoms which were trapped on <strong>in</strong>terstitial sites dur<strong>in</strong>g growth, rather than <strong>in</strong>terstitial Zn<br />

defects which were generated by the kick-out mechanism, were probably the ma<strong>in</strong> cause<br />

of the carry<strong>in</strong>g over of Zn <strong>in</strong>to subsequent layers. The <strong>in</strong>terstitial Zn <strong>in</strong>corporation<br />

seemed to be due to the saturation of In substitution, rather than of Ga substitution. It was<br />

possible that the use of pauses <strong>in</strong> the growth sequence, both before <strong>and</strong> after the growth<br />

of a Zn-doped layer, could control these effects.<br />

S.J.Taylor, B.Beaumont, J.C.Guillaume: Semiconductor Science <strong>and</strong> Technology, 1993,<br />

8[12], 2193-6<br />

[446-115/116-143]<br />

In<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

Secondary ion mass spectrometry <strong>and</strong> electrochemical profil<strong>in</strong>g studies were made of the<br />

saturation of Zn dop<strong>in</strong>g <strong>and</strong> diffusion <strong>in</strong> In 0.53 Ga 0.47 As which had been grown onto InP by<br />

us<strong>in</strong>g organometallic vapor-phase epitaxy. It was found that the results were consistent<br />

with a so-called kick-out mechanism. It was proposed that the diffus<strong>in</strong>g species was<br />

probably a neutral Zn <strong>in</strong>terstitial. Accumulation of Zn at the <strong>in</strong>terface with a highly n-<br />

doped layer <strong>in</strong>dicated the possible formation of Zn-donor complexes.<br />

S.J.Taylor, B.Beaumont, J.C.Guillaume: Semiconductor Science <strong>and</strong> Technology, 1993,<br />

8[5], 643-6<br />

[446-115/116-143]<br />

GaInAs: Zn <strong>Diffusion</strong><br />

A systematic study of Zn <strong>in</strong>corporation showed that the surface Zn concentration <strong>in</strong><br />

metalorganic vapor phase epitaxial material could be <strong>in</strong>creased from a grown-<strong>in</strong><br />

maximum of 2 x 10 19 , to 10 20 /cm 3 , by diffusion from a spun-on glass source. These<br />

values were found to depend upon the type of p-dopant element which was used <strong>in</strong> asgrown<br />

double heterostructures.<br />

D.L.Murrell: Semiconductor Science <strong>and</strong> Technology, 1990, 5[5], 414-20<br />

[446-74-030]<br />

GaInAs/<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The enhancement of In <strong>and</strong> Ga <strong>in</strong>terdiffusion, due to Zn diffusion, was studied <strong>in</strong> stra<strong>in</strong>ed<br />

In x Ga 1-x As/<strong>GaAs</strong> s<strong>in</strong>gle quantum well structures, where x was between 0.20 <strong>and</strong> 0.24. The<br />

structures were grown by means of metalorganic vapor base epitaxy, <strong>and</strong> consisted of a<br />

1000 to 2000nm-thick <strong>GaAs</strong> buffer layer plus an 8 to 10nm GaInAs layer, <strong>and</strong> a 50 to<br />

100nm <strong>GaAs</strong> cap layer. Shallow Zn <strong>in</strong>-diffusion was carried out at temperatures of<br />

between 585 <strong>and</strong> 620C, <strong>in</strong> order to vary the Zn content of the quantum wells. The samples<br />

were then annealed at temperatures of between 650 <strong>and</strong> 785C (<strong>in</strong> AsH 3 -H 2 mixtures) <strong>in</strong><br />

416


Zn (Ga,In)As Zn<br />

order to determ<strong>in</strong>e the enhanced In-Ga <strong>in</strong>terdiffusion coefficient. The results for a typical<br />

case could be described by:<br />

D(cm 2 /s) = 3 x 10 -6 exp[-2.33(eV)/kT]<br />

The <strong>in</strong>terdiffusion enhancement was expla<strong>in</strong>ed <strong>in</strong> terms of an <strong>in</strong>terstitial migration<br />

process.<br />

M.T.Furtado, E.A.Sato, M.A.Sacilotti: Superlattices <strong>and</strong> Microstructures, 1991, 10[2],<br />

225-30<br />

[446-88/89-042]<br />

Table 36<br />

Interdiffusion Coefficients for In x Ga 1-x As/<strong>GaAs</strong><br />

x<br />

D (cm 2 /s)<br />

0.23 1.5 x 10 -18<br />

0.24 7.9 x 10 -17<br />

0.21 3.2 x 10 -16<br />

In<strong>GaAs</strong>/AlGaInAs: Zn <strong>Diffusion</strong><br />

<strong>Diffusion</strong>-<strong>in</strong>duced disorder<strong>in</strong>g of multiple quantum wells was <strong>in</strong>vestigated as a possible<br />

new process<strong>in</strong>g technique for long-wavelength opto-electronic devices. Complete<br />

disorder<strong>in</strong>g of the multiple quantum well structure was confirmed by an observed<br />

shorten<strong>in</strong>g of the photolum<strong>in</strong>escence peak wavelength <strong>and</strong> by secondary ion mass<br />

spectrometry data. Lattice-matched disorder<strong>in</strong>g was also detected by us<strong>in</strong>g X-ray<br />

diffraction techniques. The first long-wavelength buried multiple quantum well laser was<br />

fabricated, <strong>in</strong> which carrier conf<strong>in</strong>ement was obta<strong>in</strong>ed by disorder<strong>in</strong>g.<br />

K.Goto, F.Uesugi, S.Takahashi, T.Takiguchi, E.Omura, Y.Mihashi: Japanese Journal of<br />

Applied Physics, 1994, 33[1-10], 5774-8<br />

[446-117/118-184]<br />

336 In<strong>GaAs</strong>/<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

An <strong>in</strong>vestigation was made, us<strong>in</strong>g a 2-stage Zn diffusion <strong>and</strong> thermal anneal<strong>in</strong>g process,<br />

of impurity-<strong>in</strong>duced disorder<strong>in</strong>g <strong>in</strong> stra<strong>in</strong>ed samples of In x Ga 1-x As/<strong>GaAs</strong> s<strong>in</strong>gle quantum<br />

well heterostructures, where x ranged from 0.21 to 0.24. The samples were grown by<br />

means of metalorganic vapor-phase epitaxy, <strong>and</strong> various shallow Zn diffusion depths<br />

were produced <strong>in</strong> the <strong>GaAs</strong> cap layer <strong>in</strong> order to vary the Zn concentration <strong>in</strong> the quantum<br />

wells. Thermal anneal<strong>in</strong>g (785C, 600s) <strong>in</strong> an AsH 3 /H 2 atmosphere then produced<br />

impurity-<strong>in</strong>duced disorder<strong>in</strong>g. Partially disordered <strong>and</strong> completely disordered quantum<br />

well heterostructures were studied, <strong>and</strong> In-Ga <strong>in</strong>terdiffusion was monitored via the<br />

photolum<strong>in</strong>escence spectroscopy of ground state emissions from the quantum wells. The<br />

<strong>in</strong>terdiffusion coefficients (table 36) were determ<strong>in</strong>ed by apply<strong>in</strong>g the envelope function<br />

to shifts <strong>in</strong> the photolum<strong>in</strong>escence peak position dur<strong>in</strong>g anneal<strong>in</strong>g. It was found that the<br />

<strong>in</strong>terdiffusion coefficient was very strongly dependent upon the Zn diffusion depth, <strong>and</strong><br />

417


Zn (Ga,In)As Zn<br />

therefore upon the Zn concentration <strong>in</strong> the quantum well layer. A model was proposed<br />

that used an <strong>in</strong>terstitial migration process to expla<strong>in</strong> the enhancement of In-Ga<br />

<strong>in</strong>terdiffusion by Zn diffusion.<br />

M.T.Furtado, M.S.S.Loural, E.A.Sato, M.A.Sacilotti: Semiconductor Science <strong>and</strong><br />

Technology, 1992, 7[6], 744-51<br />

[446-99/100-092]<br />

In<strong>GaAs</strong>/In<strong>GaAs</strong>P: Zn <strong>Diffusion</strong><br />

The stability of the Zn profile <strong>in</strong> modulation-doped multiple quantum well structures,<br />

which had been grown by means of low-pressure metalorganic vapor-phase epitaxy, was<br />

<strong>in</strong>vestigated by apply<strong>in</strong>g secondary ion mass spectrometric <strong>and</strong> transmission electron<br />

microscopic techniques to wedge-shaped samples. Although an excellent stability of the<br />

Zn profile was observed <strong>in</strong> as-grown samples with modulation dop<strong>in</strong>g (3nm, 10 18 Zn/cm 3 ),<br />

the modulation-doped structure faded after the second epitaxial re-growth of a p-type InP<br />

layer (10 18 Zn/cm 3 ) us<strong>in</strong>g either liquid-phase epitaxial or metalorganic vapor-phase<br />

epitaxial techniques. However, the modulation-dop<strong>in</strong>g profile was successfully preserved<br />

even after re-growth of the p-type InP layer for 1.5h (<strong>in</strong> a sample that comprised an<br />

undoped InP-clad layer, <strong>in</strong>stead of a p-type InP clad layer, superposed on the modulationdoped<br />

multiple quantum well structure layers). The Zn diffusion coefficient <strong>in</strong> the<br />

modulation-doped region was less than 7 x 10 -18 cm 2 /s. The maximum Zn concentration,<br />

for obta<strong>in</strong><strong>in</strong>g a stable modulation-dop<strong>in</strong>g structure <strong>in</strong> the modulation-doped region of<br />

barrier layers, was found to be 2 x 10 18 /cm 3 . It was proposed that the suppression of<br />

<strong>in</strong>terstitial Zn atoms, <strong>and</strong> of subsequently produced <strong>in</strong>terstitial group-<strong>III</strong> atoms (which<br />

were generated <strong>in</strong> the p-type InP clad layer via a kick-out mechanism <strong>and</strong> diffused <strong>in</strong>to<br />

the multiple quantum well region), was important <strong>in</strong> preserv<strong>in</strong>g the modulation-doped<br />

structure.<br />

N.Otsuka, M.Ish<strong>in</strong>o, Y.Matsui: Journal of Applied Physics, 1996, 80[3], 1405-13<br />

[446-138/139-097]<br />

In<strong>GaAs</strong>/InP: Zn <strong>Diffusion</strong><br />

Photodiodes were fabricated via selective Zn diffusion us<strong>in</strong>g a dimethylz<strong>in</strong>c source. The<br />

results showed that this method was a useful process for fabricat<strong>in</strong>g such devices.<br />

M.Wada, M.Seko, K.Sakakibara, Y.Sekiguchi: Japanese Journal of Applied Physics,<br />

1990, 29[3], L401-4<br />

[446-76/77-024]<br />

In<strong>GaAs</strong>/InP: Zn <strong>Diffusion</strong><br />

A closed-ampoule technique was used to study the simultaneous diffusion of Cd <strong>and</strong> Zn<br />

<strong>in</strong>to material which was lattice-matched to InP. It was found that the Cd atoms, whose<br />

<strong>in</strong>terstitial diffusion coefficient was small when compared with that of Zn <strong>in</strong>terstitials,<br />

penetrated <strong>in</strong>to the crystal as deeply as the Zn atoms. These data agreed with the results of<br />

418


Zn (Ga,In)As Zn<br />

numerical simulations of simultaneous diffusion, <strong>in</strong> <strong>III</strong>-V compounds, which were based<br />

upon an <strong>in</strong>terstitial-substitutional diffusion mechanism.<br />

U.Wielsch, P.Ambrée, B.Gruska: Semiconductor Science <strong>and</strong> Technology, 1990, 5[9],<br />

923-7<br />

[446-76/77-024]<br />

In<strong>GaAs</strong>/InP: Zn <strong>Diffusion</strong><br />

The diffusion of Zn from spun-on films, <strong>and</strong> <strong>in</strong>to InP/In<strong>GaAs</strong>/InP heterostructures, was<br />

studied. Marked segregation occurred at the In<strong>GaAs</strong>/InP heterojunctions <strong>and</strong> <strong>in</strong>creased<br />

the Zn concentration <strong>in</strong> In<strong>GaAs</strong> by about an order of magnitude. <strong>Diffusion</strong> <strong>and</strong><br />

segregation parameters were deduced from the Zn concentration profiles.<br />

F.Dildey, M.C.Amann, R.Treichler: Japanese Journal of Applied Physics, 1990, 29[5],<br />

810-2<br />

[446-76/77-024]<br />

In<strong>GaAs</strong>/InP: Zn <strong>Diffusion</strong><br />

Results on diffusion across In<strong>GaAs</strong>/InP <strong>and</strong> InP/In<strong>GaAs</strong> hetero-<strong>in</strong>terfaces were described.<br />

It was found that marked changes <strong>in</strong> the group-<strong>III</strong> sub-lattice occurred, near to the<br />

<strong>in</strong>terface, when Zn diffused across the heterojunction from an In<strong>GaAs</strong> top layer. The<br />

getter<strong>in</strong>g of Zn at the <strong>in</strong>terface was analyzed <strong>in</strong> terms of kick-out <strong>and</strong> vacancy<br />

mechanisms. The activation energy for Zn-stimulated Ga <strong>in</strong>terdiffusion across the<br />

In<strong>GaAs</strong>/InP heterojunction was estimated to be 3.8eV.<br />

P.Ambrée, A.Hangleiter, M.H.Pilkuhn, K.W<strong>and</strong>el: Applied Physics Letters, 1990, 56[10],<br />

931-3<br />

[446-74-039]<br />

In<strong>GaAs</strong>/InP: Zn <strong>Diffusion</strong><br />

The Zn was <strong>in</strong>troduced, us<strong>in</strong>g spun-on films, <strong>in</strong>to n-InP/p + -In<strong>GaAs</strong>/n-InP heterostructures<br />

which had been grown via metalorganic vapor phase epitaxy (with Mg as a p-dopant).<br />

After diffusion, the Mg was completely replaced by Zn <strong>and</strong> was enriched <strong>in</strong> the spun-on<br />

film. In the presence of Mg, the <strong>in</strong>-diffusion of Zn was strongly enhanced. By vary<strong>in</strong>g the<br />

dopant level <strong>and</strong> diffusion conditions, the underly<strong>in</strong>g mechanism was compared with that<br />

which operated <strong>in</strong> Be-doped Al<strong>GaAs</strong>/<strong>GaAs</strong> heterostructures.<br />

F.Dildey, R.Treichler, M.C.Amann, M.Schier, G.Ebb<strong>in</strong>ghaus: Applied Physics Letters,<br />

1989, 55[9], 876-8<br />

[446-70/71-118]<br />

In<strong>GaAs</strong>/InP: Zn <strong>Diffusion</strong><br />

It was found that Zn-<strong>in</strong>duced mix<strong>in</strong>g ma<strong>in</strong>ly affected the cation sub-lattice of the<br />

superlattice <strong>and</strong> was consistent with a so-called <strong>in</strong>terstitial kick-out mechanism. The Zndiffused<br />

superlattice rema<strong>in</strong>ed sharply def<strong>in</strong>ed. The absence of stra<strong>in</strong>-related growth<br />

419


Zn (Ga,In)As General<br />

defects <strong>in</strong> the un-stra<strong>in</strong>ed start<strong>in</strong>g material permitted the formation of a high-quality<br />

stra<strong>in</strong>ed-layer superlattice by mix<strong>in</strong>g.<br />

S.A.Schwarz, P.Mei, T.Venkatesan, R.Bhat, D.M.Hwang, C.L.Schwartz, M.Koza,<br />

L.Nazar, B.J.Skromme: Applied Physics Letters, 1988, 53[12], 1051-3<br />

[446-62/63-223]<br />

In<strong>GaAs</strong>/InP: Zn <strong>Diffusion</strong><br />

The <strong>in</strong>termix<strong>in</strong>g of multiple quantum well structures by Zn diffusion at 550C was<br />

<strong>in</strong>vestigated. Secondary ion mass spectroscopy <strong>and</strong> X-ray analysis revealed that Zn<br />

diffusion promoted the <strong>in</strong>termix<strong>in</strong>g of group-<strong>III</strong> atoms, but had little effect upon group-V<br />

profiles. However, the resultant group-<strong>III</strong> atom profiles were not completely uniform;<br />

even after Zn diffusion. The results suggested that a large lattice mismatch suppressed the<br />

Zn diffusion <strong>in</strong>termix<strong>in</strong>g process.<br />

K.Nakashima, Y.Kawaguchi, Y.Kawamura, Y.Imamura, H.Asahi: Applied Physics<br />

Letters, 1999, 52[17], 1383-5<br />

[446-62/63-223]<br />

In<strong>GaAs</strong>/InP: Zn <strong>Diffusion</strong><br />

A systematic study was made of the effects of Zn dop<strong>in</strong>g <strong>and</strong> diffusion <strong>in</strong> capped mesa<br />

buried heterostructure lasers which had been grown by means of metalorganic chemical<br />

vapor deposition. It <strong>in</strong>volved vary<strong>in</strong>g the Zn content (7 x 10 17 to 3.1 x 10 18 /cm 3 ) of the p-<br />

type InP cladd<strong>in</strong>g layer <strong>in</strong> the base epitaxial structure, while keep<strong>in</strong>g the growth<br />

conditions constant dur<strong>in</strong>g 2 subsequent re-growth steps. Secondary ion mass<br />

spectrometry was used to make quantitative determ<strong>in</strong>ations of the Zn depth profiles,<br />

follow<strong>in</strong>g re-growth, by us<strong>in</strong>g test sites on 50mm round wafers which conta<strong>in</strong>ed the<br />

appropriate epitaxial layers. Clear evidence of Zn diffusion was found, such as the<br />

penetration of Zn <strong>in</strong>to the active layer <strong>and</strong> the presence of <strong>in</strong>flection po<strong>in</strong>ts (accumulation<br />

<strong>and</strong> depletion of Zn near to the p-n heterojunction) <strong>in</strong> the depth profiles. It was observed<br />

that the diffusion of Zn dur<strong>in</strong>g the third growth step dom<strong>in</strong>ated the Zn profile <strong>in</strong> the base<br />

growth part of the p-type InP layer, <strong>and</strong> the f<strong>in</strong>al amount of Zn <strong>in</strong> this region was<br />

<strong>in</strong>dependent of the <strong>in</strong>itial dopant level. Above a Zn threshold level of about 2.2 x<br />

10 18 /cm 3 , the Zn diffusion <strong>in</strong>creased significantly <strong>and</strong> resulted <strong>in</strong> the presence of 5 x<br />

10 18 /cm 3 or more of Zn <strong>in</strong> the active layer. The threshold for the onset of diffusion was<br />

found to be <strong>in</strong> accord with the substitutional-<strong>in</strong>terstitial diffusion of Zn.<br />

V.Swam<strong>in</strong>athan, C.L.Reynolds, M.Geva: Applied Physics Letters, 1995, 66[20], 2685-7<br />

[446-121/122-079]<br />

General<br />

In<strong>GaAs</strong>/In<strong>GaAs</strong>P/InP: <strong>Diffusion</strong><br />

Laser structures, with In<strong>GaAs</strong> quantum wells which were about 1.85µ below the surface,<br />

were implanted with ions that had energies of up to 8.6MeV. Intermix<strong>in</strong>g of the quantum<br />

wells, dur<strong>in</strong>g rapid thermal anneal<strong>in</strong>g, was monitored via changes <strong>in</strong> the energy, l<strong>in</strong>e-<br />

420


General (Ga,In)As Surface<br />

width <strong>and</strong> <strong>in</strong>tensity of the photolum<strong>in</strong>escence peaks from the quantum wells. When<br />

diffusion occurred ma<strong>in</strong>ly through InP, the photolum<strong>in</strong>escence data correlated well with<br />

the calculated total number of vacancies that were created <strong>in</strong> the sample. This suggested<br />

that defect diffusion was very efficient <strong>in</strong> InP.<br />

P.J.Poole, S.Charbonneau, G.C.Aers, T.E.Jackman, M.Buchanan, M.Dion,<br />

R.D.Goldberg, I.V.Mitchell: Journal of Applied Physics, 1995, 78[4], 2367-71<br />

[446-123/124-177]<br />

In<strong>GaAs</strong>: <strong>Diffusion</strong><br />

A theoretical model was proposed for the calculation of the layer thickness of a <strong>III</strong>-V<br />

ternary crystal which was grown by us<strong>in</strong>g the source-current controlled method. The<br />

latter could be used to control the supply of depleted solute elements to a solution dur<strong>in</strong>g<br />

growth. That is, solute elements were cont<strong>in</strong>uously supplied to the growth solution from a<br />

source material <strong>and</strong> were transported to a substrate (through the temperature gradient <strong>in</strong><br />

the growth solution) via diffusion <strong>and</strong> electromigration. By us<strong>in</strong>g the theoretical model,<br />

analytical calculations could be made of the diffusion-limited <strong>and</strong> electromigrationlimited<br />

growth of In<strong>GaAs</strong> <strong>in</strong> temperature gradient In-Ga-As ternary solutions. The<br />

calculations <strong>in</strong>dicated that the thickness depended upon growth parameters such as the<br />

growth temperature, the cool<strong>in</strong>g rates of the substrate <strong>and</strong> source, the solution length, the<br />

temperature difference between substrate <strong>and</strong> source, the mobility of the migrat<strong>in</strong>g solute<br />

element, <strong>and</strong> the electric field <strong>in</strong> the solution.<br />

K.Nakajima: Journal of Crystal Growth, 1989, 98[3], 329-40<br />

[446-72/73-035]<br />

In<strong>GaAs</strong>/<strong>GaAs</strong>/Al<strong>GaAs</strong>: <strong>Diffusion</strong><br />

Laser structures, with In<strong>GaAs</strong> quantum wells which were about 1.85µ below the surface,<br />

were implanted with ions that had energies of up to 8.6MeV. Intermix<strong>in</strong>g of the quantum<br />

wells, dur<strong>in</strong>g rapid thermal anneal<strong>in</strong>g, was monitored via changes <strong>in</strong> the energy, l<strong>in</strong>ewidth<br />

<strong>and</strong> <strong>in</strong>tensity of the photolum<strong>in</strong>escence peaks from the quantum wells. When the<br />

defects had to diffuse ma<strong>in</strong>ly through Al 0.71 Ga 0.29 As, these quantities were closely<br />

related, for short anneal<strong>in</strong>g times, to the predicted vacancy generation <strong>and</strong> ion deposition<br />

at the depth of the quantum well before anneal<strong>in</strong>g. This suggested that the defect<br />

diffusion length <strong>in</strong> Al<strong>GaAs</strong> <strong>and</strong>/or <strong>GaAs</strong> was quite low.<br />

P.J.Poole, S.Charbonneau, G.C.Aers, T.E.Jackman, M.Buchanan, M.Dion,<br />

R.D.Goldberg, I.V.Mitchell: Journal of Applied Physics, 1995, 78[4], 2367-71<br />

[446-123/124-177]<br />

Surface <strong>Diffusion</strong><br />

Ga<br />

In<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

Surface diffusion dur<strong>in</strong>g molecular beam epitaxy was studied. Firstly, the mode transition<br />

between 2-dimensional nucleation <strong>and</strong> step flow dur<strong>in</strong>g molecular beam epitaxial growth<br />

421


Surface (Ga,In)As Interdiffusion<br />

on vic<strong>in</strong>al surfaces was studied theoretically <strong>and</strong> experimentally. The basis of the theory<br />

was to assume that the transition occurred when the surface supersaturation on the step<br />

terrace became identical to the critical supersaturation for 2-dimensional nucleation. This<br />

permitted the diffusion length of Ga to be calculated at the experimentally determ<strong>in</strong>ed<br />

critical temperature for the mode transition. It was found that the diffusion length<br />

<strong>in</strong>creased, as the temperature decreased, due to an <strong>in</strong>creased residence time. Also, the<br />

diffusion length on (111)B was longer than that on (001) when the same formation energy<br />

for 2-dimensional nuclei was assumed for both surfaces. The theory was then used to<br />

elucidate the dependence of the In<strong>GaAs</strong> composition upon the growth temperature, the<br />

substrate orientation, <strong>and</strong> the degree of misorientation. The theory gave good agreement<br />

with the experimental data, <strong>and</strong> it was concluded that surface diffusion was one of the<br />

most important processes which controlled molecular beam epitaxial growth <strong>and</strong> impurity<br />

<strong>in</strong>corporation.<br />

T.Nish<strong>in</strong>aga, T.Shitara, K.Mochizuki, K.I.Cho: Journal of Crystal Growth, 1990, 99, 482-<br />

90<br />

[446-76/77-009]<br />

In<br />

GaInAs/<strong>GaAs</strong>: In Surface <strong>Diffusion</strong><br />

The lateral profiles of In <strong>in</strong> a 1.5µ-thick Ga 1-x In x As layers, where x was approximately<br />

equal to 0.2, were grown onto <strong>GaAs</strong> channelled substrates with (411)A side-slopes by the<br />

use of molecular beam epitaxy <strong>and</strong> were <strong>in</strong>vestigated with the use of energy-dispersive X-<br />

ray spectroscopy. The observed profiles of the In content suggested that the In atoms<br />

migrated preferentially <strong>in</strong> the [12¯2¯] direction on the (411)A plane dur<strong>in</strong>g molecular beam<br />

epitaxial growth. This preferential migration of In atoms was confirmed by compar<strong>in</strong>g the<br />

observed lateral profiles of In <strong>in</strong> GaInAs layers which had been grown onto <strong>GaAs</strong><br />

channelled substrates with simulated In profiles that had been calculated by tak<strong>in</strong>g<br />

account of an additional one-way flow of In atoms along [12¯2¯].<br />

T.Kitada, A.Wakejima, N.Tomita, S.Shimomura, A.Adachi, N.Sano, S.Hiyamizu: Journal<br />

of Crystal Growth, 1995, 150[1-4], 487-91<br />

[446-127/128-141]<br />

Interdiffusion<br />

GaInAs/AlGaInAs: Interdiffusion<br />

Investigations were made of the <strong>in</strong>terdiffusion behavior <strong>and</strong> thermal stability of n-doped<br />

<strong>and</strong> undoped multiple quantum-well structures that were lattice-matched to InP by means<br />

of molecular beam epitaxy. The activation energy for the ma<strong>in</strong> <strong>in</strong>terdiffusion process was<br />

deduced to be 2.5eV for doped structures <strong>and</strong> 2.9eV for undoped structures. The differ<strong>in</strong>g<br />

<strong>in</strong>terdiffusion processes were monitored by us<strong>in</strong>g photolum<strong>in</strong>escence spectroscopy at 8K,<br />

after rapid thermal anneal<strong>in</strong>g. The effect of dop<strong>in</strong>g was studied by compar<strong>in</strong>g the results<br />

422


Interdiffusion (Ga,In)As Interdiffusion<br />

for n-doped <strong>and</strong> undoped structures. Photolum<strong>in</strong>escence excitation spectroscopic data for<br />

2K confirmed the differ<strong>in</strong>g <strong>in</strong>terdiffusion processes.<br />

V.Hofsäss, J.Kuhn, H.Schweizer, H.Hillmer, R.Lösch, W.Schlapp: Journal of Applied<br />

Physics, 1995, 78[5], 3534-6<br />

[446-123/124-172]<br />

GaInAs/AlInAs: Interdiffusion<br />

The <strong>in</strong>teraction of lattice-matched heterostructures, grown by means of molecular beam<br />

epitaxy, was studied by us<strong>in</strong>g electron microscopy. An appreciable amount of<br />

<strong>in</strong>terdiffusion was observed at temperatures as low as 700C. The use of X-ray microanalysis<br />

revealed that <strong>in</strong>terdiffusion occurred along a non-l<strong>in</strong>ear (non lattice-matched)<br />

path.<br />

R.E.Mallard, N.J.Long, G.R.Booker, E.G.Scott, M.Hockly, M.Taylor: Journal of Applied<br />

Physics, 1991, 70[1], 182-92<br />

[446-91/92-018]<br />

1.0E-14<br />

D (cm 2 /s)<br />

1.0E-15<br />

1.0E-16<br />

1.0E-17<br />

table 34<br />

table 37<br />

table 38<br />

table 39<br />

table 40<br />

table 41<br />

table 42<br />

1.0E-18<br />

1.0E-19<br />

7 8 9 10 11<br />

10 4 /T(K)<br />

Figure 8: Interdiffusivity <strong>in</strong> In<strong>GaAs</strong>/<strong>GaAs</strong><br />

423


Interdiffusion (Ga,In)As Interdiffusion<br />

GaInAs/<strong>GaAs</strong>: Interdiffusion<br />

The dependence of impurity-free <strong>in</strong>terdiffusion upon the properties of a dielectric cap<br />

layer was studied <strong>in</strong> pseudomorphic multi-quantum well structures that had been grown<br />

by means of molecular beam epitaxy. Electron-beam evaporated SiO 2 films, chemical<br />

vapor deposited SiO x N y films, <strong>and</strong> spun-on SiO 2 films were used as cap layers dur<strong>in</strong>g<br />

rapid thermal anneal<strong>in</strong>g at temperatures of between 850 <strong>and</strong> 950C. The<br />

photolum<strong>in</strong>escence at 10K was used to monitor <strong>in</strong>terdiffusion-<strong>in</strong>duced b<strong>and</strong>-gap shifts,<br />

<strong>and</strong> to calculate the correspond<strong>in</strong>g In-Ga <strong>in</strong>terdiffusion coefficients. The latter were found<br />

to <strong>in</strong>crease with cap layer thickness (electron-beam SiO 2 ) up to a limit which was<br />

governed by saturation of the out-diffused Ga concentration <strong>in</strong> the SiO 2 caps. A<br />

maximum concentration of between 4 x 10 19 <strong>and</strong> 7 x 10 19 /cm 3 <strong>in</strong> the SiO 2 caps was found<br />

by us<strong>in</strong>g secondary ion mass spectroscopic profil<strong>in</strong>g. Larger b<strong>and</strong>-edge shifts were also<br />

obta<strong>in</strong>ed when the O content of SiO x N y cap layers was <strong>in</strong>creased, but the differences were<br />

<strong>in</strong>sufficient to suggest a laterally selective <strong>in</strong>terdiffusion process that was based upon<br />

variations <strong>in</strong> cap layer composition alone. Much larger differences were obta<strong>in</strong>ed by<br />

us<strong>in</strong>g various deposition techniques for the cap layers. This <strong>in</strong>dicated that the porosity of<br />

the cap layer was a much more important factor than was the film composition <strong>in</strong><br />

obta<strong>in</strong><strong>in</strong>g a laterally selective <strong>in</strong>terdiffusion process. In the case of Ga 0.8 In 0.2 As/<strong>GaAs</strong><br />

<strong>in</strong>terdiffusion, the activation energies <strong>and</strong> pre-factors were estimated to range from 3.04<br />

to 4.74eV <strong>and</strong> from 5 x 10 -3 to 2 x 10 5 cm 2 /s, respectively; depend<strong>in</strong>g upon the cap layer<br />

deposition technique <strong>and</strong> the depth of the multi-quantum well below the sample surface.<br />

S.Bürkner, M.Maier, E.C.Lark<strong>in</strong>s, W.Rothemund, E.P.O’Reilly, J.D.Ralston: Journal of<br />

Electronic Materials, 1995, 24[7], 805-12<br />

[446-125/126-131]<br />

337 GaInAs/<strong>GaAs</strong>: Interdiffusion<br />

The effects of Si <strong>and</strong> Be, at dop<strong>in</strong>g levels of up to 10 19 /cm 3 , upon the <strong>in</strong>terdiffusion of<br />

quantum wells after anneal<strong>in</strong>g were studied by us<strong>in</strong>g photolum<strong>in</strong>escence techniques (table<br />

37). It was found that, for Be concentrations of up to 2.5 x 10 19 /cm 3 , <strong>and</strong> for Si<br />

concentrations of up to 10 18 /cm 3 , no change <strong>in</strong> the <strong>in</strong>terdiffusion coefficients could be<br />

measured. At a Si dopant concentration of 6 x 10 18 /cm 3 , there was a dramatic degradation<br />

of the material quality after anneal<strong>in</strong>g (750C, 15s). This caused the lum<strong>in</strong>escence from<br />

the well to disappear, while a deep-level lum<strong>in</strong>escence that was related to donor-Ga<br />

vacancy complexes <strong>and</strong> As antisite defects appeared. On the basis of these results, it was<br />

suggested that the position of the Fermi level played no role <strong>in</strong> the <strong>in</strong>termix<strong>in</strong>g of <strong>III</strong>-V<br />

heterostructures. It was also concluded that most of the enhanced <strong>in</strong>termix<strong>in</strong>g which was<br />

observed <strong>in</strong> Si-doped <strong>GaAs</strong>/Al<strong>GaAs</strong> structures was related to Si relocation at very high<br />

dop<strong>in</strong>g levels.<br />

W.P.Gill<strong>in</strong>, I.V.Bradley, L.K.Howard, R.Gwilliam, K.P.Homewood: Journal of Applied<br />

Physics, 1993, 73[11], 7715-9<br />

[446-106/107-082]<br />

424


Interdiffusion (Ga,In)As Interdiffusion<br />

Table 37<br />

Interdiffusion Data for Ga 0.8 In 0.2 As/<strong>GaAs</strong><br />

Dopant Amount (/cm 3 ) Temperature (C) Coefficient (cm 2 /s)<br />

- - 900 4.70 x 10 -17<br />

- - 950 1.96 x 10 -16<br />

- - 1000 5.50 x 10 -16<br />

- - 1050 1.00 x 10 -15<br />

Si 10 17 900 3.50 x 10 -17<br />

Si 10 17 950 1.26 x 10 -16<br />

Si 10 17 1000 3.90 x 10 -16<br />

Si 10 17 1050 7.50 x 10 -16<br />

Si 10 18 900 1.40 x 10 -16<br />

Si 10 18 950 4.00 x 10 -16<br />

Si 10 18 1000 1.20 x 10 -15<br />

Si 10 18 1050 2.80 x 10 -15<br />

Be 10 17 900 5.40 x 10 -17<br />

Be 10 17 950 1.50 x 10 -16<br />

Be 10 17 1000 2.50 x 10 -16<br />

Be 10 17 1050 1.60 x 10 -15<br />

Be 10 18 900 1.00 x 10 -17<br />

Be 10 18 950 1.02 x 10 -16<br />

Be 10 18 1000 4.20 x 10 -16<br />

Be 10 18 1050 1.33 x 10 -15<br />

Be 2.5 x 10 19 900 3.70 x 10 -17<br />

Be 2.5 x 10 19 950 1.64 x 10 -16<br />

Be 2.5 x 10 19 1000 5.00 x 10 -16<br />

Be 2.5 x 10 19 1050 2.40 x 10 -15<br />

GaInAs/<strong>GaAs</strong>: Interdiffusion<br />

Simple analytical expressions were derived for the approximate estimation of the<br />

<strong>in</strong>terdiffusion coefficient, of partially disordered quantum-well heterostructures, directly<br />

from measurements of the photolum<strong>in</strong>escence phase shift which was associated with layer<br />

<strong>in</strong>terdiffusion. The phase shift was calculated as a function of the <strong>in</strong>terdiffusion length,<br />

(Dt) ½ , <strong>in</strong> the stra<strong>in</strong>ed-layer system, Ga 0.8 In 0.2 As/<strong>GaAs</strong>. The calculations were performed<br />

with<strong>in</strong> the framework of the envelope function approximation <strong>and</strong> Fick's law. A simple<br />

relationship was derived for the variation <strong>in</strong> phase shift as a function of the dimensionless<br />

425


Interdiffusion (Ga,In)As Interdiffusion<br />

parameter, (Dt) ½ /L, where L was the quantum-well thickness. This satisfactorily<br />

accounted for most of the published <strong>in</strong>terdiffusivity values, to with<strong>in</strong> a factor of 2.<br />

M.T.Furtado, M.S.S.Loural: Superlattices <strong>and</strong> Microstructures, 1993, 14[1], 21-5<br />

[446-113/114-029]<br />

GaInAs/<strong>GaAs</strong>: Interdiffusion<br />

Molecular-beam epitaxially grown highly stra<strong>in</strong>ed Ga 0.65 In 0.35 As/<strong>GaAs</strong> multiple quantumwell<br />

structures were <strong>in</strong>vestigated. Interdiffusion was carried out via rapid thermal anneal<strong>in</strong>g,<br />

at temperatures of between 700 <strong>and</strong> 950C, by us<strong>in</strong>g <strong>GaAs</strong> proximity caps <strong>and</strong> electron-beam<br />

evaporated SiO 2 cap layers, respectively. Reduced photolum<strong>in</strong>escence l<strong>in</strong>e-widths <strong>and</strong><br />

<strong>in</strong>creased photolum<strong>in</strong>escence <strong>in</strong>tensities were observed after diffusion-<strong>in</strong>duced b<strong>and</strong>-gap<br />

shifts that ranged from 0.006 to 0.220eV. Microscopic photolum<strong>in</strong>escence methods were<br />

used to study the onset of stra<strong>in</strong> relaxation due to dislocation generation. Two types of l<strong>in</strong>e<br />

defect were found <strong>in</strong> samples which had been annealed us<strong>in</strong>g proximity caps; depend<strong>in</strong>g<br />

upon the anneal<strong>in</strong>g temperature <strong>and</strong> the number of quantum wells. These were misfit<br />

dislocations with their l<strong>in</strong>es parallel to directions, <strong>and</strong> -oriented l<strong>in</strong>e defects.<br />

No dislocations were observed, <strong>in</strong> samples which had been annealed us<strong>in</strong>g a SiO 2 cap, over<br />

the entire temperature range which was <strong>in</strong>vestigated here. Resonant Raman scatter<strong>in</strong>g<br />

measurements of the 1LO/2LO phonon <strong>in</strong>tensity ratio were used to make semi-quantitative<br />

assessments of the total defect content; <strong>in</strong>clud<strong>in</strong>g po<strong>in</strong>t defects. It was found that, whereas<br />

<strong>in</strong>creas<strong>in</strong>g po<strong>in</strong>t defect densities, <strong>and</strong> the formation of l<strong>in</strong>e defects, were observed <strong>in</strong><br />

proximity-capped samples as the anneal<strong>in</strong>g temperature was <strong>in</strong>creased, no deterioration of<br />

structural quality due to an <strong>in</strong>creased po<strong>in</strong>t defect density was observed <strong>in</strong> samples which<br />

had been annealed us<strong>in</strong>g a SiO 2 cap.<br />

S.Bürkner, M.Baeumler, J.Wagner, E.C.Lark<strong>in</strong>s, W.Rothemund, J.D.Ralston: Journal of<br />

Applied Physics, 1996, 79[9], 6818-25<br />

[446-134/135-139]<br />

In<strong>GaAs</strong>/<strong>GaAs</strong>/Si: Interdiffusion<br />

Heterostructures of the form, In x Ga 1-x As(5nm)/<strong>GaAs</strong>(5nm)/Si, were fabricated via<br />

metalorganic chemical vapor deposition <strong>and</strong> were annealed at 750C. A transmission electron<br />

microscopic study was made of the heterostructure, with particular attention be<strong>in</strong>g paid to<br />

<strong>in</strong>terdiffusion between group-<strong>III</strong> elements. Plan-view transmission electron microscopy<br />

revealed that the epitaxial In x Ga 1-x As /<strong>GaAs</strong> layer consisted basically of small isl<strong>and</strong><br />

crystals. The morphology of the isl<strong>and</strong>s varied as a function of the In content. When x was<br />

greater than 0.1, the isl<strong>and</strong>s coalesced <strong>in</strong>to larger ones; leav<strong>in</strong>g small regions of so-called<br />

sea. The spac<strong>in</strong>gs of 022 moiré fr<strong>in</strong>ges varied spatially as a result of variations <strong>in</strong> In content<br />

<strong>in</strong> the epilayers. Cross-sectional transmission electron microscopic observations showed that<br />

there was no sharp heteroboundary between the In x Ga 1-x As layer <strong>and</strong> the <strong>GaAs</strong> layer. The<br />

contrast of the 002 dark-field image was sensitive to the In content <strong>and</strong> revealed<br />

<strong>in</strong>terdiffusion between In <strong>and</strong> Ga.<br />

K.Kamei, K.Fujita, Y.Shiba, H.Katahama, Y.Maehara: Defect <strong>and</strong> <strong>Diffusion</strong> Forum,<br />

1993, 95-98, 977-82<br />

[446-95/98-977]<br />

426


Interdiffusion (Ga,In)As Interdiffusion<br />

GaInAs/GaInAsP: Interdiffusion<br />

Electron microscopy was used to characterize metalorganic chemical vapor deposited<br />

multiple quantum well structures which underwent a so-called blue shift <strong>in</strong> lum<strong>in</strong>escence<br />

dur<strong>in</strong>g thermal process<strong>in</strong>g. The sample exhibited a shift, towards shorter wavelengths, of<br />

more than 100nm dur<strong>in</strong>g anneal<strong>in</strong>g at 750C. The structural modifications which led to the<br />

blue shift <strong>in</strong>cluded the elim<strong>in</strong>ation of atomic order<strong>in</strong>g <strong>in</strong> the quaternary barrier layers of<br />

the material, plus appreciable layer <strong>in</strong>terdiffusion. A method was described by which<br />

quantitative analyses of the layer composition <strong>and</strong> lattice parameter could be obta<strong>in</strong>ed, at<br />

a spatial resolution of better than 2nm, by mak<strong>in</strong>g energy-dispersive X-ray microanalyses<br />

<strong>in</strong> a scann<strong>in</strong>g transmission electron microscope. Such analyses showed that <strong>in</strong>terdiffusion<br />

occurred along a non-l<strong>in</strong>ear (non lattice-matched) path, where the group-V diffusivities<br />

exceeded those of group-<strong>III</strong> elements. This resulted <strong>in</strong> the <strong>in</strong>corporation of excess<br />

coherency stra<strong>in</strong>s, of up to 0.5%, <strong>in</strong> the quantum-well regions.<br />

R.E.Mallard, N.J.Long, E.J.Thrush, K.Scarrott, A.G.Norman, G.R.Booker: Journal of<br />

Applied Physics, 1993, 73[9], 4297-304<br />

[446-106/107-091]<br />

In<strong>GaAs</strong>/<strong>GaAs</strong>: Interdiffusion<br />

The compositional disorder<strong>in</strong>g of superlattices which had been grown by means of<br />

molecular beam epitaxy, us<strong>in</strong>g a low-temperature <strong>GaAs</strong> cap layer, was studied.<br />

Disorder<strong>in</strong>g of the superlattice was verified by photolum<strong>in</strong>escence <strong>and</strong> double-crystal X-<br />

ray rock<strong>in</strong>g curve measurements. The disorder<strong>in</strong>g mechanism was found to <strong>in</strong>volve Ga<br />

vacancy-enhanced <strong>in</strong>terdiffusion, due to the presence of the low-temperature <strong>GaAs</strong>. The<br />

diffusion <strong>and</strong> Schröd<strong>in</strong>ger’s equations were solved numerically <strong>in</strong> order to obta<strong>in</strong> the<br />

compositional profile <strong>and</strong> the transition energies <strong>in</strong> the disordered quantum well,<br />

respectively. The simulated energy shifts for samples under various anneal<strong>in</strong>g conditions<br />

agreed well with experimental data. The calculated effective diffusivity for In-Ga<br />

<strong>in</strong>terdiffusion <strong>in</strong>volved an activation energy of 1.63eV. This was smaller than the<br />

activation energy, of 1.93eV, for <strong>in</strong>tr<strong>in</strong>sic <strong>in</strong>terdiffusion. The diffusivity for the enhanced<br />

In-Ga <strong>in</strong>terdiffusion, due to the presence of low-temperature <strong>GaAs</strong>, was some 2 orders of<br />

magnitude larger than the <strong>in</strong>tr<strong>in</strong>sic In-Ga diffusivity.<br />

J.S.Tsang, C.P.Lee, S.H.Lee, K.L.Tsai, C.M.Tsai, J.C.Fan: Journal of Applied Physics,<br />

1996, 79[2], 664-70<br />

[446-131/132-179]<br />

In<strong>GaAs</strong>/<strong>GaAs</strong>: Interdiffusion<br />

The effect of stra<strong>in</strong> upon cation <strong>in</strong>terdiffusion <strong>in</strong> quantum wells was described. It was<br />

found that Fick’s diffusion equation did not correctly describe <strong>in</strong>terdiffusion <strong>in</strong> a<br />

heterostructure with stra<strong>in</strong>ed layers. It was suggested that the stra<strong>in</strong> altered the crystal<br />

defect concentration, <strong>and</strong> that the diffusivity was therefore affected by the stra<strong>in</strong>. A<br />

diffusion equation which <strong>in</strong>cluded the effects of stra<strong>in</strong> was derived <strong>and</strong> solved<br />

numerically. Experimental photolum<strong>in</strong>escence peak shifts, as a function of anneal<strong>in</strong>g<br />

427


Interdiffusion (Ga,In)As Interdiffusion<br />

time, were closely fitted by this analysis <strong>and</strong> useful parameters such as a coefficient<br />

which described In<strong>GaAs</strong>/<strong>GaAs</strong> quantum well <strong>in</strong>terdiffusion were deduced.<br />

S.W.Ryu, I.Kim, B.D.Choe, W.G.Jeong: Applied Physics Letters, 1995, 67[10], 1417-9<br />

[446-125/126-141]<br />

Table 38<br />

Interdiffusivity <strong>in</strong> Kr-Implanted In<strong>GaAs</strong>/<strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

1050 7.71 x 10 -15<br />

1000 3.11 x 10 -15<br />

950 5.90 x 10 -16<br />

925 3.60 x 10 -16<br />

900 1.10 x 10 -16<br />

875 9.00 x 10 -17<br />

825 1.50 x 10 -17<br />

750 8.00 x 10 -19<br />

In<strong>GaAs</strong>/<strong>GaAs</strong>: Interdiffusion<br />

Low-pressure metalorganic vapor phase epitaxially grown stra<strong>in</strong>ed quantum well<br />

structures were characterized by us<strong>in</strong>g photolum<strong>in</strong>escence <strong>and</strong> X-ray diffraction<br />

techniques. It was shown that, beyond the pseudomorphic limit, these structures exhibited<br />

considerable Ga/In <strong>in</strong>terdiffusion at the <strong>in</strong>terfaces, <strong>and</strong> partial stra<strong>in</strong> relaxation <strong>in</strong> the<br />

quantum well layers.<br />

A.K.Srivastava, B.M.Arora, S.Banerjee: Journal of Electronic Materials, 1994, 23[2],<br />

191-4<br />

[446-113/114-036]<br />

338,39,40,41 In<strong>GaAs</strong>/<strong>GaAs</strong>: Interdiffusion<br />

Photolum<strong>in</strong>escence techniques <strong>and</strong> repeated anneal<strong>in</strong>g were used to determ<strong>in</strong>e the<br />

diffusion coefficients for <strong>in</strong>termix<strong>in</strong>g <strong>in</strong> quantum wells <strong>and</strong> to study the subsequent<br />

effects of ion implantation upon <strong>in</strong>termix<strong>in</strong>g. It was shown that, after ion implantation, a<br />

very fast <strong>in</strong>terdiffusion process occurred which was <strong>in</strong>dependent of the nature of the<br />

implanted ion <strong>and</strong> was thought to be due to the rapid diffusion of <strong>in</strong>terstitials which were<br />

created dur<strong>in</strong>g implantation. Follow<strong>in</strong>g this rapid process, it was found that neither Ga<br />

nor Kr ions had any effect upon the subsequent <strong>in</strong>terdiffusion coefficient (tables 38 <strong>and</strong><br />

39). After As implantation, <strong>in</strong> addition to the <strong>in</strong>itial damage-related process, an enhanced<br />

region of <strong>in</strong>terdiffusion was observed; with a diffusion coefficient which was an order of<br />

magnitude greater than that of an non-implanted control wafer (tables 40 <strong>and</strong> 41). This<br />

enhancement was suggested to be due to the creation of group-<strong>III</strong> vacancies by As atoms<br />

which moved <strong>in</strong>to group-V lattice sites. The fast process cont<strong>in</strong>ued until the structure had<br />

broadened by about 7.5nm, whereupon the diffusion coefficient returned to the non-<br />

428


Interdiffusion (Ga,In)As Interdiffusion<br />

implanted control value. The activation energy for <strong>in</strong>terdiffusion was measured at<br />

temperatures rang<strong>in</strong>g from 1050 to 750C, <strong>and</strong> a value of 3.7eV was deduced. This value<br />

was found to be <strong>in</strong>dependent of the nature of the implanted ion.<br />

I.V.Bradley, W.P.Gill<strong>in</strong>, K.P.Homewood, R.P.Webb: Journal of Applied Physics, 1993,<br />

73[4], 1686-92<br />

[446-109/110-041]<br />

Table 39<br />

Interdiffusivity <strong>in</strong> Ga-Implanted In<strong>GaAs</strong>/<strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

1050 2.68 x 10 -15<br />

1000 9.10 x 10 -16<br />

950 2.20 x 10 -16<br />

925 1.30 x 10 -16<br />

900 4.10 x 10 -17<br />

875 3.10 x 10 -17<br />

825 4.30 x 10 -18<br />

750 2.00 x 10 -19<br />

342 In<strong>GaAs</strong>/<strong>GaAs</strong>: Interdiffusion<br />

Interdiffusion <strong>in</strong> quantum wells was studied by us<strong>in</strong>g photolum<strong>in</strong>escence methods to<br />

monitor the temporal development of the diffusion process <strong>in</strong> a s<strong>in</strong>gle sample (table 42).<br />

Two dist<strong>in</strong>ct regimes were detected: a fast <strong>in</strong>itial diffusion <strong>and</strong> a second, steady-state,<br />

diffusion. The steady-state diffusion was found to depend upon the depth of the quantum<br />

well from the surface, <strong>and</strong> could be correlated with published data on the diffusion of Ga<br />

vacancies <strong>in</strong>to <strong>GaAs</strong>.<br />

W.P.Gill<strong>in</strong>, D.J.Dunstan, K.P.Homewood, L.K.Howard, B.J.Sealy: Journal of Applied<br />

Physics, 1993, 73[8], 3782-6<br />

[446-109/110-042]<br />

In<strong>GaAs</strong>/<strong>GaAs</strong>: Interdiffusion<br />

The <strong>in</strong>terdiffusion of s<strong>in</strong>gle quantum wells was studied as a function of temperature for<br />

both p-type (Be) <strong>and</strong> n-type (Si) dop<strong>in</strong>g. The <strong>in</strong>terdiffusion of group-<strong>III</strong> elements was<br />

monitored via photolum<strong>in</strong>escence from the ground states of valence <strong>and</strong> conduction b<strong>and</strong><br />

quantum wells. Intermix<strong>in</strong>g was modelled by us<strong>in</strong>g a Green's function method to solve the<br />

diffusion equation that described the evolution of well shapes dur<strong>in</strong>g process<strong>in</strong>g. It was<br />

deduced that the activation energy for <strong>in</strong>terdiffusion was 3.4eV.<br />

W.P.Gill<strong>in</strong>, K.P.Homewood, L.K.Howard, M.T.Emeny: Superlattices <strong>and</strong><br />

Microstructures, 1991, 9[1], 39-42<br />

[446-78/79-042]<br />

429


Interdiffusion (Ga,In)As Interdiffusion<br />

Table 40<br />

Interdiffusivity <strong>in</strong> Non-Implanted In<strong>GaAs</strong>/<strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

1050 3.42 x 10 -15<br />

1000 8.45 x 10 -16<br />

950 1.97 x 10 -16<br />

925 1.77 x 10 -16<br />

900 4.25 x 10 -17<br />

875 2.27 x 10 -17<br />

825 1.00 x 10 -17<br />

750 2.00 x 10 -19<br />

In<strong>GaAs</strong>/<strong>GaAs</strong>: Interdiffusion<br />

It was shown that the Si-dop<strong>in</strong>g of quantum-well materials to 10 19 /cm 3 promoted a time<strong>and</strong><br />

temperature-dependent diffusion process which was related to group-<strong>III</strong> vacancy<br />

formation. The effect of the formation of such vacancies upon subsequent <strong>in</strong>terdiffusion<br />

was modelled <strong>and</strong> was shown to reproduce variations, <strong>in</strong> the diffusion coefficient as a<br />

function of depth, without <strong>in</strong>vok<strong>in</strong>g a Fermi-level model. Experiments which were<br />

performed on layers that were doped to between 10 17 <strong>and</strong> 10 18 /cm 3 did not reveal an<br />

enhanced <strong>in</strong>terdiffusion; contrary to the predictions of the Fermi-level model. Other<br />

results suggested that the reason for this was that <strong>in</strong>terdiffusion <strong>in</strong> <strong>III</strong>-V materials was not<br />

governed by thermal equilibrium vacancy concentrations but rather by the vacancy<br />

concentrations which were grown <strong>in</strong>to the substrate materials. It was demonstrated that<br />

the position of the Fermi level played no role <strong>in</strong> <strong>III</strong>-V <strong>in</strong>termix<strong>in</strong>g.<br />

Z.H.Jafri, W.P.Gill<strong>in</strong>: Journal of Applied Physics, 1997, 81[5], 2179-84<br />

[446-148/149-182]<br />

In<strong>GaAs</strong>/InAlAs: Interdiffusion<br />

Quantum well structures were grown on InP(Fe) substrates by us<strong>in</strong>g metalorganic vapor<br />

phase epitaxial techniques. Each sample conta<strong>in</strong>ed 3 In<strong>GaAs</strong> wells, which were 2.6, 5.9,<br />

or 17.6nm <strong>in</strong> thickness <strong>and</strong> were separated by 24nm-thick InAlAs barrier layers. The<br />

samples were implanted with Si ions to uniform densities which ranged from 1.8 x 10 17 to<br />

3.9 x 10 19 /cm 3 over the quantum wells, <strong>and</strong> were then annealed under various conditions.<br />

A photolum<strong>in</strong>escence peak energy for each well was monitored <strong>in</strong> order to study<br />

<strong>in</strong>termix<strong>in</strong>g at the <strong>in</strong>terface. Blue shifts <strong>in</strong> the photolum<strong>in</strong>escence peak energy were found<br />

to occur with<strong>in</strong> the first 15s of thermal anneal<strong>in</strong>g when the Si dose exceeded a critical<br />

value of 2 x 10 18 to 3 x 10 18 /cm 3 . The saturation value of the energy shift was governed<br />

ma<strong>in</strong>ly by the Si density, but hardly depended upon the anneal<strong>in</strong>g temperature <strong>and</strong> time. It<br />

was concluded that the defects which were formed by Si ion implantation enhanced the<br />

430


Interdiffusion (Ga,In)As Interdiffusion<br />

thermal <strong>in</strong>terdiffusion of Ga <strong>and</strong> Al atoms at the In<strong>GaAs</strong>/InAlAs <strong>in</strong>terface. This ended<br />

when the implantation-<strong>in</strong>duced defects annealed out.<br />

S.Yamamura, R.Saito, S.Yugo, T.Kimura, M.Murata, T.Kamiya: Journal of Applied<br />

Physics, 1994, 75[5], 2410-4<br />

[446-117/118-184]<br />

Table 41<br />

Interdiffusivity <strong>in</strong> As-Implanted In<strong>GaAs</strong>/<strong>GaAs</strong><br />

Temperature (C) <strong>Diffusion</strong> Type D (cm 2 /s)<br />

1050 steady-state 6.42 x 10 -15<br />

1000 steady-state 3.90 x 10 -15<br />

950 steady-state 6.96 x 10 -16<br />

950 enhanced 3.50 x 10 -15<br />

925 steady-state 4.50 x 10 -16<br />

925 enhanced 1.73 x 10 -15<br />

900 steady-state 3.05 x 10 -16<br />

900 enhanced 1.01 x 10 -15<br />

750 enhanced 3.30 x 10 -18<br />

In<strong>GaAs</strong>/In<strong>GaAs</strong>P: Interdiffusion<br />

An <strong>in</strong>vestigation was made of the effect of <strong>in</strong>terdiffusion upon the photolum<strong>in</strong>escence <strong>and</strong><br />

Raman spectra of a s<strong>in</strong>gle quantum well of In<strong>GaAs</strong> (8nm wide) which was s<strong>and</strong>wiched<br />

between 2 large In<strong>GaAs</strong>P barriers. Firstly, an <strong>in</strong>vestigation was made of a blue shift of the<br />

recomb<strong>in</strong>ation l<strong>in</strong>e (2K) which occurred after anneal<strong>in</strong>g at 650 or 750C for times which<br />

ranged from 0.25 to 2h. A s<strong>in</strong>gle diffusivity coefficient was assumed for all of the atomic<br />

species, <strong>and</strong> the amount of <strong>in</strong>termix<strong>in</strong>g was deduced by means of model calculations.<br />

Average diffusivity coefficients of 0.0095 <strong>and</strong> 0.2Ų/s were found at 650 <strong>and</strong> 750C,<br />

respectively. This agreed well with published data on In<strong>GaAs</strong>/InP, <strong>and</strong> suggested that the<br />

activation energy was 2.54eV.<br />

H.Peyre, F.Als<strong>in</strong>a, J.Camassel, J.Pascual, R.W.Glew: Journal of Applied Physics, 1993,<br />

73[8], 3760-8<br />

[446-106/107-117]<br />

In<strong>GaAs</strong>/In<strong>GaAs</strong>P: Interdiffusion<br />

Thermal <strong>in</strong>terdiffusion on the group-V sub-lattice <strong>in</strong> quantum-well structures was studied<br />

by us<strong>in</strong>g samples which were annealed under silicon nitride encapsulation, or under<br />

various phosph<strong>in</strong>e over-pressures. It was found that the <strong>in</strong>terdiffusion length was<br />

comparable, under all of these conditions, with only small effects of the phosph<strong>in</strong>e overpressure<br />

be<strong>in</strong>g observed. It was suggested that <strong>in</strong>terdiffusion results which were obta<strong>in</strong>ed<br />

431


Interdiffusion (Ga,In)As Interdiffusion<br />

for nitride-capped samples could be applied to the <strong>in</strong>terdiffusion which occurred dur<strong>in</strong>g<br />

growth.<br />

W.P.Gill<strong>in</strong>, S.D.Perr<strong>in</strong>, K.P.Homewood: Journal of Applied Physics, 1995, 77[4], 1463-5<br />

[446-121/122-078]<br />

Table 42<br />

Interdiffusion <strong>in</strong> In<strong>GaAs</strong>/<strong>GaAs</strong><br />

Temperature (C) D (cm 2 /s)<br />

900 4.6 x 10 -17<br />

950 2.6 x 10 -16<br />

1000 6.4 x 10 -16<br />

1050 1.6 x 10 -15<br />

In<strong>GaAs</strong>/InP: Interdiffusion<br />

A formula was derived which described the <strong>in</strong>terdiffusion profiles of quantum wells. It<br />

was shown that it accurately modelled <strong>in</strong>terdiffusion <strong>in</strong> quantum wells of lattice-matched<br />

In<strong>GaAs</strong>. The formula took account of the differ<strong>in</strong>g <strong>in</strong>terdiffusion coefficients between<br />

layers, <strong>and</strong> of the <strong>in</strong>terfacial discont<strong>in</strong>uity of <strong>in</strong>terdiffused species. The formula expla<strong>in</strong>ed<br />

how quantum energy shifts due to <strong>in</strong>terdiffusion varied with anneal<strong>in</strong>g time <strong>and</strong> anneal<strong>in</strong>g<br />

temperature <strong>in</strong> various wide-well layers of both In<strong>GaAs</strong>P/InP <strong>and</strong> <strong>GaAs</strong>/Al<strong>GaAs</strong> quantum<br />

wells. The quantitative difference between the <strong>in</strong>terdiffusion profiles <strong>in</strong> these two<br />

materials was also demonstrated.<br />

K.Mukai, M.Sugawara, S.Yamazaki: Physical Review B, 1994, 50[4], 2273-82<br />

[446-115/116-130]<br />

In<strong>GaAs</strong>/InP: Interdiffusion<br />

An <strong>in</strong>vestigation was made of <strong>in</strong>termix<strong>in</strong>g effects, <strong>in</strong> In 0.53 Ga 0.47 As s<strong>in</strong>gle quantum wells,<br />

which were caused by 30keV Ar + ion beam implantation to doses which ranged from 10 12<br />

to 10 14 /cm 2 <strong>and</strong> by subsequent rapid thermal anneal<strong>in</strong>g at temperatures of between 600<br />

<strong>and</strong> 900C. After implantation <strong>and</strong> rapid thermal anneal<strong>in</strong>g at 600C, a significant <strong>in</strong>crease<br />

was observed (<strong>in</strong> the photolum<strong>in</strong>escence emission energy of about 0.06eV) as compared<br />

with non-implanted heterostructures. This <strong>in</strong>dicated that the <strong>in</strong>termix<strong>in</strong>g was produced by<br />

implantation. However, <strong>in</strong> the case of rapid thermal anneal<strong>in</strong>g at temperatures above<br />

850C, the energy shifts (of up to 0.2eV) which were observed <strong>in</strong> implanted samples were<br />

similar to the shifts that occurred <strong>in</strong> non-implanted samples. This <strong>in</strong>dicated that a<br />

prom<strong>in</strong>ent contribution arose from thermal <strong>in</strong>terdiffusion. A significant decrease <strong>in</strong> the Ga<br />

concentration after <strong>in</strong>terdiffusion was confirmed by Raman data.<br />

J.Osh<strong>in</strong>owo, J.Dreybrodt, A.Forchel, N.Mestres, J.M.Calleja, I.Gyuro, P.Speier,<br />

E.Ziel<strong>in</strong>ski: Journal of Applied Physics, 1993, 74[3], 1983-6<br />

[446-106/107-117]<br />

432


Interdiffusion (Ga,In)As Interdiffusion<br />

In<strong>GaAs</strong>/InP: Interdiffusion<br />

The effects of differ<strong>in</strong>g cation <strong>and</strong> anion <strong>in</strong>terdiffusion rates upon the disorder<strong>in</strong>g of<br />

In 0.53 Ga 0.47 As/InP s<strong>in</strong>gle quantum-wells were <strong>in</strong>vestigated by us<strong>in</strong>g an erf distribution to<br />

model the concentration profile after <strong>in</strong>terdiffusion. The early stages of disorder<strong>in</strong>g<br />

caused a spatially dependent stra<strong>in</strong> build-up, which could be compressive or tensile. This<br />

stra<strong>in</strong> profile, <strong>and</strong> the concentration distribution, gave rise to <strong>in</strong>terest<strong>in</strong>g carrier<br />

conf<strong>in</strong>ement profiles after disorder<strong>in</strong>g. A significantly higher cation <strong>in</strong>terdiffusion rate<br />

produced a red-shift of the ground-state transition energy which, with prolonged<br />

<strong>in</strong>terdiffusion, saturated <strong>and</strong> then decreased. A significantly higher anion <strong>in</strong>terdiffusion<br />

rate caused a blue-shift <strong>in</strong> the ground-state transition energy, <strong>and</strong> shifted the light-hole<br />

ground-state to above the heavy-hole ground-state.<br />

W.C.Shiu, J.Micallef, I.Ng, E.H.Li: Japanese Journal of Applied Physics, 1995, 34[1-<br />

4A], 1778-83<br />

[446-121/122-078]<br />

In<strong>GaAs</strong>/InP: Interdiffusion<br />

The <strong>in</strong>terdiffusion of superlattice structures, dur<strong>in</strong>g anneal<strong>in</strong>g at temperatures of between<br />

700 <strong>and</strong> 850C, was studied by means of Raman spectroscopy. The peak <strong>in</strong>tensities <strong>and</strong><br />

peak energies of InAs-, <strong>GaAs</strong>-, <strong>and</strong> InP-like longitud<strong>in</strong>al optical phonon modes varied as<br />

a function of anneal<strong>in</strong>g temperature <strong>and</strong> time. The depth profiles of group-<strong>III</strong> <strong>and</strong> group-V<br />

atoms were estimated quantitatively by monitor<strong>in</strong>g variations <strong>in</strong> the peak energies of the<br />

longitud<strong>in</strong>al optical phonon modes, <strong>and</strong> <strong>in</strong> the ratios of the mode <strong>in</strong>tensities. This revealed<br />

that the <strong>in</strong>terdiffused superlattice consisted of In<strong>GaAs</strong>P of uniform composition, <strong>and</strong> InP<br />

barrier layers with sharp <strong>in</strong>terfaces. It was found that the resultant In<strong>GaAs</strong>P well was<br />

roughly lattice-matched to InP (to with<strong>in</strong> 0.5%). The diffusion coefficient <strong>in</strong> the well<br />

region was larger than that <strong>in</strong> the barrier region, <strong>and</strong> the <strong>in</strong>terdiffusion could be described<br />

by:<br />

D(cm 2 /s) = 8.56 x 10 10 exp[-5.82(eV)/kT]<br />

Interdiffusion <strong>in</strong> this superlattice was governed by diffusion <strong>in</strong> the InP region.'<br />

S.J.Yu, H.Asashi, S.Emura, S.Gonda, K.Nakashima: Journal of Applied Physics, 1991,<br />

70[1], 204-8<br />

[446-93/94-039]<br />

In<strong>GaAs</strong>/InP: Interdiffusion<br />

The thermal stability <strong>and</strong> <strong>in</strong>terdiffusion of In 0.53 Ga 0.47 As/InP surface quantum wells was<br />

<strong>in</strong>vestigated. Well-def<strong>in</strong>ed high-<strong>in</strong>tensity photolum<strong>in</strong>escence emission spectra were<br />

obta<strong>in</strong>ed. After rapid thermal anneal<strong>in</strong>g (500 to 900C, 60s), strong emission energy shifts<br />

of up to 0.316eV were detected. By assum<strong>in</strong>g a simple model for ion <strong>in</strong>termix<strong>in</strong>g, the<br />

<strong>in</strong>terdiffusion coefficient was estimated to be equal to 1.7 x 10 -14 cm 2 /s at 900C; with an<br />

activation energy of 1.3eV.<br />

J.Osh<strong>in</strong>owo, A.Forchel, D.Grützmacher, M.Stollenwerk, M.Heuken, K.Heime: Applied<br />

Physics Letters, 1992, 60[21], 2660-2<br />

[446-88/89-044]<br />

433


Interdiffusion (Ga,In)As Interdiffusion<br />

In<strong>GaAs</strong>/InP: Interdiffusion<br />

Quantum-well structures were studied us<strong>in</strong>g magneto-optical transmission spectroscopy.<br />

The effects of dopants, overgrowth <strong>and</strong> anneal<strong>in</strong>g were <strong>in</strong>vestigated. The blue-shift effect,<br />

which was often observed <strong>in</strong> multiple quantum-well structures that were subjected to<br />

heat-treatment, was here attributed to a dom<strong>in</strong>ant group-V <strong>in</strong>terdiffusion which could be<br />

suppressed by high defect densities <strong>in</strong> the substrate. The presence of Zn <strong>in</strong> an overgrown<br />

layer on top of the multiple quantum-well structures caused a counteractive red-shift<br />

effect after long anneal<strong>in</strong>g times. This was due to group-<strong>III</strong> diffusion. On the <strong>other</strong> h<strong>and</strong>,<br />

<strong>in</strong> situ Zn or S produced no observable shift <strong>in</strong> transition energies due to <strong>in</strong>terdiffusion.<br />

This was attributed to an enhanced group-<strong>III</strong> <strong>in</strong>terdiffusion that was <strong>in</strong>duced by Zn<br />

diffusion <strong>in</strong>to the multiple quantum-wells. It was concluded that very different<br />

<strong>in</strong>terdiffusion mechanisms operated <strong>in</strong> the case of group-<strong>III</strong> <strong>and</strong> group-V elements; thus<br />

support<strong>in</strong>g the suggestion of vacancy-related group-V <strong>in</strong>terdiffusion rather than the<br />

<strong>in</strong>terstitialcy mechanism of group-<strong>III</strong> <strong>in</strong>terdiffusion.<br />

S.L.Wong, R.J.Nicholas, R.W.Mart<strong>in</strong>, J.Thompson, A.Wood, A.Moseley, N.Carr: Journal<br />

of Applied Physics, 1996, 79[9], 6826-33<br />

[446-134/135-148]<br />

434


(Ga,In)(As,P)<br />

As<br />

In<strong>GaAs</strong>P/InP: As <strong>Diffusion</strong><br />

The effect of diffusion upon quantum-well structures was studied by us<strong>in</strong>g secondary ion<br />

mass spectrometry, Auger electron spectroscopy, capacitance-voltage measurements,<br />

photolum<strong>in</strong>escence, <strong>and</strong> X-ray diffraction methods. It was found that the <strong>in</strong>terdiffusion of<br />

As was negligible.<br />

G.J.Van Gurp, W.M.Van de Wijgert, G.M.Fontijn, P.J.A.Thijs: Journal of Applied<br />

Physics, 1990, 67[6], 2919-26<br />

[446-74-040]<br />

Be<br />

In<strong>GaAs</strong>P/InP: Be <strong>Diffusion</strong><br />

A study was made of dopant redistribution between a chemical beam epitaxially grown<br />

In<strong>GaAs</strong>P laser structure, <strong>and</strong> a metalorganic vapor phase epitaxially over-grown InP<br />

layer. Secondary ion mass spectroscopic data revealed that Zn <strong>and</strong> Be atoms <strong>in</strong>terdiffused<br />

deeply <strong>in</strong>to the adjacent InP layers, at a Zn dop<strong>in</strong>g level of 10 18 /cm 3 . A fraction of the Zn<br />

atoms went through the chemical beam epitaxial InP, <strong>and</strong> penetrated the laser structure<br />

guide layer. It was found that Zn out-diffusion was appreciably suppressed by reduc<strong>in</strong>g<br />

the Be dopant concentration from 10 18 to 5 x 10 17 /cm 3 .<br />

H.Sugiura, S.Kondo, M.Mitsuhara, S.Matsumoto, M.Itoh: Applied Physics Letters, 1997,<br />

70[21], 2846-8<br />

[446-152-0435]<br />

Ga<br />

In<strong>GaAs</strong>P/InP: Ga <strong>Diffusion</strong><br />

The effect of diffusion upon quantum-well structures was studied by us<strong>in</strong>g secondary ion<br />

mass spectrometry, Auger electron spectroscopy, capacitance-voltage measurements,<br />

435


Ga (Ga,In)(As,P) Si<br />

photolum<strong>in</strong>escence, <strong>and</strong> X-ray diffraction methods. It was found that there was marked<br />

<strong>in</strong>terdiffusion of Ga. In a multiple quantum well, Zn diffusion at 500C caused Ga<br />

<strong>in</strong>termix<strong>in</strong>g <strong>and</strong> the Auger electron spectroscopic profiles became flat. At higher<br />

temperatures, an order<strong>in</strong>g was found such that the Ga concentration was greatest <strong>in</strong> the<br />

orig<strong>in</strong>al InP layers. This was attributed to the effect of m<strong>in</strong>imization of the free energy;<br />

balanced by an <strong>in</strong>crease <strong>in</strong> the mismatch stra<strong>in</strong> energy. Photolum<strong>in</strong>escence data revealed<br />

that <strong>in</strong>terdiffusion began at temperatures above 420C.<br />

G.J.Van Gurp, W.M.Van de Wijgert, G.M.Fontijn, P.J.A.Thijs: Journal of Applied<br />

Physics, 1990, 67[6], 2919-26<br />

[446-74-040]<br />

In<br />

In<strong>GaAs</strong>P/InP: In <strong>Diffusion</strong><br />

The effect of diffusion upon quantum-well structures was studied by us<strong>in</strong>g secondary ion<br />

mass spectrometry, Auger electron spectroscopy, capacitance-voltage measurements,<br />

photolum<strong>in</strong>escence, <strong>and</strong> X-ray diffraction methods. It was found that there was marked<br />

<strong>in</strong>terdiffusion of In. In a multiple quantum well, Zn diffusion at 500C caused In<br />

<strong>in</strong>termix<strong>in</strong>g <strong>and</strong> the Auger electron spectroscopic profiles became flat.<br />

G.J.Van Gurp, W.M.Van de Wijgert, G.M.Fontijn, P.J.A.Thijs: Journal of Applied<br />

Physics, 1990, 67[6], 2919-26<br />

[446-74-040]<br />

P<br />

In<strong>GaAs</strong>P/InP: P <strong>Diffusion</strong><br />

The effect of diffusion upon quantum-well structures was studied by us<strong>in</strong>g secondary ion<br />

mass spectrometry, Auger electron spectroscopy, capacitance-voltage measurements,<br />

photolum<strong>in</strong>escence, <strong>and</strong> X-ray diffraction methods. It was found that the <strong>in</strong>terdiffusion of<br />

P was negligible.<br />

G.J.Van Gurp, W.M.Van de Wijgert, G.M.Fontijn, P.J.A.Thijs: Journal of Applied<br />

Physics, 1990, 67[6], 2919-26<br />

[446-74-040]<br />

Si<br />

GaInAsP/<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The effect of concurrent Si <strong>and</strong> Zn diffusion, upon <strong>in</strong>terdiffusion between the cation <strong>and</strong><br />

anion sub-lattices, was studied <strong>in</strong> Ga 0.95 In 0.05 As 0.95 P 0.05 /<strong>GaAs</strong> heterostructures which had<br />

been grown by us<strong>in</strong>g liquid-phase epitaxy techniques. The diffusion sources were<br />

equilibrium ternary tie-triangle compositions. The extent of <strong>in</strong>terdiffusion of both group-<br />

<strong>III</strong> <strong>and</strong> group-V atoms was determ<strong>in</strong>ed by depth profil<strong>in</strong>g In <strong>and</strong> P, respectively, us<strong>in</strong>g<br />

436


Si (Ga,In)(As,P) Zn<br />

secondary ion mass spectrometry. It was found that Si diffusion enhanced both cation <strong>and</strong><br />

anion <strong>in</strong>terdiffusion to the same extent. A mono-vacancy mechanism was used to expla<strong>in</strong><br />

the effect of Si. It was concluded that the impurity diffusion mechanism was a major<br />

factor which affected the degree of enhancement.<br />

K.H.Lee, H.H.Park, D.A.Stevenson: Journal of Applied Physics, 1989, 65[3], 1048-52<br />

[446-72/73-038]<br />

In<strong>GaAs</strong>P/<strong>GaAs</strong>: Si <strong>Diffusion</strong><br />

The diffusion of Si <strong>III</strong> -Si V neutral pairs versus the diffusion of Si <strong>III</strong> -V <strong>III</strong> complexes <strong>in</strong> <strong>III</strong>-V<br />

crystals was considered with regard to experimental data which revealed the effect of Si<br />

diffusion upon the self-diffusion of column-<strong>III</strong> <strong>and</strong> column-V lattice atoms. Secondary<br />

ion mass spectroscopy was used to compare the enhanced diffusion of column-<strong>III</strong> or<br />

column-V atoms <strong>in</strong> various Si-diffused heterostructures which were closely latticematched<br />

to <strong>GaAs</strong>. An enhancement of lattice atom self-diffusion, due to impurity<br />

diffusion, was found to occur predom<strong>in</strong>antly on the column-<strong>III</strong> lattice. The data supported<br />

the Si <strong>III</strong> -V <strong>III</strong> diffusion model <strong>and</strong> <strong>in</strong>dicated that the ma<strong>in</strong> native defects which<br />

accompanied Si diffusion were column-<strong>III</strong> vacancies. These diffused directly on the<br />

column-<strong>III</strong> sub-lattice.<br />

D.G.Deppe, W.E.Plano, J.E.Baker, N.Holonyak, M.J.Ludowise, C.P.Kuo, R.M.Fletcher,<br />

T.D.Osentowski, M.G.Craford: Applied Physics Letters, 1988, 53[22], 2211-3<br />

[446-64/65-157]<br />

Zn<br />

In<strong>GaAs</strong>P: Zn <strong>Diffusion</strong><br />

The ampoule diffusion of Zn <strong>in</strong>to liquid-phase epitaxial layers was studied, at<br />

temperatures of between 425 <strong>and</strong> 525C, by means of secondary ion mass spectrometry. It<br />

was found that the <strong>in</strong>corporation <strong>and</strong> diffusion of Zn could be described <strong>in</strong> terms of the<br />

<strong>in</strong>terstitial-substitutional model. The difference between the acceptor <strong>and</strong> Zn<br />

concentrations was attributed to compensation by Zn <strong>in</strong>terstitial donors or by neutral Znvacancy<br />

complexes. The diffusion depth was slightly smaller than that <strong>in</strong> InP <strong>and</strong> was<br />

much larger than that <strong>in</strong> <strong>GaAs</strong>. In the case of n-type In<strong>GaAs</strong>P, the profiles exhibited a<br />

cut-off which was like that seen <strong>in</strong> InP.<br />

G.J.Van Gurp, D.L.A.Tjaden, G.M.Fontijn, P.R.Boudewijn: Journal of Applied Physics,<br />

1988, 64[7], 3468-71<br />

[446-72/73-037]<br />

In<strong>GaAs</strong>P: Zn <strong>Diffusion</strong><br />

Closed-ampoule diffusion led to a net acceptor concentration which was lower than the<br />

Zn concentration. Upon anneal<strong>in</strong>g <strong>in</strong> an atmosphere without Zn, the Zn <strong>and</strong> net acceptor<br />

concentrations became almost identical. This was attributed to a decreased Zn<br />

concentration <strong>and</strong> an <strong>in</strong>creased net acceptor concentration. The results were quantitatively<br />

expla<strong>in</strong>ed by assum<strong>in</strong>g that the Zn was <strong>in</strong>corporated as both substitutional acceptors <strong>and</strong><br />

437


Zn (Ga,In)(As,P) Zn<br />

<strong>in</strong>terstitial donors, <strong>and</strong> that only the <strong>in</strong>terstitial Zn was driven out by anneal<strong>in</strong>g; due to its<br />

large diffusion coefficient. The profiles which were calculated by us<strong>in</strong>g this <strong>in</strong>terstitialsubstitutional<br />

model could be fitted to experimentally determ<strong>in</strong>ed profiles by assum<strong>in</strong>g<br />

that the Zn diffused as s<strong>in</strong>gly-ionized <strong>in</strong>terstitial donors. The present model also<br />

expla<strong>in</strong>ed published data, on diffusion <strong>in</strong> n-type InP, <strong>in</strong> which a profile cut-off was found<br />

at a depth where the acceptor concentration equalled the background donor concentration.<br />

G.J.Van Gurp, T.Van Dongen, G.M.Fontijn, J.M.Jacobs, D.L.A.Tjaden: Journal of<br />

Applied Physics, 1989, 65[2], 553-60<br />

[446-72/73-037]<br />

In<strong>GaAs</strong>P: Zn <strong>Diffusion</strong><br />

Shallow diffusion was improved by us<strong>in</strong>g a new spun-on source which was based upon<br />

Zn-doped alum<strong>in</strong>a. In this case, the thermal expansion coefficients of the diffusion source<br />

<strong>and</strong> the semiconductor were better matched than when us<strong>in</strong>g Zn-doped silica films. As<br />

well as excellent mechanical stability of the spun-on films over a wide temperature range,<br />

the effect of mechanical stresses upon the diffusion process was effectively reduced.<br />

M.C.Amann, G.Franz: Journal of Applied Physics, 1987, 62[4], 1541-3<br />

[446-55/56-029]<br />

GaInAsP/<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The effect of concurrent Zn diffusion upon <strong>in</strong>terdiffusion <strong>in</strong> an Ga 0.94 In 0.06 As 0.95 P 0.05 -<br />

<strong>GaAs</strong> heterostructure, grown by means of liquid phase epitaxy, was <strong>in</strong>vestigated.<br />

<strong>Diffusion</strong> anneal<strong>in</strong>g (700C, 25h) was performed by us<strong>in</strong>g an equilibrium ternary diffusion<br />

source, <strong>and</strong> In <strong>and</strong> P profiles were measured by us<strong>in</strong>g secondary ion mass spectrometry. It<br />

was found that Zn diffusion selectively enhanced cation (In-Ga) <strong>in</strong>terdiffusion. In the case<br />

of concurrent Zn diffusion, the <strong>in</strong>terdiffusion coefficient for the In-Ga components was<br />

about 5 x 10 -14 cm 2 /s; as compared with about 6 x 10 -16 cm 2 /s for anions (As-P). A kick-out<br />

mechanism was suggested to expla<strong>in</strong> the results.<br />

H.H.Park, K.H.Lee, D.A.Stevenson: Applied Physics Letters, 1988, 53[23], 2299-301<br />

[446-64/65-174]<br />

GaInAsP/<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

The effect of concurrent Si <strong>and</strong> Zn diffusion, upon <strong>in</strong>terdiffusion between the cation <strong>and</strong><br />

anion sub-lattices, was studied <strong>in</strong> Ga 0.95 In 0.05 As 0.95 P 0.05 /<strong>GaAs</strong> heterostructures which had<br />

been grown by us<strong>in</strong>g liquid-phase epitaxy techniques. The diffusion sources were<br />

equilibrium ternary tie-triangle compositions. The extent of <strong>in</strong>terdiffusion of both group-<br />

<strong>III</strong> <strong>and</strong> group-V atoms was determ<strong>in</strong>ed by depth profil<strong>in</strong>g In <strong>and</strong> P, respectively, us<strong>in</strong>g<br />

secondary ion mass spectrometry. It was found that Zn diffusion selectively enhanced<br />

cation (In, Ga) <strong>in</strong>terdiffusion. A kick-out mechanism was used to expla<strong>in</strong> the selective<br />

enhancement of cation <strong>in</strong>terdiffusion by Zn. It was concluded that the impurity diffusion<br />

mechanism was a major factor which affected the degree of enhancement.<br />

K.H.Lee, H.H.Park, D.A.Stevenson: Journal of Applied Physics, 1989, 65[3], 1048-52<br />

[446-72/73-038]<br />

438


Zn (Ga,In)(As,P) Zn<br />

GaInAsP/InP: Zn <strong>Diffusion</strong><br />

It was shown that Zn diffusion <strong>in</strong>to a Ga x In 1-x As y P 1-y -InP quantum well structure <strong>and</strong><br />

superlattice, with a thickness of 10nm, completely disordered the quantum well <strong>and</strong> the<br />

superlattice layers. It was found that the photolum<strong>in</strong>escence wavelength of the quantum<br />

well <strong>and</strong> the superlattice had <strong>in</strong>creased after Zn diffusion. This was attributed to In-Ga<br />

<strong>in</strong>terdiffusion at the <strong>in</strong>terface, due to an <strong>in</strong>terchange process between <strong>in</strong>terstitials <strong>and</strong><br />

substitutional Zn atoms.<br />

M.Razeghi, O.Acher, F.Launay: Semiconductor Science <strong>and</strong> Technology, 1987, 2[12],<br />

793-6<br />

[446-61-077]<br />

In<strong>GaAs</strong>P/InP: Zn <strong>Diffusion</strong><br />

The effect of Zn diffusion upon quantum-well structures was studied by us<strong>in</strong>g secondary<br />

ion mass spectrometry, Auger electron spectroscopy, capacitance-voltage measurements,<br />

photolum<strong>in</strong>escence, <strong>and</strong> X-ray diffraction methods. The structures were stable to<br />

anneal<strong>in</strong>g, <strong>in</strong> the absence of Zn. In a multiple quantum well, Zn diffusion at 500C caused<br />

In <strong>and</strong> Ga <strong>in</strong>termix<strong>in</strong>g <strong>and</strong> the Auger electron spectroscopic profiles became flat.<br />

Photolum<strong>in</strong>escence data revealed that <strong>in</strong>terdiffusion began at temperatures above 420C.<br />

The diffusion of Zn changed the lattice parameter values of In<strong>GaAs</strong>P <strong>and</strong> InP so that the<br />

average value decreased.<br />

G.J.Van Gurp, W.M.Van de Wijgert, G.M.Fontijn, P.J.A.Thijs: Journal of Applied<br />

Physics, 1990, 67[6], 2919-26<br />

[446-74-040]<br />

In<strong>GaAs</strong>P/InP: Zn <strong>Diffusion</strong><br />

The effect of concurrent Zn diffusion upon <strong>in</strong>terdiffusion <strong>in</strong> a In 0.72 Ga 0.28 As 0.61 P 0.39 /InP<br />

heterostructure was <strong>in</strong>vestigated by us<strong>in</strong>g Auger electron spectroscopy <strong>and</strong> secondary ion<br />

mass spectrometry. The measured profiles showed that Zn diffusion (600C, 1 to 4h)<br />

ma<strong>in</strong>ly enhanced cation (In, Ga) <strong>in</strong>terdiffusion. This could not be expla<strong>in</strong>ed <strong>in</strong> terms of<br />

the Zn-vacancy complex model. Results which were obta<strong>in</strong>ed under conditions of group-<br />

V element over-pressure suggested that cation <strong>in</strong>terstitials could control both the rate of<br />

Zn diffusion <strong>and</strong> the mix<strong>in</strong>g of group-<strong>III</strong> sub-lattices <strong>in</strong> the InP-based alloy system.<br />

H.H.Park, B.K.Kang, E.S.Nam, Y.T.Lee, J.H.Kim, O.Kwon: Applied Physics Letters,<br />

1989, 55[17], 1768-70<br />

[446-72/73-029]<br />

In<strong>GaAs</strong>P/InP: Zn <strong>Diffusion</strong><br />

The degradation of lattice-matched In 0.72 Ga 0.28 As 0.61 P 0.39 /InP hetero-<strong>in</strong>terfaces dur<strong>in</strong>g Zn<br />

diffusion was <strong>in</strong>vestigated by us<strong>in</strong>g high-resolution transmission electron microscopy <strong>and</strong><br />

Auger electron spectroscopy. <strong>Diffusion</strong>-<strong>in</strong>duced <strong>in</strong>termix<strong>in</strong>g of In <strong>and</strong> Ga across the<br />

GaInAsP/InP <strong>in</strong>terface produced tensile stresses <strong>in</strong> the Ga-mixed InP side <strong>and</strong><br />

compressive stresses <strong>in</strong> the In-mixed GaInAsP side. The effects of localized <strong>in</strong>terfacial<br />

stresses upon the nucleation of misfit dislocations <strong>and</strong> upon their stra<strong>in</strong> accommodation<br />

439


Zn (Ga,In)(As,P) Zn<br />

behavior were clearly revealed throughout the <strong>in</strong>termixed region, <strong>and</strong> reached several<br />

thous<strong>and</strong> Å on both sides of the <strong>in</strong>terface. The <strong>in</strong>terfacial stra<strong>in</strong> was relaxed by the<br />

generation of paired dislocations, with anti-parallel Burgers vectors, which arose from the<br />

<strong>in</strong>termixed GaInAsP/GaInP <strong>in</strong>terface. The dislocation morphologies revealed strik<strong>in</strong>g<br />

contrasts across the <strong>in</strong>termixed <strong>in</strong>terface; <strong>in</strong>volv<strong>in</strong>g stack<strong>in</strong>g faults <strong>in</strong> the tensile layer <strong>and</strong><br />

perfect dislocation tangles <strong>in</strong> the compressive layer. The dislocation l<strong>in</strong>es were<br />

concentrated at the GaInAsP/GaInP <strong>in</strong>terface <strong>and</strong> along misfit boundaries <strong>in</strong> the frontal<br />

areas of the <strong>in</strong>termixed region. A model was proposed, <strong>in</strong> order to expla<strong>in</strong> the stra<strong>in</strong><br />

relaxation behavior <strong>in</strong> the <strong>in</strong>termixed region, which <strong>in</strong>voked an homogeneous nucleation<br />

mechanism <strong>and</strong> splitt<strong>in</strong>g of the paired dislocations from the <strong>in</strong>termixed <strong>in</strong>terface.<br />

H.H.Park, K.H.Lee, J.K.Lee, Y.T.Lee, E.H.Lee, J.Y.Lee, S.K.Hong, O.Kwon: Journal of<br />

Applied Physics, 1992, 72[9], 4063-72<br />

[446-106/107-118]<br />

In<strong>GaAs</strong>P/InP: Zn <strong>Diffusion</strong><br />

An open-tube Zn diffusion method that <strong>in</strong>volved the use of solid Zn 3 P 2 or InP was<br />

developed, <strong>and</strong> was used to fabricate heterojunction devices. The diffusion profiles were<br />

determ<strong>in</strong>ed by means of secondary ion mass spectrometry. It was found that the diffusion<br />

depth was proportional to the square root of the diffusion time, <strong>and</strong> that a sharp change <strong>in</strong><br />

the Zn concentration was observed at the diffusion front. The activation energies were<br />

equal to 1.2eV for undoped InP, 1.7eV for Sn-doped InP, <strong>and</strong> 1.1eV for In<strong>GaAs</strong>.<br />

T.Ohishi, K.Ohtsuka, Y.Abe, H.Sugimoto, T.Matsui: Japanese Journal of Applied<br />

Physics, 1990, 29[2], L213-6<br />

[446-76/77-024]<br />

In<strong>GaAs</strong>P/InP: Zn <strong>Diffusion</strong><br />

The diffusion of Zn <strong>in</strong>to heterostructures, from spun-on polymer films which conta<strong>in</strong>ed<br />

no Si-organic compounds, was <strong>in</strong>vestigated. It was shown that the <strong>in</strong>itial non-equilibrium<br />

stage of diffusion, <strong>and</strong> the defect relaxation process, determ<strong>in</strong>ed the f<strong>in</strong>al profile of the Zn<br />

distribution; as well as its anomalies. The results demonstrated the important role which<br />

was played by the kick-out mechanism <strong>in</strong> the formation of a sharp diffusion profile slope<br />

as well as <strong>in</strong> Zn getter<strong>in</strong>g at a hetero-<strong>in</strong>terface.<br />

B.J.Ber, L.A.Busyg<strong>in</strong>a, A.T.Gorelenok, A.V.Kaman<strong>in</strong>, A.V.Merkulov, I.A.Mok<strong>in</strong>a,<br />

N.M.Shmidt, I.J.Yakimenko, T.A.Yurre: Materials Science Forum, 1994, 143-147, 1415-<br />

20<br />

[446-113/114-037]<br />

In<strong>GaAs</strong>P/InP: Zn <strong>Diffusion</strong><br />

A study was made of dopant redistribution between a chemical beam epitaxially grown<br />

In<strong>GaAs</strong>P laser structure, <strong>and</strong> a metalorganic vapor phase epitaxially over-grown InP<br />

layer. Secondary ion mass spectroscopic data revealed that Zn <strong>and</strong> Be atoms <strong>in</strong>terdiffused<br />

deeply <strong>in</strong>to the adjacent InP layers, at a Zn dop<strong>in</strong>g level of 10 18 /cm 3 . A fraction of the Zn<br />

atoms went through the chemical beam epitaxial InP, <strong>and</strong> penetrated the laser structure<br />

440


Zn (Ga,In)(As,P) Interdiffusion<br />

guide layer. It was found that Zn out-diffusion was appreciably suppressed by reduc<strong>in</strong>g<br />

the Be dopant concentration from 10 18 to 5 x 10 17 /cm 3 .<br />

H.Sugiura, S.Kondo, M.Mitsuhara, S.Matsumoto, M.Itoh: Applied Physics Letters, 1997,<br />

70[21], 2846-8<br />

[446-152-0441]<br />

Interdiffusion<br />

In<strong>GaAs</strong>P/In<strong>GaAs</strong>P: Interdiffusion<br />

The <strong>in</strong>terdiffusion behavior of In 1-x Ga x As 1-y P y /In 1-x' Ga x' As 1-y' P y' multiple quantum-well<br />

heterostructures, with various built-<strong>in</strong> stra<strong>in</strong>s <strong>and</strong> layer thicknesses, was <strong>in</strong>vestigated by<br />

monitor<strong>in</strong>g their photolum<strong>in</strong>escence. All of the samples which were annealed at 850C<br />

consistently exhibited a blue-shift of the heavy hole exciton l<strong>in</strong>e as a result of<br />

<strong>in</strong>terdiffusion across the hetero-<strong>in</strong>terfaces. Data on a nearly stra<strong>in</strong>-compensated structure,<br />

with a constant P/As ratio <strong>and</strong> In-rich wells, showed that the blue-shift of the excitonic<br />

l<strong>in</strong>e was the result of group-<strong>III</strong> atom <strong>in</strong>terdiffusion alone, as <strong>in</strong> the <strong>GaAs</strong>/GaAlAs system.<br />

The In-Ga <strong>in</strong>terdiffusion could be described by a coefficient of about 4.72 x 10 -16 cm 2 /s.<br />

In the case of lattice-matched <strong>and</strong> compressively stra<strong>in</strong>ed structures, simultaneous<br />

<strong>in</strong>terdiffusion on the group-<strong>III</strong> <strong>and</strong> group-V sub-lattices yielded an effective <strong>in</strong>terdiffusion<br />

coefficient which ranged from 3.83 x 10 -16 to 5.51 x 10 -16 cm 2 /s.<br />

A.Hamoudi, A.Ougazzaden, P.Krauz, K.Rao, M.Juhel, H.Thibierge: Japanese Journal of<br />

Applied Physics, 1995, 34[1-1], 36-41<br />

[446-119/120-215]<br />

In<strong>GaAs</strong>P/In<strong>GaAs</strong>P: Interdiffusion<br />

By us<strong>in</strong>g quaternary/quaternary multiple quantum wells, with constant P/As ratios <strong>and</strong> Inrich<br />

wells, it was shown that it was possible to produce a blue-shift of the heavy hole<br />

exciton l<strong>in</strong>e by <strong>in</strong>terdiffus<strong>in</strong>g only group-<strong>III</strong> atoms. The use of this type of structure,<br />

which was particularly suitable for group-<strong>III</strong> diffusion studies, yielded a value (for the In-<br />

Ga <strong>in</strong>terdiffusion coefficient) of about 4.72 x 10 -16 cm 2 /s at 850C.<br />

A.Hamoudi, A.Ougazzaden, P.Krauz, E.V.K.Rao, M.Juhel, H.Thibierge: Applied Physics<br />

Letters, 1995, 66[6], 718-20<br />

[446-121/122-078]<br />

441


(Ga,In)P<br />

Zn<br />

InGaP: Zn <strong>Diffusion</strong><br />

The migration of Zn was studied by us<strong>in</strong>g angle-lapp<strong>in</strong>g <strong>and</strong> sta<strong>in</strong>-etch<strong>in</strong>g techniques. It<br />

was found that the Zn diffusion depth was proportional to the square root of the diffusion<br />

time, <strong>and</strong> depended upon the composition of In 1-x Ga x P. That is, the effective diffusion<br />

coefficient at a given diffusion temperature decreased accord<strong>in</strong>g to:<br />

D(cm 2 /s) = 3.935 x 10 -8 exp[-6.84x]<br />

while the activation energy <strong>in</strong>creased accord<strong>in</strong>g to:<br />

Q(eV) = 1.28 + 2.38x<br />

This was attributed to a decreas<strong>in</strong>g contribution from <strong>in</strong>terstitial diffusion with <strong>in</strong>creas<strong>in</strong>g<br />

Ga content.<br />

S.T.Kim, D.C.Moon: Japanese Journal of Applied Physics, 1990, 29[4], 627-9<br />

[446-76/77-025]<br />

InGaP: Zn <strong>Diffusion</strong><br />

The diffusion of Zn was studied, us<strong>in</strong>g photolum<strong>in</strong>escence techniques, <strong>in</strong> liquid-phase<br />

epitaxial In 0.5 Ga 0.5 P layers which had been grown onto semi-<strong>in</strong>sulat<strong>in</strong>g <strong>GaAs</strong> substrates.<br />

The photolum<strong>in</strong>escence exhibited a characteristic emission peak at 1.934eV after<br />

diffusion. It was found that this peak behaved like donor-acceptor pair recomb<strong>in</strong>ation <strong>and</strong><br />

was associated with the <strong>in</strong>terstitial Zn donor to substitutional Zn acceptor b<strong>and</strong> transition.<br />

The calculated activation energy of the substitutional acceptor was 0.047eV.<br />

I.T.Yoon, B.S.Jeong, H.L.Park: Th<strong>in</strong> Solid Films, 1997, 300[1-2], 284-8<br />

[446-157/159-442]<br />

GaInP/<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

Heterojunction Zn-doped bipolar transistors were prepared by means of metal-organic<br />

chemical vapor deposition. The characteristics of heterojunction bipolar transistors with<br />

an undoped spacer layer <strong>and</strong> an n-type <strong>GaAs</strong> set-back layer (heterostructure-emitter)<br />

442


Zn (Ga,In)P|(Ga,In)Sb Interdiffusion<br />

between base <strong>and</strong> emitter were <strong>in</strong>vestigated. The results showed that base dopant outdiffusion<br />

was effectively prevented <strong>in</strong> heterostructure-emitter bipolar transistors.<br />

Y.F.Yang, C.C.Hsu, E.S.Yang: Semiconductor Science <strong>and</strong> Technology, 1995, 10[3],<br />

339-43<br />

[446-121/122-069]<br />

(Ga,In)Sb<br />

Interdiffusion<br />

GaInSb/InAs: Interdiffusion<br />

Stra<strong>in</strong>ed superlattices which exhibited a high degree of structural perfection were grown<br />

onto GaSb substrates. The superlattices exhibited ideal defect-free structures. Crosssectional<br />

micrographs revealed that the layers were highly planar, regular <strong>and</strong> coherently<br />

stra<strong>in</strong>ed with respect to the substrate. No crystall<strong>in</strong>e defects were observed by us<strong>in</strong>g<br />

transmission electron microscopy, <strong>in</strong> spite of the existence of a lattice mismatch of almost<br />

2%. The planarity of the layers was confirmed by the observation of Pendellösung fr<strong>in</strong>ges<br />

<strong>in</strong> high-resolution X-ray diffraction patterns, while the observation of numerous sharp<br />

satellite peaks <strong>in</strong>dicated that essentially no <strong>in</strong>terdiffusion occurred with<strong>in</strong> the<br />

superlattices.<br />

R.H.Miles, D.H.Chow, W.J.Hamilton: Journal of Applied Physics, 1992, 71[1], 211-4<br />

[446-86/87-033]<br />

443


GaN<br />

D<br />

GaN: D <strong>Diffusion</strong><br />

The out-diffusion of H was studied, us<strong>in</strong>g 2 H plasma-treated (250 or 400C, 0.5h) or 2 H + -<br />

implanted samples, dur<strong>in</strong>g anneal<strong>in</strong>g at temperatures rang<strong>in</strong>g from 300 to 900C.<br />

Secondary ion mass spectrometry was used to measure the resultant distributions. In the<br />

case of plasma-treated samples, 2 populations of 2 H were found. At concentrations that<br />

were greater than 10 20 /cm 3 , there was a near-surface (less than 0.3µ) region that was<br />

probably due to the formation of platelet defects. At concentrations of about 10 18 /cm 3 , a<br />

plateau region was present which extended throughout the film thickness of about 1µ.<br />

This was attributed to the pair<strong>in</strong>g of 2 H with po<strong>in</strong>t defects. The D <strong>in</strong> the former region<br />

began to out-diffuse at 300C. In the latter region, out-diffusion did not beg<strong>in</strong> until the<br />

temperature exceeded 800C. In implanted samples, 2 H redistribution occurred <strong>in</strong> the same<br />

manner as the bulk population <strong>in</strong> plasma-treated material. The thermal stability of the D<br />

profiles <strong>in</strong> the nitride was much higher than that <strong>in</strong> <strong>GaAs</strong> <strong>and</strong> similar compounds.<br />

R.G.Wilson, S.J.<strong>Pearton</strong>, C.R.Abernathy, J.M.Zavada: Journal of Vacuum Science <strong>and</strong><br />

Technology A, 1995, 13[3], 719-23<br />

[446-140-029]<br />

H<br />

GaN: H <strong>Diffusion</strong><br />

An <strong>in</strong>vestigation was made of <strong>in</strong>teractions between H <strong>and</strong> dopant impurities, <strong>in</strong> this<br />

material, by us<strong>in</strong>g first-pr<strong>in</strong>ciples calculations. The results revealed a fundamental<br />

difference <strong>in</strong> the behaviors of H <strong>in</strong> p-type <strong>and</strong> n-type samples. In particular, it was<br />

expla<strong>in</strong>ed why the H concentrations <strong>in</strong> n-type material were low, <strong>and</strong> why H had a<br />

beneficial effect upon acceptor <strong>in</strong>corporation <strong>in</strong> p-type material. Overall, the conclusions<br />

supported the potential use of H as a means of improv<strong>in</strong>g the dop<strong>in</strong>g of wide-b<strong>and</strong>gap<br />

semiconductors, although the approach was not generally applicable. In order to be able to<br />

exploit H passivation as a method for enhanc<strong>in</strong>g dop<strong>in</strong>g, the H had to be the predom<strong>in</strong>ant<br />

compensat<strong>in</strong>g defect. That is, its formation energy had to be lower than that of all native<br />

444


H GaN Surface<br />

defects, <strong>and</strong> be comparable to the formation energy of the dopant impurity. Also, the<br />

activation energy which was required to dissociate the H-impurity complex, <strong>and</strong> to<br />

remove or neutralize H, had to be lower than the activation energies for native defect<br />

formation <strong>and</strong> lower than the diffusion barrier of the impurity. F<strong>in</strong>ally, the dissociated H<br />

atom had to be highly mobile.<br />

J.Neugebauer, C.G.Van de Walle: Applied Physics Letters, 1996, 68[13], 1829-31<br />

[446-134/135-263]<br />

GaN: H <strong>Diffusion</strong><br />

A study was made, of the electronic structure, energetics, <strong>and</strong> migration of H <strong>and</strong> H-<br />

complexes, on the basis of first-pr<strong>in</strong>ciples total-energy calculations. The latter revealed a<br />

number of features which were very different to those which were exhibited by H <strong>in</strong> more<br />

traditional semiconductors such as Si or <strong>GaAs</strong>. These <strong>in</strong>cluded a very large negative-U<br />

effect (of about 2.4eV), an <strong>in</strong>stability of the bond-center site, high energies of H<br />

molecules, <strong>and</strong> an unusual geometry of the Mg-H complex. All of these features were<br />

shown to be a result of the particular properties of the present material, such as its<br />

strongly ionic nature <strong>and</strong> the high strength of the Ga-N bond. A simple model was<br />

proposed, for the negative-U behavior, which was expected to be valid for H <strong>in</strong> any<br />

semiconductor.<br />

J.Neugebauer, C.G.Van de Walle: Physical Review Letters, 1995, 75[24], 4452-5<br />

[446-127/128-227]<br />

Surface <strong>Diffusion</strong><br />

GaN/<strong>GaAs</strong>: Ga Surface <strong>Diffusion</strong><br />

Metalorganic vapor phase epitaxial growth of cubic GaN on patterned <strong>GaAs</strong>(100)<br />

substrates which had (111)A or (111)B facets was studied. It was found that the growth<br />

features depended strongly upon the configuration of the pattern. It was deduced that<br />

these features arose from an orientation-dependent growth rate, which varied <strong>in</strong> the order:<br />

(111)B > (100) > (111)A, together with Ga adatom diffusion on the surface. By tak<strong>in</strong>g<br />

advantage of the growth on patterned substrates, the diffusion length of Ga adatoms on<br />

the <strong>GaAs</strong>(100) surface was estimated to be typically equal to several microns.<br />

M.Nagahara, S.Miyoshi, H.Yaguchi, K.Onabe, Y.Shiraki, R.Ito: Journal of Crystal<br />

Growth, 1994, 145[1-4], 197-202<br />

[446-119/120-306]<br />

- miscellaneous<br />

GaN: Surface <strong>Diffusion</strong><br />

The growth k<strong>in</strong>etics of GaN/(00?1)sapphire hetero-epitaxial films were studied, at<br />

substrate temperatures rang<strong>in</strong>g from 560 to 640C, by us<strong>in</strong>g reflection high-energy<br />

electron diffraction specular reflection <strong>in</strong>tensity monitor<strong>in</strong>g techniques. An alternat<strong>in</strong>gelement<br />

exposure growth method was used <strong>in</strong> which Ga <strong>and</strong> N atoms were supplied<br />

445


Surface GaN|GaP Ga<br />

separately (rather than simultaneously as usual) to the substrate, with the <strong>in</strong>sertion of a<br />

time delay between successive Ga-flux <strong>and</strong> N-flux exposures. A time-dependent recovery<br />

of the reflection high-energy electron diffraction specular reflection <strong>in</strong>tensity, dur<strong>in</strong>g the<br />

time delay, was associated with Ga-N surface molecule migration on Ga-term<strong>in</strong>ated<br />

surfaces. The activation energy for this migration process was deduced to be 1.45eV.<br />

H.Liu, J.G.Kim, M.H.Ludwig, R.M.Park: Applied Physics Letters, 1997, 71[3], 347-9<br />

[446-152-0446]<br />

GaP<br />

Fe<br />

GaP: Fe <strong>Diffusion</strong><br />

The structural properties of samples which had been implanted with 150 or 400keV Fe or<br />

Ti, to doses of between 10 12 <strong>and</strong> 10 15 /cm 2 , were studied. The depth distributions of the<br />

implants were compared before <strong>and</strong> after anneal<strong>in</strong>g with, or without, a Si 3 N 4 cap.<br />

Rutherford back-scatter<strong>in</strong>g, X-ray double-crystal diffractometry, <strong>and</strong> secondary ion mass<br />

spectroscopy results <strong>in</strong>dicated that Fe was markedly redistributed <strong>in</strong> all of the materials<br />

dur<strong>in</strong>g anneal<strong>in</strong>g. On the <strong>other</strong> h<strong>and</strong>, Ti did not redistribute at all. The driv<strong>in</strong>g force for<br />

the redistribution of Fe was thought to be not classical diffusion, but reaction with<br />

implantation-<strong>in</strong>duced defects <strong>and</strong> stoichiometric imbalances. The defect chemistry of asimplanted<br />

arsenides was found to be fundamentally different to that of as-implanted<br />

phosphides s<strong>in</strong>ce, <strong>in</strong> the latter case, the mass ratio of the constituents was much greater<br />

<strong>and</strong> the specific energy for amorphization was much lower.<br />

H.Ullrich, A.Knecht, D.Bimberg, H.Kräutle, W.Schlaak: Journal of Applied Physics,<br />

1992, 72[8], 3514-21<br />

[446-106/107-034]<br />

Ga<br />

GaP: Ga <strong>Diffusion</strong><br />

The self-diffusion of Ga was measured directly <strong>in</strong> isotopically controlled heterostructures.<br />

Secondary ion mass spectroscopy was used to monitor the <strong>in</strong>termix<strong>in</strong>g of 69 Ga <strong>and</strong> 71 Ga<br />

446


Ga GaP General<br />

between isotopically pure GaP epilayers which had been grown, by molecular beam<br />

epitaxy, onto GaP substrates. It was found that Ga self-diffusivity <strong>in</strong> undoped GaP could<br />

be described by:<br />

D (cm 2 /s) = 2.0 exp[-4.5(eV)/kT]<br />

at temperatures of between 1000 <strong>and</strong> 1190C, under P-rich condition. The entropy of selfdiffusion<br />

was estimated to be about 4k.<br />

L.Wang, J.A.Wolk, L.Hsu, E.E.Haller, J.W.Erickson, M.Cardona, T.Ruf, J.P.Silveira,<br />

F.Briones: Applied Physics Letters, 1997, 70[14], 1831-3<br />

[446-150/151-144]<br />

H<br />

GaP: H <strong>Diffusion</strong><br />

The H passivation effect of Zn was studied by monitor<strong>in</strong>g donor-acceptor pair<br />

lum<strong>in</strong>escence b<strong>and</strong>s. The existence of a critical pair-separation distance was detected,<br />

with<strong>in</strong> which neutralization by H was almost entirely suppressed by the presence of a<br />

nearby donor. This phenomenon was reflected by a relative enhancement of the<br />

<strong>in</strong>tensities of the discrete pair l<strong>in</strong>es, as compared to that of the remote pair b<strong>and</strong> <strong>in</strong> S-Zn<br />

pair emission. Further evidence was provided by the fact that the <strong>in</strong>tensity reduction <strong>in</strong><br />

the O-Zn b<strong>and</strong> was much smaller than that predicted by calculations which assumed a<br />

r<strong>and</strong>om neutralization of Zn acceptors. It was suggested that the suppression was due to<br />

the Coulomb potential of a nearby donor atom, <strong>and</strong> was consistent with the concept that<br />

the pr<strong>in</strong>cipal diffus<strong>in</strong>g form of H was positively charged.<br />

Y.Mochizuki, M.Mizuta: Materials Science Forum, 1992, 83-87, 575-80<br />

[446-99/100-085]<br />

Ti<br />

GaP: Ti <strong>Diffusion</strong><br />

The structural properties of samples which had been implanted with 150 or 400keV Ti, to<br />

doses of between 10 12 <strong>and</strong> 10 15 /cm 2 , were studied. The depth distributions of the implants<br />

were compared before <strong>and</strong> after anneal<strong>in</strong>g with, or without, a Si 3 N 4 cap. Rutherford backscatter<strong>in</strong>g,<br />

X-ray double-crystal diffractometry, <strong>and</strong> secondary ion mass spectroscopy<br />

results <strong>in</strong>dicated that the Ti did not redistribute at all.<br />

H.Ullrich, A.Knecht, D.Bimberg, H.Kräutle, W.Schlaak: Journal of Applied Physics,<br />

1992, 72[8], 3514-21<br />

[446-106/107-034]<br />

General<br />

GaP/InP: <strong>Diffusion</strong><br />

A theoretical study was made of the defect characteristics of a stra<strong>in</strong>ed-layer superlattice.<br />

On the basis of the calculated charge states, some closely-bound <strong>in</strong>terstitial-antisite pairs<br />

447


General GaP Interdiffusion<br />

were found to form at a distance of one bond length <strong>in</strong> a stable GaP/InP (1 x 1) structure.<br />

The defect-related states that were localized <strong>in</strong> the energy gap suggested that the<br />

formation mechanism of the EL2 complex <strong>in</strong> the <strong>GaAs</strong> system was <strong>in</strong>applicable to<br />

GaP/InP multilayer structures. A slow self-diffusion process was proposed for both<br />

group-<strong>III</strong> <strong>and</strong> group-V atoms <strong>in</strong> this superlattice. Kick-out reactions were suggested to be<br />

favored at higher dopant concentrations.<br />

E.G.Wang: Physica Status Solidi B, 1992, 174[2], 367-74<br />

[446-106/107-096]<br />

Surface <strong>Diffusion</strong><br />

GaP: Surface <strong>Diffusion</strong><br />

A scann<strong>in</strong>g tunnell<strong>in</strong>g microscope tip-<strong>in</strong>duced migration of di-vacancies was observed on<br />

the (110) surface of n-doped material. Such motion occurred only for a negative polarity<br />

of the tunnell<strong>in</strong>g voltage, <strong>and</strong> was directed along the [1¯10] direction; regardless of the<br />

scann<strong>in</strong>g orientation. It was therefore purely non-thermal. It was suggested that a field<strong>in</strong>duced<br />

reduction of the barrier to migration <strong>in</strong>itiated the motion. A quantitative analysis<br />

of the migration <strong>in</strong>dicated the existence of long-lived excited states. This was consistent<br />

with charged trap states <strong>in</strong> doped semiconductors.<br />

P.Ebert, M.G.Lagally, K.Urban: Physical Review Letters, 1993, 70[10], 1437-40<br />

[446-106/107-093]<br />

Interdiffusion<br />

GaP/InP: Interdiffusion<br />

The <strong>in</strong>terdiffusion of lateral composition-modulated (GaP) 2 /(InP) 2 short-period<br />

superlattices was studied. Lateral compositional modulation was obta<strong>in</strong>ed by us<strong>in</strong>g the<br />

stra<strong>in</strong>-<strong>in</strong>duced lateral layer order<strong>in</strong>g process. A blue-shift of the <strong>in</strong>ter-b<strong>and</strong> transition was<br />

observed, by means of photolum<strong>in</strong>escence spectroscopy, <strong>in</strong> cap-less <strong>and</strong> SiO 2 -<br />

encapsulated annealed (800C, 5.5h) short-period superlattices. The <strong>in</strong>tensity <strong>and</strong><br />

wavelength of Si 3 N 4 -encapsulated annealed short-period superlattices were only slightly<br />

perturbed. Transmission electron microscopy showed that capless-annealed (800C, 5.5h)<br />

short-period superlattices reta<strong>in</strong>ed their lateral composition modulation. However, the<br />

(00½) satellite reflections disappeared. After long (48h) anneal<strong>in</strong>g, the <strong>in</strong>ter-b<strong>and</strong><br />

transition corresponded to that of an In 0.50 Ga 0.50 P alloy. This suggested that the lateral<br />

composition modulation disappeared. The observed lateral <strong>in</strong>terdiffusion coefficient<br />

exceeded the vertical one by a factor of about 30; thus suggest<strong>in</strong>g that short-period<br />

superlattice <strong>in</strong>terdiffusion was enhanced by native po<strong>in</strong>t defects.<br />

J.I.Mal<strong>in</strong>, A.C.Chen, J.E.Bonkowski, J.E.Baker, K.Y.Cheng, K.C.Hsieh: Journal of<br />

Applied Physics, 1996, 80[2], 1233-5<br />

[446-136/137-120]<br />

448


GaSb<br />

H<br />

GaSb: H <strong>Diffusion</strong><br />

The use of spread<strong>in</strong>g resistance <strong>and</strong> capacitance-voltage methods showed that atomic H<br />

passivated shallow acceptors <strong>and</strong> donors <strong>in</strong> this material. Deep-level passivation by H<br />

also occurred, as revealed by deep-level transient spectroscopic measurements of<br />

Schottky diode structures. Effective H diffusion coefficients were deduced for both n + -<br />

type <strong>and</strong> p + -type samples. In the former case, the diffusion was thermally activated <strong>and</strong><br />

could be described by:<br />

D(cm 2 /s) = 3.4 x 10 -5 exp[-0.55(eV)/kT]<br />

at temperatures of between 100 <strong>and</strong> 250C. In the latter case, the diffusivity could be<br />

described by:<br />

D(cm 2 /s) = 1.5 x 10 -6 exp[-0.45(eV)/kT]<br />

Reactivation of passivated shallow <strong>and</strong> deep levels occurred at temperatures of between<br />

250 <strong>and</strong> 300C.<br />

A.Y.Polyakov, S.J.<strong>Pearton</strong>, R.G.Wilson, P.Rai-Choudhury, R.J.Hillard, X.J.Bao,<br />

M.Stam, A.G.Milnes, T.E.Schles<strong>in</strong>ger, J.Lopata: Applied Physics Letters, 1992, 60[11],<br />

1318-20<br />

[446-86/87-036]<br />

Sb<br />

GaSb: Sb <strong>Diffusion</strong><br />

Recovery <strong>in</strong> crystals which had been implanted with Ga ions, <strong>and</strong> then subjected to rapid<br />

thermal anneal<strong>in</strong>g or furnace anneal<strong>in</strong>g, was studied by us<strong>in</strong>g Raman scatter<strong>in</strong>g<br />

techniques. The <strong>in</strong>tensity of the LO phonon mode decreased with <strong>in</strong>creas<strong>in</strong>g ion<br />

implantation fluence. It was found that the threshold fluence for the amorphization of Gaimplanted<br />

GaSb was 5 x 10 13 /cm 2 . This value was much lower than that for InP<br />

(10 14 /cm 2 ). In the case of furnace anneal<strong>in</strong>g, the recovery processes <strong>in</strong> Ga-implanted<br />

material were very different above <strong>and</strong> below a fluence of 5 x 10 14 /cm 2 . No recovery was<br />

observed at this critical fluence. At lower fluences, the LO mode <strong>in</strong>tensity <strong>in</strong>creased with<br />

449


Sb GaSb Zn<br />

<strong>in</strong>creas<strong>in</strong>g anneal<strong>in</strong>g temperatures <strong>and</strong> times of up to 400C <strong>and</strong> of up to 0.25h. However,<br />

the damage recovery was very poor when compared with that of <strong>GaAs</strong>, InP, or GaP. In<br />

the case of Si 3 N 4 -capped rapid thermal anneal<strong>in</strong>g, recovery was observed even for a<br />

fluence of 5 x 10 14 /cm 2 . New modes were observed, at about 114 <strong>and</strong> 150/cm, <strong>in</strong><br />

implanted <strong>and</strong> annealed samples. These 2 modes were related to the E g <strong>and</strong> A lg modes of<br />

Sb-Sb bond vibrations, respectively, <strong>and</strong> were attributed to the out-diffusion of Sb atoms.<br />

It was found that furnace anneal<strong>in</strong>g enhanced Sb out-diffusion, <strong>and</strong> that the capped rapid<br />

thermal anneal<strong>in</strong>g process was superior to the furnace anneal<strong>in</strong>g process for the heal<strong>in</strong>g of<br />

damaged layers. These anomalous behaviors were thought to be closely related to the<br />

weak bond-strength of Sb-conta<strong>in</strong><strong>in</strong>g materials. The degree of recovery, as a function of<br />

anneal<strong>in</strong>g temperature, anneal<strong>in</strong>g time, <strong>and</strong> fluence was also <strong>in</strong>vestigated.<br />

S.G.Kim, H.Asahi, M.Seta, J.Takizawa, S.Emura, R.K.Soni, S.Gonda, H.Tanoue: Journal<br />

of Applied Physics, 1993, 74[1], 579-85<br />

[446-106/107-096]<br />

Te<br />

GaSb: Te <strong>Diffusion</strong><br />

The diffusion of Te <strong>in</strong>to undoped p-type material was carried out, <strong>and</strong> the samples were<br />

studied us<strong>in</strong>g cathodolum<strong>in</strong>escence <strong>and</strong> photolum<strong>in</strong>escence techniques. The lum<strong>in</strong>escence<br />

centers <strong>in</strong> Te-diffused samples were identified <strong>and</strong> were compared with those <strong>in</strong> Tedoped<br />

bulk material. Essential differences <strong>in</strong> the radiative levels, between diffused <strong>and</strong><br />

as-grown doped samples, were observed. Evidence of the presence of self-compensat<strong>in</strong>g<br />

acceptor complexes was found <strong>in</strong> the case of diffused samples. At short <strong>and</strong> medium<br />

diffusion times, a compensat<strong>in</strong>g acceptor complex, V Ga Ga Sb Te Sb , was observed. At long<br />

diffusion times, the predom<strong>in</strong>ant acceptor center was suggested to be the antisite defect,<br />

Ga Sb , or a related complex.<br />

P.S.Dutta, B.Méndez, J.Piqueras, E.Dieguez, H.L.Bhat: Journal of Applied Physics,<br />

1996, 80[2], 1112-5<br />

[446-136/137-120]<br />

Zn<br />

GaSb: Zn <strong>Diffusion</strong><br />

It was recalled that Zn diffusion was usually performed <strong>in</strong> an evacuated quartz ampoule,<br />

with one source for the dopant <strong>and</strong> with a second antimonide source be<strong>in</strong>g used to<br />

provide an over-pressure over the GaSb dur<strong>in</strong>g diffusion. Strict control of the system<br />

partial pressure was required <strong>in</strong> order to produce reproducible diffusion <strong>and</strong> damage-free<br />

surfaces. An alternative technique was open-tube diffusion from doped spun-on films.<br />

Such dopant emulsions had been successfully applied to some <strong>III</strong>/V semiconductors. The<br />

present work was the first report of the reproducible diffusion of Zn <strong>in</strong>to GaSb from a<br />

spun-on diffusion source <strong>in</strong> an open-tube system. A Zn/SiO 2 film was used as the<br />

450


Zn GaSb Zn<br />

diffusion source at temperatures of 630 or 680C. The concentration <strong>in</strong> the semiconductor<br />

could fall below the substrate dopant level due to the depletion effects of the source.<br />

C.He<strong>in</strong>z: Journal of the Electrochemical Society, 1988, 135[1], 250-2<br />

[446-61-078]<br />

GaSb: Zn <strong>Diffusion</strong><br />

The diffusion of Zn <strong>in</strong>to Te-doped samples was studied as a function of time, temperature<br />

<strong>and</strong> Sb over-pressure. The overall Zn profiles, as well as carrier concentration profiles,<br />

were determ<strong>in</strong>ed. The results <strong>in</strong>dicated an <strong>in</strong>verse dependence of the diffusivity upon the<br />

Sb over-pressure, <strong>and</strong> reflected the operation of an <strong>in</strong>terstitial-substitutional vacancy<br />

mechanism. At high Zn concentrations, the profiles <strong>in</strong>dicated the existence of an<br />

additional component which was associated with a non-electrically active Zn species that<br />

had a small, but highly temperature-dependent, diffusion coefficient.<br />

G.J.Conibeer, A.F.W.Willoughby, C.M.Hard<strong>in</strong>gham, V.K.M.Sharma: Optical Materials,<br />

1996, 6[1-2], 21-5<br />

[446-141/142-104]<br />

GaSb: Zn <strong>Diffusion</strong><br />

The diffusion of Zn <strong>in</strong>to Te-doped samples was studied as a function of time,<br />

temperature, <strong>and</strong> Sb over-pressure. The overall Zn profiles, as well as carrier<br />

concentration profiles, were measured. The results <strong>in</strong>dicated the operation of a<br />

substitutional-<strong>in</strong>terstitial vacancy (Frank-Turnbull) or kick-out (Gosele-Morehead)<br />

mechanism; although there was <strong>in</strong>sufficient evidence to decide between them. There was<br />

also an <strong>in</strong>verse dependence of the diffusivity upon the Sb over-pressure. This was<br />

expla<strong>in</strong>ed <strong>in</strong> terms of a Zn diffusion that was superposed on Ga vacancy diffusion. It was<br />

noted that Te dop<strong>in</strong>g appeared to have little effect upon diffusion, due to its low level<br />

when compared to that of Zn. Moreover, at high Zn concentrations, the profiles <strong>in</strong>dicated<br />

the presence of an additional component that was associated with a non-electrically active<br />

Zn species which had a small strongly temperature-dependent diffusion coefficient.<br />

G.J.Conibeer, A.F.W.Willoughby, C.M.Hard<strong>in</strong>gham, V.K.M.Sharma: Journal of<br />

Electronic Materials, 1996, 25[7], 1108-12<br />

[446-141/142-105]<br />

GaSb: Zn <strong>Diffusion</strong><br />

Sp<strong>in</strong>-on diffusion of Zn <strong>in</strong>to samples <strong>in</strong> an open tube-furnace was <strong>in</strong>vestigated by us<strong>in</strong>g a<br />

Zn-doped silica film as a diffusion source. Theoretical calculations were carried out which<br />

assumed the silica film to be an exhaustible source. <strong>Diffusion</strong> treatments were performed<br />

at 630 or 680C, us<strong>in</strong>g undiluted or diluted Zn-conta<strong>in</strong><strong>in</strong>g silica films. It was found that the<br />

junction depth exhibited a l<strong>in</strong>ear dependence upon the square root of the diffusion time for<br />

undiluted films with no source depletion, whereas results which were obta<strong>in</strong>ed by diffusion<br />

from diluted films clearly reflected depletion effects <strong>in</strong> an exhaustible diffusion<br />

451


Zn GaSb General<br />

source. The method was also applicable to Ga 0.935 Al 0.065 Sb, due to its very small Al<br />

content.<br />

C.He<strong>in</strong>z: Solid-State Electronics, 1993, 36[12], 1685-8<br />

[446-115/116-134]<br />

GaSb: Zn <strong>Diffusion</strong><br />

A closed-ampoule technique was used to <strong>in</strong>troduce Zn <strong>in</strong>to Te-doped material. Anneal<strong>in</strong>g<br />

was carried out for various times; with or without an Sb over-pressure. The total Zn<br />

profiles were measured by us<strong>in</strong>g secondary ion mass spectrometry, <strong>and</strong> carrier profiles<br />

were deduced from <strong>in</strong>cremental sheet resistance data. It was found that the diffusivity<br />

varied, with Zn concentration, from 6.3 x 10 -12 cm 2 /s for a concentration of 3 x 10 20 /cm 3<br />

at 500C.<br />

G.J.Conibeer, A.F.W.Willoughby, C.M.Hard<strong>in</strong>gham, V.K.M.Sharma: Materials Science<br />

Forum, 1994, 143-147, 1427-32<br />

[446-113/114-031]<br />

GaSb: Zn <strong>Diffusion</strong><br />

The diffusion of Zn <strong>in</strong> n-type material, at temperatures rang<strong>in</strong>g from 450 to 540C, was<br />

studied. The diffusion was carried out <strong>in</strong> a closed system, us<strong>in</strong>g a Zn-Ga source. The<br />

diffusion profiles were measured by us<strong>in</strong>g secondary ion mass spectrometry. On the basis<br />

of the diffusion profiles, concentration-dependent diffusion coefficients were calculated<br />

by us<strong>in</strong>g Boltzmann-Matano analyses. These data were qualitatively <strong>in</strong>terpreted <strong>in</strong> terms<br />

of an <strong>in</strong>terstitial-substitutional diffusion model which had orig<strong>in</strong>ally been proposed for Zn<br />

diffusion <strong>in</strong> <strong>GaAs</strong>.<br />

V.S.Sundaram, P.E.Gruenbaum: Journal of Applied Physics, 1993, 73[8], 3787-9<br />

[446-109/110-036]<br />

General<br />

GaSb: <strong>Diffusion</strong><br />

An extensive review was presented of advances <strong>in</strong> the use of GaSb-based systems. It<br />

detailed all aspects: from bulk crystal growth or epitaxy, post-growth process<strong>in</strong>g to device<br />

design. Some current areas of research <strong>and</strong> development were critically assessed, <strong>and</strong><br />

their significance with respect to the underst<strong>and</strong><strong>in</strong>g of basic physical phenomena <strong>and</strong><br />

practical applicability was addressed. These areas <strong>in</strong>cluded the role played by defects <strong>and</strong><br />

impurities <strong>in</strong> affect<strong>in</strong>g the structural <strong>and</strong> <strong>other</strong> properties of the material, <strong>and</strong> the<br />

techniques which were used for surface- <strong>and</strong> bulk-defect passivation. It was concluded<br />

that current knowledge concern<strong>in</strong>g this material was sufficient to expla<strong>in</strong> its basic<br />

properties <strong>and</strong> permit its further application. Among the topics which were listed were:<br />

defects <strong>and</strong> impurities, extended defects, native defects, isotopic effects, self- <strong>and</strong><br />

impurity diffusion, ion bombardment-<strong>in</strong>duced defects, surface <strong>and</strong> bulk defect<br />

452


General GaSb|InAs H<br />

passivation, H-plasma passivation, <strong>and</strong> the passivation of amorphous hydrogenated<br />

material.<br />

P.S.Dutta, H.L.Bhat, V.Kumar: Journal of Applied Physics, 1997, 81[9], 5821-70<br />

[446-150/151-145]<br />

InAs<br />

Fe<br />

InAs: Fe <strong>Diffusion</strong><br />

The structural properties of samples which had been implanted with 150 or 400keV Fe or<br />

Ti, to doses of between 10 12 <strong>and</strong> 10 15 /cm 2 , were studied. The depth distributions of the<br />

implants were compared before <strong>and</strong> after anneal<strong>in</strong>g with, or without, a Si 3 N 4 cap.<br />

Rutherford back-scatter<strong>in</strong>g, X-ray double-crystal diffractometry, <strong>and</strong> secondary ion mass<br />

spectroscopy results <strong>in</strong>dicated that Fe was markedly redistributed <strong>in</strong> all of the materials<br />

dur<strong>in</strong>g anneal<strong>in</strong>g. On the <strong>other</strong> h<strong>and</strong>, Ti did not redistribute at all. The driv<strong>in</strong>g force for<br />

the redistribution of Fe was thought to be not classical diffusion, but reaction with<br />

implantation-<strong>in</strong>duced defects <strong>and</strong> stoichiometric imbalances. The defect chemistry of asimplanted<br />

arsenides was found to be fundamentally different to that of as-implanted<br />

phosphides s<strong>in</strong>ce, <strong>in</strong> the latter case, the mass ratio of the constituents was much greater<br />

<strong>and</strong> the specific energy for amorphization was much lower.<br />

H.Ullrich, A.Knecht, D.Bimberg, H.Kräutle, W.Schlaak: Journal of Applied Physics,<br />

1992, 72[8], 3514-21<br />

[446-106/107-034]<br />

H<br />

InAs: H <strong>Diffusion</strong><br />

Accidentally doped InAs layers were grown, by molecular beam epitaxy, onto semi<strong>in</strong>sulat<strong>in</strong>g<br />

<strong>GaAs</strong> substrates, <strong>and</strong> were exposed to H plasma. It was found that the<br />

453


H InAs Zn<br />

diffusivity of H <strong>in</strong> these layers was high. Hydrogenation led to an order of magnitude<br />

<strong>in</strong>crease <strong>in</strong> the free carrier density. At the same time, defect-related peaks disappeared<br />

from the near-b<strong>and</strong>edge lum<strong>in</strong>escence spectrum. On the <strong>other</strong> h<strong>and</strong>, the properties of bulk<br />

InAs or those of InAs-on-InAs layers were not altered by exposure to the H plasma. A<br />

model which was based upon the passivation, by H, of electronic states which were<br />

<strong>in</strong>duced by dislocations <strong>in</strong> InAs-on-<strong>GaAs</strong> layers was proposed <strong>in</strong> order to expla<strong>in</strong> these<br />

effects.<br />

B.Theys, S.Kalem, A.Lusson, J.Chevallier, C.Grattepa<strong>in</strong>, M.Stutzmann: Materials<br />

Science Forum, 1992, 83-87, 629-34<br />

[446-99/100-090]<br />

InAs/<strong>GaAs</strong>: H <strong>Diffusion</strong><br />

Atomic H was <strong>in</strong>troduced, <strong>in</strong>to molecular beam epitaxial InAs layers on <strong>GaAs</strong> substrates,<br />

from a plasma source. It was found that the H diffused very rapidly <strong>in</strong>to the material. Its<br />

presence modified the electronic transport properties, the near-b<strong>and</strong>edge lum<strong>in</strong>escence<br />

spectra, <strong>and</strong> the far-<strong>in</strong>frared reflectivity. The free-carrier density had <strong>in</strong>creased by an<br />

order of magnitude after hydrogenation. These effects could be removed by thermal<br />

anneal<strong>in</strong>g.<br />

B.Theys, A.Lusson, J.Chevallier, C.Grattepa<strong>in</strong>, S.Kalem, M.Stutzmann: Journal of<br />

Applied Physics, 1991, 70[3], 1461-6<br />

[446-91/92-023]<br />

Ti<br />

InAs: Ti <strong>Diffusion</strong><br />

The structural properties of samples which had been implanted with 150 or 400keV Ti, to<br />

doses of between 10 12 <strong>and</strong> 10 15 /cm 2 , were studied. The depth distributions of the implants<br />

were compared before <strong>and</strong> after anneal<strong>in</strong>g with, or without, a Si 3 N 4 cap. Rutherford backscatter<strong>in</strong>g,<br />

X-ray double-crystal diffractometry, <strong>and</strong> secondary ion mass spectroscopy<br />

results <strong>in</strong>dicated that the Ti did not redistribute at all.<br />

H.Ullrich, A.Knecht, D.Bimberg, H.Kräutle, W.Schlaak: Journal of Applied Physics,<br />

1992, 72[8], 3514-21<br />

[446-106/107-034]<br />

Zn<br />

InAs: Zn <strong>Diffusion</strong><br />

Elemental Zn was diffused <strong>in</strong>to (100)-oriented samples by us<strong>in</strong>g the closed-tube method.<br />

The diffusion temperatures ranged from 350 to 500C. It was found that the results could<br />

be described by:<br />

D(cm 2 /s) = 0.00016 exp[-1.07(eV)/kT]<br />

454


Zn InAs Surface<br />

H.Khald, H.Mani, A.Joullie: Journal of Applied Physics, 1988, 64[9], 4768-70<br />

[446-72/73-034]<br />

Surface <strong>Diffusion</strong><br />

In<br />

InAs: In Surface <strong>Diffusion</strong><br />

The <strong>in</strong>ter-surface diffusion of In adatoms between a (111)A <strong>and</strong> a (001) surface on an<br />

InAs (111)A-(001) non-planar substrate was <strong>in</strong>vestigated for the first time by us<strong>in</strong>g<br />

microprobe reflection high-energy electron diffraction techniques. It was found that the<br />

surface diffusion of In adatoms depended strongly upon the growth temperature <strong>and</strong> the<br />

As pressure, but was <strong>in</strong>dependent of the In flux. It was also observed that the migration<br />

direction of the In lateral flux between the (111)A <strong>and</strong> (001) InAs surfaces changed;<br />

depend<strong>in</strong>g upon the growth conditions.<br />

X.Q.Shen, T.Nish<strong>in</strong>aga: Journal of Crystal Growth, 1995, 146[1-4], 374-8<br />

[446-127/128-148]<br />

- miscellaneous<br />

InAs: Surface <strong>Diffusion</strong><br />

The sensitivity, of the 2- to 3-dimensional growth transition of InAs self-assembled<br />

isl<strong>and</strong>s, to the InAs coverage was used to demonstrate the growth of self-aligned InAs<br />

isl<strong>and</strong>s on etched <strong>GaAs</strong> ridges by molecular beam epitaxy. The differ<strong>in</strong>g migration<br />

behavior of In adatoms upon the various crystal planes of etched ridges was used to<br />

modulate spatially the supply of In adatoms. The ridges were oriented along the [011] <strong>and</strong><br />

[01¯1] directions on (100) substrates with grat<strong>in</strong>g spac<strong>in</strong>gs of 0.28, 1 or 5µ. Atomic force<br />

microscopy revealed that the InAs isl<strong>and</strong>s were self-aligned along the ridges <strong>and</strong> were<br />

typically 40nm <strong>in</strong> diameter <strong>and</strong> 12nm <strong>in</strong> height. In samples with [011]-oriented ridges, the<br />

isl<strong>and</strong>s were located on the side-walls. On the <strong>other</strong> h<strong>and</strong>, <strong>in</strong> the case of [01¯1]-oriented<br />

ridges, the isl<strong>and</strong>s were on the (100) planes on, <strong>and</strong> at the foot of, the mesa. On samples<br />

with a grat<strong>in</strong>g pitch of 0.28µ, all of the isl<strong>and</strong>s were located either on the side-walls or at<br />

the bottom of the so-called V-groove: for both grat<strong>in</strong>g orientations.<br />

D.S.L.Mui, D.Leonard, L.A.Coldren, P.M.Petroff: Applied Physics Letters, 1995, 66[13],<br />

1620-2<br />

[446-121/122-074]<br />

InAs: Surface <strong>Diffusion</strong><br />

The surface diffusion of group-<strong>III</strong> atom <strong>in</strong>corporation dur<strong>in</strong>g molecular beam epitaxial<br />

growth was considered. Firstly, the diffusion length for <strong>in</strong>corporation on the (001) top<br />

surface, with (111)A or (411)A side surfaces on V grooves, was studied. It was shown<br />

that the diffusion length took the same value for both cases <strong>and</strong> was <strong>in</strong>versely<br />

455


Surface InAs Interdiffusion<br />

proportional to the As pressure. The same relationship was also found for the diffusion of<br />

In <strong>in</strong> InAs dur<strong>in</strong>g molecular beam epitaxy. However, the diffusion length of Ga on (111)B<br />

exhibited an <strong>in</strong>verse parabolic dependence of the As pressure. It was suggested that, on<br />

the (001) surface, two As 4 molecules met to furnish active As atoms for growth. On the<br />

<strong>other</strong> h<strong>and</strong>, the behavior of the As 4 molecule on the (111)B surface rema<strong>in</strong>ed unclear. The<br />

ratio of the surface diffusion coefficients on (111)B <strong>and</strong> (001) was calculated. It was<br />

found that the ratio took a value of about 140. Us<strong>in</strong>g this ratio, the <strong>in</strong>corporation lifetimes<br />

on (111)B <strong>and</strong> (001) surfaces were calculated as functions of the As pressure. It was<br />

found that the curves of <strong>in</strong>corporation lifetime <strong>in</strong>tersected at the As pressure where flow<br />

<strong>in</strong>version occurred.<br />

T.Nish<strong>in</strong>aga, X.Q.Shen, D.Kishimoto: Journal of Crystal Growth, 1996, 163[1], 60-6<br />

[446-138/139-078]<br />

Interdiffusion<br />

InAs/<strong>GaAs</strong>: Interdiffusion<br />

The existence of <strong>in</strong>terdiffusion, between self-assembled InAs quantum dots <strong>and</strong> a <strong>GaAs</strong><br />

substrate, was <strong>in</strong>vestigated by us<strong>in</strong>g Rutherford back-scatter<strong>in</strong>g techniques. These were<br />

also useful for determ<strong>in</strong><strong>in</strong>g the value of the average InAs layer thickness. As a result, data<br />

on the diffusion of Ga atoms <strong>in</strong>to the dot were obta<strong>in</strong>ed.<br />

T.Haga, M.Kataoka, N.Matsumura, S.Muto, Y.Nakata, N.Yokoyama: Japanese Journal of<br />

Applied Physics, 1997, 36[2-8B], L1113-5<br />

[446-157/159-456]<br />

InAs/InP: Interdiffusion<br />

The <strong>in</strong>termix<strong>in</strong>g of ultra-th<strong>in</strong> stra<strong>in</strong>ed quantum-well structures, dur<strong>in</strong>g anneal<strong>in</strong>g at<br />

temperatures of 730 to 830C, was <strong>in</strong>vestigated by means of photolum<strong>in</strong>escence<br />

measurements. Upon analyz<strong>in</strong>g the results, us<strong>in</strong>g a microscopic model, the <strong>in</strong>terdiffusion<br />

process was found to be characterized by an activation energy of about 3.8eV. The<br />

<strong>in</strong>terdiffusion coefficient was close to 7 x 10 -7 cm 2 /s at 830C.<br />

J.M.Sallese, S.Taylor, H.J.Bühlmann, J.F.Carl<strong>in</strong>, A.Rudra, R.Houdré, M.Ilegems:<br />

Applied Physics Letters, 1994, 65[3], 341-3<br />

[446-125/126-140]<br />

456


In(As,P)<br />

Interdiffusion<br />

InAsP/InP: Interdiffusion<br />

Evidence was presented for the occurrence of anomalously high stra<strong>in</strong>-dependent<br />

<strong>in</strong>terdiffusion <strong>in</strong> InAsP layers which had been grown onto InP(001) substrates by means<br />

of organometallic vapor phase epitaxy at 620C. In particular, there were clear <strong>in</strong>dications<br />

of the existence of a critical stra<strong>in</strong>. If the stra<strong>in</strong> was equal to about 1.9% or more, marked<br />

P-As mix<strong>in</strong>g occurred. At smaller stra<strong>in</strong>s, the degree of mix<strong>in</strong>g was greatly reduced. The<br />

<strong>in</strong>cidence of <strong>in</strong>terdiffusion was also highly sensitive to the temperature. A set of samples<br />

which was prepared at 580C exhibited an approximately 2-fold decrease <strong>in</strong> P-As mix<strong>in</strong>g,<br />

as compared with samples that were prepared at 620C.<br />

D.J.Tweet, H.Matsuhata, P.Fons, H.Oyanagi, H.Kamei: Applied Physics Letters, 1997,<br />

70[25], 3410-2<br />

[446-152-0457]<br />

In(As,Sb)<br />

Zn<br />

InAsSb: Zn <strong>Diffusion</strong><br />

Elemental Zn was diffused <strong>in</strong>to (100)-oriented samples of InAs 1-x Sb x , where x ranged<br />

from 0.10 to 0.12, by us<strong>in</strong>g the closed-tube method. The diffusion temperatures ranged<br />

from 350 to 500C. It was found that the results could be described by:<br />

D(cm 2 /s) = 0.00016 exp[-1.07(eV)/kT]<br />

H.Khald, H.Mani, A.Joullie: Journal of Applied Physics, 1988, 64[9], 4768-70<br />

[446-72/73-034]<br />

457


InN<br />

D<br />

InN: D <strong>Diffusion</strong><br />

The out-diffusion of H was studied, us<strong>in</strong>g 2 H plasma-treated (250 or 400C, 0.5h) or 2 H + -<br />

implanted samples, dur<strong>in</strong>g anneal<strong>in</strong>g at temperatures rang<strong>in</strong>g from 300 to 900C.<br />

Secondary ion mass spectrometry was used to measure the resultant distributions. At<br />

concentrations that were greater than 10 20 /cm 3 , there was a near-surface (less than 0.3µ)<br />

region that was probably due to the formation of platelet defects. At concentrations of<br />

about 10 18 /cm 3 , a plateau region was present which extended throughout the film<br />

thickness of about 1µ. This was attributed to the pair<strong>in</strong>g of 2 H with po<strong>in</strong>t defects. In<br />

implanted samples, 2 H redistribution occurred <strong>in</strong> the same manner as the bulk population<br />

<strong>in</strong> plasma-treated material. The thermal stability of the D profiles <strong>in</strong> the nitride was much<br />

higher than that <strong>in</strong> <strong>GaAs</strong> <strong>and</strong> similar compounds.<br />

R.G.Wilson, S.J.<strong>Pearton</strong>, C.R.Abernathy, J.M.Zavada: Journal of Vacuum Science <strong>and</strong><br />

Technology A, 1995, 13[3], 719-23<br />

[446-140-029]<br />

InP<br />

Au<br />

InP: Au <strong>Diffusion</strong><br />

Migration of Au at temperatures rang<strong>in</strong>g from 400 to 700C was studied by us<strong>in</strong>g<br />

secondary ion mass spectrometry. Low values were found for the diffusion coefficient;<br />

458


Au InP Cd<br />

which was equal to 2 x 10 -12 cm 2 /s at 550C. By us<strong>in</strong>g deep-level transient spectroscopy,<br />

Au was found to behave as a shallow donor with a level that was situated at 0.55eV from<br />

the conduction b<strong>and</strong>. It was concluded that Au thermal migration from contacts was not<br />

the mechanism which was responsible for device degradation.<br />

V.Parguel, P.N.Favennec, M.Gauneau, Y.Rihet, R.Chapla<strong>in</strong>, H.L'Haridon, C.Vaudry:<br />

Journal of Applied Physics, 1987, 62[3], 824-7<br />

[446-55/56-029]<br />

Be<br />

InP: Be <strong>Diffusion</strong><br />

Implantation (1.5 x 10 14 /cm 2 ) of 30keV Be <strong>in</strong>to (x11)A-oriented semi-<strong>in</strong>sulat<strong>in</strong>g InP<br />

substrates (where x took values of up to 4) was carried out. For comparison, (110)- <strong>and</strong><br />

(100)-oriented substrates were also implanted. The <strong>in</strong>-diffusion of Be <strong>in</strong> (311)A-oriented<br />

substrates was lower than that <strong>in</strong> (100) material.<br />

M.V.Rao, H.B.Dietrich, P.B.Kle<strong>in</strong>, A.Fathimulla, D.S.Simons, P.H.Chi: Journal of<br />

Applied Physics, 1994, 75[12], 7774-8<br />

[446-117/118-166]<br />

InP: Be <strong>Diffusion</strong><br />

Models were presented for the distribution profiles of Be <strong>in</strong> ion-doped layers after<br />

implantation <strong>and</strong> anneal<strong>in</strong>g. The possibility of predict<strong>in</strong>g the mean free path of Be <strong>in</strong> <strong>III</strong>-<br />

V compounds was considered. The effect of defect-impurity <strong>in</strong>teractions upon Be<br />

diffusion was also exam<strong>in</strong>ed. It was found that a flux of impurities towards the surface<br />

occurred which was not diffusive <strong>in</strong> nature.<br />

G.I.Koltsov, V.V.Makarov, S.J.Yurchuk: Fizika i Tekhnika Poluprovodnikov, 1996,<br />

30[10], 1907-16 (<strong>Semiconductors</strong>, 1996, 30[10], 996-1000)<br />

[446-148/149-171]<br />

Cd<br />

InP: Cd <strong>Diffusion</strong><br />

The so-called leaky tube method was used to diffuse elemental Cd <strong>in</strong>to InP at 500C, <strong>and</strong><br />

the junction depths were determ<strong>in</strong>ed after various times. It was found that the diffusion<br />

coefficient was concentration-dependent, <strong>and</strong> ranged from about 10 -14 to 10 -10 cm 2 /s.<br />

C.B.Wheeler, R.J.Roedel, R.W.Nelson, S.N.Schauer, P.Williams: Journal of Applied<br />

Physics, 1990, 68[3], 969-72<br />

[446-86/87-042]<br />

InP: Cd <strong>Diffusion</strong><br />

The Cd was diffused <strong>in</strong>to InP by us<strong>in</strong>g Cd 3 P 2 plus P or Cd 3 P 2 plus Cd 3 As 2 as diffusion<br />

sources. Two diffusion fronts were observed. The diffusion characteristics of Cd 3 P 2 plus P<br />

459


Cd InP D<br />

sources were expla<strong>in</strong>ed <strong>in</strong> terms of the <strong>in</strong>terstitial-substitutional model or the vacancy<br />

complex model. The charge state of the diffus<strong>in</strong>g <strong>in</strong>terstitial Cd atom was a s<strong>in</strong>gly ionized<br />

donor. The chemical species of P which reacted with InP was P 2 , <strong>and</strong> gaseous Cd<br />

orig<strong>in</strong>ated from solid-phase CdP 2 . In the case of Cd 3 P 2 plus Cd 3 As 2 diffusion sources, the<br />

effective diffusion coefficient <strong>and</strong> the surface acceptor concentration decreased with<br />

<strong>in</strong>creas<strong>in</strong>g weight fraction of Cd 3 As 2 . The relative depth of the deeper diffusion front<br />

<strong>in</strong>creased when the supply of vacancies was suppressed.<br />

K.I.Ohtsuka, T.Matsui, H.Ogata: Japanese Journal of Applied Physics, 1988, 27[2], 253-9<br />

[446-60-009]<br />

InP: Cd <strong>Diffusion</strong><br />

Phase diagrams between Cd <strong>and</strong> <strong>III</strong>-V compounds were <strong>in</strong>vestigated <strong>in</strong> order to develop<br />

new acceptor diffusion sources. On the basis of thermodynamic data <strong>and</strong> experimental<br />

studies of the Cd-Ga-As <strong>and</strong> Cd-In-P phase diagrams, it was found that Zn 3 Cd 3 B 2 <strong>and</strong><br />

ZnCdB 2 were suitable acceptor sources <strong>in</strong> that they did not erode the semiconductor<br />

surface.<br />

S.F.Marenk<strong>in</strong>, O.N.Pashkova, V.N.Ravich, I.Z.Babievskaya, N.N.Kazar<strong>in</strong>a,<br />

J.A.Poroikov: Izvestiya Akademii Nauk SSSR - Neorganicheskie Materialy, 1990, 26[9],<br />

1814-8. (Inorganic Materials, 1991, 26[9], 1552-5)<br />

[446-84/85-054]<br />

Cu<br />

InP: Cu <strong>Diffusion</strong><br />

A study of Cu diffusion <strong>in</strong> both p-type <strong>and</strong> n type samples showed that this material<br />

exhibited a transition to semi-<strong>in</strong>sulat<strong>in</strong>g behavior at relatively low Cu diffusion<br />

temperatures. It was found that all, or most, of the Cu precipitates formed a Cu-In<br />

compound, that both orig<strong>in</strong>ally n-type <strong>and</strong> p-type material became semi-<strong>in</strong>sulat<strong>in</strong>g, <strong>and</strong><br />

that there was a negligibly low concentration of deep-level defects. It was observed that<br />

there was an abnormal reduction <strong>in</strong> both electron <strong>and</strong> hole mobilities, which resulted from<br />

the <strong>in</strong>troduction of Cu, <strong>and</strong> that there were isolated pockets of highly conductive material<br />

<strong>in</strong> <strong>other</strong>wise semi-<strong>in</strong>sulat<strong>in</strong>g material. It was concluded that all of these experimental<br />

observations could be best expla<strong>in</strong>ed by the buried Schottky barrier model <strong>in</strong>stead of<br />

compensation by deep levels.<br />

R.P.Leon, M.Kam<strong>in</strong>ska, K.M.Yu, Z.Liliental-Weber, E.R.Weber: Materials Science<br />

Forum, 1992, 83-87, 723-8<br />

[446-99/100-093]<br />

D<br />

InP: D <strong>Diffusion</strong><br />

Data on the effusion of D from various samples were presented. It was shown that the<br />

effusion was limited by surface phenomena, <strong>and</strong> that decomposition of the sample surface<br />

played a major role.<br />

B.Theys, J.Chevallier, M.Miloche, B.Rose: Materials Science Forum, 1994, 143-147,<br />

945-50<br />

[446-113/114-037]<br />

460


D InP Fe<br />

InP: D <strong>Diffusion</strong><br />

The problem of hydrogenat<strong>in</strong>g this material without caus<strong>in</strong>g surface degradation was<br />

solved by expos<strong>in</strong>g the surface to a plasma via a th<strong>in</strong> SiN x (H) cap. This layer was<br />

permeable at the hydrogenation pressure of 250C, but was impermeable to P or PH 3 . It<br />

was found that shallow acceptors were heavily passivated, whereas shallow donors were<br />

only weakly affected. The presence of acceptors impeded D <strong>in</strong>-diffusion. Thus, D<br />

diffusion under the same conditions occurred to a depth of 0.018mm <strong>in</strong> p-type (2 x 10 16<br />

Zn/cm 3 ) material, but to a depth of 0.035mm <strong>in</strong> n-type (S, Sn) material.<br />

W.C.Dautremont-Smith, J.Lopata, S.J.<strong>Pearton</strong>, L.A.Koszi, M.Stavola, V.Swam<strong>in</strong>athan:<br />

Journal of Applied Physics, 1989, 66[5], 1993-6<br />

[446-74-040]<br />

Fe<br />

InP: Fe <strong>Diffusion</strong><br />

The diffusion of Fe was found to occur via the kick-out mechanism. A published Fe<br />

diffusion profile <strong>in</strong> InP was simulated by us<strong>in</strong>g the complete set of 3 partial differential<br />

equations for the kick-out mechanism. A value for the contribution which In self<strong>in</strong>terstitials<br />

made to the self-diffusion coefficient of InP was deduced <strong>and</strong> was found to be<br />

much smaller than was suggested by the known self-diffusion coefficients which had<br />

been determ<strong>in</strong>ed from In tracer diffusion measurements.<br />

H.Zimmermann, U.Gösele, T.Y.Tan: Applied Physics Letters, 1993, 62[1], 75-7<br />

[446-106/107-120]<br />

InP: Fe <strong>Diffusion</strong><br />

The dop<strong>in</strong>g <strong>and</strong> diffusion characteristics of Fe <strong>in</strong> semi-<strong>in</strong>sulat<strong>in</strong>g layers were assessed by<br />

us<strong>in</strong>g secondary ion mass spectrometry. Fairly flat Fe depth profiles, <strong>and</strong> a l<strong>in</strong>ear dop<strong>in</strong>g<br />

curve, were obta<strong>in</strong>ed at concentrations of up to 10 17 /cm 3 . Accumulation of Fe at the<br />

substrate/layer <strong>in</strong>terface was found <strong>in</strong> some samples; thus reveal<strong>in</strong>g a getter<strong>in</strong>g effect of<br />

the substrate. Very little, <strong>and</strong> probably negligible, diffusion was observed on alternately<br />

Fe-doped <strong>and</strong> undoped structures; even after high-temperature heat treatment (provided<br />

that the Fe content was about 10 16 /cm 3 or less).<br />

D.Franke, P.Harde, P.Wolfram, N.Grote: Journal of Crystal Growth, 1990, 100[3], 309-<br />

12<br />

[446-76/77-026]<br />

InP: Fe <strong>Diffusion</strong><br />

The structural properties of samples which had been implanted with 150 or 400keV Fe or<br />

Ti, to doses of between 10 12 <strong>and</strong> 10 15 /cm 2 , were studied. The depth distributions of the<br />

implants were compared before <strong>and</strong> after anneal<strong>in</strong>g with, or without, a Si 3 N 4 cap.<br />

Rutherford back-scatter<strong>in</strong>g, X-ray double-crystal diffractometry, <strong>and</strong> secondary ion mass<br />

spectroscopy results <strong>in</strong>dicated that Fe was markedly redistributed <strong>in</strong> all of the materials<br />

dur<strong>in</strong>g anneal<strong>in</strong>g. On the <strong>other</strong> h<strong>and</strong>, Ti did not redistribute at all. The driv<strong>in</strong>g force for<br />

461


Fe InP Fe<br />

the redistribution of Fe was thought to be not classical diffusion, but reaction with<br />

implantation-<strong>in</strong>duced defects <strong>and</strong> stoichiometric imbalances. The defect chemistry of asimplanted<br />

arsenides was found to be fundamentally different to that of as-implanted<br />

phosphides s<strong>in</strong>ce, <strong>in</strong> the latter case, the mass ratio of the constituents was much greater<br />

<strong>and</strong> the specific energy for amorphization was much lower.<br />

H.Ullrich, A.Knecht, D.Bimberg, H.Kräutle, W.Schlaak: Journal of Applied Physics,<br />

1992, 72[8], 3514-21<br />

[446-106/107-034]<br />

InP: Fe <strong>Diffusion</strong><br />

It was po<strong>in</strong>ted out that high-resistivity material could be grown by metalorganic vapor<br />

phase epitaxy, us<strong>in</strong>g ferrocene as a dopant source. Adjacent Zn-doped layers removed the<br />

resistivity of Fe-doped material. The presence of Zn markedly enhanced the out-diffusion<br />

of Fe from Fe-doped layers, <strong>and</strong> <strong>in</strong>to Zn-doped material. It was suggested that <strong>in</strong>terstitial<br />

Zn took over the lattice sites of substitutional Fe. The Fe then became <strong>in</strong>terstitial <strong>and</strong><br />

mobile.<br />

E.W.A.Young, G.M.Fontijn: Applied Physics Letters, 1990, 56[2], 146-7<br />

[446-74-040]<br />

InP: Fe <strong>Diffusion</strong><br />

A new technique, which <strong>in</strong>volved a Zn 3 P 2 diffusion source <strong>and</strong> rapid thermal anneal<strong>in</strong>g,<br />

was studied. A p + -type layer could be obta<strong>in</strong>ed only at temperatures of between 500 <strong>and</strong><br />

550C by us<strong>in</strong>g a 15s diffusion time. The diffusivity was calculated <strong>and</strong> was compared<br />

with the results of furnace anneal<strong>in</strong>g. The present diffusivity was deduced to be equal to<br />

2.6 x 10 -12 cm 2 /s. In order to form a shallow layer, it was necessary to avoid any<br />

treatments which might redistribute Fe or dopants. Anneal<strong>in</strong>g (850C, 15s) before<br />

diffusion shifted the carrier profile from 300 to 600nm <strong>in</strong> depth. The second diffusion<br />

front extended to 0.0024mm <strong>in</strong> a semi-<strong>in</strong>sulat<strong>in</strong>g substrate. <strong>Diffusion</strong> treatments which<br />

were performed on samples without pre-anneal<strong>in</strong>g resulted <strong>in</strong> 2 diffusion fronts. A<br />

0.0028mm-deep second diffusion front was found when the treatment was applied to a<br />

pre-annealed epi-wafer. The diffusivity under these conditions was deduced to be equal to<br />

1.4 x 10 -11 cm 2 /s.<br />

K.W.Wang, S.M.Parker, C.L.Cheng, J.Long: Journal of Applied Physics, 1988, 63[6],<br />

2104-9<br />

[446-72/73-038]<br />

InP: Fe <strong>Diffusion</strong><br />

The diffusivity was measured by us<strong>in</strong>g secondary ion mass spectrometry. Deliberately<br />

doped metalorganic vapor phase epitaxial layers, as well as ion-implanted samples, were<br />

<strong>in</strong>vestigated. In addition, resistivity measurements were performed on Fe-doped layers. It<br />

was found that the diffusion behavior of Fe was strongly affected by the presence of Zn,<br />

<strong>and</strong> vice versa. In adjacent regions of Fe-doped <strong>and</strong> Zn-doped layers, there was a marked<br />

<strong>in</strong>terdiffusion of the dopants. The <strong>in</strong>terdiffusion process could be described <strong>in</strong> terms of a<br />

kick-out mechanism <strong>in</strong> which Fe <strong>in</strong>terstitials kicked out substitutional Zn. The diffusion<br />

of Fe <strong>in</strong>terstitials was an extremely fast transport process, but the concentration of Fe<br />

462


Fe InP H<br />

<strong>in</strong>terstitials rema<strong>in</strong>ed below 5 x 10 14 /cm 3 . Due to the rapid transport, <strong>in</strong>terdiffusion<br />

proceeded even through barrier layers of (undoped) InP. In the barrier layer itself, the Fe<br />

concentration rema<strong>in</strong>ed below the secondary ion mass spectrometric detection limit of 5 x<br />

10 14 /cm 3 . It was found that a S-doped n-type InP layer prevented the diffusion of Fe. The<br />

semi-<strong>in</strong>sulat<strong>in</strong>g properties of Fe-doped InP were affected by the <strong>in</strong>terdiffusion of Fe <strong>and</strong><br />

Zn. S<strong>in</strong>ce S-doped InP <strong>in</strong>hibited <strong>in</strong>terdiffusion, such a layer could be used as a barrier <strong>in</strong><br />

order to separate Zn-doped <strong>and</strong> Fe-doped regions, <strong>and</strong> thus preserve the semi-<strong>in</strong>sulat<strong>in</strong>g<br />

character of the Fe-doped InP.<br />

E.W.A.Young, G.M.Fontijn, C.J.Vriezema, P.C.Zalm: Journal of Applied Physics, 1991,<br />

70[7], 3593-9<br />

[446-93/94-040]<br />

Ge<br />

InP: Ge <strong>Diffusion</strong><br />

Implantation of Ge <strong>in</strong>to (x11)A-oriented semi-<strong>in</strong>sulat<strong>in</strong>g InP substrates (where x took<br />

values of up to 4) was carried out. For comparison, (110)- <strong>and</strong> (100)-oriented substrates<br />

were also implanted. Follow<strong>in</strong>g 200keV implantation (3 x 10 13 /cm 2 ), after anneal<strong>in</strong>g<br />

(850C, 7s), it was found that the InP was always n-type <strong>and</strong> had a similar sheet resistance<br />

regardless of the substrate orientation. No <strong>in</strong>-diffusion of Ge was observed after anneal<strong>in</strong>g<br />

substrates of any orientation.<br />

M.V.Rao, H.B.Dietrich, P.B.Kle<strong>in</strong>, A.Fathimulla, D.S.Simons, P.H.Chi: Journal of<br />

Applied Physics, 1994, 75[12], 7774-8<br />

[446-117/118-166]<br />

H<br />

InP: H <strong>Diffusion</strong><br />

The H passivation of Zn acceptors, <strong>and</strong> Zn-H dissociation k<strong>in</strong>etics, were compared for the<br />

cases of homo-epitaxial <strong>and</strong> lattice-mismatched hetero-epitaxial n + p structures. Dop<strong>in</strong>g<br />

profile measurements revealed a marked <strong>in</strong>crease <strong>in</strong> the depth <strong>and</strong> degree of passivation<br />

<strong>in</strong> the p-type region of hetero-epitaxial samples. This <strong>in</strong>dicated an enhanced diffusion of<br />

H along dislocations, followed by additional Zn deactivation. Also, the strong aff<strong>in</strong>ity<br />

between H <strong>and</strong> extended defects was found to promote the subsequent dissociation of Zn-<br />

H complexes. This was revealed by reverse bias anneal<strong>in</strong>g studies which showed that the<br />

Zn-H dissociation energy decreased, from 1.19eV <strong>in</strong> homo-epitaxial samples, to 1.12eV<br />

<strong>in</strong> hetero-epitaxial samples. An<strong>other</strong> <strong>in</strong>dicator was the enhanced passivation of extended<br />

defect-related traps by H that was liberated from Zn acceptors dur<strong>in</strong>g the reverse bias<br />

anneal<strong>in</strong>g process; as determ<strong>in</strong>ed by means of deep level transient spectroscopy.<br />

B.Chatterjee, S.A.R<strong>in</strong>gel: Applied Physics Letters, 1996, 69[6], 839-41<br />

[446-138/139-097]<br />

463


H InP In<br />

InP: H <strong>Diffusion</strong><br />

The migration of H was much higher <strong>in</strong> p-type than <strong>in</strong> n-type samples. It was concluded<br />

that H was a deep donor <strong>in</strong> this material.<br />

E.M.Omeljanovsky, A.V.Pakhomov, A.Y.Polyakov, O.M.Borod<strong>in</strong>a, E.A.Kozhukhova,<br />

A.Y.Nashelskii, S.V.Yakobson, V.V.Novikova: Solid State Communications, 1989,<br />

72[5], 409-11<br />

[446-72/73-035]<br />

InP: H <strong>Diffusion</strong><br />

The problem of hydrogenat<strong>in</strong>g this material without caus<strong>in</strong>g surface degradation was<br />

solved by expos<strong>in</strong>g the surface to a H plasma via a th<strong>in</strong> SiN x (H) cap. This layer was<br />

permeable to H at the hydrogenation pressure of 250C, but was impermeable to P or PH 3 .<br />

It was found that shallow acceptors were heavily passivated, whereas shallow donors<br />

were only weakly affected. The presence of acceptors impeded H <strong>in</strong>-diffusion.<br />

W.C.Dautremont-Smith, J.Lopata, S.J.<strong>Pearton</strong>, L.A.Koszi, M.Stavola, V.Swam<strong>in</strong>athan:<br />

Journal of Applied Physics, 1989, 66[5], 1993-6<br />

[446-74-040]<br />

In<br />

InP: In <strong>Diffusion</strong><br />

S<strong>in</strong>gle crystals with a [123]-type orientation were deformed by us<strong>in</strong>g constant stra<strong>in</strong> rates<br />

at temperatures of between 540 <strong>and</strong> 780C. The resultant stress-stra<strong>in</strong> curves were rather<br />

similar to those which were observed for Ge, Si <strong>and</strong> InSb. Two stages of dynamic<br />

recovery could be clearly identified. From the stra<strong>in</strong>-rate <strong>and</strong> temperature dependences of<br />

the stress at the beg<strong>in</strong>n<strong>in</strong>g of the first recovery stage, an activation energy of 2.3eV was<br />

deduced. This was regarded as be<strong>in</strong>g a lower bound on the activation energy for selfdiffusion<br />

of the slowest-mov<strong>in</strong>g species. A value of between 0.001 <strong>and</strong> 0.01cm 2 /s was<br />

estimated for the pre-exponential factor.<br />

H.Siethoff, K.Ahlborn, H.G.Brion, J.Völkl: Philosophical Magaz<strong>in</strong>e A, 1988, 57[2], 235-<br />

44<br />

[446-60-009]<br />

InP: In <strong>Diffusion</strong><br />

The slow evaporation of In from wafers was <strong>in</strong>vestigated, <strong>and</strong> was identified as be<strong>in</strong>g a<br />

major cause of surface roughen<strong>in</strong>g dur<strong>in</strong>g thermal anneal<strong>in</strong>g <strong>and</strong> mass transport under<br />

adequate P-vapor protection. Analysis showed that In evaporation could be prevented by<br />

cover<strong>in</strong>g the wafer dur<strong>in</strong>g anneal<strong>in</strong>g. However, the usual graphite cover alone was an<br />

<strong>in</strong>adequate protection because the In vapor was able to permeate the graphite. On the<br />

<strong>other</strong> h<strong>and</strong>, the use of an InP cover alone resulted <strong>in</strong> problems which were caused by the<br />

mass transport that occurred between the InP wafer <strong>and</strong> the cover. A scheme was<br />

464


In InP Mg<br />

developed that used InP covers <strong>and</strong> a quartz enclosure <strong>in</strong> addition to a graphite cover.<br />

This arrangement permitted smooth wafer surfaces to be reproducibly obta<strong>in</strong>ed.<br />

Z.L.Liau: Applied Physics Letters, 1991, 58[17], 1869-71<br />

[446-81/82-040]<br />

Mg<br />

InP: Mg <strong>Diffusion</strong><br />

The diffusion mechanism of Mg was studied dur<strong>in</strong>g low-pressure metalorganic vaporphase<br />

epitaxial growth. The Mg dopant profiles were measured by means of secondary<br />

ion mass spectroscopy. The analysis revealed that abrupt Mg dopant profiles were<br />

possible. However, the Mg diffusivity depended markedly upon the Mg concentration <strong>in</strong><br />

the crystal lattice. Simultaneous dop<strong>in</strong>g with Si led to a dist<strong>in</strong>ct decrease <strong>in</strong> Mg diffusion.<br />

This behavior was consistent with a model which assumed that the Mg diffused as a<br />

complex which <strong>in</strong>volved a deep donor.<br />

E.Veuhoff, H.Baumeister, R.Treichler, O.Br<strong>and</strong>t: Applied Physics Letters, 1989, 55[10],<br />

1017-9<br />

[446-72/73-038]<br />

InP: Mg <strong>Diffusion</strong><br />

Samples of Fe-doped material were implanted with 80keV Mg, Mg <strong>and</strong> P, or Mg <strong>and</strong> Ar<br />

<strong>in</strong> order to produce shallow p + layers. After rapid thermal anneal<strong>in</strong>g (850 or 875C, 5 or<br />

10s), activations of between 10 <strong>and</strong> 50% <strong>and</strong> mobilities of up to 110cm 2 /Vs were<br />

obta<strong>in</strong>ed. Secondary ion mass spectrometry profiles revealed a pile-up of Mg at the<br />

surface, <strong>and</strong> <strong>in</strong>-diffusion tails which were deeper than 2µ. The use of P or Ar coimplantation<br />

reduced Mg <strong>in</strong>-diffusion <strong>and</strong> <strong>in</strong>creased the activation, but not as much as <strong>in</strong><br />

the case of Be implantation. Photolum<strong>in</strong>escence measurements revealed good crystall<strong>in</strong>e<br />

quality after anneal<strong>in</strong>g. Narrow emissions close to the gap wavelength, <strong>and</strong> 2 broad b<strong>and</strong>s<br />

which were centered at about 1.3 <strong>and</strong> 0.87eV, were found <strong>in</strong> the photolum<strong>in</strong>escence<br />

spectra. The b<strong>and</strong>s were the predom<strong>in</strong>ant emission <strong>in</strong> the photolum<strong>in</strong>escence spectra of<br />

layers with higher implanted doses. This b<strong>and</strong> was tentatively attributed to complexes that<br />

<strong>in</strong>volved Mg <strong>and</strong> a defect.<br />

J.M.Mart<strong>in</strong>, S.García, F.Calle, I.Mártil, G.González-Díaz: Journal of Electronic<br />

Materials, 1995, 24[1], 59-67<br />

[446-119/120-217]<br />

InP: Mg <strong>Diffusion</strong><br />

The anneal<strong>in</strong>g behavior of implanted Mg was studied. It was found that the activated<br />

fraction of dopants depended markedly upon the implant dose, <strong>and</strong> upon the substrate<br />

temperature dur<strong>in</strong>g implantation. Low activation of high-dose (10 15 /cm 2 ) implants was<br />

465


Mg InP S<br />

attributed to the effect of pronounced (80%) out-diffusion. A large variation was found,<br />

<strong>in</strong> the apparent activation energy, for implantation temperatures between ambient <strong>and</strong><br />

300C.<br />

W.H.Van Berlo, M.Ghaffari, G.L<strong>and</strong>gren: Journal of Electronic Materials, 1992, 21[4],<br />

431-6<br />

[446-93/94-040]<br />

P<br />

InP: P <strong>Diffusion</strong><br />

S<strong>in</strong>gle crystals with a [123]-type orientation were deformed by us<strong>in</strong>g constant stra<strong>in</strong> rates<br />

at temperatures of between 540 <strong>and</strong> 780C. The resultant stress-stra<strong>in</strong> curves were rather<br />

similar to those which were observed for Ge, Si <strong>and</strong> InSb. Two stages of dynamic<br />

recovery could be clearly identified. From the stra<strong>in</strong>-rate <strong>and</strong> temperature dependences of<br />

the stress at the beg<strong>in</strong>n<strong>in</strong>g of the first recovery stage, an activation energy of 2.3eV was<br />

deduced. This was regarded as be<strong>in</strong>g a lower bound on the activation energy for selfdiffusion<br />

of the slowest-mov<strong>in</strong>g species. A value of between 0.001 <strong>and</strong> 0.01cm 2 /s was<br />

estimated for the pre-exponential factor.<br />

H.Siethoff, K.Ahlborn, H.G.Brion, J.Völkl: Philosophical Magaz<strong>in</strong>e A, 1988, 57[2], 235-<br />

44<br />

[446-60-009]<br />

Rh<br />

InP: Rh <strong>Diffusion</strong><br />

A secondary-ion mass spectroscopic <strong>in</strong>vestigation was made of the thermally <strong>in</strong>duced<br />

redistribution of Rh <strong>in</strong> low-pressure metalorganic chemical vapor-deposited InP<br />

structures. Control measurements were performed on Fe-doped structures. In the case of<br />

alternately Rh-doped InP <strong>and</strong> undoped InP structures, an upper limit on the Rh diffusion<br />

coefficient of about 10 -14 cm 2 /s (at 800C) was established. This was much smaller than the<br />

Fe diffusivity of about 10 -11 cm 2 /s at 750C. No exchange reactions were observed at the<br />

<strong>in</strong>terfaces of p-InP <strong>and</strong> Rh-doped InP structures. Only Rh which was implanted <strong>in</strong>to InP<br />

exhibited defect-<strong>in</strong>duced redistribution <strong>in</strong>to amorphous areas.<br />

A.Näser, A.Dadgar, M.Kuttler, R.Heitz, D.Bimberg, J.Y.Hyeon, H.Schumann: Applied<br />

Physics Letters, 1995, 67[4], 479-81<br />

[446-123/124-177]<br />

S<br />

InP[l]: S <strong>Diffusion</strong><br />

Macro-segregation <strong>and</strong> micro-segregation of S <strong>in</strong> crystals which had been grown from In<br />

solutions by us<strong>in</strong>g the travell<strong>in</strong>g heater method (under micro-gravity or normal gravity<br />

conditions) were analyzed by us<strong>in</strong>g spatially resolved photolum<strong>in</strong>escence methods. It was<br />

466


S InP Si<br />

found that, whereas macro-segregation <strong>in</strong> both terrestrially-grown <strong>and</strong> space-grown<br />

crystals could be expla<strong>in</strong>ed by conventional steady-state models which were based upon<br />

the Burton-Prim-Slichter theory, micro-segregation could be expla<strong>in</strong>ed only <strong>in</strong> terms of<br />

the non steady-state step-exchange model.<br />

A.N.Danilewsky, Y.Okamoto, K.W.Benz, T.Nish<strong>in</strong>aga: Japanese Journal of Applied<br />

Physics, 1992, 31[1-7], 2195-201<br />

[446-93/94-231]<br />

Si<br />

InP: Si <strong>Diffusion</strong><br />

Implantation of Si <strong>in</strong>to (x11)A-oriented semi-<strong>in</strong>sulat<strong>in</strong>g InP substrates (where x took<br />

values of up to 4) was carried out. For comparison, (110)- <strong>and</strong> (100)-oriented substrates<br />

were also implanted. Follow<strong>in</strong>g 200keV implantation (5 x 10 13 /cm 2 ), after anneal<strong>in</strong>g<br />

(850C, 7s), it was found that the InP was always n-type <strong>and</strong> had a similar sheet resistance<br />

regardless of the substrate orientation. No <strong>in</strong>-diffusion of Si was observed after anneal<strong>in</strong>g<br />

substrates of any orientation. A similar behavior was observed for Si/B co-implants <strong>in</strong><br />

InP.<br />

M.V.Rao, H.B.Dietrich, P.B.Kle<strong>in</strong>, A.Fathimulla, D.S.Simons, P.H.Chi: Journal of<br />

Applied Physics, 1994, 75[12], 7774-8<br />

[446-117/118-166]<br />

InP: Si <strong>Diffusion</strong><br />

Semi-<strong>in</strong>sulat<strong>in</strong>g material was implanted with 200keV Xe + ions to 10 14 /cm 2 at room<br />

temperature, <strong>and</strong> with 100keV Hg + ions to 10 14 /cm 2 at 200C or room temperature.<br />

Implanted <strong>and</strong> non-implanted substrates were encapsulated <strong>in</strong> AlN/Si 3 N 4 <strong>and</strong> were<br />

subjected to rapid (60s) thermal anneal<strong>in</strong>g cycles at 650 to 900C. Electrical measurements<br />

<strong>and</strong> secondary ion mass spectrometry were used to correlate an observed n-type behavior<br />

with the presence of Si <strong>in</strong> the near-surface region. An enhanced near-surface Si<br />

concentration was found after the rapid thermal anneal<strong>in</strong>g of Xe + -implanted samples, <strong>and</strong><br />

n-type surface layers (approximately 80nm thick) were formed. Non-implanted samples<br />

exhibited no measurable electrical behavior, <strong>and</strong> there was no enhanced Si concentration<br />

after rapid thermal anneal<strong>in</strong>g. Hot Hg + implantation led to p-type behavior <strong>and</strong> to low Si<br />

concentrations. Room-temperature Hg implantation led to semi-<strong>in</strong>sulat<strong>in</strong>g behavior <strong>and</strong><br />

high Si levels. It was concluded that the presence of implantation damage enhanced Si <strong>in</strong>diffusion<br />

from the Si-based encapsulants which were used to protect the InP surface<br />

dur<strong>in</strong>g post-implantation anneal<strong>in</strong>g.<br />

J.H.Wilkie, B.J.Sealy: Th<strong>in</strong> Solid Films, 1988, 162, 49-57<br />

[446-70/71-119]<br />

InP: Si <strong>Diffusion</strong><br />

Migration was <strong>in</strong>vestigated by us<strong>in</strong>g two configurations. These were diffusion from an<br />

external source <strong>in</strong>to uniformly n-doped substrates, <strong>and</strong> diffusion between the layers of n-<br />

467


Si InP Sn<br />

p-n-p-n structures which had been grown via metalorganic chemical vapor deposition.<br />

Alternat<strong>in</strong>g layers of n-type material (0.0005mm, [Si] = 10 16 to 3 x 10 19 /cm 3 ) were grown<br />

by us<strong>in</strong>g low-pressure metalorganic chemical vapor deposition at 625C. The distributions<br />

of Si were determ<strong>in</strong>ed by means of secondary ion mass spectrometry. No diffusion of Si<br />

across the grown dopant <strong>in</strong>terface was detected. Electrochemical capacitance-voltage<br />

profil<strong>in</strong>g <strong>in</strong>dicated that the Si was electrically active.<br />

C.Blaauw, F.R.Shepherd, D.Eger: Journal of Applied Physics, 1989, 66[2], 605-10<br />

[446-74-041]<br />

InP: Si <strong>Diffusion</strong><br />

Material was deposited by means of metalorganic chemical vapor deposition, <strong>and</strong> was<br />

simultaneously doped with Si (donor) <strong>and</strong> Zn (acceptor) species dur<strong>in</strong>g growth. It was<br />

found that the <strong>in</strong>corporation of Si was not affected by the presence of Zn, whereas Zn<br />

<strong>in</strong>corporation was markedly enhanced by the presence of Si. The results were consistent<br />

with the formation of donor-acceptor pairs.<br />

C.Blaauw, L.Hobbs: Applied Physics Letters, 1991, 59[6], 674-6<br />

[446-84/85-054]<br />

InP/Si: Si <strong>Diffusion</strong><br />

The spatial distribution of the charge concentration of InP layers which had been grown<br />

onto Si substrates by metalorganic vapor-phase epitaxy was <strong>in</strong>vestigated. The<br />

concentrations near to the surface, <strong>and</strong> with<strong>in</strong> the bulk of the layer, were found to be<br />

governed by Si dop<strong>in</strong>g from the ambient gas. The diffusion of Si across the hetero<strong>in</strong>terface,<br />

which could be partially assisted by dislocations, predom<strong>in</strong>ated <strong>in</strong> a region near<br />

to the InP/Si <strong>in</strong>terface. In the vic<strong>in</strong>ity of the hetero-<strong>in</strong>terface, the charge concentration <strong>in</strong><br />

the InP layer was determ<strong>in</strong>ed by a strong compensation which was attributed to defects<br />

that were caused by a mismatch between the lattice parameters <strong>and</strong> thermal expansion<br />

coefficients of the InP <strong>and</strong> Si.<br />

A.Bartels, E.Pe<strong>in</strong>er, R.Klockenbr<strong>in</strong>k, A.Schlachetzki: Journal of Applied Physics, 1995,<br />

78[1], 224-8<br />

[446-123/124-179]<br />

Sn<br />

InP: Sn <strong>Diffusion</strong><br />

The lattice location <strong>and</strong> electrical activity of ion implanted Sn, after rapid thermal<br />

anneal<strong>in</strong>g, were determ<strong>in</strong>ed by means of Mössbauer spectroscopic (us<strong>in</strong>g 119m Sn) <strong>and</strong><br />

differential Hall resistivity methods, respectively. It was found that the Sn was<br />

preferentially located on the In sub-lattice at concentrations below 2 x 10 19 /cm 3 ; result<strong>in</strong>g<br />

<strong>in</strong> high electrical activation <strong>and</strong> mobility. At Sn concentrations which were above this<br />

468


Sn InP Zn<br />

concentration, various electrically <strong>in</strong>active Sn complexes were also observed. No sign<br />

was found of Sn on P sub-lattice sites.<br />

P.Kr<strong>in</strong>ghøj, G.Weyer: Applied Physics Letters, 1993, 62[16], 1973-5<br />

Ti<br />

[446-99/100-093]<br />

InP: Ti <strong>Diffusion</strong><br />

The structural properties of samples which had been implanted with 150 or 400keV Ti, to<br />

doses of between 10 12 <strong>and</strong> 10 15 /cm 2 , were studied. The depth distributions of the implants<br />

were compared before <strong>and</strong> after anneal<strong>in</strong>g with, or without, a Si 3 N 4 cap. Rutherford backscatter<strong>in</strong>g,<br />

X-ray double-crystal diffractometry, <strong>and</strong> secondary ion mass spectroscopy<br />

results <strong>in</strong>dicated that the Ti did not redistribute at all.<br />

H.Ullrich, A.Knecht, D.Bimberg, H.Kräutle, W.Schlaak: Journal of Applied Physics,<br />

1992, 72[8], 3514-21<br />

[446-106/107-034]<br />

InP: Ti <strong>Diffusion</strong><br />

Samples of n-type InP were implanted with Co at 200C. Dur<strong>in</strong>g high-temperature<br />

anneal<strong>in</strong>g, out-diffusion of the implant was as severe as that for room-temperature<br />

implants. In-diffusion of the implant also occurred, but it was not as severe as the outdiffusion.<br />

High-temperature anneal<strong>in</strong>g of Ti-implanted material resulted <strong>in</strong> slight Ti <strong>in</strong>diffusion,<br />

with m<strong>in</strong>imal redistribution or out-diffusion. In the case of high-temperature<br />

implants, the lattice quality of the annealed material was close to that of virg<strong>in</strong> material.<br />

Regardless of the ion type, resistivities that were close to the <strong>in</strong>tr<strong>in</strong>sic limit were<br />

measured <strong>in</strong> implanted <strong>and</strong> annealed materials.<br />

M.V.Rao, S.M.Gulwadi, S.Mulpuri, D.S.Simons, P.H.Chi, C.Caneau, W.P.Hong,<br />

O.W.Holl<strong>and</strong>, H.B.Dietrich: Journal of Electronic Materials, 1992, 21[9], 923-8<br />

[446-93/94-038]<br />

Zn<br />

InP: Zn <strong>Diffusion</strong><br />

Spun-on SiO 2 films which conta<strong>in</strong>ed 1, 2, 6, 22 or 36at%Zn were prepared <strong>and</strong> were used<br />

for p-diffusion <strong>in</strong>to undoped n-type material. It was found that an <strong>in</strong>creas<strong>in</strong>g Zn<br />

concentration of the spun-on film led to a higher atomic Zn concentration <strong>and</strong> diffusion<br />

depth. On the basis of the experimental data, the InP/film distribution coefficient was<br />

deduced to be 0.012. A higher Zn concentration <strong>in</strong> the InP resulted <strong>in</strong> a lower acceptor<br />

concentration. The electrical activity of Zn could be significantly <strong>in</strong>creased by means of<br />

additional heat treatment.<br />

C.Lauterbach: Semiconductor Science <strong>and</strong> Technology, 1995, 10[4], 500-3<br />

[446-121/122-080]<br />

469


Zn InP Zn<br />

InP: Zn <strong>Diffusion</strong><br />

The formation of defects dur<strong>in</strong>g Zn diffusion <strong>in</strong>to undoped or semi-<strong>in</strong>sulat<strong>in</strong>g Fe-doped<br />

s<strong>in</strong>gle crystals at 700C was observed by means of transmission electron microscopy under<br />

various diffusion conditions. Agglomerates of predom<strong>in</strong>antly perfect <strong>in</strong>terstitial-type<br />

dislocation loops, dislocations, <strong>and</strong> small In precipitates with<strong>in</strong> voids were observed <strong>in</strong><br />

the Zn-diffused region. Also, large planar arrays of precipitates were formed by climb<strong>in</strong>g<br />

dislocations. From these observations, it was deduced that the <strong>in</strong>corporation of Zn at In<br />

sub-lattice sites created a supersaturation of In self-<strong>in</strong>terstitials which was removed by<br />

dislocation loop formation that led to a supersaturation of P vacancies <strong>and</strong> to void<br />

formation.<br />

D.Wittorf, A.Rucki, W.Jäger, R.H.Dixon, K.Urban, H.G.Hettwer, N.A.Stolwijk,<br />

H.Mehrer: Journal of Applied Physics, 1995, 77[6], 2843-5<br />

[446-121/122-080]<br />

InP: Zn <strong>Diffusion</strong><br />

The diffusivity was measured by us<strong>in</strong>g secondary ion mass spectrometry. Deliberately<br />

doped metalorganic vapor phase epitaxial layers, as well as ion-implanted samples, were<br />

<strong>in</strong>vestigated. In addition, resistivity measurements were performed on Fe-doped layers. It<br />

was found that the diffusion behavior of Fe was strongly affected by the presence of Zn,<br />

<strong>and</strong> vice versa. In adjacent regions of Fe-doped <strong>and</strong> Zn-doped layers, there was a marked<br />

<strong>in</strong>terdiffusion of the dopants. The <strong>in</strong>terdiffusion process could be described <strong>in</strong> terms of a<br />

kick-out mechanism <strong>in</strong> which Fe <strong>in</strong>terstitials kicked out substitutional Zn.<br />

E.W.A.Young, G.M.Fontijn, C.J.Vriezema, P.C.Zalm: Journal of Applied Physics, 1991,<br />

70[7], 3593-9<br />

[446-93/94-040]<br />

InP: Zn <strong>Diffusion</strong><br />

A photolum<strong>in</strong>escence study was made of Zn-diffused <strong>and</strong> annealed material. A new peak<br />

was found near to 1.33eV. After anneal<strong>in</strong>g, the peak energy of the lum<strong>in</strong>escence shifted<br />

towards higher energies. This reflected the out-diffusion of Zn. By depth profil<strong>in</strong>g of this<br />

lum<strong>in</strong>escence it was deduced that recomb<strong>in</strong>ation, due to the <strong>in</strong>terstitial Zn donor,<br />

predom<strong>in</strong>ated near to the surface.<br />

J.S.Choi, H.J.Lim, J.I.Lee, S.K.Chang, H.L.Park: Physica Status Solidi B, 1991, 164[2],<br />

K69-72<br />

[446-93/94-041]<br />

InP: Zn <strong>Diffusion</strong><br />

The diffusivity was studied by us<strong>in</strong>g dimethylz<strong>in</strong>c as the source <strong>in</strong> a PH 3 /H 2 /N 2<br />

atmosphere at 500 to 600C. The results were compared with the Zn dop<strong>in</strong>g of material<br />

which had been grown via metalorganic vapor-phase epitaxial growth us<strong>in</strong>g<br />

triethyl<strong>in</strong>dium, PH 3 <strong>and</strong> dimethylz<strong>in</strong>c. It was found that the carrier concentration <strong>in</strong> Zndiffused<br />

material was lower, by about 2 orders of magnitude, than that <strong>in</strong> Zn-doped<br />

material which had been grown at the same mole fraction. The activation energy for Zn<br />

470


Zn InP Zn<br />

<strong>in</strong>corporation was deduced to be the same for both diffusion <strong>and</strong> dop<strong>in</strong>g. The observed<br />

dependence of the surface carrier concentration upon the diffusion temperature <strong>and</strong> the<br />

mole fraction of dimethylz<strong>in</strong>c was qualitatively expla<strong>in</strong>ed by consider<strong>in</strong>g an InP surface<br />

condensation limit for Zn adsorption. Slightly <strong>in</strong>creased carrier concentrations, after<br />

subsequent anneal<strong>in</strong>g (500C, PH 3 /H 2 /N 2 atmosphere) <strong>in</strong>dicated the existence of a low<br />

concentration of Zn <strong>in</strong>terstitials <strong>in</strong> as-diffused InP under a high PH 3 flow-rate. The lower<br />

carrier concentration of shallow Zn-diffused material, after such anneal<strong>in</strong>g, suggested the<br />

occurrence of a large out-diffusion of <strong>in</strong>terstitials near to the surface region dur<strong>in</strong>g<br />

anneal<strong>in</strong>g.<br />

M.Wada, K.Sakakibara, M.Higuchi, Y.Sekiguchi: Journal of Crystal Growth, 1991,<br />

114[3], 321-6<br />

[446-91/92-025]<br />

InP: Zn <strong>Diffusion</strong><br />

A wafer of InP, with an evaporated th<strong>in</strong> layer of metallic Zn, was used as a diffusion<br />

source <strong>and</strong> was placed 0.15mm away from the sample. In this way, it was possible to<br />

obta<strong>in</strong> good diffusion-front planarity, perfect surface quality, <strong>and</strong> a free-hole<br />

concentration of about 8 x 10 18 /cm 3 . It was suggested that the first stage of the diffusion<br />

process was characterized by the <strong>in</strong>corporation of a large quantity of Zn, such that<br />

<strong>in</strong>terstitial Zn atoms compensated the acceptors. The Zn concentration then levelled out,<br />

but most of the <strong>in</strong>terstitials left the crystal. Due to the small free volume which existed<br />

between sample <strong>and</strong> source, P loss was decreased <strong>and</strong> the formation of P vacancies was<br />

dim<strong>in</strong>ished.<br />

T.Krieghoff, E.Nowak, G.Kühn, B.Schumann, A.Höpner: Crystal Research <strong>and</strong><br />

Technology, 1992, 27[1], 49-57<br />

[446-88/89-044]<br />

InP: Zn <strong>Diffusion</strong><br />

The migration of Zn dur<strong>in</strong>g the growth of InP epitaxial layers was <strong>in</strong>vestigated <strong>in</strong><br />

structures which consisted of Zn-InP epilayers that had been grown onto S-InP <strong>and</strong> Fe-<br />

InP substrates, <strong>and</strong> onto undoped InP epilayers. The layers were grown by means of<br />

metalorganic chemical vapor deposition at 625C, under a pressure of 75torr. The dopant<br />

diffusion profiles were measured by us<strong>in</strong>g secondary ion mass spectrometry. At high Zn<br />

dopant levels (greater than 8 x 10 17 /cm 3 ), diffusion <strong>in</strong>to S-InP substrates took place, with<br />

Zn accumulation <strong>in</strong> the substrate at a concentration which was similar to [S]. <strong>Diffusion</strong><br />

<strong>in</strong>to undoped InP epilayers produced a diffusion tail at low [Zn] levels. This was<br />

suggested to be associated with <strong>in</strong>terstitial Zn diffusion. In the case of diffusion <strong>in</strong>to Fe-<br />

InP, this low-level diffusion produced a region of constant Zn concentration when the<br />

dopant concentration was equal to 3 x 10 16 /cm 3 . This was attributed to kick<strong>in</strong>g-out of the<br />

orig<strong>in</strong>al Fe species from substitutional sites. <strong>Diffusion</strong> out of (Zn,Si) co-doped InP<br />

epilayers which had been grown onto Fe-InP substrates was also <strong>in</strong>vestigated. The<br />

secondary ion mass spectrometry profiles were characterized by a sharp decrease <strong>in</strong> [Zn]<br />

at the epilayer/substrate <strong>in</strong>terface. The magnitude of this decrease corresponded to that of<br />

the Si donor level <strong>in</strong> the epilayer. When [Si] was greater than [Zn] <strong>in</strong> the epilayer, no Zn<br />

471


Zn InP Zn<br />

diffusion was observed. Hall measurements <strong>in</strong>dicated that the donor <strong>and</strong> acceptor species<br />

<strong>in</strong> those samples were electrically active. All of the results were consistent with the<br />

presence of donor-acceptor <strong>in</strong>teractions; result<strong>in</strong>g <strong>in</strong> the formation of ionized donoracceptor<br />

pairs which were immobile <strong>and</strong> did not contribute to the diffusion process.<br />

C.Blaauw, B.Emmerstorfer, D.Kreller, L.Hobbs, A.J.Spr<strong>in</strong>gthorpe: Journal of Electronic<br />

Materials, 1992, 21[2], 173-9<br />

[446-88/89-045]<br />

InP: Zn <strong>Diffusion</strong><br />

A low-pressure open-tube system, <strong>in</strong>volv<strong>in</strong>g diethylz<strong>in</strong>c <strong>and</strong> PH 3 , was used to study the<br />

diffusion of Zn. The Zn <strong>and</strong> hole concentrations were measured by us<strong>in</strong>g secondary ion<br />

mass spectrometry <strong>and</strong> capacitance-voltage etch profil<strong>in</strong>g. It was found that anneal<strong>in</strong>g of<br />

the samples <strong>in</strong>creased the hole concentration, due to the out-diffusion of <strong>in</strong>terstitial Zn<br />

donors.<br />

J.Wisser, M.Glade, H.J.Schmidt, K.He<strong>in</strong>e: Journal of Applied Physics, 1992, 71[7],<br />

3234-7<br />

[446-86/87-042]<br />

InP: Zn <strong>Diffusion</strong><br />

Phase diagrams between Zn <strong>and</strong> <strong>III</strong>-V compounds were <strong>in</strong>vestigated <strong>in</strong> order to develop<br />

new acceptor diffusion sources. On the basis of thermodynamic data <strong>and</strong> experimental<br />

studies of the Cd-Ga-As <strong>and</strong> Cd-In-P phase diagrams, it was found that Zn 3 Cd 3 B 2 <strong>and</strong><br />

ZnCdB 2 were suitable acceptor sources <strong>in</strong> that they did not erode the semiconductor<br />

surface.<br />

S.F.Marenk<strong>in</strong>, O.N.Pashkova, V.N.Ravich, I.Z.Babievskaya, N.N.Kazar<strong>in</strong>a,<br />

J.A.Poroikov: Izvestiya Akademii Nauk SSSR - Neorganicheskie Materialy, 1990, 26[9],<br />

1814-8. (Inorganic Materials, 1991, 26[9], 1552-5)<br />

[446-84/85-054]<br />

InP: Zn <strong>Diffusion</strong><br />

Material was deposited by means of metalorganic chemical vapor deposition, <strong>and</strong> was<br />

simultaneously doped with Si (donor) <strong>and</strong> Zn (acceptor) species dur<strong>in</strong>g growth. It was<br />

found that the <strong>in</strong>corporation of Si was not affected by the presence of Zn, whereas Zn<br />

<strong>in</strong>corporation was markedly enhanced by the presence of Si. The results were consistent<br />

with the formation of donor-acceptor pairs; a concept which had previously been used to<br />

expla<strong>in</strong> Zn diffusion profiles <strong>in</strong> Si-doped InP.<br />

C.Blaauw, L.Hobbs: Applied Physics Letters, 1991, 59[6], 674-6<br />

[446-84/85-054]<br />

InP: Zn <strong>Diffusion</strong><br />

Diethylz<strong>in</strong>c was used as a p-type dopant source dur<strong>in</strong>g the chemical beam epitaxial<br />

growth of InP. Secondary ion mass spectrometry measurements <strong>in</strong>dicated that very<br />

472


Zn InP Zn<br />

marked Zn diffusion occurred when the Zn concentration appeared to reduce the pyrolytic<br />

efficiency of trimethyl<strong>in</strong>dium.<br />

W.T.Tsang, F.S.Choa, N.T.Ha: Journal of Electronic Materials, 1991, 20[7], 541-4<br />

[446-84/85-054]<br />

InP: Zn <strong>Diffusion</strong><br />

The non-<strong>in</strong>tentional diffusion of Zn acceptors was <strong>in</strong>vestigated dur<strong>in</strong>g low-pressure<br />

metalorganic vapor-phase epitaxial growth at 550 or 640C. It was found that the diffusion<br />

of Zn dur<strong>in</strong>g deposition could be described by an <strong>in</strong>terstitial-substitutional model. The<br />

diffusivity <strong>in</strong> the non-<strong>in</strong>tentionally doped material was lower than that <strong>in</strong> <strong>in</strong>tentionally<br />

doped material. This was attributed to the low concentration of <strong>in</strong>terstitial Zn atoms <strong>in</strong><br />

samples which were doped dur<strong>in</strong>g growth. Some deposition parameters, such as a high<br />

temperature <strong>and</strong> a high V/<strong>III</strong> ratio, m<strong>in</strong>imized diffusion. In this way, a normalized<br />

diffusivity which could be as low as 6.5 x 10 -14 cm 2 /s could be obta<strong>in</strong>ed at a dopant level<br />

of 10 18 /cm 3 .<br />

M.Glade, J.Hergeth, D.Grützmacher, K.Masseli, P.Balk: Journal of Crystal Growth,<br />

1991, 108[3-4], 449-54<br />

[446-84/85-054]<br />

InP: Zn <strong>Diffusion</strong><br />

The saturation behavior of the free carrier concentrations <strong>in</strong> p-type InP monocrystals<br />

which had been doped by Zn diffusion was <strong>in</strong>vestigated. The maximum free-hole<br />

concentration appeared at about 5 x 10 18 /cm 3 . The difference <strong>in</strong> saturation hole<br />

concentrations of materials was <strong>in</strong>vestigated by study<strong>in</strong>g the <strong>in</strong>corporation <strong>and</strong> lattice<br />

location of Zn. The latter was an acceptor when located on a group-<strong>III</strong> atom site. Z<strong>in</strong>c<br />

was diffused <strong>in</strong>to <strong>III</strong>-V wafers <strong>in</strong> a sealed quartz ampoule. Particle-<strong>in</strong>duced X-ray<br />

emission <strong>and</strong> ion-channell<strong>in</strong>g techniques were then used to determ<strong>in</strong>e the exact lattice<br />

location of Zn atoms. In InP, the substitutional state of Zn depended upon the cool<strong>in</strong>g rate<br />

of the sample after high-temperature diffusion. In slowly cooled samples, a large fraction<br />

(about 90%) of the Zn atoms formed r<strong>and</strong>om precipitates of Zn 3 P 2 <strong>and</strong> elemental Zn.<br />

However, after rapid cool<strong>in</strong>g, only 60% of the Zn atoms formed such precipitates while<br />

the rema<strong>in</strong>der occupied specific sites. The results were analyzed <strong>in</strong> terms of the<br />

amphoteric native defect model. It was shown that differences <strong>in</strong> the electrical activities<br />

of Zn atoms were a consequence of differ<strong>in</strong>g locations of the Fermi-level stabilization<br />

energy.<br />

L.Y.Chan, K.M.Yu, M.Ben-Tzur, E.E.Haller, J.M.Jaklevic, W.Walukiewicz,<br />

C.M.Hanson: Journal of Applied Physics, 1991, 69[5], 2998-3006<br />

[446-78/79-015]<br />

InP: Zn <strong>Diffusion</strong><br />

The characteristics of Fe-doped semi-<strong>in</strong>sulat<strong>in</strong>g layers, with overgrown Zn-doped p-type<br />

layers, were <strong>in</strong>vestigated by means of scann<strong>in</strong>g electron microscopy, secondary ion mass<br />

spectrometry, capacitance-voltage, <strong>and</strong> current-voltage measurements. The resistivity<br />

which was deduced from the current-voltage characteristics was found to be strongly<br />

dependent upon the Zn dopant concentration. The secondary ion mass spectrometry depth<br />

473


Zn InP Zn<br />

profiles revealed the occurrence of Zn accumulation at the semi-<strong>in</strong>sulat<strong>in</strong>g/p-type<br />

<strong>in</strong>terface, <strong>and</strong> the peak concentration of Zn accumulation <strong>in</strong>creased with dop<strong>in</strong>g level <strong>and</strong><br />

overgrowth time of the p-type layers. The accumulation of Zn at the semi-<strong>in</strong>sulat<strong>in</strong>g/ptype<br />

<strong>in</strong>terface was related to a reduction <strong>in</strong> semi-<strong>in</strong>sulat<strong>in</strong>g layer resistivity. The<br />

accumulation of Zn at the <strong>in</strong>terface could be m<strong>in</strong>imized by us<strong>in</strong>g a short growth time,<br />

together with low or medium dop<strong>in</strong>g of the p-type layers. Such growth conditions led to<br />

higher semi-<strong>in</strong>sulat<strong>in</strong>g layer resistivity.<br />

W.H.Cheng, H.Kuwamoto, A.Appelbaum, D.Renner, S.W.Zehr: Journal of Applied<br />

Physics, 1991, 69[4], 1862-5<br />

[446-78/79-045]<br />

InP: Zn <strong>Diffusion</strong><br />

The migration of Zn was <strong>in</strong>vestigated by us<strong>in</strong>g two configurations. These were diffusion<br />

from an external source <strong>in</strong>to uniformly n-doped substrates, <strong>and</strong> diffusion between the<br />

layers of n-p-n-p-n structures which had been grown via metalorganic chemical vapor<br />

deposition. Alternat<strong>in</strong>g layers of p-type material (0.0005mm, [Zn] = 4 x 10 17 to 2 x 10 18<br />

/cm 3 ) <strong>and</strong> n-type material (0.0005mm, [Si] = 10 16 to 3 x 10 19 /cm 3 ) were grown by us<strong>in</strong>g<br />

low-pressure metalorganic chemical vapor deposition at 625C. The distributions of Zn<br />

were determ<strong>in</strong>ed by means of secondary ion mass spectrometry. In the case of un-doped<br />

spacer layers (with n approximately equal to 10 16 /cm 3 ), the diffusion profiles depended<br />

markedly upon the Zn dopant level. Little Zn out-diffusion was observed when [Zn] was<br />

equal to 4 x 10 17 /cm 3 . When [Zn] was greater than 10 18 /cm 3 , the Zn diffused completely<br />

across the spacer layers dur<strong>in</strong>g growth times of 1 to 2h. In the case of doped spacer<br />

layers, the dop<strong>in</strong>g level of Si had a marked effect upon the Zn diffusion profiles. The total<br />

Zn diffusion across the grown dopant <strong>in</strong>terface was not substantially affected, but<br />

accumulation of Zn occurred <strong>in</strong> the Si-doped layers; with the formation of Zn spikes for<br />

which the <strong>in</strong>crease <strong>in</strong> Zn level - as compared to that (about 10 18 /cm 3 ) of the Zn-doped<br />

layer - was similar to [Si]. Electrochemical capacitance-voltage profil<strong>in</strong>g <strong>in</strong>dicated that<br />

the Zn was electrically active. The results were expla<strong>in</strong>ed <strong>in</strong> terms of a model <strong>in</strong> which<br />

the mobile Zn species that diffused <strong>in</strong>to the Si-doped layers were immobilized by the<br />

formation of Zn-donor pairs. This model was shown to be consistent with the profiles<br />

which were obta<strong>in</strong>ed for Zn diffusion <strong>in</strong>to n-type material from an external ZnGaCdIn<br />

source.<br />

C.Blaauw, F.R.Shepherd, D.Eger: Journal of Applied Physics, 1989, 66[2], 605-10<br />

[446-74-041]<br />

InP: Zn <strong>Diffusion</strong><br />

The ampoule diffusion of Zn gave rise to donor-acceptor photolum<strong>in</strong>escence transitions<br />

with peak positions that depended upon the cool<strong>in</strong>g rate after diffusion. Subsequent<br />

anneal<strong>in</strong>g <strong>in</strong> a Zn-free ambient caused a shift <strong>in</strong> the peak position. Lum<strong>in</strong>escence peaks<br />

were found between 1.30 <strong>and</strong> 1.38eV. These peaks were attributed to transitions between<br />

various Zn <strong>in</strong>terstitial donor levels <strong>and</strong> the Zn substitutional acceptor level. The<br />

lum<strong>in</strong>escence data were correlated with secondary ion mass spectrometry <strong>and</strong> Schottky<br />

barrier capacitance-voltage measurements, <strong>and</strong> were found to be consistent with an earlier<br />

474


Zn InP Zn<br />

model <strong>in</strong> which Zn was supposed to diffuse as both an <strong>in</strong>terstitial donor <strong>and</strong> a<br />

substitutional acceptor.<br />

E.A.Montie, G.J.Van Gurp: Journal of Applied Physics, 1989, 66[11], 5549-53<br />

[446-74-041]<br />

InP: Zn <strong>Diffusion</strong><br />

The diffusion of Zn was studied by us<strong>in</strong>g boat diffusion, diffusion from As- or P-doped<br />

spun-on films, or diffusion from In-doped spun-on films. The depth profiles were<br />

deduced from junction positions. It was found that the p + /p - junction position depended<br />

upon the diffusion method which was used, but not upon the sample growth technique.<br />

The p - /n junction position depended upon both factors. Because the amounts of In <strong>and</strong> P<br />

(or of the respective vacancies) differed, it was possible to identify diffusion mechanisms.<br />

It was proposed that <strong>in</strong>terstitially diffus<strong>in</strong>g Zn was <strong>in</strong>dependently trapped by 2 immobile<br />

vacancy centers. These consisted of Zn on V In <strong>in</strong> the p + region, <strong>and</strong> Zn on V P ZnV P <strong>in</strong> the<br />

p - region.<br />

U.König, H.Haspeklo, P.Marschall, M.Kuisl: Journal of Applied Physics, 1989, 65[2],<br />

548-52<br />

[446-72/73-035]<br />

InP: Zn <strong>Diffusion</strong><br />

Closed-ampoule diffusion led to a net acceptor concentration which was lower than the<br />

Zn concentration. Upon anneal<strong>in</strong>g <strong>in</strong> an atmosphere without Zn, the Zn <strong>and</strong> net acceptor<br />

concentrations became almost identical. This was attributed to a decreased Zn<br />

concentration <strong>and</strong> an <strong>in</strong>creased net acceptor concentration. The results were quantitatively<br />

expla<strong>in</strong>ed by assum<strong>in</strong>g that the Zn was <strong>in</strong>corporated as both substitutional acceptors <strong>and</strong><br />

<strong>in</strong>terstitial donors, <strong>and</strong> that only the <strong>in</strong>terstitial Zn was driven out by anneal<strong>in</strong>g; due to its<br />

large diffusion coefficient. The profiles which were calculated by us<strong>in</strong>g this <strong>in</strong>terstitialsubstitutional<br />

model could be fitted to experimentally determ<strong>in</strong>ed profiles by assum<strong>in</strong>g<br />

that the Zn diffused as s<strong>in</strong>gly-ionized <strong>in</strong>terstitial donors. The present model also<br />

expla<strong>in</strong>ed published data, on diffusion <strong>in</strong> n-type InP, <strong>in</strong> which a profile cut-off was found<br />

at a depth where the acceptor concentration equalled the background donor concentration.<br />

G.J.Van Gurp, T.Van Dongen, G.M.Fontijn, J.M.Jacobs, D.L.A.Tjaden: Journal of<br />

Applied Physics, 1989, 65[2], 553-60<br />

[446-72/73-037]<br />

InP: Zn <strong>Diffusion</strong><br />

The Zn was diffused, from a dimethylz<strong>in</strong>c source, at temperatures rang<strong>in</strong>g from 400 to<br />

570C us<strong>in</strong>g surface concentrations rang<strong>in</strong>g from 10 17 to 10 18 /cm 3 . The resultant diffusion<br />

profiles were determ<strong>in</strong>ed by us<strong>in</strong>g secondary ion mass spectrometry, electrolytic etch<strong>in</strong>g,<br />

<strong>and</strong> capacitance-voltage measurements. The results <strong>in</strong>dicated that an <strong>in</strong>terstitialsubstitutional<br />

mechanism operated at the above concentrations.<br />

M.Wada, M.Seko, K.Sakakibara, Y.Sekiguchi: Japanese Journal of Applied Physics,<br />

1989, 28[10], L1700-3<br />

[446-72/73-039]<br />

475


Zn InP Zn<br />

InP: Zn <strong>Diffusion</strong><br />

It was po<strong>in</strong>ted out that the use of models, which did not <strong>in</strong>volve dopant effects upon the<br />

<strong>in</strong>terstitial-substitutional <strong>in</strong>terchange, could lead to the identification of an apparently<br />

larger charge state for the <strong>in</strong>terstitial. This was expected to <strong>in</strong>crease the apparent<br />

activation energy for diffusion. This <strong>in</strong>crease was approximately equivalent to the Zn<br />

solubility activation energy for substrate dop<strong>in</strong>g near to the substitutional Zn<br />

concentration at the surface.<br />

C.Kazmierski: Journal of Applied Physics, 1988, 64[11], 6573-5<br />

[446-72/73-039]<br />

InP: Zn <strong>Diffusion</strong><br />

The Zn was diffused <strong>in</strong>to an unpassivated surface, from an open gas flow system, at<br />

temperatures of between 733 <strong>and</strong> 773K. In the region where the carrier concentration<br />

profile could be described by an erfc function, the diffusivity was given by:<br />

D(m 2 /s) = 3 x 10 -7 exp[-120(kJ/mol)/RT]<br />

It was shown that thermal processes caused changes <strong>in</strong> the charge state of Zn <strong>in</strong> InP.<br />

These resulted <strong>in</strong> a variation of the carrier profile.<br />

T.O.Budko, E.V.Gushch<strong>in</strong>skaya, J.S.Emelyanenko, S.A.Malyshev: Physica Status Solidi<br />

A, 1989, 111[2], 451-6<br />

[446-64/65-174]<br />

InP: Zn <strong>Diffusion</strong><br />

The profiles of Zn <strong>in</strong> n-type [100] wafers after ampoule diffusion were measured us<strong>in</strong>g<br />

secondary-ion mass spectrometry, Auger electron spectrometry, differential Hall-effect<br />

measurements, capacitance measurements, <strong>and</strong> scann<strong>in</strong>g electron microscopy. The results<br />

could be expla<strong>in</strong>ed <strong>in</strong> terms of an <strong>in</strong>terstitial-substitutional mechanism <strong>in</strong> which the Zn<br />

diffused as a s<strong>in</strong>gly ionized <strong>in</strong>terstitial <strong>and</strong> was <strong>in</strong>corporated <strong>in</strong>to the In sub-lattice as an<br />

electrically active substitutional acceptor or as an electrically <strong>in</strong>active complex. At Zn<br />

concentrations which were lower than the background donor concentration, the profile<br />

was cut off as <strong>in</strong>terstitial diffusion broke down. The activation energies for diffusion <strong>and</strong><br />

solubility were found to be 1.40 <strong>and</strong> 1.0eV, respectively.<br />

G.J.Van Gurp, P.R.Boudewijn, M.N.C.Kempeners, D.L.A.Tjaden: Journal of Applied<br />

Physics, 1987, 61[5], 1846-55<br />

[446-60-009]<br />

InP: Zn <strong>Diffusion</strong><br />

The results of open-tube Zn diffusion <strong>in</strong>to undoped or S-doped n-type material at<br />

temperatures rang<strong>in</strong>g from 550 to 675C were presented. The results were consistent with<br />

<strong>in</strong>terstitial-substitutional diffusion. In the case of undoped samples, the results were<br />

described by:<br />

D (cm 2 /s) = 0.049 exp[-1.52(eV)/kT]<br />

In the case of heavily S-doped samples, the results were described by:<br />

D (cm 2 /s) = 1400 exp[-2.34(eV)/kT]<br />

476


Zn InP Zn<br />

The difference <strong>in</strong> the activation energies was comparable to the Fermi level difference for<br />

the two substrate types, <strong>and</strong> was consistent with the differ<strong>in</strong>g diffusion mechanisms<br />

which occurred <strong>in</strong> these two types of InP.<br />

H.S.Marek, H.B.Serreze: Applied Physics Letters, 1987, 51[24], 2031-3<br />

[446-60-010]<br />

InP: Zn <strong>Diffusion</strong><br />

The substitutional fraction of Zn atoms which was diffused <strong>in</strong>to s<strong>in</strong>gle crystals was<br />

measured by us<strong>in</strong>g the proton-<strong>in</strong>duced X-ray excitation technique. The diffusion times<br />

ranged from 0.25 to 1h at 425 to 650C. For several samples with diffusion depths rang<strong>in</strong>g<br />

from 0.00075 to 0.0037mm (as determ<strong>in</strong>ed us<strong>in</strong>g secondary ion mass spectrometry), it<br />

was found that the Zn impurity atoms resided almost entirely on lattice sites <strong>and</strong> that the<br />

substitutional fraction was equal to 0.9. There was no evidence of precipitation <strong>in</strong> the<br />

diffused layers. Only 1 to 10% of the Zn was electrically active. This was consistent with<br />

the existence of neutral V P Zn In V P complexes.<br />

W.N.Lennard, M.L.Swanson, D.Eger, A.J.Spr<strong>in</strong>gthorpe, F.R.Shepherd: Journal of<br />

Electronic Materials, 1988, 17[1], 1-4<br />

[446-60-010]<br />

InP: Zn <strong>Diffusion</strong><br />

It was recalled that, when p-n junctions were formed by dop<strong>in</strong>g with an element that<br />

diffused via a dissociative mechanism, dopant diffusion was suppressed <strong>and</strong> dopants<br />

could pile up near to the junction; at well above their orig<strong>in</strong>al concentration. Calculations<br />

confirmed this behavior, if no local neutrality was assumed. The results agreed well with<br />

published experimental data on Zn diffusion <strong>in</strong> the present material. It was noted that the<br />

<strong>in</strong>creased built-<strong>in</strong> electric field, due to this pile-up, was expelled almost entirely to the<br />

side of the junction without the pile-up. It was suggested that this effect had important<br />

implications for devices, which conta<strong>in</strong>ed th<strong>in</strong> <strong>and</strong>/or small regions that were doped with<br />

such elements, because such regions might become completely depleted.<br />

I.Lyubomirsky, V.Lyahovitskaya, D.Cahen: Applied Physics Letters, 1997, 70[5], 613-5<br />

[446-150/151-148]<br />

InP: Zn <strong>Diffusion</strong><br />

The <strong>in</strong>itial stages of Zn diffusion <strong>in</strong>to InP from a polymer sp<strong>in</strong>-on film were <strong>in</strong>vestigated.<br />

It was found that there were high concentrations of micro-defects <strong>and</strong> extended defects <strong>in</strong><br />

the near-surface region (down to 1µ), an anomalously deep penetration of Zn atoms (with<br />

a diffusivity that was almost <strong>in</strong>dependent of temperature) <strong>and</strong> a low degree of activation<br />

of the diffused Zn. A kick-out mechanism was thought to predom<strong>in</strong>ate <strong>in</strong> the <strong>in</strong>itial<br />

stages.<br />

A.V.Kaman<strong>in</strong>, I.A.Mok<strong>in</strong>a, N.M.Shmidt: Solid-State Electronics, 1996, 39[10], 1441-4<br />

[446-141/142-111]<br />

477


Zn InP Zn<br />

InP: Zn <strong>Diffusion</strong><br />

It was noted that substitutional Zn was l<strong>in</strong>early <strong>in</strong>corporated <strong>in</strong>to device-quality material<br />

under low Zn-source flow-rates dur<strong>in</strong>g atmospheric-pressure metalorganic vapor-phase<br />

epitaxy at 625C. It saturated at about 4 x 10 18 /cm 3 under high Zn-source flow-rates. An<br />

<strong>in</strong>crease, <strong>in</strong> the Zn-source flow-rate, to beyond saturation significantly <strong>in</strong>creased the<br />

amount of <strong>in</strong>terstitial <strong>in</strong>corporation. The excess <strong>in</strong>terstitials diffused <strong>in</strong>to the undoped<br />

region via an <strong>in</strong>terstitial-substitutional diffusion mechanism, <strong>and</strong> revealed themselves via<br />

an enhanced diffusivity. It was recalled that a model had previously been proposed, for<br />

surface adsorption-desorption trapp<strong>in</strong>g dur<strong>in</strong>g substitutional Zn <strong>in</strong>corporation, <strong>in</strong> which<br />

the saturation level was assumed to be governed by surface <strong>in</strong>corporation sites for<br />

substitutional Zn. This model was applied here, to <strong>in</strong>terstitial Zn <strong>in</strong>corporation at Zn<br />

source flow rates which were above the saturation level for substitutional Zn, <strong>in</strong> order to<br />

expla<strong>in</strong> the enhanced Zn diffusion. The analysis was extended so as to <strong>in</strong>clude the<br />

<strong>in</strong>corporation of neutral Zn <strong>in</strong> the presence of excess P vacancies. It was concluded that<br />

this model could be used for the simultaneous <strong>in</strong>corporation of Zn of all 3 types dur<strong>in</strong>g<br />

epitaxy; provided that the <strong>in</strong>corporation processes were <strong>in</strong>dependent.<br />

S.N.G.Chu, R.A.Logan, M.Geva, N.T.Ha, R.F.Karlicek: Journal of Applied Physics,<br />

1996, 80[6], 3221-7<br />

[446-138/139-098]<br />

InP: Zn <strong>Diffusion</strong><br />

The diffusion <strong>and</strong> <strong>in</strong>corporation characteristics of Zn dopants <strong>in</strong> organometallic vaporphase<br />

epitaxially grown material were studied. The Zn diffusion coefficient depended<br />

strongly upon the concentration, <strong>and</strong> <strong>in</strong>creased by 4 orders of magnitude for Zn<br />

concentrations of between 2 x 10 18 <strong>and</strong> 8 x 10 18 /cm 3 . This marked concentration<br />

dependence of the Zn diffusion coefficient was shown to govern Zn <strong>in</strong>corporation dur<strong>in</strong>g<br />

organometallic vapor-phase epitaxial growth. The spread of Zn dopants <strong>in</strong>to <strong>in</strong>tentionally<br />

undoped regions could result <strong>in</strong> high Zn dopant concentrations.<br />

E.F.Schubert, C.J.P<strong>in</strong>zone, M.Geva: Applied Physics Letters, 1995, 67[5], 700-2<br />

[446-123/124-178]<br />

InP: Zn <strong>Diffusion</strong><br />

The concentration-dependent diffusion of Zn dur<strong>in</strong>g metalorganic vapor-phase epitaxy<br />

from a Zn-doped InP layer, <strong>and</strong> <strong>in</strong>to the adjacent undoped InP buffer layer, was studied<br />

by means of secondary ion mass spectroscopy <strong>and</strong> carrier concentration profil<strong>in</strong>g. If the<br />

growth rate of the Zn-doped film was faster than the <strong>in</strong>terdiffusion of Zn <strong>in</strong>to the<br />

underly<strong>in</strong>g undoped buffer layer, the diffusion problem could be treated as a 1-<br />

dimensional couple between 2 semi-<strong>in</strong>f<strong>in</strong>ite media. Also, Zn diffusion under optimum<br />

growth conditions completely elim<strong>in</strong>ated the thermal decomposition problem which was<br />

encountered when us<strong>in</strong>g sealed-ampoule or open-tube methods, <strong>and</strong> also reta<strong>in</strong>ed all of<br />

the <strong>in</strong>tr<strong>in</strong>sic po<strong>in</strong>t defects <strong>in</strong> their thermodynamic equilibrium concentrations. When us<strong>in</strong>g<br />

an optimum growth temperature of 625C, <strong>and</strong> a maximum Zn flow that was below the<br />

<strong>in</strong>corporation limit for substitutional Zn (<strong>in</strong> order to ensure that the Zn was <strong>in</strong>corporated<br />

478


Zn InP Zn<br />

substitutionally), the diffusion profiles of Zn across the <strong>in</strong>terface could be simulated by<br />

assum<strong>in</strong>g a concentration-dependent diffusivity. A third-power concentration dependence<br />

of the effective diffusion coefficient was found. This applied to both Frank-Turnbull <strong>and</strong><br />

kick-out equilibrium mechanisms for an <strong>in</strong>terstitial-substitutional diffusion model. This<br />

<strong>in</strong>dicated a 2+ charge state for the fast-diffus<strong>in</strong>g Zn <strong>in</strong>terstitials. Extrapolations <strong>in</strong>to the<br />

high-concentration regime of sealed-ampoule experiments generally agreed with<br />

published data, although the predom<strong>in</strong>ant Zn atoms which were found <strong>in</strong> the highconcentration<br />

regime formed complexes with P vacancies <strong>in</strong> a neutral state.<br />

S.N.G.Chu, R.A.Logan, M.Geva, N.T.Ha: Journal of Applied Physics, 1995, 78[5], 3001-<br />

7<br />

[446-123/124-178]<br />

InP: Zn <strong>Diffusion</strong><br />

A new sp<strong>in</strong>-on solution was proposed for Zn diffusion <strong>in</strong>to this material. The solution<br />

consisted of a Zn-SiO 2 sol with a very long shelf-life. After sp<strong>in</strong>n<strong>in</strong>g-on the sol, a<br />

ZnO/SiO 2 film was produced on the InP wafer. It was found that, for a thickness of<br />

150nm, the film acted as an <strong>in</strong>exhaustible source for short-time (30s, 650C) diffusions.<br />

U.Schade, B.Unger: Semiconductor Science <strong>and</strong> Technology, 1993, 8[12], 2048-52<br />

[446-115/116-145]<br />

InP: Zn <strong>Diffusion</strong><br />

The defects which were <strong>in</strong>troduced by Zn diffusion were studied by measur<strong>in</strong>g the<br />

photolum<strong>in</strong>escence <strong>and</strong> photo-emission spectra of Zn-diffused samples which had been<br />

fabricated by us<strong>in</strong>g a new diffusion technique. The results <strong>in</strong>dicated that Zn diffusion<br />

generated broad emission b<strong>and</strong>s, with energies rang<strong>in</strong>g from 0.7 to 1eV, only <strong>in</strong> a surface<br />

layer with a thickness of less than about 100nm. It also left a P-rich layer with a very high<br />

Zn concentration <strong>and</strong> a thickness of less than about 20nm. It was suggested that Zn<br />

diffusion from a high-Zn concentration source, under P-rich conditions, occurred near to<br />

the surface <strong>and</strong> <strong>in</strong>troduced deep centers which were responsible for the b<strong>and</strong>s.<br />

M.Wada, K.Sakakibara: Japanese Journal of Applied Physics, 1993, 32[2-4A], L469-72<br />

[446-109/110-042]<br />

InP: Zn <strong>Diffusion</strong><br />

The electrical activity <strong>and</strong> lattice-site locations of Zn atoms which had been diffused <strong>in</strong>to<br />

InP were studied by us<strong>in</strong>g various characterization techniques. Particle-<strong>in</strong>duced X-ray<br />

emission channell<strong>in</strong>g showed that, <strong>in</strong> InP, most of the Zn atoms were situated <strong>in</strong><br />

<strong>in</strong>terstitial sites or formed r<strong>and</strong>om Zn precipitates which were electrically <strong>in</strong>active. The<br />

distribution of Zn was shown to depend upon the cool<strong>in</strong>g rate after high-temperature<br />

diffusion. The difference between the behaviors of Zn <strong>in</strong> <strong>GaAs</strong> <strong>and</strong> InP could be<br />

understood <strong>in</strong> terms of the amphoteric native defect model. It was also shown that the<br />

479


Zn InP Zn<br />

Fermi level stabilization energy provided a convenient energy reference for the treatment<br />

of dopant diffusion at semiconductor hetero-<strong>in</strong>terfaces.<br />

W.Walukiewicz, K.M.Yu, L.Y.Chan, J.Jaklevic, E.E.Haller: Materials Science Forum,<br />

1992, 83-87, 941-6<br />

[446-99/100-065]<br />

InP: Zn <strong>Diffusion</strong><br />

A stripe heater was used to diffuse Zn <strong>in</strong>to semi-<strong>in</strong>sulat<strong>in</strong>g, or n-type, (001) samples from<br />

a th<strong>in</strong> spun-on silica film. The diffusion profiles were determ<strong>in</strong>ed by means of secondary<br />

ion mass spectrometry <strong>and</strong> capacitance-voltage measurements. The diffusion depth <strong>and</strong><br />

activation energy of the effective diffusion coefficient were compared with published data<br />

on Zn ampoule diffusion. An activation energy of 1.35eV was found for the diffusion of<br />

Zn <strong>in</strong> semi-<strong>in</strong>sulat<strong>in</strong>g InP. An expression for the effective diffusion coefficient of Zn <strong>in</strong><br />

InP was derived <strong>and</strong> was compared with the results of a Boltzmann-Matano analysis of<br />

diffusion profiles. It was concluded that the Zn diffusion could be expla<strong>in</strong>ed by an<br />

<strong>in</strong>terstitial-substitutional model, <strong>in</strong> which Zn diffused as a s<strong>in</strong>gly positively charged<br />

<strong>in</strong>terstitial <strong>and</strong> acted as an acceptor by fill<strong>in</strong>g an In vacancy.<br />

U.Schade, P.Enders: Semiconductor Science <strong>and</strong> Technology, 1992, 7[6], 752-7<br />

[446-99/100-093]<br />

InP/In<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

It was shown that the growth of emitter layers of InP/In<strong>GaAs</strong>/InP double heterojunction<br />

bipolar transistors could result <strong>in</strong> significant Zn diffusion from the base <strong>and</strong> <strong>in</strong>to the<br />

collector. The extent of the diffusion depended upon the n-dop<strong>in</strong>g level of the emitter.<br />

This behavior was expla<strong>in</strong>ed <strong>in</strong> terms of non-equilibrium po<strong>in</strong>t defects which were<br />

<strong>in</strong>troduced by a comb<strong>in</strong>ation of surface p<strong>in</strong>n<strong>in</strong>g of the Fermi level, <strong>and</strong> n-dop<strong>in</strong>g. It was<br />

also shown that Zn diffusion could be greatly reduced by us<strong>in</strong>g AlInAs, <strong>in</strong>stead of InP, as<br />

the emitter layer. The difference <strong>in</strong> behavior was shown to be at least partly due to the<br />

lower diffusivity of group-<strong>III</strong> <strong>in</strong>terstitials <strong>in</strong> AlInAs. Moreover, it was shown that the<br />

<strong>in</strong>troduction of only 50nm of AlInAs between emitter <strong>and</strong> base resulted <strong>in</strong> a significant<br />

reduction of Zn diffusion <strong>in</strong>to the collector.<br />

R.Bhat, M.A.Koza, J.I.Song, S.A.Schwarz, C.Caneau, W.P.Hong: Applied Physics<br />

Letters, 1994, 65[3], 338-40<br />

[446-119/120-218]<br />

InP/In<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

It was recalled that high n + -dop<strong>in</strong>g, of the cap layers of heterojunction structures,<br />

produced anomalous Zn diffusion <strong>in</strong> the base region dur<strong>in</strong>g metalorganic vapor phase<br />

epitaxial growth. This was attributed to non-equilibrium group-<strong>III</strong> <strong>in</strong>terstitials that were<br />

generated <strong>in</strong> the cap layer, <strong>and</strong> created highly diffusive Zn <strong>in</strong>terstitials via the kick-out<br />

mechanism. It was shown here that low-temperature (550C) growth was effective <strong>in</strong><br />

reduc<strong>in</strong>g the effect of the n + cap layer. Due to a large time constant for the recovery of<br />

thermal po<strong>in</strong>t defect equilibrium, the last-to-grow n + cap layer could not <strong>in</strong>ject excessive<br />

numbers of group-<strong>III</strong> <strong>in</strong>terstitials <strong>in</strong>to the base region dur<strong>in</strong>g growth. However, dur<strong>in</strong>g<br />

480


Zn InP Zn<br />

low-temperature growth the first-to-grow n + sub-collector produced group-<strong>III</strong> <strong>in</strong>terstitials<br />

<strong>and</strong> thus caused anomalous Zn diffusion. In order to prevent this effect, it was suggested<br />

that growth should be <strong>in</strong>terrupted for 0.5h before grow<strong>in</strong>g the base layer, <strong>and</strong> that growth<br />

of the n + sub-collector should be carried out at 600C. These changes were effective <strong>in</strong><br />

remov<strong>in</strong>g undesirable group-<strong>III</strong> <strong>in</strong>terstitials.<br />

K.Kurishima, T.Kobayashi, H.Ito, U.Gösele: Journal of Applied Physics, 1996, 79[8],<br />

4017-23<br />

[446-134/135-152]<br />

InP/In<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

It was recalled that highly n + -doped sub-collector layers <strong>in</strong> such heterojunction bipolar<br />

transistor structures led to markedly enhanced Zn diffusion <strong>in</strong> the subsequently grown<br />

base layer. It was shown that this abnormal Zn diffusion could be suppressed by<br />

<strong>in</strong>terrupt<strong>in</strong>g growth before the Zn-doped layer was grown. It was speculated that this<br />

<strong>in</strong>terruption of growth permitted excess non-equilibrium group-<strong>III</strong> self-<strong>in</strong>terstitials,<br />

com<strong>in</strong>g from the n + -doped sub-collector layer, to disappear before they could enhance Zn<br />

diffusion <strong>in</strong> the base layer.<br />

T.Kobayashi, K.Kurishima, U.Gösele: Applied Physics Letters, 1993, 62[3], 284-5<br />

[446-106/107-127]<br />

InP/In<strong>GaAs</strong>: Zn <strong>Diffusion</strong><br />

An <strong>in</strong>vestigation was made of the behavior of Zn impurities <strong>in</strong> heterojunction bipolar<br />

transistor structures that had been grown by us<strong>in</strong>g a low-pressure metalorganic chemical<br />

vapor deposition technique. In this technique, Zn was anomalously diffused <strong>in</strong>to<br />

InP/In<strong>GaAs</strong> heterojunction bipolar transistors with a heavily Si-doped (2 x 10 19 /cm 3 ) subcollector<br />

layer when the growth temperature before base-layer growth was higher or<br />

lower than about 600C. On the <strong>other</strong> h<strong>and</strong>, an abrupt Zn profile <strong>in</strong> the same<br />

heterojunction bipolar transistor structure was obta<strong>in</strong>ed when the growth temperature of<br />

the sub-collector layer was 600C. Two types of po<strong>in</strong>t defect reaction, which depended<br />

strongly upon the growth temperature, were proposed. In one type, excess group-<strong>III</strong><br />

<strong>in</strong>terstitials which were produced <strong>in</strong> a heavily Si-doped sub-collector layer were easily<br />

removed. In the <strong>other</strong> type, the equilibrium concentration of charged group-<strong>III</strong> vacancies<br />

decreased as the growth temperature was <strong>in</strong>creased. Accord<strong>in</strong>g to this model, complete<br />

suppression of unwanted Zn diffusion could be achieved by us<strong>in</strong>g a growth <strong>in</strong>terruption<br />

technique to obta<strong>in</strong> po<strong>in</strong>t defect equilibrium before base-layer growth at 550C when the<br />

growth temperature of the sub-collector layer was either lower or higher than 600C. It<br />

was concluded that <strong>in</strong>terruption of the growth for a suitable period of time before baselayer<br />

growth, or the use of a suitable growth temperature for the sub-collector, was<br />

essential <strong>in</strong> order to obta<strong>in</strong> an abrupt Zn profile <strong>in</strong> a heterojunction bipolar transistor<br />

structure with a heavily doped sub-collector layer.<br />

T.Kobayashi, K.Kurishima, U.Gösele: Journal of Crystal Growth, 1995, 146[1-4], 533-7<br />

[446-127/128-152]<br />

481


Zn InP Interdiffusion<br />

InP/In<strong>GaAs</strong>P: Zn <strong>Diffusion</strong><br />

In order to prevent carriers <strong>in</strong> multi-layer heterostructures from be<strong>in</strong>g redistributed, a new<br />

method for open-tube Zn diffusion, us<strong>in</strong>g an In-Zn alloy as the source <strong>and</strong> a<br />

polycrystall<strong>in</strong>e InP cover to limit surface thermo-damage at temperature as low as 500C,<br />

was developed. It was found that the diffusion rate was proportional to the square of the P<br />

content. When us<strong>in</strong>g proper mask<strong>in</strong>g, the ratio of the lateral width to the diffusion depth<br />

was about 0.6.<br />

W.Li, H.Pan: Journal of the Electrochemical Society, 1987, 134[9], 2329-32<br />

[446-55/56-030]<br />

General<br />

InP: <strong>Diffusion</strong><br />

Facet growth near to SiO 2 mask edges, dur<strong>in</strong>g metalorganic molecular beam epitaxy, was<br />

studied for various V/<strong>III</strong> ratios on (100) substrates with a 2°misorientation towards (110).<br />

It was found that, whereas ideal vertical layer growth occurred at high V/<strong>III</strong> ratios (even<br />

after 2µ of growth), oblique (111) planes were k<strong>in</strong>etically favored near to mask edges at<br />

lower V/<strong>III</strong> ratios. The V/<strong>III</strong> ratio was a key parameter s<strong>in</strong>ce it determ<strong>in</strong>ed the facets with<br />

the lowest k<strong>in</strong>etically limited growth rate at the border of the grow<strong>in</strong>g layer. Also, the<br />

diffusion length of mobile adsorbed species, which expla<strong>in</strong>ed the presence of additional<br />

features near to mask edges <strong>and</strong> corners, decreased with the V/<strong>III</strong> ratio. As well as <strong>in</strong>terfacet<br />

diffusion which was driven by concentration gradients between facets with differ<strong>in</strong>g<br />

growth rates, there was also evidence for the occurrence of anisotropic diffusion along<br />

[0¯11] on (100) InP. It was suggested that this was the cause of the f<strong>in</strong>e surface ripples<br />

which were observed on one side, near to the SiO 2 masks.<br />

R.Matz, H.He<strong>in</strong>ecke, B.Baur, R.Primig, C.Cremer: Journal of Crystal Growth, 1993,<br />

127[1-4], 230-6<br />

[446-106/107-120]<br />

Interdiffusion<br />

InP/GaInAsP: Interdiffusion<br />

Interdiffusion experiments <strong>and</strong> results for InP/GaInAs(P) heterostructures were<br />

considered <strong>in</strong> terms of a thermodynamic model. Important factors which affected<br />

<strong>in</strong>terdiffusion <strong>in</strong> the GaInAsP system were shown to <strong>in</strong>clude a miscibility gap, differ<strong>in</strong>g<br />

diffusivities on each of the sub-lattices of the 2 materials, Fermi level or impurity-<strong>in</strong>duced<br />

changes <strong>in</strong> diffusivity or diffusion mechanism, <strong>and</strong> the type of experiment. When a<br />

miscibility gap was present, the activity coefficients <strong>and</strong> solubilities of all of the species<br />

varied near to a heterojunction <strong>and</strong> caused the <strong>in</strong>terdiffusion to become strongly<br />

composition-dependent. At the usual growth <strong>and</strong> anneal<strong>in</strong>g temperatures, many<br />

superlattices were expected to equilibrate as 2 quaternary superlattices rather than as an<br />

homogeneous alloy. Differ<strong>in</strong>g diffusivities on the sub-lattices of a superlattice could lead<br />

482


Interdiffusion InP Interdiffusion<br />

to widen<strong>in</strong>g or narrow<strong>in</strong>g of quantum wells. When this occurred, optical measurements of<br />

the b<strong>and</strong>-gap energy were likely to be mislead<strong>in</strong>g, because of quantum size effects. The<br />

diffusivity on each sub-lattice could be altered by the presence of group-II, -IV, or -VI<br />

dopants. <strong>Diffusion</strong> on the group-<strong>III</strong> sub-lattice <strong>in</strong> p-type GaInAsP was found to be<br />

consistent with an <strong>in</strong>terstitialcy mechanism. The mechanism rema<strong>in</strong>ed unknown for n-<br />

type dop<strong>in</strong>g <strong>and</strong> for the group-V sub-lattice. Poorly designed <strong>and</strong> controlled experiments<br />

were found to be associated with large discrepancies <strong>in</strong> the observed diffusivities, with<br />

unreliable concentration profiles, <strong>and</strong> with the appearance of new condensed phases <strong>in</strong><br />

the solid. Experiments <strong>in</strong>dicated that the ordered Cu-Pt structure which was often found<br />

<strong>in</strong> GaIn(As)P epilayers was unstable, <strong>and</strong> was not stra<strong>in</strong>-stabilized relative to the<br />

disordered structure at normally used growth <strong>and</strong> anneal<strong>in</strong>g temperatures.<br />

R.M.Cohen: Journal of Applied Physics, 1993, 73[10], 4903-15<br />

[446-106/107-127]<br />

343 InP/GaP: Interdiffusion<br />

The formation of a solid solution was monitored by means of X-ray studies of annealed<br />

powder mixtures of the components. The results (table 43) <strong>in</strong>dicated that the overall<br />

activation energy for <strong>in</strong>terdiffusion at temperatures of between 650 <strong>and</strong> 725C was equal<br />

to 3.15eV.<br />

U.Vol<strong>and</strong>, R.Cerny, P.Deus, D.Bergner, G.Fenn<strong>in</strong>ger: Crystal Research <strong>and</strong> Technology,<br />

1989, 24[11], 1177-85<br />

[446-72/73-040]<br />

Table 43<br />

Interdiffusion Parameters for InP/GaP Mixtures<br />

Temperature (C) D o (cm 2 /s) Q(eV)<br />

650 - 800 0.007 2.1<br />

650 - 700 1000 3.2<br />

675 - 725 400 3.1<br />

InP/In<strong>GaAs</strong>: Interdiffusion<br />

A study was made of the <strong>in</strong>terfacial quality <strong>and</strong> thermal <strong>in</strong>terdiffusion of quantum wells<br />

which had been grown us<strong>in</strong>g hydride vapor phase epitaxy. It was deduced that isl<strong>and</strong> <strong>and</strong><br />

valley structures, with a height of one monolayer <strong>and</strong> a lateral extent which was of the<br />

order of one third of an exciton radius, existed at the <strong>in</strong>terface. The <strong>in</strong>terdiffusivity<br />

coefficient was estimated from the photolum<strong>in</strong>escence peak energy shift at 77K. Values<br />

of 2.5 x 10 -19 <strong>and</strong> 1.5 x 10 -18 cm 2 /s were deduced for temperatures of 700 <strong>and</strong> 750C,<br />

respectively. These values were more than 100 times higher than those for Al<strong>GaAs</strong>/<strong>GaAs</strong><br />

quantum well structures, <strong>and</strong> more than 100 times smaller than those for InAlAs/In<strong>GaAs</strong><br />

quantum well structures.<br />

K.Makita, K.Taguti: Superlattices <strong>and</strong> Microstructures, 1988, 4[1], 101-5<br />

[446-61-081]<br />

483


InSb<br />

Bi<br />

InSb: Bi <strong>Diffusion</strong><br />

Bulk s<strong>in</strong>gle crystals of In 1-x Ga x Sb 1-y Bi y (where x was between 0 <strong>and</strong> 0.21 <strong>and</strong> y was<br />

between 0 <strong>and</strong> 0.005) were grown onto InSb seed crystals by us<strong>in</strong>g a rotary Bridgman<br />

method. The quality of the crystals was assessed by us<strong>in</strong>g optical microscopic, X-ray<br />

topographic, 4-crystal X-ray diffractometric, electron-probe micro-analytical, energydispersive<br />

spectroscopic, <strong>and</strong> secondary-ion mass spectroscopic techniques. Due to<br />

segregation, the compositional ratio of Bi <strong>in</strong>creased as the crystals grew. Dur<strong>in</strong>g the<br />

growth of InGaSbBi, Bi diffused <strong>in</strong>to the InSb seed, <strong>and</strong> doma<strong>in</strong>s of InBi appeared. For<br />

comparison, InSb 1-y Bi y (where y was between 0 <strong>and</strong> 0.05) <strong>and</strong> In 1-x Ga x Sb (where x was<br />

between 0 <strong>and</strong> 0.16) were grown on InSb. It was found that Bi did not diffuse <strong>in</strong>to InSb<br />

without Ga, whereas Ga diffused without Bi. The <strong>in</strong>corporation of Ga produced excess In<br />

<strong>and</strong> led to the formation of InBi doma<strong>in</strong>s.<br />

Y.Hayakawa, M.Ando, T.Matsuyama, E.Hamakawa, T.Koyama, S.Adachi, K.Takahashi,<br />

V.G.Lifshits, M.Kumagawa: Journal of Applied Physics, 1994, 76[2], 858-64<br />

[446-117/118-189]<br />

Cd<br />

InSb: Cd <strong>Diffusion</strong><br />

A two-temperature zone method, applied to an InSb substrate plus Cd source system, was<br />

used. The diffusion profiles were determ<strong>in</strong>ed by us<strong>in</strong>g capacitance-voltage measurements,<br />

<strong>and</strong> were similar to those which were habitually found for <strong>other</strong> <strong>III</strong>-V systems. L<strong>in</strong>ear<br />

plots of junction-depth versus the square root of the diffusion time did not pass through<br />

the orig<strong>in</strong>. This suggested that the usual <strong>in</strong>terstitial-substitutional model <strong>and</strong> the<br />

conventional Boltzmann-Matano method could not be used to analyze the results. The<br />

diffusivity exhibited a maximum as a function of carrier concentration.<br />

S.L.Tu, K.F.Huang, S.J.Yang: Japanese Journal of Applied Physics, 1990, 29[3], 463-7<br />

[446-76/77-029]<br />

484


Ga InSb In<br />

Ga<br />

InSb: Ga <strong>Diffusion</strong><br />

Bulk s<strong>in</strong>gle crystals of In 1-x Ga x Sb 1-y Bi y (where x was between 0 <strong>and</strong> 0.21 <strong>and</strong> y was<br />

between 0 <strong>and</strong> 0.005) were grown onto InSb seed crystals by us<strong>in</strong>g a rotary Bridgman<br />

method. The quality of the crystals was assessed by us<strong>in</strong>g optical microscopic, X-ray<br />

topographic, 4-crystal X-ray diffractometric, electron-probe micro-analytical, energydispersive<br />

spectroscopic, <strong>and</strong> secondary-ion mass spectroscopic techniques. Due to<br />

segregation, the compositional ratio of Ga decreased, as the crystals grew. Dur<strong>in</strong>g the<br />

growth of InGaSbBi, Ga diffused <strong>in</strong>to the InSb seed. For comparison, InSb 1-y Bi y (where y<br />

was between 0 <strong>and</strong> 0.05) <strong>and</strong> In 1-x Ga x Sb (where x was between 0 <strong>and</strong> 0.16) were grown<br />

on InSb. It was found that Bi did not diffuse <strong>in</strong>to InSb without Ga, whereas Ga diffused<br />

without Bi. The <strong>in</strong>corporation of Ga produced excess In <strong>and</strong> led to the formation of InBi<br />

doma<strong>in</strong>s.<br />

Y.Hayakawa, M.Ando, T.Matsuyama, E.Hamakawa, T.Koyama, S.Adachi, K.Takahashi,<br />

V.G.Lifshits, M.Kumagawa: Journal of Applied Physics, 1994, 76[2], 858-64<br />

[446-117/118-189]<br />

InSb: Ga <strong>Diffusion</strong><br />

Permeation of Ga was studied by plac<strong>in</strong>g an In-Ga-Sb solution <strong>in</strong> contact with InSb<br />

substrates under conditions of zero crystal growth. It was found that the permeation<br />

distances were equal to 770, 1270 <strong>and</strong> 2750µ at 380, 430 <strong>and</strong> 480C, respectively. It was<br />

deduced that the apparent diffusivity of Ga ranged from 3 x 10 -7 to 5 x 10 -6 cm 2 /s. Rapid<br />

permeation occurred when the substrate came <strong>in</strong>to contact with the solution.<br />

M.Kumagawa, H.Ohtsu, E.Hamakawa, T.Koyama, M.Masaki, K.Takahashi, V.G.Lifshits,<br />

Y.Hayakawa: Materials Science <strong>and</strong> Eng<strong>in</strong>eer<strong>in</strong>g B, 1997, 44[1-3], 301-3<br />

InSb: Ga <strong>Diffusion</strong><br />

A 10mm-thick InGaSb crystal was grown onto an InSb seed by us<strong>in</strong>g the rotary Bridgman<br />

method. It was found that Ga diffused rapidly <strong>in</strong>to the seed <strong>and</strong> displaced some of the In.<br />

The apparent Ga diffusion coefficient was between 10 -8 <strong>and</strong> 10 -7 cm 2 /s. These values were<br />

much higher than the self-diffusion coefficients of In or Sb. Rapid diffusion occurred<br />

only when the substrate was <strong>in</strong> contact with the solution. The diffusion distance of Ga<br />

<strong>in</strong>creased upon <strong>in</strong>creas<strong>in</strong>g the hold<strong>in</strong>g temperature or time.<br />

Y.Hayakawa, E.Hamakawa, T.Koyama, M.Kumagawa: Journal of Crystal Growth, 1996,<br />

163, 220-5<br />

[446-136/137-125]<br />

In<br />

344 InSb: In <strong>Diffusion</strong><br />

Self-diffusion <strong>in</strong> Bridgman-type s<strong>in</strong>gle crystals was studied, at temperatures rang<strong>in</strong>g from<br />

400 to 500C, by us<strong>in</strong>g 114m In radiotracers. An anodic oxidation technique was used for<br />

485


In InSb Li<br />

serial section<strong>in</strong>g, <strong>and</strong> the penetration profiles were fitted to an erf solution of the diffusion<br />

equations. It was found that the self-diffusion of In (table 44) could be described by:<br />

D(cm 2 /s) = 6.0 x 10 -7 [-1.45(eV)/kT]<br />

The migration enthalpy of In atoms was estimated to be equal to 0.66eV, <strong>and</strong> the<br />

correspond<strong>in</strong>g formation enthalpy for an In vacancy was 0.79eV.<br />

A.Rastogi, K.V.Reddy: Journal of Applied Physics, 1994, 75[10], 4984-9<br />

[446-117/118-190]<br />

Table 44<br />

Diffusivity of 114m In <strong>in</strong> InSb<br />

Temperature (C) Diffusivity (cm 2 /s)<br />

414 1.5 x 10 -17<br />

435 3.6 x 10 -17<br />

449 4.8 x 10 -17<br />

457 5.6 x 10 -17<br />

477 8.9 x 10 -17<br />

500 2.4 x 10 -16<br />

345 InSb: In Gra<strong>in</strong> Boundary <strong>Diffusion</strong><br />

The self-diffusion of In was studied (table 45) <strong>in</strong> polycrystall<strong>in</strong>e films by us<strong>in</strong>g neutron<br />

activation tracer scann<strong>in</strong>g methods. The gra<strong>in</strong> boundary diffusion parameters were<br />

evaluated at temperatures rang<strong>in</strong>g from 200 to 400C. The data could be described by:<br />

D (cm 2 /s) = 1.17 x 10 -6 exp[-0.84(eV)/kT]<br />

The In diffused via gra<strong>in</strong> boundaries with<strong>in</strong> the temperature range which was studied. The<br />

gra<strong>in</strong> boundary energy <strong>and</strong> its temperature dependence was also deduced.<br />

A.Rastogi, K.V.Reddy: Semiconductor Science <strong>and</strong> Technology, 1994, 9[11], 2067-72<br />

[446-119/120-219]<br />

Li<br />

InSb: Li <strong>Diffusion</strong><br />

The lithiation of n-type monocrystall<strong>in</strong>e (111)-oriented specimens, via direct reaction<br />

with n-butyllithium <strong>in</strong> hexane solution, was carried out at room temperature. This process<br />

was monitored by us<strong>in</strong>g X-ray diffraction <strong>and</strong> scann<strong>in</strong>g electron microscope techniques.<br />

Resistivity <strong>and</strong> Hall coefficient data permitted the change <strong>in</strong> Li concentration to be<br />

evaluated. It was deduced that the Li diffusivity at 298K was 1.09 x 10 -8 cm 2 /s.<br />

G.Herren, N.E.Walsöe de Reca: Solid State Ionics, 1991, 47[1-2], 57-61<br />

[446-84/85-057]<br />

486


Li InSb Sb<br />

Table 45<br />

Gra<strong>in</strong> Boundary Diffusivity of In <strong>in</strong> InSb<br />

Temperature (C) D (cm 2 /s)<br />

218 2.67 x 10 -15<br />

248 7.89 x 10 -15<br />

282 2.42 x 10 -14<br />

302 4.91 x 10 -14<br />

353 1.78 x 10 -13<br />

390 4.53 x 10 -13<br />

413 7.75 x 10 -13<br />

Sb<br />

346 InSb: Sb <strong>Diffusion</strong><br />

Self-diffusion <strong>in</strong> Bridgman-type s<strong>in</strong>gle crystals was studied, at temperatures rang<strong>in</strong>g from<br />

400 to 500C, by us<strong>in</strong>g 125 Sb radiotracers. An anodic oxidation technique was used for<br />

serial section<strong>in</strong>g, <strong>and</strong> the penetration profiles were fitted to an erf solution of the diffusion<br />

equations. It was found that the self-diffusion of Sb (table 46) could be described by:<br />

D(cm 2 /s) = 5.35 x 10 -4 [-1.91(eV)/kT]<br />

The migration enthalpy of Sb atoms was estimated to be equal to 0.70eV, <strong>and</strong> the<br />

correspond<strong>in</strong>g formation enthalpy for an Sb vacancy was 1.21eV.<br />

A.Rastogi, K.V.Reddy: Journal of Applied Physics, 1994, 75[10], 4984-9<br />

[446-117/118-190]<br />

Table 46<br />

Diffusivity of 125 Sb <strong>in</strong> InSb<br />

Temperature (C) Diffusivity (cm 2 /s)<br />

408 3.9 x 10 -18<br />

418 1.2 x 10 -17<br />

436 1.5 x 10 -17<br />

455 3.2 x 10 -17<br />

467 5.3 x 10 -17<br />

479 8.7 x 10 -17<br />

347 InSb: Sb Gra<strong>in</strong> Boundary <strong>Diffusion</strong><br />

The self-diffusion of Sb was studied (table 47) <strong>in</strong> polycrystall<strong>in</strong>e films by us<strong>in</strong>g neutron<br />

activation tracer scann<strong>in</strong>g methods. The gra<strong>in</strong> boundary diffusion parameters were<br />

evaluated at temperatures rang<strong>in</strong>g from 200 to 400C. The data could be described by:<br />

D (cm 2 /s) = 1.32 x 10 -4 exp[-1.11(eV)/kT]<br />

487


Sb InSb Te<br />

The Sb diffused via gra<strong>in</strong> boundaries with<strong>in</strong> the temperature range which was studied.<br />

The gra<strong>in</strong> boundary energy <strong>and</strong> its temperature dependence was also deduced.<br />

A.Rastogi, K.V.Reddy: Semiconductor Science <strong>and</strong> Technology, 1994, 9[11], 2067-72<br />

[446-119/120-219]<br />

Table 47<br />

Gra<strong>in</strong> Boundary Diffusivity of Sb <strong>in</strong> InSb<br />

Temperature (C) D (cm 2 /s)<br />

218 5.15 x 10 -16<br />

248 2.41 x 10 -15<br />

256 3.44 x 10 -15<br />

282 1.09 x 10 -14<br />

310 3.34 x 10 -14<br />

350 1.34 x 10 -13<br />

Te<br />

InSb: Te <strong>Diffusion</strong><br />

A PbTe source was used to grow n-type InSb by means of molecular beam epitaxy. Auger<br />

electron spectroscopic data showed that no surface segregation of Te occurred at dop<strong>in</strong>g<br />

levels of up to about 10 19 /cm 3 . Secondary ion mass spectrometry did not reveal the<br />

presence of Pb <strong>in</strong> the films, even at growth temperatures which were as low as 280C. This<br />

suggested that the Pb rapidly evaporated from the surface dur<strong>in</strong>g growth. The secondary<br />

ion mass spectrometric depth profiles for Te revealed signs of solid-state diffusion at<br />

360C; with a diffusion coefficient of about 10 -13 cm 2 /s.<br />

D.L.Part<strong>in</strong>, J.Heremans, C.M.Thrush: Journal of Applied Physics, 1992, 71[5], 2328-32<br />

[446-86/87-046]<br />

488


Author Index<br />

Abe, Y. [440]<br />

Abernathy, C.R. [269, 294, 444, 458]<br />

Abrosimova, V.N. [292, 323]<br />

Acher, O. [439]<br />

Adachi, A. [422]<br />

Adachi, S. [484, 485]<br />

Aers, G.C. [421]<br />

Ahlborn, K. [464, 466]<br />

Ahlgren, T. [326]<br />

Alex<strong>and</strong>re, F. [398]<br />

Algora, C. [250, 355]<br />

Allen, E.L. [234, 239, 248, 307,<br />

327, 330, 342, 350]<br />

Als<strong>in</strong>a, F. [431]<br />

Alvarez, A.L. [211]<br />

Alwan, J.J. [242, 248]<br />

Amann, M.C. [373, 419, 438]<br />

Amaratunga, G. [318]<br />

Ambreé, P. [406, 407, 410, 419]<br />

Anderson, G.B. [248, 396]<br />

Anderson, S.G. [352]<br />

Andersson, M. [277, 304]<br />

Andersson, Y. [277, 304]<br />

Ando, M. [484, 485]<br />

Ansaldo, E.J. [372]<br />

Anthony, J.M. [308]<br />

Anthony, P.J. [238]<br />

Appelbaum, A. [474]<br />

Araujo, D. [356, 368]<br />

Araujo, G.L. [250, 355]<br />

Arent, D.J. [382]<br />

Armiento, C.A. [392]<br />

Arora, B.M. [428]<br />

Asahi, H. [420, 450]<br />

489<br />

Asashi, H. [433]<br />

Ashw<strong>in</strong>, M.J. [344]<br />

Baba, T. [274]<br />

Baba-Ali, N. [227, 229, 245, 328,<br />

390, 394]<br />

Babievskaya, I.Z. [460, 472]<br />

Baeumler, M. [426]<br />

Baiocchi, F. [335]<br />

Baird, R.J. [412]<br />

Baker, J.E. [230, 242, 244, 247, 248,<br />

252, 253, 257, 258, 263, 288,<br />

338, 339, 437, 448]<br />

Balk, P. [473]<br />

Bamba, Y. [236]<br />

Ban, Y. [263]<br />

B<strong>and</strong>urko, V.V. [272]<br />

Banerjee, S. [428]<br />

Bao, X.J. [449]<br />

Baranowski, J.M. [399]<br />

Baratte, H. [281, 373]<br />

Barcz, A.J. [399]<br />

Baribeau, J.M. [325]<br />

Barker, R.C. [261]<br />

Baroni, S. [390]<br />

Bársony, I. [343]<br />

Bartels, A. [468]<br />

Baumann, F.H. [267]<br />

Baumeister, E. [247]<br />

Baumeister, H. [465]<br />

Baur, B. [482]<br />

Beall, R.B. [236, 246, 248, 280, 282,<br />

327, 333, 335, 349]<br />

Beaumont, B. [416]<br />

Beern<strong>in</strong>k, K.J. [248, 262, 264]


Author Index<br />

Beggy, J.C. [261]<br />

Benchimol, J.L. [405]<br />

Bénière, F. [342]<br />

Ben-Tzur, M. [359, 473]<br />

Benz, K.W. [467]<br />

Ber, B.J. [440]<br />

Bergner, D. [483]<br />

Bernholc, J. [343, 347, 366, 367,<br />

368, 389]<br />

Besson, J.M. [313]<br />

Bhat, H.L. [450, 453]<br />

Bhat, K.N. [363]<br />

Bhat, R. [406, 415, 420, 480]<br />

Bhattacharya, P. [274, 405]<br />

Bhattacharya, P.K. [412]<br />

Bimberg, D. [297, 332, 352, 446, 447,<br />

453, 454, 462, 466, 469]<br />

Bisberg, J.E. [251, 358, 400]<br />

Biswas, D. [274]<br />

Bithell, E.G. [392]<br />

Blaauw, C. [468, 472, 474]<br />

Blanchard, B. [365, 366, 368, 370]<br />

Bockstedte, M. [301]<br />

Boer<strong>in</strong>ger, D.W. [305]<br />

Boesker, G. [363]<br />

Bonapasta, A.A. [133]<br />

Bonkowski, J.E. [448]<br />

Booker, G.R. [423, 427]<br />

Borenste<strong>in</strong>, J.T. [311]<br />

Borod<strong>in</strong>a, O.M. [313, 410, 464]<br />

Bösker, G. [298, 363, 364]<br />

Boudewijn, P.R. [437, 476]<br />

Boudreau, R. [245, 249, 339, 361]<br />

Bour, B.P. [262, 264]<br />

Bour, D. [314]<br />

Bour, D.P. [399]<br />

Bowen, T. [245, 249, 339, 361]<br />

Bradley, I.V. [397, 403, 414, 424, 429]<br />

Br<strong>and</strong>t, O. [465]<br />

Braunste<strong>in</strong>, G. [331]<br />

Bravman, J.C. [237, 285, 341, 343]<br />

Breitenste<strong>in</strong>, O. [340, 372, 398, 401]<br />

Brewer, J.H. [372]<br />

Briggs, A.T.R. [403, 414]<br />

Brillouet, F. [391, 398]<br />

Brion, H.G. [464, 466]<br />

Briones, F. [447]<br />

Bronner, W. [237]<br />

490<br />

Bruk, A.S. [313]<br />

Brum, J.A. [408, 412]<br />

Buchanan, M. [421]<br />

Budko, T.O. [476]<br />

Bühlmann, H.J. [456]<br />

Burke, P.T. [395]<br />

Bürkner, S. [396, 424, 426]<br />

Burnham, R.D. [234, 243, 244, 245,<br />

247, 249, 252, 256, 338]<br />

Burri, G. [368]<br />

Burton, R.S. [243, 245]<br />

Busyg<strong>in</strong>a, L.A. [440]<br />

Butherus, A.D. [335]<br />

Cahen, D. [477]<br />

Cai, D.Q. [395]<br />

Calawa, A.R. [388]<br />

Calle, F. [211, 465]<br />

Calleja, J.M. [432]<br />

Camassel, J. [431]<br />

Caneau, C. [266, 267, 415, 469, 480]<br />

Cannelli, G. [294]<br />

Cantelli, R. [294]<br />

Capizzi, M. [133, 294]<br />

Cardona, M. [302, 447]<br />

Cardone, F. [281, 373]<br />

Carl<strong>in</strong>, J.F. [456]<br />

Carr, N. [434]<br />

Caruso, R. [277]<br />

Castagné, J. [236, 246, 248, 280, 282,<br />

327, 333, 335, 349]<br />

Cerny, R. [483]<br />

Chan, L.Y. [359, 366, 473, 480]<br />

Chan, W.K. [406]<br />

Ch<strong>and</strong>, N. [238]<br />

Chang, C.C. [406]<br />

Chang, C.S. [37]<br />

Chang, J.C.P. [233, 346]<br />

Chang, K.H. [274, 287, 345]<br />

Chang, K.J. [289]<br />

Chang, S.K. [470]<br />

Chang, T.C. [287]<br />

Chang, T.Y. [267]<br />

Chang, Y.A. [400]<br />

Chapla<strong>in</strong>, R. [342, 459]<br />

Charbonneau, S. [421]<br />

Chase, M.P. [302, 362]<br />

Chatterjee, B. [463]<br />

Chatterjee, S. [363]


Author Index<br />

Chavignon, J. [391]<br />

Chayahara, A. [319, 324, 325]<br />

Chen, A.C. [448]<br />

Chen, B. [343, 347, 389]<br />

Chen, C.H. [175]<br />

Chen, C.Y. [364]<br />

Chen, E.I. [253]<br />

Chen, H.R. [394]<br />

Chen, P. [398]<br />

Chen, S. [331]<br />

Chen, W.C. [37]<br />

Chen, Y. [364, 391]<br />

Chen, Y.C. [405]<br />

Cheng, C.L. [462]<br />

Cheng, K.Y. [448]<br />

Cheng, W.H. [474]<br />

Cheong, B.H. [289]<br />

Chester, M.J. [324]<br />

Chevallier, J. [238, 241, 294, 295,<br />

310, 312, 313, 314, 454, 460]<br />

Chi, P.H. [266, 267, 286, 309, 315, 341,<br />

364, 415, 459, 463, 467, 469]<br />

Chia, V.F.K. [234, 239, 248]<br />

Ch<strong>in</strong>, A.K. [251, 355, 358, 400]<br />

Chiu, T.H. [336, 337, 338, 357, 386]<br />

Cho, K.I. [377, 422]<br />

Choa, F.S. [473]<br />

Choe, B.D. [83, 428]<br />

Choi, J.S. [470]<br />

Chow, D.H. [443]<br />

Chow, K. [372]<br />

Chow, K.H. [371]<br />

Christen, J. [332]<br />

Chu, S.N.G. [478, 479]<br />

Chui, T.H. [331]<br />

Cibert, J. [392]<br />

Clarke, R. [274]<br />

Clegg, J.B. [236, 246, 248, 280, 282,<br />

327, 333, 335, 349]<br />

Clemencon, J.J. [355]<br />

Cockayne, D.J.H. [395]<br />

Cockerill, T.M. [242, 248]<br />

Cohen, D.D. [350]<br />

Cohen, P.I. [231, 374, 375]<br />

Cohen, R.M. [17, 298, 315, 364,<br />

388, 393, 395, 483]<br />

Colas, E. [228, 254]<br />

Coldren, L. [250, 359]<br />

491<br />

Coldren, L.A. [244, 250, 455]<br />

Coleman, J.J. [242, 248]<br />

Conibeer, G.J. [451, 452]<br />

Corbett, J.W. [311]<br />

Cordero, F. [294]<br />

Corz<strong>in</strong>e, S. [244, 250]<br />

Cox, H.M. [406]<br />

Cox, S.F.J. [372]<br />

Craford, M.G. [230, 263, 437]<br />

Craighead, H.G. [275]<br />

Cremer, C. [482]<br />

Crook, A. [242, 248]<br />

Cunn<strong>in</strong>gham, B.T. [256, 262, 288]<br />

Cunn<strong>in</strong>gham, J.E. [331, 336, 337, 338]<br />

Cunn<strong>in</strong>gham, T.J. [261]<br />

Cutlerywala, H. [364]<br />

Dabiran, A.M. [231, 374, 375]<br />

Dabkowski, F.P. [251, 358, 400]<br />

Dabrowski, J. [304, 342, 344]<br />

Dadgar, A. [466]<br />

Dagata, J.A. [324]<br />

Dallesasse, J.M. [243, 249, 257]<br />

D<strong>and</strong>rea, R.G. [389]<br />

Danilewsky, A.N. [467]<br />

Dautremont-Smith, W.C. [294, 461, 464]<br />

Davies, G.J. [405]<br />

Däweritz, L. [384]<br />

De Cremoux, B. [361]<br />

De Souza, J.P. [281]<br />

De, T.J. [359]<br />

Deal, M.D. [234, 237, 239, 248, 282,<br />

283, 285, 286, 302, 303, 307, 318,<br />

327, 329, 330, 334, 341, 342, 343,<br />

350, 362]<br />

Deicher, M. [290]<br />

Deppe, D.G. [230, 252, 253, 256, 257,<br />

258, 263, 338, 339, 360, 437]<br />

Descouts, B. [391]<br />

De Temple, T.A. [242, 248]<br />

Deus, P. [483]<br />

Dhanasekaran, R. [278]<br />

Dieckmann, R. [270]<br />

Dieguez, E. [450]<br />

Dietrich, H.B. [286, 309, 341, 415,<br />

459, 463, 467, 469]<br />

Dildey, F. [419]<br />

Dion, M. [421]<br />

Dixon, R.H. [470]


Author Index<br />

Dodds, S.A. [372]<br />

Dos Passos, W. [274]<br />

Dreybrodt, J. [432]<br />

Dreyer, K. [386]<br />

Dubon-Chevallier, C. [405]<br />

Duke, C.B. [389]<br />

Dunsiger, S.R. [371]<br />

Dunstan, D.J. [429]<br />

Dutta, P.S. [450, 453]<br />

Duvarney, R.C. [372]<br />

Dzhumakulov, D.T. [321]<br />

Eastman, L.F. [283]<br />

Ebb<strong>in</strong>ghaus, G. [419]<br />

Eberl, K. [302]<br />

Ebert, P. [448]<br />

Eccleston, R. [392]<br />

Eger, D. [468, 474, 477]<br />

Egger, U. [340, 372, 398, 401]<br />

Egilsson, T. [317]<br />

Eizenberg, M. [290]<br />

Elman, B. [392]<br />

El-Ze<strong>in</strong>, N. [243, 249, 319]<br />

Emanuel, M.A. [248]<br />

Emelyanenko, J.S. [476]<br />

Emeny, M.T. [429]<br />

Emmerstorfer, B. [472]<br />

Emura, S. [433, 450]<br />

Enders, P. [480]<br />

Endicott, F.J. [347, 396, 399]<br />

Endoh, A. [236]<br />

Enquist, P. [359]<br />

Epler, J.E. [234, 256, 347]<br />

Erickson, J.W. [302, 447]<br />

Erofeeva, E.A. [321]<br />

Estle, T.L. [314, 371, 372]<br />

Fahy, M.R. [344]<br />

Fan, J.C. [287, 345, 427]<br />

Fareed, R.S.Q. [278]<br />

Farley, C.W. [308]<br />

Fathimulla, A. [286, 309, 341, 459,<br />

463, 467]<br />

Favennec, P.N. [459]<br />

Fedotov, A.B. [321]<br />

Fenn<strong>in</strong>ger, G. [483]<br />

Fewster, P.F. [344]<br />

Fischer, A. [302]<br />

Flat, A. [332]<br />

Fleischman, A.J. [373]<br />

492<br />

Fletcher, R.M. [230, 263, 437]<br />

Fons, P. [457]<br />

Fonstad, C.G. [406, 413, 414]<br />

Fontijn, G.M. [435, 436, 437, 438,<br />

439, 462, 463, 470, 475]<br />

Forchel, A. [432, 433]<br />

Fornari, R. [103]<br />

Forster, T. [382]<br />

Franke, D. [461]<br />

Franz, G. [373, 438]<br />

Franzheld, R. [398, 401]<br />

Freundlich, A. [348]<br />

Frova, A. [294]<br />

Fujii, E. [346]<br />

Fujii, M. [377]<br />

Fujii, N. [369]<br />

Fujii, S. [371]<br />

Fujii, T. [236]<br />

Fujii, Y. [324]<br />

Fujita, K. [340, 377, 426]<br />

Fukunaga, T. [252, 332]<br />

Furtado, M.T. [396, 409, 411, 417,<br />

418, 426]<br />

Furuya, A. [258]<br />

Gailhanou, M. [368]<br />

Gal, M. [395]<br />

Ganière, J.D. [245, 328, 356, 365,<br />

366, 368, 370]<br />

Gao, Y. [391, 398]<br />

García, S. [465]<br />

Gauneau, M. [342, 459]<br />

Gautam, D.K. [290, 365]<br />

Gautier, S. [404]<br />

Gav<strong>and</strong>, M. [294]<br />

Gavrilovic, P. [244, 247, 258, 338]<br />

Gaymann, A. [237]<br />

Genut, M. [290]<br />

George, T. [348]<br />

Geva, M. [238, 420, 478, 479]<br />

Ghaffari, M. [466]<br />

Giannelis, E.P. [351]<br />

Giannozzi, P. [311]<br />

Gibala, R. [274]<br />

Gibart, P. [241, 310]<br />

Gibbons, J.F. [360]<br />

Gilderman, V.K. [269]<br />

Gill<strong>in</strong>, W.P. [397, 399, 403, 414,<br />

424, 429, 430, 432]


Author Index<br />

Giov<strong>in</strong>e, E. [294]<br />

Gisdakis, S. [319]<br />

Gislason, H.P. [293, 317]<br />

Glade, M. [472, 473]<br />

Glew, R.W. [431]<br />

Goldberg, R.D. [421]<br />

Gonda, S. [433, 450]<br />

González-Díaz, G. [465]<br />

Gorelenok, A.T. [440]<br />

Gösele, U. [237, 241, 252, 261, 277,<br />

280, 283, 291, 297, 298, 300, 301,<br />

305, 333, 336, 344, 345, 346, 354,<br />

360, 367, 372, 398, 401, 461, 481]<br />

Goto, K. [417]<br />

Goto, S. [238, 281, 379, 380, 387]<br />

Govorkov, A.V. [313]<br />

Gr<strong>and</strong>jean, N. [387]<br />

Grant, J. [390]<br />

Grattepa<strong>in</strong>, C. [348, 454]<br />

Grattepa<strong>in</strong>, C.M. [238, 294]<br />

Grenet, J.C. [348]<br />

Griehl, S. [293]<br />

Griff<strong>in</strong>, P.B. [318]<br />

Grigorev, N.N. [361]<br />

Grodkiewicz, W.H. [335]<br />

Grote, N. [461]<br />

Gruen, N. [237]<br />

Gruenbaum, P.E. [452]<br />

Grundmann, M. [332]<br />

Gruska, B. [406, 407, 419]<br />

Grützmacher, D. [433, 473]<br />

Gudmundsson, J.T. [317]<br />

Guersen, R. [1]<br />

Guido, L.J. [234, 244, 247, 252, 256,<br />

259, 261, 262, 288, 338]<br />

Guillaume, J.C. [416]<br />

Gulwadi, S.M. [415, 469]<br />

Guoba, L.B. [278, 300]<br />

Gushch<strong>in</strong>skaya, E.V. [476]<br />

Gwilliam, R. [392, 397, 424]<br />

Gyuro, I. [432]<br />

Ha, J.S. [386]<br />

Ha, N.T. [357, 473, 478, 479]<br />

Haddara, Y.M. [285]<br />

Hafich, M.J. [265]<br />

Haga, D. [357]<br />

Haga, T. [456]<br />

Hailemariam, E. [266, 267, 405, 414]<br />

493<br />

Hallali, P.E. [281]<br />

Haller, E.E. [297, 302, 359, 366,<br />

372, 447, 473, 480]<br />

Hamakawa, E. [484, 485]<br />

Hamakawa, Y. [292]<br />

Hamilton, W.J. [443]<br />

Hamoudi, A. [441]<br />

Han, H.T. [245, 249, 339, 361]<br />

Hangleiter, A. [407, 410, 419]<br />

Hanson, C.M. [359, 473]<br />

Hara, N. [242, 306, 345, 390]<br />

Hara, T. [289]<br />

Harada, Y. [346]<br />

Harbison, J.P. [227]<br />

Harde, P. [461]<br />

Hard<strong>in</strong>gham, C.M. [451, 452]<br />

Harper, J. [386]<br />

Harris, G.L. [283]<br />

Harris, J.J. [236, 246, 248, 280, 282,<br />

327, 333, 335, 349]<br />

Harris, J.S. [407, 411]<br />

Harrison, I. [227, 229, 230, 245,<br />

328, 390, 394]<br />

Hart, L. [344]<br />

Harun-ur Rashid, A.B.M. [351]<br />

Hashimoto, A. [252, 332]<br />

Hashimoto, H. [258, 284, 338]<br />

Haspeklo, H. [415, 475]<br />

Hata, M. [376, 387]<br />

Haukson, I.S. [317]<br />

Hawk<strong>in</strong>s, R.L. [279]<br />

Hayakawa, Y. [484, 485]<br />

Haynes, T.E. [342]<br />

Heime, K. [433]<br />

He<strong>in</strong>e, K. [472]<br />

He<strong>in</strong>ecke, H. [482]<br />

He<strong>in</strong>z, C. [451, 452]<br />

Heitz, R. [466]<br />

Helms, C.R. [340]<br />

Hen<strong>in</strong>i, M. [227, 229, 230, 245, 328]<br />

Heremans, J. [488]<br />

Hergeth, J. [473]<br />

Herms, M. [293]<br />

Hern<strong>and</strong>es, C.S. [351]<br />

Herren, G. [486]<br />

Herr<strong>in</strong>g, C. [314, 315]<br />

Herz<strong>in</strong>ger, C.M. [242, 248]<br />

Herzog, L. [340]


Author Index<br />

Hesse, R. [354]<br />

Hettwer, H.G. [340, 353, 363, 364, 470]<br />

Heuken, M. [433]<br />

Heurtel, A. [312]<br />

Higashisaka, A. [289]<br />

Higuchi, M. [471]<br />

Hill, D.M. [352]<br />

Hillard, R.J. [449]<br />

Hillmer, H. [423]<br />

Hira, K. [319]<br />

Hirai, K. [323, 324, 325]<br />

Hirai, M. [340]<br />

Hitti, B. [371]<br />

Hiyamizu, S. [422]<br />

Ho, H.P. [227, 229, 230, 394]<br />

Hobbs, L. [468, 472]<br />

Hobson, W.S. [266, 267, 405, 414]<br />

Hockly, M. [423]<br />

Hoepfner, A. [319]<br />

Höfler, G.E. [243, 261, 319]<br />

Hofsäss, V. [423]<br />

Holl<strong>and</strong>, O.W. [415, 469]<br />

Holmgren, D.J. [243, 245]<br />

Holonyak, N. [230, 234, 243, 244, 247,<br />

249, 252, 253, 256, 257, 258, 261,<br />

262, 263, 288, 319, 338, 339,<br />

360, 437]<br />

Homewood, K.P. [397, 399, 403, 414,<br />

424, 429, 432]<br />

Hong, C.Y. [389]<br />

Hong, J.M. [275]<br />

Hong, S.K. [440]<br />

Hong, W.P. [415, 469, 480]<br />

Hopk<strong>in</strong>s, L.C. [279]<br />

Höpner, A. [471]<br />

Horikoshi, Y. [254, 315, 375, 377, 384]<br />

Hor<strong>in</strong>o, Y. [319, 325]<br />

Houde-Walter, S.N. [256, 258, 259, 260]<br />

Houdré, R. [456]<br />

Hovel, H.J. [373]<br />

Howard, A.J. [265]<br />

Howard, L.K. [397, 424, 429]<br />

Hsieh, K.C. [234, 243, 247, 252, 256,<br />

257, 258, 261, 319, 320, 360, 448]<br />

Hsieh, K.Y. [260, 409, 412]<br />

Hsu, C.C. [443]<br />

Hsu, J.T. [292]<br />

Hsu, L. [302, 447]<br />

494<br />

Hu, E. [250, 359]<br />

Hu, E.L. [175, 244, 250]<br />

Hu, J.C. [285, 286, 303]<br />

Hu, P. [393]<br />

Huang, J.H. [267]<br />

Huang, K.F. [484]<br />

Huang, L.J. [325]<br />

Hughes, O.H. [230]<br />

Hult, M. [277, 304]<br />

Humer-Hager, T. [288, 369]<br />

Hutchby, J.A. [359]<br />

Hwang, D.M. [415, 420]<br />

Hwang, Y. [355]<br />

Hwang, Y.L. [409, 412]<br />

Hyeon, J.Y. [466]<br />

Ichimura, K. [250]<br />

Iguchi, H. [233]<br />

Iikawa, F. [408, 412]<br />

Ikarashi, N. [274]<br />

Ikeda, H. [295]<br />

Ilegems, M. [456]<br />

Imamura, Y. [420]<br />

Inai, M. [286, 342]<br />

Inohara, H. [319]<br />

Inoue, Y. [299]<br />

Ishibashi, A. [263]<br />

Ishii, K. [236]<br />

Ishikawa, H. [236]<br />

Ish<strong>in</strong>o, M. [418]<br />

Isu, T. [376, 380, 387]<br />

Ito, H. [242, 398, 401, 411, 481]<br />

Ito, R. [445]<br />

Ito, T. [255, 378]<br />

Itoh, H. [242]<br />

Itoh, M. [435, 441]<br />

Ivanov, S.V. [235, 281]<br />

Jach, T. [324]<br />

Jackman, T.E. [421]<br />

Jackson, T.N. [373]<br />

Jacobs, J.M. [438, 475]<br />

Jaeger, W. [363]<br />

Jafri, Z.H. [430]<br />

Jagadish, C. [395]<br />

Jäger, W. [298, 305, 353, 363, 364, 470]<br />

Jaklevic, J. [366, 480]<br />

Jaklevic, J.M. [359, 473]<br />

Jalil, A. [312]<br />

James, D.J. [326]


Author Index<br />

Jan, W. [331, 338]<br />

Janzén, E. [317]<br />

Jeanjean, P. [347]<br />

Jeong, B.S. [442]<br />

Jeong, W.G. [428]<br />

Jiménez, J. [103]<br />

Johnson, M.B. [284]<br />

Johnson, N.M. [314, 315, 407, 411]<br />

Jones, K.S. [307, 330, 341, 342, 350]<br />

Jordan, A.S. [238]<br />

Joullie, A. [455, 457]<br />

Joyce, B.A. [231]<br />

Juhel, M. [441]<br />

Kadono, R. [371, 372]<br />

Kahen, K.B. [239, 249, 334, 358,<br />

360, 367, 393]<br />

Kak<strong>in</strong>oki, H. [295]<br />

Kalem, S. [454]<br />

Kalish, R. [290]<br />

Kaliski, R.W. [234, 252, 256]<br />

Kaman<strong>in</strong>, A.V. [440, 477]<br />

Kamei, H. [457]<br />

Kamei, K. [426]<br />

Kamejima, T. [307]<br />

Kameneckas, J.V. [278, 300]<br />

Kam<strong>in</strong>ska, M. [460]<br />

Kamiya, T. [431]<br />

Kanber, H. [283, 291, 320]<br />

Kaneko, Y. [242]<br />

Kang, B.K. [439]<br />

Kang, T.W. [389]<br />

Kano, H. [232, 387]<br />

Karen<strong>in</strong>a, L.S. [269]<br />

Karlicek, R.F. [478]<br />

Karttunen, V. [312]<br />

Kasai, K. [306, 345]<br />

Kasu, M. [232, 380, 385, 386]<br />

Katahama, H. [426]<br />

Kataoka, M. [456]<br />

Kataoka, Y. [236]<br />

Katayama, M. [299]<br />

Katayama, Y. [379, 380, 387]<br />

Katoda, T. [351, 390]<br />

Katsumata, H. [113]<br />

Kavanagh, K.L. [316, 322, 346, 348]<br />

Kawabe, M. [280, 329]<br />

Kawaguchi, Y. [420]<br />

Kawai, T. [316]<br />

495<br />

Kawamura, Y. [420]<br />

Kawashima, M. [254, 375, 377, 384]<br />

Kazar<strong>in</strong>a, N.N. [460, 472]<br />

Kazmierski, C. [476]<br />

Kazmierski, K. [361]<br />

Kehr, K.W. [374]<br />

Ke<strong>in</strong>onen, J. [312, 326]<br />

Keller, R. [290]<br />

Keller, R.C. [340]<br />

Kempeners, M.N.C. [476]<br />

Ketata, K. [403, 404, 405]<br />

Ketata, M. [403, 404, 405]<br />

Khald, H. [455, 457]<br />

Khoo, G.S. [382]<br />

Khreis, O.M. [399]<br />

Kidoguchi, I. [263]<br />

Kiefl, R.F. [371, 372]<br />

Kim, I. [428]<br />

Kim, J.G. [446]<br />

Kim, J.H. [439]<br />

Kim, M.S. [333]<br />

Kim, S.G. [450]<br />

Kim, S.K. [389]<br />

Kim, S.T. [442]<br />

Kim, T.S. [308]<br />

Kim, T.W. [389]<br />

Kim, Y. [333, 395]<br />

Kim, Y.H. [389]<br />

Kimura, T. [369, 431]<br />

Kirchner, P.D. [346]<br />

Kish, F.A. [243, 247, 249, 258, 339]<br />

Kishi, M. [351]<br />

Kishimoto, D. [381, 385, 456]<br />

Kisielowski, C. [388]<br />

Kitada, T. [422]<br />

Kle<strong>in</strong>, P.B. [286, 309, 341, 459,<br />

463, 467]<br />

Klem, J.F. [265]<br />

Klöber, J. [293]<br />

Klockenbr<strong>in</strong>k, R. [468]<br />

Knecht, A. [297, 352, 446, 447, 453,<br />

454, 462, 469]<br />

Ko, K.Y. [331]<br />

Kobayashi, J. [284, 338]<br />

Kobayashi, N. [232, 380, 385, 386]<br />

Kobayashi, T. [481]<br />

Koch, M.W. [288]<br />

Koehler, K. [237]


Author Index<br />

Koenraad, P.M. [284, 343]<br />

Kohnke, G.E. [288]<br />

Kohzu, H. [289]<br />

Kolbas, R.M. [260, 355, 409, 412]<br />

Koltsov, G.I. [284, 459]<br />

Komeno, J. [306, 345]<br />

Kondo, E. [376]<br />

Kondo, S. [435, 441]<br />

König, U. [415, 475]<br />

Koo, J.Y. [386]<br />

Kopev, P.S. [235, 281]<br />

Kopf, R.F. [244, 279]<br />

Koponen, I. [312]<br />

Korbutyak, D.N. [293]<br />

Korobkov, N.N. [292, 323]<br />

Koshiba, S. [232, 387]<br />

Koszi, L.A. [461, 464]<br />

Koteles, E.S. [392]<br />

Kothiyal, G.P. [412]<br />

Koumetz, S. [403, 404, 405]<br />

Koyama, T. [484, 485]<br />

Koza, M. [415, 420]<br />

Koza, M.A. [480]<br />

Kozhukhova, E.A. [410, 464]<br />

Kozlovskii, V.V. [292, 323]<br />

Krames, M.R. [253]<br />

Kraus, G.T. [351]<br />

Kräutle, H. [297, 352, 446, 447, 453,<br />

454, 462, 469]<br />

Krauz, P. [391, 398, 402, 410, 413, 441]<br />

Kreitzman, S.R. [372]<br />

Kreller, D. [472]<br />

Krieghoff, T. [471]<br />

Kr<strong>in</strong>ghøj, P. [469]<br />

Kristjánsson, S. [317]<br />

Kubena, R. [393]<br />

Kucera, J. [321]<br />

Kudyk<strong>in</strong>a, T.A. [361]<br />

Kuech, T.F. [235, 240]<br />

Kühn, G. [354, 355, 471]<br />

Kuhn, J. [423]<br />

Kuisl, M. [415, 475]<br />

Kumagai, M. [323, 324]<br />

Kumagawa, M. [484, 485]<br />

Kumagaya, M. [319, 325]<br />

Kumar, V. [453]<br />

Kuo, C.P. [230, 263, 437]<br />

Kuo, J.M. [244, 279]<br />

496<br />

Kuo, T.Y. [331]<br />

Kupka, R.K. [391]<br />

Kurishima, K. [481]<br />

Kuriyama, Y. [276, 301]<br />

Kuroda, S. [242]<br />

Kusano, C. [238, 281]<br />

Kuttler, M. [466]<br />

Kuwamoto, H. [474]<br />

Kuzuhara, M. [307]<br />

Kwiatkowski, S. [399]<br />

Kwon, H.K. [83]<br />

Kwon, O. [439, 440]<br />

Kwon, Y.S. [251]<br />

Kwon1, S.D. [83]<br />

Ky, N.H. [245, 328, 356, 365,<br />

366, 368, 370]<br />

Lagally, M.G. [448]<br />

Lahiri, I. [1, 233]<br />

Lakdawala, V.K. [292]<br />

L<strong>and</strong>gren, G. [253, 317, 353, 466]<br />

L<strong>and</strong>heer, D. [325]<br />

Lanzillotto, A.M. [246, 334]<br />

Lark<strong>in</strong>s, E.C. [396, 424, 426]<br />

Laruelle, F. [393]<br />

Lashkevich, L.N. [293]<br />

Lau, W.M. [325]<br />

Launay, F. [361, 439]<br />

Launay, P. [403, 404, 405]<br />

Launois, H. [391]<br />

Lauterbach, C. [469]<br />

Lawrence, D.J. [239, 251, 357, 358, 400]<br />

Ledentsov, N.N. [235, 281]<br />

Lee, C. [333, 342]<br />

Lee, C.C. [237, 285, 341, 342, 343]<br />

Lee, C.P. [287, 345, 394, 427]<br />

Lee, E. [386]<br />

Lee, E.H. [440]<br />

Lee, J.C. [235, 240]<br />

Lee, J.H. [260, 409, 412]<br />

Lee, J.I. [470]<br />

Lee, J.K. [440]<br />

Lee, J.L. [280, 329]<br />

Lee, J.Y. [440]<br />

Lee, K.H. [329, 437, 438, 440]<br />

Lee, S. [386]<br />

Lee, S.C. [37, 287]<br />

Lee, S.H. [394, 427]<br />

Lee, S.T. [251, 261, 305, 331, 357, 400]


Author Index<br />

Lee, Y.T. [439, 440]<br />

Lefebvre, F. [404]<br />

Leitch, A.W.R. [310, 311]<br />

Lengel, G. [386]<br />

Lennard, W.N. [325, 477]<br />

Leon, R.P. [460]<br />

Leonard, D. [455]<br />

Lester, S.D. [308]<br />

Lesunova, R.P. [269]<br />

Leycuras, A. [348]<br />

L'Haridon, H. [459]<br />

Li, E.H. [433]<br />

Li, W. [482]<br />

Li, W.M. [315]<br />

Li, Y.J. [275, 307]<br />

Liau, Z.L. [465]<br />

Lichti, R.L. [314, 371, 372]<br />

Lifshits, V.G. [484, 485]<br />

Likonen, J. [326]<br />

Liliental-Weber, Z. [388, 460]<br />

Lim, H.J. [470]<br />

Lim, H. [83]<br />

L<strong>in</strong>, H.H. [287]<br />

L<strong>in</strong>, Z. [352]<br />

L<strong>in</strong>deberg, I. [277, 304]<br />

L<strong>in</strong>narsson, M. [317]<br />

L<strong>in</strong>narsson, M.K. [253, 317, 353, 357]<br />

Liou, D.C. [287, 345]<br />

Liu, B.D. [287]<br />

Liu, D.G. [287, 345]<br />

Liu, H. [446]<br />

Lo, V.C. [339]<br />

Lo, Y.C. [260]<br />

Logan, R.A. [478, 479]<br />

Logan, R.C. [294]<br />

Lomasov, V.N. [292, 323]<br />

Long, J. [462]<br />

Long, N.J. [423, 427]<br />

Lopata, J. [294, 449, 461, 464]<br />

López, M. [379]<br />

Lord, S.M. [407, 411]<br />

Lösch, R. [423]<br />

Loural, M.S.S. [396, 418, 426]<br />

Ludowise, M.J. [230, 263, 437]<br />

Ludwig, M.H. [446]<br />

Luftman, H.S. [266, 267, 279, 357,<br />

405, 414]<br />

Luken, K.M. [277]<br />

497<br />

Lunardi, L.M. [244]<br />

Luo, Y.S. [383]<br />

Lusson, A. [454]<br />

Lyahovitskaya, V. [477]<br />

Lyubomirsky, I. [477]<br />

MacFarlane, A. [371]<br />

Machayekhi, B.[238, 241, 294, 310,<br />

314]<br />

Machayekhi, D. [313]<br />

Maehara, Y. [426]<br />

Magee, C.W. [246, 316, 322, 348]<br />

Magerle, R. [290]<br />

Maier, M. [237, 396, 424]<br />

Major, J.S. [256, 262, 288]<br />

Makarov, V.V. [284, 459]<br />

Makita, K. [483]<br />

Mal<strong>in</strong>, J.I. [448]<br />

Malkovich, R.S. [320]<br />

Mallard, R.E. [423, 427]<br />

Malyshev, S.A. [476]<br />

Mani, H. [455, 457]<br />

Mannoh, M. [263]<br />

Marbeuf, A. [294]<br />

Marcon, J. [403, 404, 405]<br />

Marek, H.S. [477]<br />

Marenk<strong>in</strong>, S.F. [460, 472]<br />

Marfa<strong>in</strong>g, Y. [312]<br />

Marschall, P. [415, 475]<br />

Marti, A. [250, 355]<br />

Mártil, I. [465]<br />

Mart<strong>in</strong>, J.M. [266, 267, 465]<br />

Mart<strong>in</strong>, P. [404]<br />

Mart<strong>in</strong>, R.W. [434]<br />

Masaki, M. [485]<br />

Masseli, K. [473]<br />

Massies, J. [387]<br />

Masut, R.A. [408, 412]<br />

Matsuhata, H. [457]<br />

Matsui, T. [406, 440, 460]<br />

Matsui, Y. [418]<br />

Matsumoto, S. [435, 441]<br />

Matsumura, N. [456]<br />

Matsushita, A. [371]<br />

Matsushita, S. [346]<br />

Matsuyama, T. [484, 485]<br />

Matyi, R.J. [247, 360]<br />

Matz, R. [482]<br />

Meglicki, Z. [350]


Author Index<br />

Mehrer, H. [353, 363, 470]<br />

Mei, P. [227, 228, 415, 420]<br />

Meier, H.P. [382]<br />

Melchior, H. [361]<br />

Melloch, M.R. [1, 233]<br />

Melman, P. [392]<br />

Méndez, B. [450]<br />

Mendonça, C.A.C. [386]<br />

Menéndez, J. [390]<br />

Merkulov, A.V. [440]<br />

Merl<strong>in</strong>, R. [274]<br />

Merz, J. [393]<br />

Merz, J.L. [244, 245, 249, 250,<br />

339, 359, 361]<br />

Mestres, N. [274, 432]<br />

Meyer, J.W. [316, 322, 348]<br />

Micallef, J. [433]<br />

Migitaka, M. [295, 296]<br />

Mihashi, Y. [417]<br />

Miles, R.H. [443]<br />

Milnes, A.G. [449]<br />

Miloche, M. [241, 310, 314, 460]<br />

M<strong>in</strong>, S.K. [333]<br />

M<strong>in</strong>erv<strong>in</strong>i, A.D. [253]<br />

Mishima, T. [238]<br />

Mitchell, I.V. [421]<br />

Mitev, P. [259]<br />

Mitra, S. [274, 306]<br />

Mitsuhara, M. [435, 441]<br />

Mitsui, S. [369]<br />

Miyauchi, E. [284, 338]<br />

Miyoshi, S. [445]<br />

Mizuta, M. [447]<br />

Mochizuki, A. [289]<br />

Mochizuki, K. [238, 281, 286, 289,<br />

377, 422]<br />

Mochizuki, Y. [447]<br />

Moise, T.S. [261]<br />

Mok<strong>in</strong>a, I.A. [440, 477]<br />

Mol<strong>in</strong>ari, E. [390]<br />

Montie, E.A. [475]<br />

Moon, D.C. [442]<br />

Moore, W.J. [279]<br />

Morgenstern, T. [355]<br />

Mori, H. [324]<br />

Mori, M. [242]<br />

Morishita, Y. [379, 380, 387]<br />

Morita, T. [284, 338]<br />

498<br />

Moriya, N. [290]<br />

Morrow, R.A. [277, 311, 322]<br />

Moseley, A. [434]<br />

Motisuke, P. [408, 412]<br />

Mui, D.S.L. [245, 249, 339, 361, 455]<br />

Mukai, K. [395, 432]<br />

Mukai, S. [242]<br />

Müller, G. [270]<br />

Mulpuri, S. [415, 469]<br />

Murata, M. [431]<br />

Murray, J.J. [307, 327, 330, 334,<br />

348, 350]<br />

Murray, R. [335]<br />

Murrell, D.L. [416]<br />

Musbah, O.A. [400]<br />

Muto, S. [456]<br />

Nadella, R.K. [266, 267]<br />

Nagahara, M. [445]<br />

Nagam<strong>in</strong>e, K. [371]<br />

Nagamune, Y. [232, 387]<br />

Nagano, J. [276, 301]<br />

Nakajima, K. [403, 408, 421]<br />

Nakamura, M. [296]<br />

Nakamura, T. [286, 289]<br />

Nakamura, Y. [232, 250, 387, 401]<br />

Nakanishi, H. [319]<br />

Nakano, Y. [290, 365]<br />

Nakashima, K. [420, 433]<br />

Nakata, Y. [456]<br />

Nam, D.W. [244, 247, 256, 338, 360]<br />

Nam, E.S. [439]<br />

Näser, A. [466]<br />

Nashelskii, A.Y. [410, 464]<br />

Nassibian, A.G. [350]<br />

Navratil, K. [321]<br />

Nazar, L. [415, 420]<br />

Neave, J.H. [231]<br />

Nekado, Y. [296]<br />

Nelson, R.W. [459]<br />

Neugebauer, J. [445]<br />

Neumann, R. [319]<br />

Newman, R.C. [322, 335, 344]<br />

Ng, I. [433]<br />

Nicholas, R.J. [434]<br />

Niedermayer, C. [372]<br />

Niklas, J.R. [293]<br />

Nilsson, S. [382]<br />

Nishida, K. [325]


Author Index<br />

Nish<strong>in</strong>aga, T. [276, 302, 374, 375, 376,<br />

377, 380, 381, 385, 422,<br />

455, 456, 467]<br />

Nish<strong>in</strong>o, T. [292]<br />

Nishiyama, K. [371]<br />

Noack, M. [374]<br />

Noda, T. [232, 387]<br />

Noge, H. [232, 387]<br />

Nolte, D.D. [1, 233]<br />

Nomura, Y. [379, 380, 387]<br />

Norcott, M. [281]<br />

Nordell, N. [253, 317, 353]<br />

Nörenberg, H. [384]<br />

Norman, A.G. [427]<br />

Northrup, J. [344]<br />

Northrup, J.E. [299, 304, 306, 342, 345]<br />

Novikova, V.V. [410, 464]<br />

Nowak, E. [354, 355, 471]<br />

Nozaki, S. [327, 348]<br />

Nozaki, T. [307]<br />

Nukui, T. [242]<br />

Nutt, H.C. [326]<br />

O’Reilly, E.P. [396, 424]<br />

Ogasawara, Y. [316]<br />

Ogata, H. [406, 460]<br />

Ogihara, M. [401]<br />

Oh, J.E. [274]<br />

Oh, Y.T. [389]<br />

Ohfuji, S. [276, 301]<br />

Ohishi, T. [440]<br />

Ohnaka, K. [263]<br />

Ohnishi, H. [340]<br />

Ohno, T. [255]<br />

Ohori, T. [306, 345]<br />

Ohsawa, J. [295, 296]<br />

Ohtsu, H. [485]<br />

Ohtsuka, K. [440]<br />

Ohtsuka, K.I. [406, 460]<br />

Oikawa, H. [289]<br />

Ojala, P. [253, 317, 353]<br />

Okamoto, Y. [467]<br />

Olmsted, B.L. [256, 258, 259, 260]<br />

Omelyanovskii, E.M. [313, 410, 464]<br />

Omura, E. [244, 250, 359, 417]<br />

Onabe, K. [445]<br />

Ong, C.K. [382]<br />

Ono, H. [274]<br />

Osentowski, T.D. [230, 263, 437]<br />

499<br />

Osgood, R.M. [321]<br />

Osh<strong>in</strong>owo, J. [432, 433]<br />

Ossart, P. [391, 398, 402, 410, 413]<br />

Ostl<strong>in</strong>g, M. [277, 304]<br />

Osvald, J. [388]<br />

Otsuka, N. [418]<br />

Ougazzaden, A. [441]<br />

Ourmazd, A. [267]<br />

Oyanagi, H. [457]<br />

Pa<strong>in</strong>e, B.M. [283]<br />

Pajot, B. [294, 295]<br />

Pak, K. [316]<br />

Pakhomov, A.V. [313, 410, 464]<br />

Palguev, S.F. [269]<br />

Palma, A. [379]<br />

Pan, H. [482]<br />

Paoli, T.L. [234, 252, 256, 347]<br />

Parguel, V. [459]<br />

Park, H.H. [437, 438, 439, 440]<br />

Park, H.L. [442, 470]<br />

Park, R.M. [446]<br />

Parker, S.M. [462]<br />

Part<strong>in</strong>, D.L. [488]<br />

Pascual, J. [431]<br />

Pashkova, O.N. [460, 472]<br />

Paska, Z.F. [357]<br />

Pass, C.J. [234, 239, 248]<br />

Pavesi, L. [245, 311, 328, 356, 368]<br />

<strong>Pearton</strong>, S.J. [63, 266, 267, 269, 277,<br />

294, 311, 405, 414, 444,<br />

449, 458, 461, 464]<br />

Pechman, R.J. [383]<br />

Pe<strong>in</strong>er, E. [468]<br />

Perenboom, J.A.A.J. [343]<br />

Perley, A.P. [266, 267, 405, 414]<br />

Perr<strong>in</strong>, S.D. [432]<br />

Peskov, N.V. [380]<br />

Peterson, D.L. [239, 393]<br />

Petrich, G.S. [231, 374, 375]<br />

Petroff, P.M. [275, 307, 392, 393, 455]<br />

Pétursson, J. [317]<br />

Peyre, H. [431]<br />

Pfeiffer, L. [246]<br />

Pfeiffer, L.N. [390]<br />

Pfeiffer, W. [290]<br />

Pfister, J.C. [366]<br />

Pfiz, T. [372]<br />

Phillips, C. [392]


Author Index<br />

Pilkuhn, M.H. [407, 410, 419]<br />

P<strong>in</strong>zone, C.J. [478]<br />

Piqueras, J. [450]<br />

Pisarev, A.A. [272]]<br />

Planel, R. [347, 391]<br />

Plano, W.E. [230, 234, 244, 247, 256,<br />

257, 258, 263, 338, 360, 437]<br />

Ploog, K. [384]<br />

Plummer, J.D. [234, 239, 248, 285, 286,<br />

302, 303, 307, 318, 330,<br />

350, 362]<br />

Polyakov, A.J. [313]<br />

Polyakov, A.Y. [410, 449, 464]<br />

Ponce, F.A. [347, 396, 399]<br />

Poole, P.J. [421]<br />

Poroikov, J.A. [460, 472]<br />

Potter, T.J. [412]<br />

Pozsgai, I. [293]<br />

Prescha, T. [310, 311]<br />

Primig, R. [482]<br />

Pross, P. [290]<br />

Pudenzi, M.A.A. [351]<br />

Quillec, M. [402, 410, 413]<br />

Qu<strong>in</strong>tana, V. [355]<br />

Rahbi, R. [238, 241, 294, 314]<br />

Rai-Choudhury, P. [449]<br />

Räisänen, J. [312, 326]<br />

Rajatora, M. [326]<br />

Rajeswaran, G. [239, 358, 367, 393]<br />

Ralston, J.D. [396, 424, 426]<br />

Ramasamy, P. [278]<br />

Rao, E.V.K. [391, 398, 402, 410,<br />

413, 441]<br />

Rao, K. [441]<br />

Rao, M.V. [266, 267, 286, 309, 341,<br />

415, 459, 463, 467, 469]<br />

Rao, P.R.S. [363]<br />

Rao, S.S. [403, 414]<br />

Rastogi, A. [486, 487, 488]<br />

Ravich, V.N. [460, 472]<br />

Razeghi, M. [439]<br />

Reddy, K.V. [486, 487, 488]<br />

Reddy, V. [342]<br />

Red<strong>in</strong>bo, G.F. [275]<br />

Rees, P.K. [326]<br />

Régrény, A. [342]<br />

Re<strong>in</strong>hart, F.K. [245, 328, 356, 365,<br />

366, 368, 370]<br />

500<br />

Ren, H.W. [302]<br />

Renner, D. [474]<br />

Rentschler, J.A. [267]<br />

Reynolds, C.L. [420]<br />

Reynolds, S. [360]<br />

Richard, T.A. [243, 249]<br />

Rihet, Y. [459]<br />

R<strong>in</strong>gel, S.A. [463]<br />

Riseman, T.M. [372]<br />

Robb<strong>in</strong>s, V.M. [257]<br />

Robert, J.L. [347]<br />

Roberts, C. [236, 246, 248, 280,<br />

327, 349]<br />

Rob<strong>in</strong>son, D. [270]<br />

Rob<strong>in</strong>son, H.G.[282, 283, 318, 341, 342]<br />

Rob<strong>in</strong>son, W. [393]<br />

Rodionov, A.I. [321]<br />

Roedel, R.J. [364, 459]<br />

Roos, G. [315, 407, 411]<br />

Rose, B. [460]<br />

Roth, A.P. [408, 412]<br />

Rothemund, W. [396, 424, 426]<br />

Rubart, W.S. [307, 330, 350]<br />

Rucki, A. [353, 363, 364, 470]<br />

Rudra, A. [456]<br />

Ruf, T. [447]<br />

Rytova, N.S. [310]<br />

Ryu, S.W. [428]<br />

Sacilotti, M.A. [408, 409, 411, 412,<br />

417, 418]<br />

Sacks, R.N. [261]<br />

Sadana, D.K. [281]<br />

Sai, N. [394]<br />

Saito, D. [316]<br />

Saito, R. [431]<br />

Sajoto, T. [335]<br />

Sakaguchi, M. [319, 323, 324]<br />

Sakai, S. [251]<br />

Sakaki, H. [232, 387]<br />

Sakakibara, K. [418, 471, 475, 479]<br />

Sakalas, A.P. [278, 300]<br />

Salem<strong>in</strong>k, H.W.M. [284]<br />

Sallese, J.M. [456]<br />

Salmeron, M. [233]<br />

S<strong>and</strong>rik, R. [388]<br />

Sano, N. [422]<br />

Santos, M. [246, 334, 335]<br />

Sapriel, J. [391, 398]


Author Index<br />

Sargünas, V.R. [278, 300]<br />

Sasa, S. [236]<br />

Satho, M. [319, 325]<br />

Sato, E.A. [409, 411, 417, 418]<br />

Sauer, N.J. [279]<br />

Scarrott, K. [427]<br />

Schad, R.G. [281]<br />

Schade, U. [479, 480]<br />

Schaff, W.J. [283]<br />

Schauer, S.N. [459]<br />

Scheffler, M. [301]<br />

Schier, M. [419]<br />

Schlaak, W. [297, 352, 446, 447, 453,<br />

454, 462, 469]<br />

Schlachetzki, A. [468]<br />

Schlapp, W. [423]<br />

Schles<strong>in</strong>ger, T.E. [235, 240, 243,<br />

245, 449]<br />

Schlotter, N.E. [406]<br />

Schmidt, H.J. [472]<br />

Schmidt, M.T. [321]<br />

Schneider, M. [391]<br />

Schneider, W. [372]<br />

Scholz, R. [372, 398]<br />

Schowalter, L.J. [381]<br />

Schubert, E.F. [244, 246, 279, 336,<br />

337, 478]<br />

Schultz, M. [277, 372, 398, 401]<br />

Schulz, K.J. [400]<br />

Schumann, B. [354, 355, 471]<br />

Schumann, D. [371]<br />

Schumann, H. [466]<br />

Schützendübe, P. [384]<br />

Schwab, C. [314, 371, 372]<br />

Schwartz, B. [335]<br />

Schwartz, C.L. [228, 415, 420]<br />

Schwarz, S.A. [227, 228, 415, 420, 480]<br />

Schweizer, H. [423]<br />

Scilla, G.J. [373]<br />

Scott, E.G. [405, 423]<br />

Sealy, B.J. [429, 467]<br />

Sekiguchi, Y. [418, 471, 475]<br />

Seko, M. [418, 475]<br />

Sellitto, P. [347]<br />

Sempr<strong>in</strong>i, E. [379]<br />

Serreze, H.B. [477]<br />

Seshadri, S. [259, 261]<br />

Seta, M. [450]<br />

501<br />

Shacham-Diam<strong>and</strong>, Y. [351]<br />

Shah, D.M. [406]<br />

Shahar, A. [254]<br />

Shapira, Y. [352]<br />

Sharma, V.K.M. [451, 452]<br />

Shayegan, M. [246, 334, 335]<br />

Sheets, J. [316, 322, 348]<br />

Shen, X.Q. [375, 380, 381, 385,<br />

455, 456]<br />

Shepherd, F.R. [468, 474, 477]<br />

Shi, S.S. [245, 249, 339, 361]<br />

Shiba, Y. [426]<br />

Shichijo, H. [247, 360]<br />

Shieh, T.H. [287]<br />

Shi<strong>in</strong>a, K. [306, 345]<br />

Shimogaki, Y. [365]<br />

Shimomura, S. [422]<br />

Sh<strong>in</strong>oda, A. [286, 342]<br />

Sh<strong>in</strong>tani, Y. [251]<br />

Shiomi, A. [323, 324]<br />

Shiraishi, K. [255, 376, 378]<br />

Shiraki, Y. [445]<br />

Shitara, T. [231, 376, 377, 422]<br />

Shiu, W.C. [433]<br />

Shmidt, N.M. [440, 477]<br />

Sicart, J. [347]<br />

Siefert, R.L. [383]<br />

Siegel, W. [293]<br />

Siethoff, H. [464, 466]<br />

Silveira, J.P. [447]<br />

Simes, R. [393]<br />

Simons, D.S. [266, 267, 286, 309, 315,<br />

341, 364, 415, 459, 463, 467, 469]<br />

S<strong>in</strong>, Y.K. [355]<br />

S<strong>in</strong>gh, J. [274, 405]<br />

S<strong>in</strong>gh, S. [335]<br />

Skoryat<strong>in</strong>a, E.A. [320]<br />

Skromme, B.J. [415, 420]<br />

Skudlik, H. [290]<br />

Slotte, J. [326]<br />

Smirnov, V.M. [272]<br />

Smith, A.D. [403, 414]<br />

Smith, F.T. [251, 357, 400]<br />

Smith, R.S. [326]<br />

Smith, S.C. [243, 245, 249]<br />

Sobotta, H. [354]<br />

Södervall, U. [298, 364]<br />

Solmon, H. [270]


Author Index<br />

Somogyi, K. [241, 293, 310]<br />

Song, J.I. [480]<br />

Soni, R.K. [450]<br />

Sonoda, T. [369]<br />

Speier, P. [432]<br />

Spence, J.P. [367]<br />

Spencer, M.G. [283]<br />

Spiller, G.D.T. [405]<br />

Spr<strong>in</strong>gthorpe, A.J. [472, 477]<br />

Srivastava, A.K. [428]<br />

Stähl, K. [277, 304]<br />

Stam, M. [449]<br />

Stark, J.B. [336]<br />

Stavola, M. [461, 464]<br />

Steckl, A.J. [398]<br />

Sternitzke, M. [270]<br />

Stevenson, D.A. [282, 318, 327, 329,<br />

334, 437, 438]<br />

Stillman, G.E. [256, 262, 288]<br />

Stobbs, W.M. [392]<br />

Stollenwerk, M. [433]<br />

Stolwijk, N.A. [298, 353, 363, 364, 470]<br />

Storz, F.G. [386]<br />

Streetman, B.G. [308]<br />

Stutius, W. [258]<br />

Stutzmann, M. [454]<br />

Suehiro, H. [242]<br />

Sugawara, M. [395, 432]<br />

Sugg, A.R. [243, 247, 249, 256, 258]<br />

Sugimoto, H. [440]<br />

Sugiura, H. [435, 441]<br />

Sun, D. [262, 264]<br />

Sun, J.Z. [339]<br />

Sundaram, V.S. [452]<br />

Suzuki, T. [374, 376]<br />

Swam<strong>in</strong>athan, V. [238, 277, 420,<br />

461, 464]<br />

Swanson, M.L. [477]<br />

Swart, J.W. [351]<br />

Syfosse, G. [313]<br />

Szafranek, I. [262]<br />

Szafranek, M. [262]<br />

Tada, K. [290, 365]<br />

Tadayon, B. [283]<br />

Tadayon, S. [283]<br />

Taguti, K. [483]<br />

Takahashi, K. [484, 485]<br />

Takahashi, S. [417]<br />

502<br />

Takamiya, S. [369]<br />

Takamori, A. [258, 284, 338]<br />

Takamori, T. [284, 338]<br />

Takano, H. [319, 323, 324, 325]<br />

Takebe, T. [286, 342, 377]<br />

Takeda, A. [364]<br />

Takeda, S. [289]<br />

Takiguchi, T. [417]<br />

Takizawa, J. [450]<br />

Talamo, A. [379]<br />

Tamura, S. [319, 325]<br />

Tan, T.Y. [237, 241, 252, 261, 277,<br />

283, 291, 297, 298, 300, 301, 303,<br />

305, 331, 333, 336, 344, 345, 346,<br />

360, 367, 372, 401, 461]<br />

Tanaka, M. [302, 374, 376]<br />

Tang, T.K. [242, 248]<br />

Tanigawa, S. [280, 329, 371]<br />

Tan<strong>in</strong>aka, M. [401]<br />

Tanoue, H. [450]<br />

Tasker, P.J. [283]<br />

Taylor, M. [423]<br />

Taylor, S. [456]<br />

Taylor, S.J. [416]<br />

Tejwani, M.J. [283]<br />

Tell, B. [336, 337, 338]<br />

Terada, S. [346]<br />

Terauchi, Y. [251]<br />

Tews, H. [288, 319, 369]<br />

Theys, B. [191, 238, 241, 294, 310,<br />

313, 314, 454, 460]<br />

Thibierge, H. [398, 402, 410, 413, 441]<br />

Thijs, P.J.A. [435, 436, 439]<br />

Thomas, L.M. [292]<br />

Thompson, J. [434]<br />

Thornton, R.L. [234, 248, 252, 256,<br />

396, 399]<br />

Thrush, C.M. [488]<br />

Thrush, E.J. [427]<br />

Thundat, T. [381]<br />

Tian, Q. [271]<br />

Tjaden, D.L.A. [437, 438, 475, 476]<br />

Tokuda, Y. [299]<br />

Tokumitsu, E. [357]<br />

Tomass<strong>in</strong>i, N. [379]<br />

Tomioka, T. [236]<br />

Tomita, N. [422]<br />

Toml<strong>in</strong>son, W.J. [254]


Author Index<br />

Towers, M. [326]<br />

Trafas, B.M. [383]<br />

Tramontana, J.C. [399]<br />

Treat, D. [399]<br />

Treat, D.W. [262, 264]<br />

Treichler, R. [247, 288, 369, 419, 465]<br />

Trequattr<strong>in</strong>i, F. [294]<br />

Tsai, C.M. [427]<br />

Tsai, K.L. [394, 427]<br />

Tsang, J.S. [394, 427]<br />

Tsang, W.T. [386, 473]<br />

Tsu, R. [305]<br />

Tsuchida, N. [296]<br />

Tsuchiya, M. [232, 275, 307, 387]<br />

Tsugami, M. [369]<br />

Tsushi, A. [250]<br />

Tu, C.W. [244, 294]<br />

Tu, S.L. [484]<br />

Tuck, B. [227, 229, 230, 245, 328, 390]<br />

Tulch<strong>in</strong>sky, D.A. [311]<br />

Tweet, D.J. [457]<br />

Uekusa, S. [113]<br />

Uematsu, M. [277, 280, 281, 287,<br />

354, 398, 401]<br />

Uesugi, F. [417]<br />

Ullrich, H. [297, 352, 446, 447, 453,<br />

454, 462, 469]<br />

Umeno, M. [348]<br />

Unger, B. [479]<br />

Urban, K. [363, 448, 470]<br />

Usami, A. [299]<br />

Uskov, W.A. [321]<br />

Vaccaro, P. [340]<br />

Van Berlo, W.H. [253, 317, 353, 466]<br />

Van de Walle, C.G. [445]<br />

Van de Wijgert, W.M. [435, 436, 439]<br />

Van der Stadt, A.F.W. [343]<br />

Van der Vleuten, W.C. [284]<br />

Van Dongen, T. [438, 475]<br />

Van Gieson, E. [382]<br />

Van Gurp, G.J. [435, 436, 437, 438,<br />

439, 475, 476]<br />

Van Uitert, L.G. [335]<br />

Varava, A.V. [272]<br />

Vaudry, C. [459]<br />

Vawter, G.A. [244, 250, 359]<br />

Venkatesan, T. [227, 228, 415, 420]<br />

Vesely, E.J. [256]<br />

503<br />

Veuhoff, E. [247, 465]<br />

Vicek, J.C. [406, 413, 414]<br />

Vieu, C. [391]<br />

Virkar, A.V. [271]<br />

Vitkauskas, A.A. [278, 300]]<br />

Viturro, R.E. [259]<br />

Vol<strong>and</strong>, U. [483]<br />

Völkl, J. [464, 466]<br />

Vook, D.W. [360]<br />

Vriezema, C.J. [463, 470]<br />

Wada, K. [280, 281, 287, 354]<br />

Wada, M. [418, 471, 475, 479]<br />

Wada, N. [251]<br />

Wada, O. [258]<br />

Wada, T. [299, 364]<br />

Wager, J.F. [275, 300]<br />

Wagner, J. [426]<br />

Wake, D. [405]<br />

Wakejima, A. [422]<br />

Walker, J. [315]<br />

Walsöe de Reca, N.E. [486]<br />

Walter, W. [382]<br />

Walters, R.J. [125]<br />

Walukiewicz, W. [359, 366, 473, 480]<br />

W<strong>and</strong>el, K. [407, 410, 419]<br />

Wang, C. [366, 367, 368, 389]<br />

Wang, E.G. [448]<br />

Wang, K.W. [462]<br />

Wang, L. [297, 302, 447]<br />

Wang, X.S. [383]<br />

Warren, A.C. [346]<br />

Watanabe, A. [376]<br />

Watanabe, M. [242]<br />

Watanabe, N. [252, 332]<br />

Watanabe, T. [286, 340, 342, 377]<br />

Weaver, J.H. [352, 383]<br />

Webb, R.P. [429]<br />

Weber, E.R. [233, 348, 460]<br />

Weber, J. [310, 311]<br />

Wei, L. [280, 329]<br />

Weill, G. [313]<br />

Weimer, M. [386]<br />

Wenzl, H. [374]<br />

Werner, J. [361]<br />

Werner, P. [277, 372, 398, 401]<br />

West, K.W. [246, 390]<br />

Weyer, G. [469]<br />

Wheeler, C.B. [459]


Author Index<br />

Whelan, J.M. [283, 291, 320]<br />

Whitlow, H.J. [277, 304]<br />

Whitney, P.S. [406, 413, 414]<br />

Wichert, T. [290]<br />

Wicks, G.W. [259, 288]<br />

Wielsch, U. [407, 419]<br />

Wilkie, J.H. [467]<br />

Willén, B. [357]<br />

Williams, J. [258]<br />

Williams, M.D. [386]<br />

Williams, P. [459]<br />

Willoughby, A.F.W. [451, 452]<br />

Wilson, R.G. [63, 269, 444, 449, 458]<br />

Wisser, J. [472]<br />

Wittorf, D. [470]<br />

Wolf, H. [290]<br />

Wolfram, P. [461]<br />

Wolk, J.A. [447]<br />

Wolter, J.H. [284, 343]<br />

Wong, S.L. [434]<br />

Wood, A. [434]<br />

Wood, C.E.C. [283, 288]<br />

Woodall, J.M. [1, 233, 346]<br />

Woodbridge, K. [236, 246, 248, 280,<br />

327, 349]<br />

Wu, A.T. [348]<br />

Wu, C.H. [243, 319, 320]<br />

Wu, M.Y. [287]<br />

Wu, X.S. [244, 250, 359]<br />

Wu, Z. [321]<br />

Wurz<strong>in</strong>ger, P. [288, 369]<br />

Xia, H. [325]<br />

Xu, F. [352]<br />

Xu, J. [394]<br />

Xu, Z. [394]<br />

Yaguchi, H. [445]<br />

Yajima, H. [242]<br />

Yakimenko, I.J. [440]<br />

Yakobson, B. [389]<br />

Yakobson, S.V. [410, 464]<br />

Yamaguchi, H. [254, 315, 375, 377, 384]<br />

Yamamoto, T. [286, 342, 377]<br />

Yamamura, S. [431]<br />

Yamazaki, S. [395, 432]<br />

Yang, B. [317]<br />

Yang, B.H. [293, 317]<br />

Yang, E.S. [443]<br />

Yang, K. [381]<br />

504<br />

Yang, S.J. [484]<br />

Yang, X. [394]<br />

Yang, Y.F. [443]<br />

Yang, Y.N. [383]<br />

Yesis, L. [335]<br />

Yi, J.Y. [386]<br />

Yokota, K. [319, 323, 324, 325]<br />

Yokoyama, N. [456]<br />

Yoneda, M. [250]<br />

Yonezu, H. [316]<br />

Yoo, H.J. [251]<br />

Yoon, H.W. [227]<br />

Yoon, I.T. [442]<br />

You, H.M. [261, 305, 344, 345, 346]<br />

Young, E.W.A. [462, 463, 470]<br />

Yu, C.F. [321]<br />

Yu, K.M. [359, 366, 460, 473, 480]<br />

Yu, S. [237, 241, 252, 291, 297,<br />

298, 305, 333]<br />

Yu, S.J. [433]<br />

Yuan, S. [395]<br />

Yugo, S. [431]<br />

Yurchuk, S.J. [284, 459]<br />

Yurre, T.A. [440]<br />

Yutani, A. [325]<br />

Zahraman, K. [241, 310]<br />

Zalm, P.C. [463, 470]<br />

Zavada, J.M. [63, 269, 444, 458]<br />

Zehr, S.W. [474]<br />

Zhang, K. [405]<br />

Zhang, P. [394]<br />

Zhang, Q. [389]<br />

Zhang, Q.M. [343, 347, 366, 367, 368]<br />

Zhang, S. [276]<br />

Zhang, S.B. [299, 306, 345]<br />

Zhang, T. [355]<br />

Zhdanov, S.K. [272]<br />

Zheng, B. [394]<br />

Zheng, J.F. [233]<br />

Zheng, L.R. [331]<br />

Ziel<strong>in</strong>ski, E. [432]<br />

Zimmermann, H. [367, 461]<br />

Zolper, J.C. [63, 265]<br />

Zou, J. [395]<br />

Zou, W.X.[244, 245, 249, 250, 339, 361]<br />

Zrenner, A. [335]<br />

Zucker, E.P. [252]<br />

Zundel, T. [310]


Author Index<br />

Zwicknagl, P. [288, 369]<br />

Zydzik, G.J. [335]<br />

Zypman, F. [305]<br />

Zytkiewicz, Z.R. [372]<br />

505


Keywords (abstracts)<br />

2-dimensional [230, 231, 232, 343,<br />

374, 375, 376, 377, 384, 385,<br />

386, 392, 421, 422]<br />

3-dimensional [316, 386, 392, 455]<br />

adatoms [254, 375, 376, 378, 381,<br />

382, 445]<br />

amorphization [297, 307, 329, 341, 349,<br />

446, 449, 453, 462]<br />

amorphous [316, 319, 322, 323,<br />

324, 325, 341, 347, 453, 466]<br />

amphoteric [244, 247, 333, 358, 366,<br />

473, 479]<br />

anelastic [294]<br />

anions [385, 414, 433, 436, 437, 438]<br />

anisotropy [232, 234, 256, 374,<br />

384, 385, 482]<br />

anneal<strong>in</strong>g [224, 227, 231, 233, 234, 235,<br />

236, 240, 242, 243, 245, 252, 253,<br />

256, 257, 258, 259, 260, 261, 263,<br />

264, 265, 266, 267, 268, 271, 274,<br />

275, 276, 277, 278, 280, 281, 282,<br />

283, 284, 285, 288, 289, 290, 291,<br />

292, 296, 299, 300, 301, 304, 305,<br />

306, 307, 309, 312, 316, 317, 318,<br />

319, 320, 321, 324, 325, 327, 329,<br />

331, 332, 334, 336, 337, 338, 340,<br />

341, 342, 343, 344, 346, 348, 349,<br />

351, 352, 353, 357, 359, 360, 362,<br />

363, 365, 366, 369, 373, 374, 388,<br />

389, 390, 391, 392, 393, 394, 395,<br />

396, 397, 398, 401, 402, 403, 404,<br />

405, 407, 408, 409, 410, 412, 413,<br />

414, 415, 417, 420, 421, 422, 424,<br />

426, 427, 428, 430, 431, 432, 433,<br />

434, 437, 438, 439, 444, 446, 447,<br />

448, 449, 450, 453, 454, 456, 458,<br />

459, 461, 462, 463, 464, 465, 467,<br />

468, 469, 470, 471, 472, 474, 475,<br />

482, 483]<br />

annihilation [287]<br />

anodic [364, 485, 487]<br />

anodization [395]<br />

anomaly [331, 332]<br />

antisites [264, 265, 296, 350, 355, 396,<br />

[424, 450]<br />

arrays [470]<br />

b<strong>and</strong>-edge [244, 247, 338, 395, 424]<br />

b<strong>and</strong>-gap [238, 241, 260, 294, 312,<br />

338, 340, 395, 424, 426, 483]<br />

b<strong>and</strong>s[316, 317, 340, 391, 447, 465, 479]<br />

barrier [231, 243, 249, 274, 276,<br />

304, 311, 342, 344, 384, 386, 397,<br />

406, 413, 418, 427, 430, 433, 445,<br />

448, 460, 463, 474]<br />

bombardment [278, 300, 351, 383]<br />

boundaries[224, 270, 276, 286, 301, 325,<br />

387, 440, 486, 488]<br />

capless [234, 256, 257, 281, 288,<br />

291, 320]<br />

carrier [236, 243, 249, 250, 251, 258,<br />

287, 288, 289, 294, 295, 299, 306,<br />

319, 320, 326, 338, 339, 340, 345,<br />

346, 350, 358, 368, 388, 393, 400,<br />

417, 433, 451, 452, 454, 462, 470,<br />

471, 473, 476, 478, 484]<br />

cathodolum<strong>in</strong>escence [259, 274, 275,<br />

306, 332, 382, 392, 401, 450]<br />

507


cation [232, 233, 254, 385, 386, 414,<br />

419, 427, 433, 436, 437, 438, 439]<br />

channell<strong>in</strong>g [325, 341, 366, 479]<br />

Czochralski [285, 293, 322, 331, 339]<br />

dangl<strong>in</strong>g-bond [386]<br />

deep-level [293, 307, 312, 316,<br />

340, 396, 424, 449, 459, 460]<br />

defect-free [443]<br />

deposition [231, 247, 251, 254, 261,<br />

276, 286, 288, 289, 301, 324, 332,<br />

335, 348, 357, 372, 373, 374, 375,<br />

377, 380, 383, 386, 395, 400, 420,<br />

421, 424, 426, 442, 468, 471, 472,<br />

473, 474, 481]<br />

diffusion-controlled [290]<br />

diffusion-<strong>in</strong>duced [253, 283, 300, 340,<br />

363, 364, 368, 369, 370, 426]<br />

diode [250, 311, 449]<br />

dislocations [224, 230, 274, 293, 306,<br />

332, 341, 342, 344, 348, 360, 363,<br />

399, 426, 440, 470]<br />

dislocation-free [251]<br />

di-vacancy [414]<br />

donor-acceptor [442, 447, 468, 472, 474]<br />

donors [238, 241, 249, 257, 258,<br />

259, 289, 294, 295, 297, 298, 303,<br />

304, 306, 307, 308, 310, 311, 312,<br />

313, 314, 315, 328, 329, 330, 333,<br />

338, 340, 344, 345, 346, 347, 350,<br />

360, 369, 370, 371,388, 398, 406,<br />

410, 437, 438, 442, 447, 449, 459,<br />

460, 461, 464, 465, 468, 470, 471,<br />

472, 474, 475, 476]<br />

DX [246, 334, 347]<br />

EL2 [226, 276, 277, 293, 311, 373, 448]<br />

electromigration [278, 372, 421]<br />

epilayer [236, 286, 287, 316, 329,<br />

344, 471]<br />

epitaxy [227, 232, 234, 237, 239,<br />

244, 248, 250, 253, 274, 281, 283,<br />

286, 287, 289, 315, 316, 317, 326,<br />

329, 335, 337, 345, 346, 347, 348,<br />

353, 355, 357, 368, 369, 376, 377,<br />

379, 380, 381, 382, 384, 387, 389,<br />

395, 406, 407, 408, 411, 413, 416,<br />

417, 418, 419, 421, 422, 423, 424,<br />

427, 436, 438, 447, 452, 453, 455,<br />

456, 457, 462, 468, 478, 482, 483, 488]<br />

etch<strong>in</strong>g [253, 276, 301, 317, 324,<br />

343, 353, 365, 372, 373, 386, 402,<br />

409, 413, 455, 472, 475]<br />

exciton [244, 247, 287, 338, 392,<br />

441, 483]<br />

faults [287, 345, 440]<br />

Fickian [403, 414]<br />

films [234, 239, 242, 244, 248, 250,<br />

251, 276, 290, 301, 316, 319, 320,<br />

322, 324, 334, 335, 347, 354, 355,<br />

357, 359, 373, 381, 395, 400, 401,<br />

415, 419, 424, 438, 440, 445, 450,<br />

451, 469, 475, 486, 487, 488]<br />

first-pr<strong>in</strong>ciples [343, 347, 378,<br />

386, 389, 444, 445]<br />

fluence [278, 392, 449, 450]<br />

free-carrier [454]<br />

Frenkel [245, 275, 300, 327]<br />

hetero-epitaxial [348, 445, 463]<br />

hetero-<strong>in</strong>terface [250, 258, 274, 316,<br />

348, 406, 413, 440, 468]<br />

heterojunction [237, 238, 254, 286,<br />

359, 383, 410, 419, 420, 440, 442,<br />

480, 481, 482]<br />

heterostructure [234, 238, 252, 256,<br />

262, 287, 388, 392, 420, 426, 427,<br />

438, 439]<br />

homo-epitaxial [381, 463]<br />

hysteresis [302]<br />

<strong>in</strong>-diffusion [234, 237, 239, 248, 252,<br />

258, 267, 276, 277, 280, 281, 286,<br />

291, 292, 294, 297, 309, 321, 339,<br />

341, 346, 354, 363, 366, 367, 389,<br />

398, 408, 411, 415, 416, 419, 459,<br />

461, 463, 464, 465, 467, 469]<br />

<strong>in</strong>terstitial [230, 236, 237, 238, 250, 253,<br />

278, 282, 283, 284, 285, 289, 300,<br />

301, 303, 304, 312, 317, 318, 339,<br />

342, 344, 353, 355, 357, 359, 360,<br />

361, 362, 364, 366, 367, 368, 369,<br />

370, 371, 373, 386, 387, 405, 406,<br />

407, 409, 411, 416, 417, 418, 419,<br />

437, 438, 442, 460, 462, 470, 471,<br />

472, 473, 474, 475, 476, 478, 479, 480]<br />

<strong>in</strong>terstitialcy [261, 298, 434, 483]<br />

<strong>in</strong>terstitial-substitutional [227, 229, 250,<br />

508


261, 297, 355, 361, 366, 401, 406,<br />

407, 419, 437, 438, 451, 452, 460,<br />

473, 475, 476, 478, 479, 480, 484]<br />

<strong>in</strong>terstitial-type [470]<br />

isl<strong>and</strong>s [335, 383, 384, 426, 455]<br />

lattice-mismatch [389, 408, 412, 463]<br />

light-emitt<strong>in</strong>g [250]<br />

L<strong>in</strong>dhard-Scharff-Schiott [334]<br />

l<strong>in</strong>e-width [395, 421]<br />

lum<strong>in</strong>escence [235, 240, 274, 281, 292,<br />

340, 356, 390, 396, 424, 427, 447,<br />

450, 454, 470, 474]<br />

metalorganic [231, 232, 247, 253, 261,<br />

286, 288, 289, 306, 317, 332, 345,<br />

348, 353, 357, 380, 385, 386, 387,<br />

408, 411, 416, 417, 418, 419, 420,<br />

426, 427, 428, 430, 435, 440, 462,<br />

465, 466, 468, 470, 471, 472, 473,<br />

474, 478, 480, 481, 482]<br />

microcathodolum<strong>in</strong>escence [312]<br />

microcrystall<strong>in</strong>e [321]<br />

mid-gap [276]<br />

migration [231, 236, 244, 245, 246, 248,<br />

253, 254, 255, 262, 264, 267, 269,<br />

280, 283, 285, 291, 292, 301, 304,<br />

312, 314, 315, 317, 320, 334, 335,<br />

336, 340, 342, 344, 350, 353, 355,<br />

374, 375, 377, 378, 380, 381, 382,<br />

383, 385, 388, 402, 409, 410, 411,<br />

413, 414, 417, 418, 422, 442, 445,<br />

446, 448, 455, 459, 464, 471, 474,<br />

486, 487]<br />

misfit [236, 426, 439, 440]<br />

mismatch [290, 420, 436, 443, 468]<br />

misorientation [231, 237, 281, 374,<br />

375, 376, 386, 422, 482]<br />

mobility [231, 241, 246, 283, 294, 295,<br />

310, 343, 372, 373, 374, 375, 381,<br />

421, 468]<br />

models [256, 278, 321, 336, 360, 367,<br />

404, 467, 476]<br />

moiré [426]<br />

monolayer [231, 232, 287, 315, 345, 374,<br />

375, 386, 483]<br />

Monte Carlo [231, 239, 249, 254, 289,<br />

341, 380, 385, 392]<br />

Mössbauer [468]<br />

multi-layers [267, 370, 372, 377,<br />

383, 448, 482]<br />

muons [370, 371, 372]<br />

muonium [313, 314, 370, 371, 372]<br />

native [230, 243, 245, 249, 257,<br />

258, 259, 263, 276, 281, 298, 299,<br />

301, 306, 321, 322, 345, 358, 366,<br />

437, 444, 445, 448, 452, 473, 479]<br />

negative-U [311, 445]<br />

neutron [320, 486, 487]<br />

non-equilibrium[260, 279, 280, 285, 305,<br />

331, 354, 365, 366, 367, 389, 404,<br />

440, 480, 481]<br />

n-type [233, 238, 250, 251, 253, 257,<br />

258, 266, 287, 294, 295, 298, 299,<br />

303, 305, 310, 311, 312, 313, 314,<br />

316, 317, 319, 321, 322, 339, 340,<br />

344, 345, 346, 355, 357, 363, 365,<br />

371, 373, 388, 393, 394, 400, 410,<br />

415, 429, 437, 438, 442, 444, 452,<br />

460, 461, 463, 464, 467, 468, 469,<br />

474, 475, 476, 480, 483, 486, 488]<br />

offsets [242, 247]<br />

organometallic [228, 254, 283, 315, 416,<br />

457, 478]<br />

out-diffusion [237, 252, 266, 267, 268,<br />

274, 281, 286, 287, 290, 291, 297,<br />

299, 305, 307, 344, 345, 346, 363,<br />

369, 398, 405, 415, 435, 441, 443,<br />

444, 450, 458, 462, 466, 469, 470,<br />

471, 472, 474]<br />

oxidation [243, 249, 321, 324, 340,<br />

485, 487]<br />

passivated [310, 311, 312, 407, 411,<br />

449, 461, 464]<br />

passivation [241, 263, 310, 312, 407,<br />

411, 444, 447, 449, 452, 453, 454, 463]<br />

patterned [242, 247, 254, 377, 382, 445]<br />

p-doped [253, 300, 311, 317, 353]<br />

Pendellösung [443]<br />

permeation [485]<br />

photo-electron [299, 324]<br />

photo-emission [276, 301, 351, 479]<br />

photo-<strong>in</strong>duced [292]<br />

photolum<strong>in</strong>escence [236, 239, 245, 248,<br />

253, 258, 259, 260, 262, 264, 277,<br />

283, 286, 287, 288, 292, 293, 316,<br />

509


317, 322, 327, 340, 350, 356, 365,<br />

366, 368, 369, 370, 389, 390, 391,<br />

392, 394, 395, 396, 403, 407, 408,<br />

409, 411, 412, 414, 417, 421, 422,<br />

424, 425, 426, 427, 428, 429, 430,<br />

431, 432, 433, 435, 436, 439, 441,<br />

442, 448, 450, 456, 465, 466, 470,<br />

474, 479, 483]<br />

photonic [259]<br />

Poole [421]<br />

positron [280, 329]<br />

pre-exponential [275, 299, 300, 464, 466]<br />

protons [292, 312, 322, 323]<br />

p-type [233, 237, 238, 243, 253, 257,<br />

263, 264, 265, 267, 279, 280, 283,<br />

287, 288, 289, 294, 297, 298, 299,<br />

300, 303, 305, 316, 317, 319, 321,<br />

322, 328, 340, 345, 354, 358, 360,<br />

363, 365, 373, 388, 393, 394, 403,<br />

406, 410, 413, 414, 415, 418, 420,<br />

429, 444, 450, 460, 461, 463, 464,<br />

467, 472, 473, 474, 483]<br />

quantum-well [234, 239, 243, 249, 252,<br />

253, 254, 256, 260, 262, 274, 287,<br />

306, 369, 370, 391, 392, 393, 395,<br />

396, 399, 422, 425, 426, 427, 430,<br />

431, 433, 434, 435, 436, 439, 441, 456]<br />

radiation-enhanced [272]<br />

radiotracers [485, 487]<br />

Raman [283, 306, 345, 350, 351, 368,<br />

390, 391, 426, 431, 432, 433, 449]<br />

roughen<strong>in</strong>g [464]<br />

roughness [380, 384]<br />

self-<strong>in</strong>terstitials [237, 240, 252, 277, 279,<br />

280, 283, 291, 297, 298, 300, 301,<br />

303, 333, 354, 367, 368, 394, 404,<br />

461, 470, 481]<br />

simulation [231, 239, 249, 277, 278,<br />

282, 285, 286, 289, 301, 303, 318,<br />

339, 342, 343, 353, 360, 362, 367,<br />

378, 380, 385, 386, 388, 392, 398,<br />

403, 404, 407, 419, 422, 427, 461, 479]<br />

s<strong>in</strong>ter<strong>in</strong>g [270, 361]<br />

stoichiometric [296, 446, 453, 462]<br />

superlattices [227, 228, 229, 230, 232,<br />

233, 234, 235, 237, 239, 240, 241,<br />

243, 244, 247, 248, 249, 252, 255,<br />

256, 257, 258, 259, 261, 283, 287,<br />

297, 300, 302, 305, 315, 335, 343,<br />

345, 347, 360, 366, 367, 388, 389,<br />

390, 391, 393, 394, 397, 398, 401,<br />

415, 419, 420, 427, , 433, 439, 443,<br />

447, 448, 482]<br />

temperature-dependent [239, 293, 316,<br />

321, 351, 430, 451]<br />

tracer [374, 461, 486, 487]<br />

transients [284, 285, 292, 293, 307, 312,<br />

314, 316, 317, 318, 373, 398, 449,<br />

459, 463]<br />

tunnell<strong>in</strong>g [231, 232, 233, 284, 376, 380,<br />

382, 383, 384, 385, 386, 448]<br />

tw<strong>in</strong>n<strong>in</strong>g [224, 274, 306]<br />

ultra-th<strong>in</strong> [390, 456]<br />

undersaturation [283, 297, 300, 301,<br />

346, 365]<br />

up-hill [282, 285, 317, 318, 339]<br />

vacancies [230, 239, 243, 245, 248,<br />

249, 252, 253, 256, 257, 258, 259,<br />

260, 263, 275, 276, 277, 280, 282,<br />

285, 286, 291, 292, 294,297, 298,<br />

299, 300, 301, 303, 304, 305, 307,<br />

308, 315, 316, 317, 318, 325, 326,<br />

327, 328, 329, 330, 331, 333, 334,<br />

336, 338, 339, 340, 341, 342, 343,<br />

344, 345, 346, 350, 351, 356, 357,<br />

359, 360, 361, 366, 367, 368, 369,<br />

370, 374, 383,385, 388, 389, 390,<br />

392, 393, 394, 396, 397, 398, 399,<br />

404, 406, 407, 410, 414, 415, 419,<br />

421, 424, 428, 429, 430, 437, 451,<br />

460, 470, 471, 475, 478, 479, 480,<br />

481, 486, 487]<br />

wafers [239, 277, 284, 292, 295,<br />

318, 321, 338, 358, 361, 354, 364,<br />

420, 428,464, 465, 471, 473,<br />

476, 479]<br />

X-ray [274, 276, 277, 278, 286, 290,<br />

296, 299, 300, 301, 304, 324, 325,<br />

343, 348, 351, 352, 358, 366, 368,<br />

388, 399, 417, 420, 422, 423, 427,<br />

428, 435, 436, 439, 443, 446, 447,<br />

453, 454, 461, 469, 473, 477, 479,<br />

483, 484, 485, 486]<br />

510

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!