23.07.2014 Views

SRC Users' Meeting - Synchrotron Radiation Center - University of ...

SRC Users' Meeting - Synchrotron Radiation Center - University of ...

SRC Users' Meeting - Synchrotron Radiation Center - University of ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Welcome to the 36 th <strong>SRC</strong> Users’ <strong>Meeting</strong>!<br />

John Joyce<br />

UM2003 Chair<br />

At this meeting you will learn <strong>of</strong> the status <strong>of</strong> the<br />

<strong>SRC</strong>, find out who is the Aladdin Lamp Award<br />

winner, listen to talks given by researchers, and<br />

participate in a poster session (and best poster<br />

contest!).<br />

The <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong> (<strong>SRC</strong>) is a<br />

national facility open to all qualified investigators.<br />

It is operated by the <strong>University</strong> <strong>of</strong> Wisconsin-<br />

Madison with funding primarily from the National<br />

Science Foundation. Use <strong>of</strong> the “Aladdin”<br />

electron storage ring is made available free <strong>of</strong><br />

charge to scientists who perform research<br />

publishable in open literature. Propriety research can be conducted based on<br />

full cost recovery.<br />

<strong>Synchrotron</strong> <strong>Radiation</strong> based research at the <strong>SRC</strong> is performed in a large<br />

variety <strong>of</strong> disciplines such as atomic and molecular physics, surface science,<br />

material science, semiconductor research, chemistry, infra-red spectroscopy,<br />

biology, nanotechnology, photoelectron microscopy, and geology.<br />

This booklet contains a few examples <strong>of</strong> the many research projects<br />

conducted at the <strong>SRC</strong>.<br />

We hope you will enjoy this Users’ <strong>Meeting</strong>!<br />

Acknowledgement<br />

We wish to thank all contributors to this booklet, in particular Users <strong>of</strong> the <strong>SRC</strong> facility.<br />

Financial support for the <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong> from the National Science<br />

Foundation under Award No. DMR-0084402 is gratefully acknowledged.


36 th <strong>SRC</strong> Users’ <strong>Meeting</strong><br />

Saturday, October 25, 2003<br />

(Location <strong>of</strong> <strong>Meeting</strong>: PSL Large Conference Room)<br />

PROGRAM<br />

8:00 - 8:30 Registration, refreshments<br />

SESSION 1 Chair: Hartmut Höchst<br />

8:30 -8:50 Welcome & Status <strong>of</strong> <strong>SRC</strong>, Joe Bisognano, Executive Director<br />

8:50 - 9:10 Mark Bissen, <strong>SRC</strong><br />

“Current and Future Status <strong>of</strong> <strong>SRC</strong> Beam Lines”<br />

9:10 – 9:25 Bob Legg, <strong>SRC</strong><br />

“Operations 2003”<br />

9:25 – 9:40 Ken Jacobs, <strong>SRC</strong><br />

“Accelerator Developments”<br />

9:40 – 10:25 Announcement <strong>of</strong> Aladdin Lamp Award Winner:<br />

Christian Ast, Physics, UW-Madison (now at Max-Planck-Institut)<br />

“Interactions in the Photoemission Process for Materials with<br />

Low Conductivity”<br />

10:25 – 10:40 Break<br />

SESSION 2 Chair: Cliff Olson<br />

10:40 - 11:20 George Sawatzky (invited), <strong>University</strong> <strong>of</strong> British Columbia, Vancouver<br />

“High Energy Research Possibilities with an Extended<br />

Photon Energy Range”<br />

11:20 – 11:35 Juan Carlos Campuzano, U <strong>of</strong> Illinois at Chicago<br />

“A High Resolution, High Photon Energy Capability at Aladdin, the<br />

Instrument and the Science”<br />

11:35 –11:50 Tom Miller, U <strong>of</strong> Illinois at Urbana-Champaign<br />

“Atomically Uniform Thin Films on Silicon”<br />

11:50 –12:05 S. H. Southworth, Argonne National Lab<br />

“Many-Electron Effects On The Xe 5s Nondipole Photoelectron<br />

Asymmetry”<br />

12:05 –12:20 Adam Kaminski, <strong>University</strong> <strong>of</strong> Wales – Swansea<br />

“New features in the phase diagram <strong>of</strong> cuprates”


12:20 - 1:10 Lunch<br />

1:10 –1:45 Posters at Aladdin, Aladdin Tours<br />

SESSION 3 Chair: Tom Miller<br />

1:45 –2:00 Scott Whitfield, U <strong>of</strong> Wisconsin – Eau Claire<br />

“Photoelectron Spectrometry <strong>of</strong> Atomic Chromium in the Region <strong>of</strong> the 3p<br />

to 3d Giant Resonance”<br />

2:00 –2:40 Jim Tobin (invited), Lawrence Livermore National Lab<br />

“Spin-Resolved PES <strong>of</strong> Complex Systems”<br />

2:40 –2:55 Tomasz Durakiewicz, Los Alamos National Lab<br />

“Photoemission Study <strong>of</strong> USb and UTe – 5f Electronic Structure and<br />

Magnetism”<br />

2:55 – 3:10 Cherice Evans, Queens College – CUNY, Flushing, NY<br />

“Field Ionization <strong>of</strong> CH 3 I in Supercritical Ar”<br />

3:10 –3:25 Break<br />

3:25 – 4:05 <strong>Users'</strong> <strong>Meeting</strong> (election, discussion <strong>of</strong> new beam lines, etc.)<br />

SESSION 4 Chair: Dave McIlroy<br />

4:05 –4:25 Astrid Jürgensen, CSRF<br />

“Canadian <strong>Synchrotron</strong> <strong>Radiation</strong> Facility”<br />

4:25 - 4:40 James W. Taylor, <strong>Center</strong> for NanoTechnology (CNTech)<br />

“Future Plans and Recent Accomplishments at CNTech”<br />

4:40 – 4:55 Paul Nealey, CNTech and Chemical Engineering, UW-Madison<br />

“Self-assembly <strong>of</strong> Block Copolymers on Lithographically Defined<br />

Nanopatterned Substrates”<br />

4:55 –5:10 Brad Frazer, UW-Madison & Institute de Physique Appliquée, Lausanne<br />

“Identification <strong>of</strong> Sub-micrometer Silicate Inclusions in Archean Zircons<br />

with X-PEEM”<br />

5:10 – 5:15 Best Poster Contest Winner<br />

6:00 – 7:00 Cocktail Hour at Dry Bean Saloon & Steakhouse - Madison<br />

(hors d’ouevres provided; cash bar)<br />

(directions to the Dry Bean are enclosed in registration packet)<br />

7:00 Dinner at Dry Bean<br />

_______________________________________


O<br />

R<br />

A<br />

D<br />

M<br />

I<br />

N<br />

G<br />

A<br />

N<br />

I<br />

Z<br />

E<br />

R<br />

S


Administration<br />

Executive Director<br />

Dr. Joseph Bisognano<br />

Scientific Directors<br />

Pr<strong>of</strong>essor Juan Carlos Campuzano<br />

Scientific Director for Condensed Matter<br />

Pr<strong>of</strong>essor Gelsomina "pupa" De Stasio<br />

Scientific Director for Multidisciplinary Science<br />

Pr<strong>of</strong>essor James Taylor<br />

Scientific Director for Educational Programs<br />

__________________________________<br />

UM2003 Organizing Committee<br />

Program Chair<br />

John Joyce, Los Alamos National Laboratory<br />

Local Committee<br />

Hartmut Hchst<br />

Pamela Layton<br />

Michelle Kirch<br />

Christopher Moore


T<br />

A<br />

L<br />

K<br />

S


STATUS OF THE <strong>SRC</strong> BEAMLINES AND INSTRUMENTATION<br />

M. Bissen<br />

<strong>SRC</strong>, Univ. <strong>of</strong> Wisconsin, 3731 Schneider Dr., Stoughton, WI-53589<br />

The Instrumentation Group maintains the <strong>SRC</strong> and <strong>SRC</strong>/PRT monochromator beamlines on<br />

Aladdin for use by the scientific community. The technical details <strong>of</strong> these instruments and the<br />

PRT and CnTech beamlines is available on-line at www.src.wisc.edu. Thus, instead <strong>of</strong> repeating<br />

that information, this report will focus on current uses <strong>of</strong> the existing beamline and progress for<br />

the new beamlines currently under development. Also, the Group operates a variety <strong>of</strong><br />

experimental endstations that are available to the User community.<br />

Initial operation <strong>of</strong> the Wadsworth beamline on the U2 undulator is planned for Fall 2003.<br />

The beamline will cover the range from 8 to 50 eV utilizing the high brightness undulator source<br />

for ARPES experiments.<br />

An additional beamline from U2, the VLS-PGM, is being constructed for PEEM experiments<br />

requiring high brightness. The proposed variable line spacing monochromator will cover the<br />

range from 60 to 2000 eV with a resolving power <strong>of</strong> nearly 1000. For this mode <strong>of</strong> operation, the<br />

U2 undulator will be operated as a wiggler at maximum k value (4.69). Commissioning is<br />

currently scheduled for summer 2004.<br />

An update on the NIM classic maintenance and a progress report on the PGM repairs will be<br />

presented as well as future plans for upgrades <strong>of</strong> some <strong>of</strong> the facility experimental chambers.<br />

In addition, results <strong>of</strong> the recent U1 NIM and Scienta SES 2002 will be shown.<br />

This Users’ <strong>Meeting</strong> is supported by the National Science Foundation, Grant No. DMR-<br />

0084402.


<strong>SRC</strong> OPERATIONS<br />

Bob Legg<br />

<strong>SRC</strong> provides beam in scheduled 3-week experimental quantums. There is a scheduled<br />

one-week “Development Week” for maintenance after every two quantums. Each week <strong>of</strong> an<br />

experimental quantum is scheduled from 8 AM Monday until 8 AM Saturday; five days, 24<br />

hours a day. Saturdays are used for holiday make-up and to provide users with additional beam<br />

time. Each day during the week has four injections, 8 AM, noon, 6 PM and midnight. The noon<br />

to 8 AM period is considered scheduled beam for users. <strong>SRC</strong> over the year ending 10/1/2003<br />

provided the normal 4271 hours <strong>of</strong> scheduled beam with a 98% reliability. The hours <strong>of</strong><br />

operation can be broken down into 3636 hours <strong>of</strong> 800 MeV beam and 634 hours <strong>of</strong> 1 GeV beam.<br />

Saturdays and morning beams accounted for an additional 887 hours <strong>of</strong> user beam. The major<br />

source <strong>of</strong> downtime for the year was intermittent equipment failure.<br />

Operations have continued to be dynamic. The group changed this year with the<br />

departure <strong>of</strong> swing shift operator, Dan Granger, and his replacement by Craig Trewartha. As a<br />

group we worked to complete several projects left from previous years (vacion readbacks,<br />

Diagnostic front ends, U3 Control interface, etc.). There was continued effort in the installation<br />

<strong>of</strong> the user utilities on the Wadsworth beamline. We also assigned a specific electronic<br />

technician to become a schematic draftsman, specifically responsible for AutoCAD support <strong>of</strong><br />

the <strong>SRC</strong> electrical drawings. This was done to convert some <strong>of</strong> the original hand sketches <strong>of</strong><br />

Aladdin systems to appropriate, dated, drawings.<br />

Operations are also working to improve operator training. With the advent <strong>of</strong> the<br />

automated “One Button” setups for Aladdin, many <strong>of</strong> the operators don’t get time doing machine<br />

setup and control during normal user operations. To maintain their ability to deal with the <strong>of</strong>f<br />

normal situations, Operator training has been instituted during the development periods. By<br />

improving Operator capabilities and addressing weaknesses in the installed machine, operations<br />

will continue to make the machine more reliable in the year to come.


ACCELERATOR DEVELOPMENTS<br />

Ken Jacobs<br />

<strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong>, <strong>University</strong> <strong>of</strong> Wisconsin - Madison<br />

3731 Schneider Drive, Stoughton, WI 53589-3097<br />

Work has continued on a variety <strong>of</strong> accelerator developments at the <strong>SRC</strong> over the<br />

past year. The low emittance operating mode “LF15” is now routinely run for all<br />

800 MeV beams. This gives a factor <strong>of</strong> three reduction in horizontal beam size, with<br />

unchanged vertical beam size and beam lifetime. The exception to routine LF15 running<br />

is when vacuum work is done in the ring, after which Base Lattice is run for the first few<br />

weeks.<br />

LF15 cannot be run at 1 GeV due to power supply, cabling, and magnet<br />

limitations. For reduced emittance at 1 GeV, an alternate lattice with emittance between<br />

LF15 and Base Lattice has been designed. Operation will require upgraded cables on<br />

some <strong>of</strong> the quadrupoles, and this is presently being pursued. However, LF15 can be run<br />

at 950 MeV, requiring only the cable upgrades. Tests <strong>of</strong> 950 MeV LF15 have been<br />

encouraging. We are also investigating other low emittance 950 MeV lattices.<br />

We continued to make improvements made in beam stability. An additional<br />

feedback loop was installed in the main RF system to reduce coherent synchrotron<br />

oscillations <strong>of</strong> the beam. This, combined with a general reduction <strong>of</strong> 60 Hz harmonics on<br />

the beam, has reduced noise levels on the infrared beamlines so that they can now be<br />

used when running low emittance beam. Undulator compensation has progressed to the<br />

point that U1 and U3 can now be scanned over their full ranges in LF15. Compensation<br />

work is continuing with U2, and in Base Lattice.<br />

In an effort to improve beam lifetime, we are continuing studies to more fully<br />

understand beam loss mechanisms. To increase lifetime, it may be necessary to open up<br />

some limiting apertures in the ring.<br />

Other projects completed over the past year include upgrading the accelerator<br />

control system CPUs, coating the injection kicker ceramics to reduce wakefields and<br />

radiated RF, and installation <strong>of</strong> a modern scripting language on the control system.<br />

Scripts are used to simplify accelerator operations, implement control algorithms to<br />

improve beam stability, and perform experiments during accelerator development<br />

periods.


INTERACTIONS IN THE PHOTOEMISSION PROCESS OF<br />

MATERIALS WITH LOW CONDUCTIVITY<br />

Christian R. Ast<br />

Max-Planck-Institut für Festkörperforschung, Stuttgart, Germany<br />

Scattering <strong>of</strong> the photoelectron is an omnipresent phenomenon in photoemission<br />

spectra. In most cases inelastic scattering manifests itself in either discrete loss structures,<br />

a featureless background or an asymmetric line shape. Two examples are presented in<br />

which increased interactions <strong>of</strong> the photoelectron with its surroundings lead to additional<br />

emission features in the photoemission spectra <strong>of</strong> the semimetal bismuth. Scattering<br />

events can be seen in core level spectra as structured loss features from interband<br />

transitions. In valence band spectra, scattering processes can contribute additional<br />

structures from secondary cone emission.<br />

Spectra <strong>of</strong> the Bi 5d core level show an additional peak split <strong>of</strong>f in energy<br />

between 160 and 260meV, which can be related to momentum dependent energy losses<br />

suffered by the photoelectron from interband transitions. Similar phenomena have been<br />

observed in the 4d core levels <strong>of</strong> Sb, which leads to the conclusion that loss features are<br />

enhanced in materials with low conductivity and low charge carrier concentration.<br />

Valence band spectra <strong>of</strong> Bi(111) at normal emission as a function <strong>of</strong> photon energy show<br />

multiple emission features from the same initial state band. These emission features can<br />

be attributed to primary and secondary cone emission. To be observed in the normal<br />

emission spectra photoelectrons from secondary cone emission have to be scattered back<br />

into the normal emission direction. Accounting for the secondary cone emission in the<br />

analysis <strong>of</strong> the final state dispersion, it is than possible to experimentally determine the<br />

deviations <strong>of</strong> the final states from the free electron band structure. With these<br />

experimentally determined final state bands on hand, it is than possible to extract the<br />

initial state band structure E i (k).


HIGH ENERGY RESEARCH POSSIBILITIES WITH AN<br />

EXTENDED PHOTON ENERGY RANGE<br />

George Sawatzky<br />

Department <strong>of</strong> Physics, <strong>University</strong> <strong>of</strong> British Columbia, Vancouver<br />

No abstract provided.


A HIGH RESOLUTION, HIGH PHOTON ENERGY CAPABILITY<br />

AT ALADDIN, THE INSTRUMENT AND THE SCIENCE<br />

J. C. Campuzano<br />

<strong>University</strong> <strong>of</strong> Illinois at Chicago, Department <strong>of</strong> Physics,<br />

845 W. Taylor, M/C 273, Chicago, IL 60607<br />

No abstract provided.


ATOMICALLY UNIFORM THIN FILMS ON SILICON<br />

M. Upton, T. Miller, and T.-C. Chiang<br />

Department <strong>of</strong> Physics, <strong>University</strong> <strong>of</strong> Illinois at Urbana-Champaign, 1110 West Green Street,<br />

Urbana, Illinois 61801-3080<br />

Frederick Seitz Materials Research Laboratory, <strong>University</strong> <strong>of</strong> Illinois at Urbana-Champaign,<br />

104 South Goodwin Avenue, Urbana, Illinois 61801-2902<br />

The continued miniaturization <strong>of</strong> silicon-based electronic devices is pushing component<br />

layer thicknesses toward the nanoscale, and a critical hurdle along this path is atomistic<br />

fluctuation. An uncertainty in layer thickness or roughness at the monolayer level can lead to<br />

substantial and unacceptable property variations. The effect is generally on the order <strong>of</strong> 1/N,<br />

where N is the thickness <strong>of</strong> the film in terms <strong>of</strong> monolayers (ML). At the nanoscale, 1/N can be<br />

as large as 20%, and for a typical valence band width <strong>of</strong> ~10 eV, the resulting changes in the<br />

electronic structure can be severe or even catastrophic for device performance. Exact control <strong>of</strong><br />

layer thickness and atomic uniformity is thus a critical issue, but so far has not been achieved for<br />

films grown on silicon. Yet this is extremely important in view <strong>of</strong> the vast technological and<br />

manufacturing base for the silicon industry. In this work, we show that Pb deposition on Si(111)<br />

at 100 K can lead to atomically uniform thin films provided that the Si substrate is pretreated<br />

appropriately. The resulting films support quantum well states [1-3] due to confinement <strong>of</strong><br />

electrons in the Pb films by the band gap in the Si substrate. A measurement <strong>of</strong> such states by<br />

angle-resolved photoemission determines the film<br />

thickness and reveals a quantum electronic structure<br />

that varies substantially as the film thickness<br />

undergoes monolayer changes. The result is a rather<br />

dramatic contrast between films consisting <strong>of</strong> even<br />

and odd numbers <strong>of</strong> monolayers. These findings<br />

illustrate the importance <strong>of</strong> atomic layer precision for<br />

controlling the electronic structure <strong>of</strong> thin films.<br />

Our angle-resolved photoemission measurements<br />

were performed at the <strong>Synchrotron</strong> <strong>Radiation</strong><br />

<strong>Center</strong>, <strong>University</strong> <strong>of</strong> Wisconsin-Madison, using a<br />

Scienta analyzer equipped with a two-dimensional<br />

detector. The Si(111) substrates were prepared from<br />

commercial n-type wafers with a resistivity <strong>of</strong> 1-60<br />

Figure 1: The three-dimensional plot<br />

above shows the photoemission<br />

intensity as a function <strong>of</strong> Pb film<br />

thickness and binding energy<br />

relative to the Fermi level. Three<br />

major quantum well peaks are seen<br />

at thicknesses N = 5, 7, and 9. A<br />

weak resonance peak at higher<br />

binding energy is seen at N = 8.<br />

ohm-cm. The Pb films were prepared by sequential,<br />

incremental deposition at 100 K, and the<br />

photoemission spectra were taken with the sample at<br />

the same temperature. The three-dimensional color<br />

plot in Fig. 1 shows the normal-emission intensity as<br />

a function <strong>of</strong> binding energy and film thickness.<br />

Three major peaks, at binding energies <strong>of</strong> 0.40, 0.26,<br />

and 0.15 eV below the Fermi level, attain their<br />

maximum intensities at film thicknesses N = 5, 7, and<br />

9, respectively, while no such peaks are observed at


the even layer thicknesses N = 6 and 8 in the same energy range. These peaks represent quantum<br />

well states formed by confinement <strong>of</strong> the Pb p-band electrons by the Si band gap. No such<br />

quantum well states exist for N = 6 and 8, as will be explained below.<br />

The appearance <strong>of</strong> intense peaks for odd N only has to do with the specific band structure<br />

<strong>of</strong> Pb and the Si band gap. From data taken over a wide thickness range, it is concluded that the<br />

Fermi level <strong>of</strong> the Pb is at 0.5 eV above the Si valence band edge. Electrons in the Pb film with<br />

binding energies within 0.5 eV <strong>of</strong> the Fermi level are thus confined by the Si band gap, giving<br />

rise to sharp and intense quantum well peaks. Electrons at higher binding energies are not<br />

confined. Nevertheless, partial reflection at the Pb-Si boundary can give rise to resonances which<br />

appear in photoemission as weak and broad peaks [3]. One such weak resonance peak at a<br />

binding energy <strong>of</strong> 0.63 eV can be seen in the three-dimensional plot <strong>of</strong> Fig. 1 at N = 8. Further<br />

evidence for the confinement edge is seen in the line scan <strong>of</strong> the peak at N = 5 in Fig. 1. This<br />

peak, with a binding energy very close to the confinement edge, is asymmetric. Its higherbinding-energy<br />

side is substantially broader because this portion <strong>of</strong> the spectral weight lies<br />

outside the confinement range. From the confinement range <strong>of</strong> 0.5 eV and the known Si band<br />

gap <strong>of</strong> 1.2 eV, a Schottky barrier height <strong>of</strong> 0.7 eV is deduced for our n-type Si substrate. This is<br />

consistent with a recent reported value <strong>of</strong> 0.72 0.02 eV based on electrical measurements [4].<br />

The binding energies <strong>of</strong> quantum well states are determined by the Bohr-Sommerfeld<br />

quantization rule [2, 3, 5]:<br />

2kNt<br />

s<br />

i<br />

2n<br />

, (1)<br />

6 where k is the wave vector, t is the monolayer<br />

0<br />

thickness, <br />

s<br />

is the phase shift at the surface, i<br />

is<br />

5<br />

the phase shift at the interface, and n is a quantum<br />

1<br />

number. For a given N and integer quantum<br />

4 numbers, this equation determines the allowed k<br />

values, which in turn determine the binding energies<br />

2<br />

3 <strong>of</strong> the quantum well states through the band<br />

dispersion relation E(k). For Pb(111), the relevant<br />

3<br />

2 band is the p valence band with a known dispersion<br />

that extends from 4.2 eV below the Fermi level to<br />

Binding Energy (eV)<br />

4 n = 0<br />

0 2 4 6 8 10 12<br />

Thickness N (ML)<br />

Figure 2: Binding energies <strong>of</strong> quantum<br />

well states (circles connected by lines)<br />

deduced from a fit to the experimental<br />

results (crosses). The band structure <strong>of</strong><br />

Pb and the phase shifts enter the model<br />

calculation. The only fitting parameter<br />

is an unknown constant <strong>of</strong>fset in the<br />

interface phase shift. The quantum<br />

number n for each branch is indicated.<br />

1<br />

8.0 eV above the Fermi level [5]. The surface phase<br />

shift <br />

s<br />

as a function <strong>of</strong> energy has been computed<br />

by a first-principles method [5], and this is used in<br />

the present analysis. The interface phase shift i<br />

is<br />

given by<br />

<br />

<br />

1<br />

E E <br />

L<br />

Re<br />

cos<br />

2<br />

1<br />

<br />

0<br />

, (2)<br />

EU<br />

EL<br />

<br />

where Re refers to the real part, E is the energy, EL<br />

is the lower edge <strong>of</strong> the Si band gap, E<br />

U<br />

is the upper<br />

edge, and 0<br />

is a constant [6, 7]. It is easy to verify<br />

that <br />

i<br />

changes by across the Si band gap. The<br />

only unknown quantity in Eq. (1) is the constant <br />

0<br />

.<br />

This is treated as a fitting parameter in a fit <strong>of</strong> the


calculated binding energies <strong>of</strong> the quantum well states to the observed values. The calculated<br />

binding energies <strong>of</strong> the quantum well state from the fit are shown as open circles connected by<br />

lines in Fig. 2, and the quantum number n for each branch is indicated. The results agree well<br />

with the experimental values shown as crosses. The experimental value for the quantum well<br />

state at N = 11 is not included in the figure because the peak is too close to the Fermi level for a<br />

precise determination <strong>of</strong> its position.<br />

The confinement edge at 0.5 eV is indicated by a horizontal dashed line in Fig. 2. In<br />

agreement with experimental observations, no quantum well states exist, or only weak<br />

resonances are observed, for even N = 4, 6, 8, and 10, while intense quantum-well states are<br />

observed for odd N = 5, 7, and 9. This difference between even- and odd-N configurations is just<br />

a consequence <strong>of</strong> the way that the quantum well states evolve as a function <strong>of</strong> thickness as seen<br />

in Fig. 2. This evolution is largely determined by the band structure [3, 5, 8, 9]. It happens that<br />

one half <strong>of</strong> the Fermi wave length in Pb is approximately two atomic layers, and a simple<br />

analysis based on Eq. (1) shows that a new quantum well state drops below the Fermi level for a<br />

film thickness increment <strong>of</strong> about two atomic layers [3, 5]. The corresponding changes in the<br />

occupied density <strong>of</strong> states should give rise to an oscillation period <strong>of</strong> about two atomic layers for<br />

essentially all physical properties <strong>of</strong> the films. Also in agreement with our experimental<br />

observation is that no quantum well states are predicted for N = 1-3.<br />

Numerous groups have experimented with Pb growth on Si [8-13], but none has achieved<br />

atomic uniformity. Prior to this experiment, we had performed extensive experimentation with<br />

direct deposition on the (7x7) reconstructed Si(111) at various temperatures, but the resulting<br />

films were inevitably rough. In the present experiment, the atomic-layer uniformity is achieved<br />

by first preparing either the or 3 3<br />

reconstructions [12, 14, 15]. Our experiment<br />

shows that deposition <strong>of</strong> Pb on the phase, the phase, or any intermediate phases at a low<br />

temperature (100 K) leads to atomically uniform films. The final results are the same for the<br />

same total amounts <strong>of</strong> Pb deposition including the initial Pb coverages. The determination <strong>of</strong><br />

film thicknesses is described in detail in [1].<br />

The stability <strong>of</strong> the films has been measured, and some interesting features were<br />

observed, including the bilayer oscillation predicted by Wei et al. [17].<br />

Why does Pb pretreatment promote uniform film formation, while direct deposition on<br />

Si(111)-(7x7) does not work? A possible explanation is that the (7x7) surface, with its<br />

complicated reconstruction involving adatoms, dimers, corner holes, and partial stacking fault, is<br />

not smooth on the atomic scale. These structural features can pin the Pb growth locally at low<br />

temperatures, resulting in small crystallites that are structurally incoherent. Increasing the<br />

substrate temperature to promote long-range diffusion and structural coherence leads to<br />

formation <strong>of</strong> islands instead <strong>of</strong> smooth films due to electronic effects [10, 16]. Pretreatment <strong>of</strong> Si<br />

by Pb leads to a smooth bulk-terminated Si substrate with a well ordered Pb overlayer. This can<br />

be a good template for smooth growth upon further deposition at low temperatures. Our results<br />

illustrate an important issue in film growth – the initial surface structure can be a deciding factor<br />

for the morphological development <strong>of</strong> films. As shown in this study, proper conditioning <strong>of</strong> the<br />

starting surface allows us to make uniform films on Si, a result <strong>of</strong> potential interest and<br />

importance for nano and quantum electronics.<br />

This work is supported by the U.S. National Science Foundation (grant DMR-02-03003).<br />

We acknowledge the Petroleum Research Fund, administered by the American Chemical<br />

Society, and the U.S. Department <strong>of</strong> Energy, Division <strong>of</strong> Materials Sciences (grant DEFG02-<br />

91ER45439), for partial support <strong>of</strong> the synchrotron beamline operation and the central facilities


<strong>of</strong> the Frederick Seitz Materials Research Laboratory. The <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong> <strong>of</strong> the<br />

<strong>University</strong> <strong>of</strong> Wisconsin-Madison is supported by the U.S. National Science Foundation (grant<br />

DMR-00-84402).<br />

References:<br />

[1] M. Upton, T. Miller, and T.-C Chiang, Phys. Rev. B, in press.<br />

[2] J. J. Paggel, T. Miller, and T.-C. Chiang, Phys. Rev. Lett. 81, 5632 (1998); Science 283,<br />

1709 (1999).<br />

[3] T.-C. Chiang, Surf. Sci. Rep. 39, 181 (2000), and references therein.<br />

[4] A Schottky barrier height <strong>of</strong> 0.93 eV was reported by D. R. Heslinga, H. H. Weitering, D. P.<br />

van der Werf, T. M. Klapwijk, and T. Hibma, Phys. Rev. Lett. 64, 1589 (1990). In a more<br />

recent study by R. F. Schmitsdorf and W. Mönch, Eur. Phys. J. B 7, 457 (1999), a value <strong>of</strong><br />

0.72 0.02 eV was reported.<br />

[5] C. M. Wei and M. Y. Chou, Phys. Rev. B 66, 233408 (2002).<br />

[6] D.-A. Luh, T. Miller, J. J. Paggel, and T.-C. Chiang, Phys. Rev. Lett. 88, 256802 (2002).<br />

[7] N. V. Smith, Phys. Rev. B 32, 3549 (1985).<br />

[8] A. Mans, J. H. Dil, A. R. H. F. Etterma, and H. H. Weitering, Phys. Rev. B 66, 195410<br />

(2002).<br />

[9] M. Jalochowski, H. Knoppe, G. Lilienkamp, and E. Bauer, Phys. Rev. B 46, 4693 (1992);<br />

M. Jalochowski, M. H<strong>of</strong>fmann, and E. Bauer, Phys. Rev. B 51, 7231 (1995); Th. Schmidt<br />

and E. Bauer, Surf. Sci. 480, 137 (2001).<br />

[10] M. Hupalo and M. C. Tringides, Phys. Rev. B 65, 115406 (2002). M. Hupalo, V. Yeh, L.<br />

Berbil-Bautista, S. Kremmer, E. Abram, and M. C. Tringides, Phys. Rev. B 64, 155307<br />

(2001).<br />

[11] W. B. Su, S. H. Chang, W. B. Jian, C. S. Chang, L. J. Chen, and T. T. Tsong, Phys. Rev.<br />

Lett. 86, 5116 (2001).<br />

[12] I.-S. Hwang, R. E. Martinez, C. Liu, and J. A. Golovchenko, Surf. Sci. 323, 241 (1995).<br />

[13] I. B. Altfeder, K. A. Matveev, and D. M. Chen, Phys. Rev. Lett. 78, 2815 (1997).<br />

[14] J. A. Carlisle, T. Miller, and T.-C. Chiang, Phys. Rev. B 45, 3400 (1992).<br />

[15] P. J. Estrup and J. Morrison, Surf. Sci. 2, 465 (1964).<br />

[16] Hawoong Hong, C. M. Wei, M. Y. Chou, Z. Wu, L. Basile, H. Chen, M. Holt, and T.-C.<br />

Chiang, Phys. Rev. Lett. 90, 076104 (2003).<br />

[17] C. M. Wei, and M. Y. Chou, Phys. Rev. B 66, 233408 (2002).


MANY-ELECTRON EFFECTS ON THE XE 5S NONDIPOLE<br />

PHOTOELECTRON ASYMMETRY<br />

S. H. Southworth 1 , E. P. Kanter 1 , B. Krässig 1 , R. Guillemin 2 , O. Hemmers 2 ,<br />

D. W. Lindle 2 , and R. Wehlitz 3<br />

1 Argonne National Laboratory, Argonne, IL 60439*<br />

2 <strong>University</strong> <strong>of</strong> Nevada, Las Vegas, NV 89154<br />

3 <strong>SRC</strong>, <strong>University</strong> <strong>of</strong> Wisconsin, Stoughton, WI 53589<br />

Photoionization <strong>of</strong> the Xe 5s subshell has been extensively studied because <strong>of</strong> its<br />

sensitivity to relativistic and many-electron interactions. Previous studies <strong>of</strong> the partial cross<br />

section and the photoelectron anisotropy parameter are sensitive to the electric-dipole<br />

photoionization amplitudes. Recently, the nondipole asymmetry parameter has been calculated<br />

and is sensitive to both electric-dipole and electric-quadrupole photoexcitation channels [1,2].<br />

We measured over the 26–140 eV photon energy range at the <strong>SRC</strong> and combined our results<br />

with measurements made over 80–197.5 eV at the ALS. Measurements over the 90–225 eV<br />

region were also reported in Ref [3]. In<br />

Asymmetry parameter γ<br />

1<br />

0.5<br />

0<br />

-0.5<br />

HF<br />

Xe 5s<br />

RRPA<br />

RPAE<br />

-1<br />

50 100 150 200<br />

Photon energy (eV)<br />

Fig. 1. Xe 5s nondipole asymmetry parameters measured at<br />

the <strong>SRC</strong> (open circles) and ALS (closed circles) compared<br />

with HF (dotted curve), RPAE (dashed curve), and RRPA<br />

(solid curve) calculations.<br />

Fig. 1, the <strong>SRC</strong> and ALS results are<br />

compared with Hartree-Fock (HF) [1],<br />

random-phase approximation with<br />

exchange (RPAE) [1], and relativistic<br />

random-phase approximation (RRPA)<br />

[2] calculations. (The curves plotted in<br />

Fig. 1 are based on more accurate<br />

calculations than were originally<br />

reported in Refs. [1,2]; see Ref. [4].)<br />

The parameter varies rapidly near<br />

threshold and passes through a<br />

minimum near 35 eV. This is due to<br />

the 5s p (dipole) amplitude passing<br />

through a "Cooper minimum," and is<br />

essentially a one-electron effect but is<br />

modified by interchannel coupling with<br />

the 5p and 4d subshells and by ionic-state satellite channels. The broad maximum near 150 eV<br />

results mainly from interchannel coupling with the 4d subshell, which is included in the RPAE<br />

and RRPA calculations but not HF. The small feature measured near 160 eV and predicted by<br />

the RRPA calculation results from coupling <strong>of</strong> the 5s d (quadrupole) channels with the 4p <br />

f quadrupole shape resonances. A full report on this study is in Ref. [4].<br />

[1] M. Ya. Amusia et al., Phys. Rev. A 63, 052506 (2001).<br />

[2] W. R. Johnson and K. T. Cheng, Phys. Rev. A 63, 022504 (2001).<br />

[3] S. Ricz et al., Phys. Rev. A 67, 012712 (2003).<br />

[4] O. Hemmers et al., Phys. Rev. Lett. 91, 053002 (2003).<br />

*The ANL group is supported by the Chemical Sciences, Geosciences, and Biosciences Division, Office <strong>of</strong> Basic<br />

Energy Sciences, Office <strong>of</strong> Science, U.S. Dept. <strong>of</strong> Energy, under Contract No. W-31-109-Eng-38.


NEW FEATURES IN THE PHASE DIAGRAM OF CUPRATES<br />

Adam Kaminski<br />

Department <strong>of</strong> Physics, <strong>University</strong> <strong>of</strong> Wales Swansea, United Kingdom<br />

One <strong>of</strong> the most intriguing properties <strong>of</strong> the cuprates is their generic phase<br />

diagram, where high temperature superconductivity occurs in the<br />

proximity <strong>of</strong> a metal-insulator transition. The understanding <strong>of</strong> the various<br />

phases and their connections is deemed to be the key to solving the<br />

problem <strong>of</strong> high temperature superconductivity. We have used Angle Resolved<br />

Photo Emission Spectroscopy (ARPES) to address this problem. By studying<br />

the temperature dependence <strong>of</strong> the bilayer splitting we have found evidence <strong>of</strong><br />

a cross over between coherent and incoherent electron behavior - long anticipated<br />

by a number <strong>of</strong> theoretical models. We have also shown that a<br />

pseudogap exists on the overdoped side <strong>of</strong> the phase diagram below T c , when<br />

the superconductivity is suppressed by application <strong>of</strong> a sufficiently high<br />

current density. However, above certain doping, the pseudogap is<br />

no longer observed under the superconducting dome. This may indicate existence<br />

<strong>of</strong> a Quantum Critical Point (QCP) at a doping level between 0.2 and 0.25 holes per Cu atom.<br />

Existence <strong>of</strong> such point could mean close relation between cuprates and<br />

in heavy fermion compounds, where QCP is usually surrounded by dome <strong>of</strong> superconductivity.<br />

These results are instrumental in choosing a viable theoretical model <strong>of</strong> high temperature<br />

superconductivity.


PHOTOELECTRON SPECTROMETRY OF ATOMIC CHROMIUM IN<br />

THE REGION OF THE 3P TO 3D GIANT RESONANCE<br />

Scott B. Whitfield 1 , Brian Krosschell 1 , and Ralf Wehlitz 2<br />

1 Department <strong>of</strong> Physics and Astronomy, <strong>University</strong> <strong>of</strong> Wisconsin, Eau Claire, WI 54701<br />

2 <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong>, <strong>University</strong> <strong>of</strong> Wisconsin, Stoughton, WI 53589<br />

Atomic Cr, by virtue <strong>of</strong> its half filled 3d and 4s subshells, [Ar]3d 5 4s 1 ( 7 S 3 ), is one <strong>of</strong> the<br />

simplest open-shell atoms to have a partially filled d subshell. This makes detailed studies <strong>of</strong><br />

atomic Cr particularly attractive both theoretically and experimentally. In fact, atomic Cr has<br />

been the object <strong>of</strong> fairly extensive experimental and theoretical investigations [1]. Nevertheless,<br />

there is presently no experimental data <strong>of</strong> the angular distribution parameters, , <strong>of</strong> the Cr<br />

mainlines. Only recently has there been any theoretical calculations <strong>of</strong> [2] and this for the 3d<br />

photoelectron only. This lack <strong>of</strong> experimental data for prompted us to investigate its photon<br />

energy dependence for both the 3d and 4s mainlines and the strongest photoelectron satellite line<br />

in the region <strong>of</strong> the 3p 3d giant resonance. A comparison between our measured values as a<br />

function <strong>of</strong> photon energy for the 3d mainline and theory is shown in the figure below. There is<br />

generally good accord between experiment and theory with the exception <strong>of</strong> certain resonance<br />

transitions which were not accounted for by theory.<br />

We have also found strong<br />

deviations from = 2.0 for the 4s<br />

mainline on most <strong>of</strong> the 3p 3d<br />

resonances, while <strong>of</strong>f resonance they have<br />

values <strong>of</strong> 2.0 as generally expected. For<br />

half-open shubshell atoms one generally<br />

expects s-subshell photoelectrons to have<br />

= 2.0 independent <strong>of</strong> the incident<br />

photon energy. Deviations from this<br />

behavior are an indication <strong>of</strong> relativistic<br />

effects. We will present a simple<br />

qualitative argument for why we observe<br />

deviations <strong>of</strong> = 2.0 for the 4s mainline<br />

on resonance.<br />

This work was supported by a Research Corporation College Cottrell Grant No. CC5243.<br />

The <strong>SRC</strong> is operated under Grant No. DMR-0084402.<br />

References:<br />

[1] Th. Dohrmann, A. von dem Borne, A. Verweyen, B. Sonntag, M. Wedowski, K. Godehusen,<br />

P. Zimmermann and V. Dolmatov. J. Phys. B, 29, 4641 (1996) and references therein.<br />

[2] V. K. Dolmatov, A. S. Baltenkov, and S. T. Manson, Phys. Rev. A, 67, 062714 (2003).


SPIN-RESOLVED PES OF COMPLEX SYSTEMS<br />

J. G. Tobin<br />

Lawrence Livermore National Laboratory<br />

Livermore, CA, USA<br />

Several examples <strong>of</strong> the utilization <strong>of</strong> spin-resolved photoelectron spectroscopy to<br />

investigate complex systems will be discussed. This will include the application <strong>of</strong> SR-PES to<br />

possible half-metallic ferromagnets at the Advanced Light Source at Lawrence Berkeley<br />

National Laboratory [1] and the potential <strong>of</strong> measurements at higher energies, as illustrated by<br />

the first spin-resolved photoelectron spectroscopy results from the Advanced Photon Source at<br />

Argonne National Laboratory [2]. The utility <strong>of</strong> the application <strong>of</strong> SR-PES to non-magnetic<br />

systems will also be discussed [3,4], including the investigation <strong>of</strong> surface states [5] and future<br />

potential studies <strong>of</strong> the correlated electron system -Pu. [6,7]<br />

1. S.A. Morton, G.D. Waddill, S. Kim, I.K. Schuller, S.A. Chambers, and J.G. Tobin,<br />

“Spin-Resolved Photoelectron Spectroscopy <strong>of</strong> Fe3O4,” Surface Science Letters 513,<br />

L451(2002).<br />

2. G.D. Waddill, T. Komesu, and S.A. Morton and J.G. Tobin, http://www-cms.llnl.gov/st/aps.html.<br />

3. Ch. Roth et al, Phys. Rev. Lett. 73, 1963 (1994).<br />

4. K. Starke et al, Phys. Rev. B 53, 10544 (1996).<br />

5. M. Hochstrasser, J.G. Tobin, E. Rotenberg and S.D. Kevan, “Spin-Resolved<br />

Photoemission <strong>of</strong> Surface States <strong>of</strong> W(110)-(1X1)H,” Phys. Rev. Lett. 89, 216802<br />

(2002).<br />

6. K.T. Moore, M.A. Wall, A.J. Schwartz, B.W. Chung, D.K. Shuh, R.K. Schulze, and J.G.<br />

Tobin, “The Failure <strong>of</strong> Russell-Saunders Coupling in the 5f States <strong>of</strong> Plutonium”, Phys.<br />

Rev. Lett. 90, 196404 (2003).<br />

7. J.G. Tobin, B.W. Chung, R. K. Schulze, J. Terry, J. D. Farr, D. K. Shuh, K. Heinzelman,<br />

E. Rotenberg, G.D. Waddill, and G. Van der Laan, “Resonant Photoemission in f-<br />

electron Systems: Pu and Gd", Phys. Rev. B 68, 1151XX (2003).


PHOTOEMISSION STUDY OF USb AND UTe – 5f ELECTRONIC<br />

STRUCTURE AND MAGNETISM<br />

T. Durakiewicz 1, 2 , G. Lander 1 , C.G. Olson 3 , J.J. Joyce 1 , M. T. Butterfield 1 , E. Guziewicz 1, 4 ,<br />

A.J. Arko 1 , J. L. Sarrao 1 , F. Wastin 4 , J. Rebizant 4 , K. Mattenberger 5 , and O. Vogt 5<br />

1 Los Alamos National Laboratory, Los Alamos NM87545, USA; 2 Institute <strong>of</strong> Physics, UMCS<br />

Lublin, 20-031 Poland; 3 Ames Laboratory, Iowa State <strong>University</strong>, Ames IA, USA; 4 Institute <strong>of</strong><br />

Physics, Polish Academy <strong>of</strong> Sciences, Warsaw, Poland; 4 European Commission, JRC, Institute<br />

<strong>of</strong> Transuranium Elements, Postfach 2340, D-76125 Karlsruhe, Germany; 5 Laboratorium fur<br />

Festkorperphysik, ETH, CH-8093 Zurich, Switzerland.<br />

USb and UTe single crystal compounds were investigated by angle-resolved<br />

photoelectron spectroscopy. Measurements were performed at the <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong>,<br />

Stoughton, WI, using the plane grating monochromator line, with an energy resolution <strong>of</strong> 20meV<br />

at 20eV light and an angular resolution <strong>of</strong> 1 degree.<br />

Uranium monoantimonide is an antiferromagnet with T N =214K. USb is an unusual<br />

monopnicitde, having a 3k magnetic ordering structure <strong>of</strong> type I, with the moments aligned<br />

towards the (111) direction [1]. Within the first eV below E F there are at least three well resolved<br />

peaks, A, B and C, positioned at approximately 55meV, 210meV and 610meV below E F ,<br />

respectively. In previous work [2] peaks A and B were not resolved, due to lower energy<br />

resolution (50meV). In our ARPES study one may notice the following: (i) all three peaks are<br />

dispersive, with about 20meV dispersion for peak A, 50meV for peak B and about 100meV for<br />

peak C, (ii) the maximum photoemission intensity is seen around the X point, (iii) none <strong>of</strong> the<br />

peaks represent a band that crosses E F .<br />

The monotelluride UTe is a ferromagnet with T C =104K <strong>of</strong> a semi-metallic character. The<br />

-X ARPES reveals a structure within the first 1eV below E F composed <strong>of</strong> two bands, where the<br />

broader band B might be a superposition <strong>of</strong> more than one feature. Both bands seem to be<br />

relatively flat. Only a minor change in intensity is seen around the X point. Major changes in<br />

photoemission intensity are seen around the FM transition, where band A crosses EF at the<br />

transition. Since the FWHM <strong>of</strong> peak A is <strong>of</strong> the order <strong>of</strong> 200 meV, and it’s BE at low<br />

temperatures is about -55 meV, one cannot state that an actual gap is formed below the<br />

transition, as suggested in [3].<br />

When compared with Np and Pu monoantimonides and monochalcogenides, a direct<br />

correlation between the binding energy <strong>of</strong> the peak bearing most <strong>of</strong> the 5f weight in the<br />

photoemission spectrum, magnetic moment and transition temperature may be seen within the<br />

series. Existence <strong>of</strong> such a correlation indicates the central role <strong>of</strong> 5f electrons in establishing the<br />

magnetic properties <strong>of</strong> these materials. Hybridization <strong>of</strong> the 5f electrons with the conduction<br />

band is found within the series and the level <strong>of</strong> localization is shown to increase from Sb to Te.<br />

The <strong>SRC</strong> is operated under Grant No. DMR-0084402. Work Supported by the US<br />

Department <strong>of</strong> Energy, Office <strong>of</strong> Science.<br />

References:<br />

[1] G.H. Lander and P. Burlet, Physica B 215, 7 (1995).<br />

[2] H. Kumigashira, T. Ito, A. Ashihara, H.D. Kim, H. Aoki, T. Suzuki, H. Yamgami and T.<br />

Takhashi, Phys. Rev. B 61, 15707 (2000).<br />

[3] B. Reihl, N. Martensson and O. Vogt, J. Appl. Phys. 53, 2008 (1982).


FIELD IONIZATION OF CH 3 I IN SUPERCRITICAL AR<br />

C. M. Evans 1 and G. L. Findley 2<br />

1 Department <strong>of</strong> Chemistry and Biochemistry, Queens College – CUNY,<br />

65-30 Kissena Blvd, Flushing, NY 11367<br />

2 Department <strong>of</strong> Chemistry, <strong>University</strong> <strong>of</strong> Louisiana at Monroe,<br />

Monroe, LA 71209<br />

Supercritical fluids have recently been shown to improve rates and modify product ratios<br />

<strong>of</strong> chemical reactions, to vary chemical shifts in NMR, and to alter lifetimes and energies <strong>of</strong><br />

molecular vibrational states. However, the detailed nature <strong>of</strong> the molecule (i.e., dopant)/fluid<br />

(i.e., perturber) interactions that lead to these effects is not well understood. In recent studies <strong>of</strong><br />

the density dependence <strong>of</strong> solvatochromic shifts <strong>of</strong> vibrational and UV-visible absorption bands,<br />

an increase in the energy shift near the critical density along the critical isotherm <strong>of</strong> the perturber<br />

was observed (cf. Fig. 1). In measurements <strong>of</strong> the field ionization <strong>of</strong> high-n CH 3 I Rydberg states<br />

in supercritical argon, we observed a decrease in the shift near the critical density along the<br />

critical isotherm (cf. Fig. 2). Such a dramatic difference in behavior is striking. The densitydependent<br />

shift <strong>of</strong> the dopant ionization energy in dense media can be written as a sum <strong>of</strong><br />

contributions, = w 0 (P) + V 0 (P), where w 0 is the shift due to the average polarization <strong>of</strong> the<br />

perturber by the ionic core, V 0 is the quasi-free electron energy in the perturbing medium and P<br />

is the perturber number density. Our preliminary analysis suggests that while w 0 shifts in a<br />

manner similar to the vibrational and UV-visible band shifts, V 0 does not. Thus, the difference in<br />

behavior between the shift <strong>of</strong> high-n Rydberg states and <strong>of</strong> vibrational (or UV-visible) absorption<br />

bands is due to the interaction <strong>of</strong> the quasi-free electron with the perturbing medium.<br />

This work was conducted at <strong>SRC</strong> (NSF DMR-0084402) and was supported by a grant from<br />

the Louisiana Board <strong>of</strong> Regents Support Fund (LEQSF (1997-00)-RD-A-14).<br />

1996<br />

0.0<br />

1994<br />

-0.2<br />

line position (cm )<br />

-1<br />

1992<br />

1990<br />

1988<br />

(eV)<br />

-0.4<br />

-0.6<br />

1986<br />

-0.8<br />

1984<br />

0 2 4 6 8 10 12 14 16 18<br />

density (mol/L)<br />

Fig. 1. Infrared absorption line peak position <strong>of</strong> the T 1u<br />

asymmetric CO stretching mode <strong>of</strong> W(CO) 6 in CO 2 vs<br />

CO 2 density at (!) the critical temperature <strong>of</strong> 33C and<br />

() 50C. The lines provide a visual aid. (Modified<br />

from R.S. Urdahl, D. J. Myers, K. D. Rector, P. H.<br />

Davis, B. J. Cherayil and M. D. Fayer, J. Chem. Phys<br />

107, 3747 (1997).)<br />

-1.0<br />

0<br />

5<br />

10<br />

15<br />

Ar<br />

(10 21 cm -3 )<br />

Fig. 2. Ionization potential <strong>of</strong> CH 3 I in argon<br />

plotted as a function <strong>of</strong> argon number density Ar.<br />

() -114C; () -118C; () various lower<br />

temperatures other than the critical temperature;<br />

() the critical temperature <strong>of</strong> -122C. The lines<br />

provide a visual aid. (C. M. Evans and G. L.<br />

Findley, to be published.)<br />

20<br />

25


CANADIAN SYNCHROTRON RADIATION FACILITY<br />

Y.F. Hu, A. Jürgensen and K.H. Tan<br />

Canadian <strong>Synchrotron</strong> <strong>Radiation</strong> Facility, Synchrotorn <strong>Radiation</strong> <strong>Center</strong>,<br />

3731 Schneider Drive, Stoughton, WI 53589 USA<br />

T.K. Sham<br />

Dept. <strong>of</strong> Chemistry, the <strong>University</strong> <strong>of</strong> Western Ontario,<br />

London, Ont. N6A 5B7 Canada<br />

The Canadian <strong>Synchrotron</strong> <strong>Radiation</strong> Facility (CSRF) is a national research facility that is<br />

owned and managed by NRC-CNRC (National Research Council Canada) and NSERCC<br />

(Natural Sciences and Engineering Research Council <strong>of</strong> Canada). It is located at the <strong>Synchrotron</strong><br />

<strong>Radiation</strong> <strong>Center</strong> (<strong>SRC</strong>), <strong>University</strong> <strong>of</strong> Wisconsin-Madison, and is operated by the <strong>University</strong> <strong>of</strong><br />

Western Ontario since 1982. CSRF provides three beamlines to the Canadian user community: a<br />

Grasshopper (20 to 250 eV), an SGM (240 to 700 eV) and a DCM (1500 to ~5000 eV). The<br />

SGM is moving to the Canadian Light Source (CLS) in October <strong>of</strong> 2003.<br />

These beamlines are regularly used for XANES and EXAFS analysis <strong>of</strong> solids with<br />

simultaneous collection <strong>of</strong> the Total Electron Yield (TEY) and Fluorescence Yield (FY) signals.<br />

This allows the determination <strong>of</strong> surface structure (TEY) and bulk structure (FY) <strong>of</strong> the sample.<br />

Other experiments performed at CSRF beamlines include X-ray Excitation Optical<br />

Luminescence (XEOL), Magnetic Circular Dichroism (MCD) and Photoemission Electron<br />

Energy Microscopy (PEEM) <strong>of</strong> solids, and PhotoElectron-PhotoIon-CoIncidence (PEPICO)<br />

spectroscopy <strong>of</strong> gases.<br />

The research done at CSRF spans over a broad range <strong>of</strong> disciplines including: materials<br />

science (semiconductors and nanoparticles), biology (plants and bacteria), environmental studies<br />

(soils and solid particles from lakewater), tribology (anti-wear oil films), agriculture (manure)<br />

and geology (rocks and minerals). Most users <strong>of</strong> the CSRF beamlines are from Canada. A few<br />

are from New Zealand and the United States. On average, over 25 papers containing work done<br />

at CSRF are published annually.


FUTURE PLANS AND RECENT ACCOMPLISHMENTS AT CNTECH<br />

James W. Taylor, Franco Cerrina, Paul Nealey, Don Thielman, and Dan Malueg<br />

<strong>Center</strong> for NanoTechnology (CNTech), 3731 Schneider Drive, Stoughton, WI 53589-3097<br />

There have been a number <strong>of</strong> new instruments added to the facilities at CNTech over this<br />

past year and a number <strong>of</strong> accomplishments in the research and development area that will be<br />

highlighted in this presentation. First, the high energy beamline providing 2,750 keV (0.45 nm)<br />

exposure radiation to the Mitsubishi Electric Company (MELCO) written diamond membrane X-<br />

ray masks has proven the concept that removing the carbon from the mask membrane and<br />

operating at higher energy will reduce the effects <strong>of</strong> diffraction and permit the printing <strong>of</strong> smaller<br />

linewidth mask features. This beamline feeds the JSAL Mod 4 stepper, and it has been upgraded<br />

to perform well at small gaps. The metrology capabilities have been upgraded with a LEO<br />

scanning microscope with 2 nm resolution and an atomic force microscope equipped with<br />

nanotube tips.<br />

New etching equipment for CNTech has arrived and has been installed on campus in the<br />

new cleanrooms <strong>of</strong> WCAM in the Engineering <strong>Center</strong>s building. These etchers will provide the<br />

capability for etching polysilicon, a material <strong>of</strong> choice for the clear phase X-ray mask involving<br />

the Bright Peak Enhanced X-ray Phase Mask (BPEXPM). The progress <strong>of</strong> the BPEXPM effort<br />

will be described, and we note that the work has progressed to the point that modeling and<br />

experiment are in agreement. The BPEXPM has been shown to produce wafer images factors <strong>of</strong><br />

3-5 smaller than the mask image and considerably larger gaps than utilized with proximity X-ray<br />

lithography. CNTech is working with BAE Systems <strong>of</strong> Nashua, NH to make devices with this<br />

technology.<br />

Negotiations are in progress with SEMATECH to install a short section undulator on the<br />

ring to power a new EUV exposure station operating at 13.4 nm. SEMATECH needs access to<br />

such a station to foster the development <strong>of</strong> EUV lithography involving multi-layer masks.<br />

Currently, CNTech operates an EUV exposure station on the U2 undulator, and recent results on<br />

that beamline will be described.<br />

CNTech has continued its work with new chemically-amplified resists and their<br />

optimization for high resolution, high contrast, and high sensitivity operation. The Design <strong>of</strong><br />

Experiments optimization has proven to be especially useful for this effort.<br />

We are expecting delivery <strong>of</strong> a stepper in June <strong>of</strong> 2004 from JSAL <strong>of</strong> Burlington, VT that<br />

will push the state-<strong>of</strong>-the-art in overlay alignment. This stepper will utilize an approach that was<br />

developed at MIT called the Interferometric Broad Band Imaging (IBBI) system.<br />

This work is supported by DARPA Grant MDA972-01-1-0039, a DARPA/NAVAIR<br />

grant N00421-03-1-0001, and a sub-contract with BAE Systems under NAVAIR N00421-02-C-<br />

3029. The <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong> is supported by the National Science Foundation under<br />

Grant DMR-0084402.


SELF-ASSEMBLY OF BLOCK COPOLYMERS ON<br />

LITHOGRAPHICALLY DEFINED NANOPATTERNED SUBSTRATES<br />

Paul Nealey<br />

<strong>Center</strong> for NanoTechnology (CNTech), 3731 Schneider Drive, Stoughton, WI 53589-3097<br />

and<br />

Dept. <strong>of</strong> Chemical Engineering, <strong>University</strong> <strong>of</strong> Wisconsin-Madison, Madison, WI 53706<br />

Top-down approaches to fabrication such as advanced lithographic techniques are<br />

designed to meet severe processing constraints, but may be prohibitively capital intensive or may<br />

not <strong>of</strong>fer sufficient control at the nanoscale. Inexpensive bottom-up approaches based on selfassembling<br />

materials such as colloidal particles and block copolymers <strong>of</strong>ten possess the required<br />

nanometer resolution, but the dimensions over which the self-assembled structures are defectfree<br />

limits potential applications. Tremendous promise exists for the development <strong>of</strong> hybrid<br />

technologies in which self-assembling materials are integrated into existing manufacturing<br />

processes to deliver molecular level control in parallel processes to meet exacting tolerances and<br />

margins, and placement <strong>of</strong> the structures, including registration and overlay, with nanometer<br />

precision. Here we demonstrate the integration <strong>of</strong> advanced lithography with the self-assembly<br />

<strong>of</strong> thin films <strong>of</strong> block copolymer to induce the epitaxial assembly <strong>of</strong> densely packed nanoscopic<br />

domains. The areas over which the patterns are defect free, oriented and registered with the<br />

underlying substrate are arbitrarily large, determined by the size and quality <strong>of</strong> the<br />

lithographically defined surface pattern rather than the inherent limitations <strong>of</strong> the self-assembly<br />

process.


IDENTIFICATION OF SUB-MICROMETER SILICATE INCLUSIONS IN<br />

ARCHEAN ZIRCONS WITH A X-RAY PHOTOELECTRON EMISSION<br />

SPECTROMICROSCOPY (X-PEEM)<br />

B. H. Frazer (1,2), G. De Stasio (1), B. Gilbert (3), A. Cavosie (4) and J. W. Valley (4)<br />

(1) <strong>University</strong> <strong>of</strong> Wisconsin-Madison, Dept. <strong>of</strong> Physics and <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong><br />

(2) Institute de Physique Appliquée, Ecole Polytechnique Fédérale de Lausanne<br />

(3) <strong>University</strong> <strong>of</strong> California-Berkeley, Earth and Planetary Science<br />

(4) <strong>University</strong> <strong>of</strong> Wisconsin, Department <strong>of</strong> Geology and Geophysics<br />

We recently optimized a new differential-thickness coating technique to analyze<br />

insulating samples with X-ray PhotoElectron Emission spectroMicroscopy (X-PEEM). X-PEEM<br />

is non-destructive, analyzes the chemical composition and crystal structure <strong>of</strong> minerals and can<br />

spatially resolve chemical species with a resolution presently reaching 35 nm. We tested the<br />

differential coating by analyzing a 4.4 billion-year-old zircon containing silicate inclusions. We<br />

observed quartz inclusions smaller than 1 µm in size, that could not be analyzed with any other<br />

non-destructive technique. We also present X-ray absorption near-edge structure (XANES)<br />

spectroscopy <strong>of</strong> 20 silicate, alumosilicate and aluminum oxide minerals and two glasses at the<br />

SiK and SiL 2,3 , and OK edges. The similar nearest-neighbor environments lead to similar spectral<br />

lineshapes at each edge, but the fine-structure differences allow individual and groups <strong>of</strong><br />

structurally similar minerals to be distinguished. By combining spectra and their first energy<br />

derivative from three absorption edges, we show that every mineral studied is distinguishable<br />

with XANES. These reference spectra, assist in the interpretation <strong>of</strong> sub-micrometer inclusions<br />

in archean zircons, studied with X-PEEM.


A<br />

B<br />

S<br />

T<br />

R<br />

A<br />

C<br />

T<br />

S


Epitaxial strain in La 2-x Sr x CuO 4 : Fermi Surface change and T c enhancement.<br />

M. Abrecht 1* , D. Ariosa 1 , D. Cloëtta 1 , G. Margaritondo 1 , M. Onellion 2 , and D. Pavuna 1 .<br />

1 IPMC, School <strong>of</strong> Basic Sciences, Ecole Polytechnique Fédérale de Lausanne (EPFL)<br />

CH-1015 Lausanne, Switzerland<br />

2 Department <strong>of</strong> Physics, <strong>University</strong> <strong>of</strong> Wisconsin-Madison, Madison, WI-53706, USA<br />

Extended abstract<br />

Using our dedicated Pulsed Laser Deposition (PLD) system and in-situ Angle Resolved<br />

Photoemission Spectroscopy (ARPES) as opposed to the traditional cleaving or scraping <strong>of</strong><br />

samples in UHV, we measured the band dispersion <strong>of</strong> epitaxially strained La 2-x Sr x CuO 4<br />

(0.1x0.2) thin films to investigate the links between strain induced increase in<br />

superconducting transition temperature (T c ) and low energy electronic features. Such<br />

measurements are the first direct probing <strong>of</strong> the electronic structure <strong>of</strong> any high T c<br />

superconductor under epitaxial strain, and were made possible only thanks to our new approach<br />

<strong>of</strong> sample growth at the <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong>.<br />

We emphasize the importance <strong>of</strong> such measurements since strain has been known to increase<br />

substantially the superconducting properties <strong>of</strong> cuprates: the critical temperature <strong>of</strong> La 2-x Sr x CuO 4<br />

(LSCO) under epitaxial strain for example can reach 50K whereas for a non strained sample, the<br />

T c value does not exceed 39K whatever the Sr (doping) content [Locquet].<br />

Our APRES results on in-plane compressed films show a band dispersing along the k x reciprocal<br />

space direction and crossing the Fermi level before the Brillouin zone boundary (figure 1), in<br />

sharp contrast to a flat band staying well below the Fermi level for unstrained samples with equal<br />

doping [Ino]. Our results were confirmed on various sampes with different doping values x<br />

[Abrecht] using the Scienta analyzers <strong>of</strong> the <strong>SRC</strong>. Very surprisingly, the critical temperatures <strong>of</strong><br />

the strained samples are enhanced by more than 6 degrees despite a tremendous reduction <strong>of</strong> the<br />

density <strong>of</strong> states due to the disappearance <strong>of</strong> the saddle point near the Fermi level (figure 2).<br />

Based on a simple 2D tight binding approximation, we were able to predict the Fermi Surface<br />

evolution with doping and strain (figure 3), a simulation that proved accurate for both Ino’s bulk<br />

data [Ino] and our strained and unstrained film samples. Implications <strong>of</strong> our findings on possible<br />

superconducting mechanisms in the cuprates include that the van-Hove singularity is not<br />

essential to obtain high critical temperatures, and that T c =T c (x, ), where is a parameter related<br />

to strain that should be included in the Superconducting phase diagram.<br />

* Present address: National Institute <strong>of</strong> Standards and Technology, Magnetic Technology Division, 325 Broadway,<br />

Boulder CO-80305.


Photoemission spectra, a.u.<br />

-0.3<br />

-0.2<br />

0.5<br />

0.6<br />

K y = 0, K x / =<br />

0.7<br />

0.8<br />

0.9<br />

0.0<br />

-0.1<br />

Binding<br />

Energy, eV<br />

Figure 1: Momentum resolved photoemission data along k x <strong>of</strong> an optimally doped (x=0.15) LSCO<br />

film under compressive strain. Note the band crossing the Fermi level for k x 0.8.<br />

Figure 2: Dispersion <strong>of</strong> the 2-dimensional band for bulk LSCO (left), and for an in-plane<br />

compressed film with identical doping (right). These figures were obtained by fitting experimental<br />

data (similar to Fig 1) but along the two high-symmetry directions -X and -M (for the unstrained<br />

sample, experimental data were taken from [Ino]). See also Fig 2 in reference attached [Abrecht].<br />

Note the disappearance <strong>of</strong> the saddle point in the strained film, and the topological transition <strong>of</strong><br />

the Fermi Surface (zero energy cut) from so-called hole-like (unstrained) to electron-like (strained).<br />

Critical temperature values were 38K (unstrained sample) and 44K (in-plane compressed sample).


1<br />

1<br />

0.8<br />

0.8<br />

0.6<br />

0.6<br />

0.4<br />

0.4<br />

0.2<br />

0.2<br />

Ky<br />

0<br />

Ky<br />

0<br />

-0.2<br />

-0.2<br />

-0.4<br />

-0.4<br />

-0.6<br />

-0.6<br />

-0.8<br />

-0.8<br />

-1<br />

-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1<br />

Kx<br />

-1<br />

-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1<br />

Kx<br />

Figure 3a: Calculated Fermi Surface<br />

change with doping (band filling) <strong>of</strong> nonstrained<br />

LSCO, assuming an unchanged<br />

band shape, ie. identical to that <strong>of</strong> Fig 2<br />

left. From the outside to the inside,<br />

doping values are x = 0.05, 0.10, 0.15,<br />

0.22, and 0.30. Note that these simulated<br />

Fermi Surfaces are in excellent<br />

agreement with the experimentally<br />

measured ones by Ino et al. [Ino]<br />

Figure 3b: Calculated Fermi Surface<br />

change (band filling) <strong>of</strong> in-plane<br />

compressed LSCO films assuming an<br />

unchanged band shape, ie. identical to<br />

that <strong>of</strong> Fig 2 right. From the outside to<br />

the inside, doping values are x = 0.05,<br />

0.08, 0.15, 0.2, 0.3.<br />

References<br />

[Loquet]: J.P. Locquet et al., Nature 394, 453 (1998)<br />

[Ino]: A. Ino et al., PRB 65, 945041 (2002)<br />

[Abrecht]: M. Abrecht et al., PRL 91 570021 (2003); reference is attached.


THE INITIAL AND FINAL STATE BANDS<br />

IN BI ALONG T<br />

Christian R. Ast 1,2 and Hartmut Höchst 1<br />

1 <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong>, <strong>University</strong> <strong>of</strong> Wisconsin-Madison, Stoughton, WI 53589<br />

2 Max-Planck-Institut für Festkörperforschung, Stuttgart, Germany<br />

Determining the dispersion <strong>of</strong> an energy band in three dimensional momentum space<br />

from angle resolved photoemission spectra requires knowledge <strong>of</strong> the final state band structure<br />

since information about the momentum <strong>of</strong> the photoelectron is lost as it is refracted at the sample<br />

surface. Commonly, a free electron final state model is assumed, however, at low final state<br />

energies the bands can still be affected by the periodic potential such that the free electron model<br />

breaks down. It has been shown from normal emission spectra <strong>of</strong> Bi(111) that for final state<br />

energies below about 60 eV deviations from the free electron approximation can be observed [1].<br />

In order to find the band dispersion in momentum space from photoemission spectra involving<br />

low final state energies both initial and final state bands have to be extracted from the<br />

photoemission spectra.<br />

Fig. 1: Normal emission spectra <strong>of</strong> Bi(111) as a function <strong>of</strong> photon energy.<br />

We report angle resolved photoemission spectra <strong>of</strong> Bi(111) measured at normal emission<br />

as a function <strong>of</strong> photon energy. Changing the photon energy in this geometry samples the<br />

Brillouin zone along the T line. The photon energy range extends from 8 eV to 100 eV<br />

covering several Brillouin zones. From the data we have extracted the initial state bands within 3<br />

eV binding energy as well as the final state bands up to a final state energy <strong>of</strong> 100 eV using a<br />

phenomenological, modified free electron model that incorporates deviations from the freeelectron<br />

dispersion. In Fig. 1 a close up <strong>of</strong> the photoemission spectra can be seen for one <strong>of</strong> the<br />

bulk bands in Bi. The solid and dashed lines trace primary and secondary cone emission features<br />

while the circles indicate peak positions in the spectra. The arrows labeled G1 to G3 indicate<br />

final state gaps away from the Brillouin zone boundaries which are visible in the photoemission<br />

spectra.


Fig. 2: Comparison <strong>of</strong> the experimental initial state bulk band structure <strong>of</strong> Bi along<br />

the T direction with various model calculations.<br />

Furthermore, scattering events can play a significant role in the interpretation <strong>of</strong><br />

photoemission spectra. They lead to various phenomena that can be observed in the<br />

photoemission spectra. It has been shown for bismuth that scattering events induce loss features<br />

in the 5d-core level spectra [2], which can be associated with interband transitions inside the<br />

crystal. In the normal emission spectra we present here secondary cone emission features, which<br />

result from transitions involving reciprocal lattice vectors not parallel to the normal emission<br />

direction, can be observed. For these features to be observed in the normal emission spectra the<br />

photoelectrons need to be scattered back into the normal emission direction. We show that for a<br />

complete data analysis secondary cone emission phenomena have to be included. The presence<br />

<strong>of</strong> secondary cone emission features in the spectra clearly shows the significance <strong>of</strong> scattering<br />

phenomena in photoemission spectra <strong>of</strong> bismuth.<br />

Figure 2 shows the extracted initial state bands (black lines) in comparison with a third<br />

neighbor tight-binding calculation [3], as well as two first principle calculations [4, 5] (dotted<br />

lines). The agreement in the dispersion is mostly qualitatively. The tight binding calculation has<br />

difficulties reproducing the dispersion <strong>of</strong> the middle band, whereas the calculation by Gonze [4]<br />

shows the best agreement with the experimental initial state bands.<br />

[1] G. Jezequel, J. Thomas, and I. Pollini, Phys. Rev. B 56, 6620 (1997)<br />

[2] C. R. Ast and H. Höchst, Phys. Rev. Lett. in press (2003).<br />

[3] Y. Liu and R. E. Allen, Phys. Rev. B 52, 1566 (1995)<br />

[4] X. Gonze, J.-P. Michenaud, and J.-P. Vigneron, Phys. Rev. B 41, 11827 (1990)<br />

[5] A. B. Shick, J. B. Ketterson, D. L. Novikov, and A. J. Freeman, Phys. Rev. B 60, 15484<br />

(1999)


TRANSITION LAYERS AT BURIED FERROMAGNETIC<br />

INTERFACES PROBED BY SOFT-X-RAY RESONANT MAGNETIC<br />

SCATTERING<br />

B. Barnes 1 ; Z. Li 2 ; D. Savage 2 ; E. Wiedemann 2 ; M. Lagally 2<br />

1 Physics, UW-Madison, Madison, WI<br />

2<br />

Materials Science and Engineering, UW-Madison, Madison, WI<br />

Understanding the effect <strong>of</strong> interfaces on the magnetic properties <strong>of</strong> thin films is<br />

critical to understanding such diverse phenomena as spin-dependent transport (e.g. giant<br />

magnetoresistance [GMR]) and coupling between magnetic films. Interfacial morphology<br />

for ferromagnetic [FM] materials may be characterized as a combination <strong>of</strong> chemical and<br />

magnetic boundaries. Previous work by Kelly, et al.[1] used the diffusely scattered<br />

component <strong>of</strong> X-ray resonant magnetic scattering (XRMS) to compare the magnetic and<br />

chemical roughness the upper interface <strong>of</strong> 70 Co films that were either bare or Fecapped.<br />

The chemical and magnetic upper boundaries within the Co differed in the<br />

absence <strong>of</strong> an adjoining Fe layer, due to a transition layer <strong>of</strong> spins that do not follow<br />

applied magnetic fields. Though the bottom interface contributed very little to the<br />

resultant scattering, its relative contribution could not be resolved. However, by<br />

performing specular XRMS over a wide range <strong>of</strong> incident angles, we now report a depth<br />

pr<strong>of</strong>ile <strong>of</strong> the magnetization within Co films 12 to 36 thick. To isolate the effect <strong>of</strong><br />

underlayer magnetism upon the lower transition layer, uncapped Co films were sputterdeposited<br />

onto both Si substrates and onto ~ 8 Ni underlayers on Si. We quantify both the<br />

upper transition layer (due to the vacuum-Co interface) and the lower transition layer, and<br />

thus illustrate the role <strong>of</strong> magnetic underlayers in maintaining magnetic order.<br />

[1] J.J. Kelly IV, et al. J. Appl. Phys. v.91 pp.9978-9986 (2002).<br />

Funding provided by ONR. Funding for the <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong> provided<br />

by NSF under Award No. DMR-0084402


TRIPLE PHOTOIONIZATION OF NEON AND ARGON NEAR<br />

THRESHOLD<br />

J.B. Bluett 1 , D. Luki 2 , S.B.Whitfield 3 , I.A. Sellin 4 , and R. Wehlitz 1<br />

1 <strong>SRC</strong>, UW-Madison, 3731 Schneider Dr., Stoughton, WI 53589<br />

2 Institute <strong>of</strong> Physics, 11001 Belgrade, Yugoslavia<br />

3 Department <strong>of</strong> Physics and Astronomy, UW-Eau Claire, WI 54702<br />

4 Department <strong>of</strong> Physics, <strong>University</strong> <strong>of</strong> Tennessee-Knoxville, TN 37996<br />

The threshold behavior <strong>of</strong> the triple photoionization cross-section <strong>of</strong> neon was<br />

investigated using monochromatized synchrotron radiation and ion time-<strong>of</strong>-flight (TOF)<br />

spectrometry. The monochromatized photon beam ionized neon or argon atoms in the<br />

experimental chamber. The ions created were accelerated by a pulsed electric field and<br />

detected by a Z-stack <strong>of</strong> microchannel plates [1].<br />

The absolute cross-section is<br />

found to follow the Wannier<br />

power law [2], i.e., the partial<br />

cross-section is proportional to<br />

the excess energy E raised to a<br />

power , E . We obtained<br />

an exponent <strong>of</strong> 2.20 0.05 that<br />

has a range <strong>of</strong> validity <strong>of</strong> ca. 5eV<br />

(green curve in Fig. 1). This<br />

result is consistent with the<br />

exponent <strong>of</strong> 2.162 predicted by<br />

theory and is also consistent with<br />

the finding <strong>of</strong> Samson and Angel<br />

[3].<br />

However, we did not find a secondary power law as in Ref. [3] but observed a smooth<br />

decrease <strong>of</strong> the exponent with increasing excess energy. Nevertheless, we could<br />

reproduce the exponent for the “second” power law as reported in [3] with an exponent <strong>of</strong><br />

1.89(4) (red curve) if we perfom the fit over the same energy range (5-9 eV) as in Ref [3].<br />

Also for argon, we can confirm the Wannier power law and determined an exponent <strong>of</strong><br />

2.21(12) in good agreement with the theoretical prediction. However, the range <strong>of</strong><br />

validity is significantly shorter than for Ne, namely ca. 2 eV only.<br />

This work was supported by NSF Grant No. 9987638. The <strong>SRC</strong> is operated under NSF<br />

Grant No. DMR-0084402.<br />

References:<br />

[1] R. Wehlitz, D. Luki, C. Koncz, and I.A. Sellin, Rev. Sci. Instrum. 73, 1671 (2002).<br />

[2] G.H. Wannier, Phy. Rev. 90, 817 (1953).<br />

[3] J.A.R. Samson and G.C. Angel, Phys. Rev. Lett. 61 1584 (1988).


LOW-DIMENSIONAL ELECTRONS AT SILICON SURFACES<br />

J. N. Crain 1 , J. L. McChesney 1 , Fan Zheng 1 , M. C. Gallagher 2 , P. C. Snijders 3 , M. Bissen 4 ,<br />

C. Gundelach 4 , S. C. Erwin 5 , and F. J. Himpsel 1<br />

1 Dept. <strong>of</strong> Physics, UW-Madison, 1150 <strong>University</strong> Ave., Madison, Wisconsin 53706<br />

2 Dept. <strong>of</strong> Physics, Lakehead <strong>University</strong>, Thunder Bay, ON P7B-5E1 Canada<br />

3 Dept. <strong>of</strong> NanoScience, Delft <strong>University</strong> <strong>of</strong> Technology,<br />

Lorentzweg, 2628 CJ Delft, The Netherlands<br />

4 <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong>, UW-Madison, 3731 Schneider Dr., Stoughton, WI 53589<br />

5 <strong>Center</strong> for Computational Materials Science, Naval Research Laboratory,<br />

Washington, DC 20375<br />

Electrons become increasingly correlated as the dimensionality is reduced from 3D to<br />

2D and 1D. Already in 2D one observes correlations between electrons and vortices in the<br />

quantum Hall effect leading to fractional charge and statistics. Predictions for a 1D metal<br />

anticipate a complete breakdown <strong>of</strong> the single electron concept and the separation <strong>of</strong> spin and<br />

charge into two collective excitations, spinons and holons.<br />

It has become possible to systematically engineer two- and one-dimensional surface<br />

structures <strong>of</strong> metals on silicon that are metallic [1,2]. Electrons near the Fermi level are decoupled<br />

from the substrate because there energy lies in the band gap. The metal atoms,<br />

however, are rigidly tied to the silicon lattice in substitutional positions according to x-ray<br />

diffraction [3] and first principles band calculations. That makes a Peierls transition to an<br />

insulator unfavorable and creates an opportunity for observing exotic states predicted for onedimensional<br />

metallic electrons.<br />

Examples <strong>of</strong> two-dimensional metals are the 3x3 and 21x21 reconstructions <strong>of</strong><br />

Ag and Au on Si(111), which exhibit an intricate pattern <strong>of</strong> Fermi surfaces induced by the<br />

superlattice [2]. Information about the Fourier components <strong>of</strong> the superlattice potential is<br />

obtained from avoided band crossings.<br />

One-dimensional atomic chains are created by growing Au and a variety <strong>of</strong> other<br />

metals at vicinal Si(111) surfaces in the 1/5 monolayer coverage regime [1]. This appears to<br />

be a rather universal phenomenon, and even the flat Si(111) surface breaks its three-fold<br />

symmetry to accommodate chain structures. This wide-open territory <strong>of</strong> 1D structures is<br />

explored in real and reciprocal space by scanning tunneling microscopy (STM) and angleresolved<br />

photoemission. The resulting energy bands and Fermi surfaces can be tuned between<br />

2D and 1D by increasing the chain spacing, with the intra-chain / inter-chain coupling ratio<br />

varying from 10:1 to >70:1. Unexpected metallic bands are found, such as a pair <strong>of</strong> nearly<br />

degenerate, half-filled bands, a quarter-filled band, and a fractional electron count <strong>of</strong> 8/3<br />

electrons per 1x1 cell [1]. A possible explanation for the fractional electron count is<br />

suggested, where extra Si atoms in a 2D cell “dope” the 1D chain embedded in it. This would<br />

be analogous to the doping in HiT c materials, where a 3D cell dopes an embedded 2D plane.<br />

References:<br />

[1] Losio, et al., Phys. Rev. Lett. 86, 4632 (2001); Altmann et al. Phys. Rev. B 64, 035406<br />

(2001); Crain et al., Phys Rev. Lett., in press (2003).<br />

[2] Crain, et al., Phys. Rev. B 66, 205302 (2002).<br />

[3] Robinson, et al., Phys. Rev. Lett. 88, 096104 (2002).


THE CANADIAN LIGHT SOURCE: A BRIGHT LIGHT SHINING<br />

IN A COLD LANDSCAPE<br />

J.N. Cutler<br />

Canadian Light Source, <strong>University</strong> <strong>of</strong> Saskatchewan, Saskatoon, Saskatchewan, S7N 5C6<br />

The Canadian Light Source (CLS), a 2.9 GeV synchrotron facility, is currently<br />

being constructed at the <strong>University</strong> <strong>of</strong> Saskatchewan. The immense 84 m x 83 m<br />

building is now complete; and major accelerator equipment (e.g. the booster ring) is<br />

currently being commissioned. The facility (with at least six beamlines) will be<br />

operational by the end <strong>of</strong> 2003.<br />

Fifteen beamlines (with very heavy involvement from the Canadian chemistry<br />

community) have now been approved to go forward; and proposals for two other<br />

beamlines are being developed. Most <strong>of</strong> these beamlines and associated endstations, will<br />

be extremely useful for research ranging from materials and environmental science to life<br />

sciences – using infrared spectroscopy and spectromicroscopy, µ-EXAFS, single crystal<br />

diffraction, and s<strong>of</strong>t and hard x-ray XANES and EXAFS. This poster will present the<br />

current status <strong>of</strong> the Canadian Light Source project and some insight into the future <strong>of</strong> the<br />

facility.


SUBCELLULAR DISTRIBUTION OF<br />

MOTEXAFIN GADOLINIUM IN TUMOR CELLS<br />

M. J. Daniels 1 , R. J. Erhardt 1 , B. H. Frazer 1 , D. Rajesh 2 ,<br />

S. P. Howard 2 , M. P. Mehta 2 , G. De Stasio 1<br />

1 UW-Madison-<strong>SRC</strong>, 3731 Schneider Dr., Stoughton WI, 53589<br />

2 Department <strong>of</strong> Human Oncology, Medical School UW-Madison, Madison WI 53792<br />

Gadolinium Neutron Capture Therapy (GdNCT) is a brain cancer therapy at the<br />

experimental stages. It requires that Gd target the nuclei <strong>of</strong> cancer cells, while sparing those <strong>of</strong><br />

healthy tissue. In our multi-year quest for the optimum Gd compound to use for this therapy, we<br />

recently tested Motexafin gadolinium (MGd). MGd is an agent with known tumor specificity, as<br />

revealed by MRI, proposed as a radiosensitizer. Macroscopic tumor uptake <strong>of</strong> MGd has been<br />

well documented, but the microscopic, subcellular distribution <strong>of</strong> MGd is not well known. We<br />

studied the subcellular distribution <strong>of</strong> MGd in two different in vitro cell lines, TB10 and U87<br />

(both human glioblastoma cells) using the Spectromicroscope for PHotoelectron Imaging <strong>of</strong><br />

Nanostructures with X-rays (SPHINX) at <strong>SRC</strong>. Our extensive data analysis revealed that 86% <strong>of</strong><br />

the cell nuclei contained Gd. This result is promising because we previously calculated that an<br />

effective GdNCT agent must target >90% <strong>of</strong> cell nuclei. Furthermore, this studied showed that<br />

the concentration <strong>of</strong> Gd was highest in the cytoplasm, than the nucleus, but with longer exposure<br />

times this nuclear Gd concentration seemed to increase.


X-RAY PHOTOELECTRON EMISSION SPECTROMICROSCOPY<br />

REVEALS THAT POLYSACCHARIDES TEMPLATE ASSEMBLY OF<br />

BIOGENIC NANOSCALE CRYSTAL FIBERS<br />

G. De Stasio 1 , C.S. Chan 2 , S.A. Welch 3 , M. Girasole 4 , B.H. Frazer 1 ,<br />

M. Nesterova 5 , L.M. Weise 1 and J.F. Banfield 2<br />

(1) UW-Madison (2) UC-Berkeley<br />

(3) Australian National <strong>University</strong> (4) Italian National Research Council<br />

Neutrophilic iron oxidizing bacteria extract metabolic energy from iron oxidation in various<br />

environments, such as groundwater seeps, streams, wetlands, the rhizosphere, and hydrothermal<br />

vents. Their physiology is not well understood, and in particular the role <strong>of</strong> extracellular<br />

polymers in iron oxide mineralization and possibly energy metabolism. Recent experiments with<br />

the Spectromicroscope for PHotoelectron Imaging <strong>of</strong> Nanostructure with X-rays (SPHINX) on<br />

precipitated Fe oxides in bi<strong>of</strong>ilms clarified that microbially extruded polysaccharide filaments<br />

provide the precipitation substrate for amorphous FeOOH. Upon aging the mineralized filaments<br />

crystallize to ferrihydrite (2-line FeOOH), with one curved pseudo-single crystal <strong>of</strong> akaganeite<br />

(-FeOOH), at the core <strong>of</strong> each filament, <strong>of</strong> aspect ratio 1:1000:1. Structure and morphology <strong>of</strong><br />

this unusual and unprecedented nanoscale crystal is therefore templated by polysaccharides.<br />

After formation <strong>of</strong> the crystal fiber, the polysaccharide structure is also altered, and C1s spectra<br />

suggest that the COO - group is involved in the templation mechanism.


LINEAR POLARIZATION MEASUREMENTS OF THE<br />

SYNCHROTRON RADIATION FROM A BENDING MAGNET<br />

B. M. Dirksen 1 and K. W. McLaughlin 2<br />

1 Dept. <strong>of</strong> Medical Physics, <strong>University</strong> <strong>of</strong> Wisconsin, Madison, WI 53713<br />

2 Dept. <strong>of</strong> Physics and Engineering, Loras College, Dubuque, IA 52001<br />

We have measured the linear<br />

polarization <strong>of</strong> the ionizing radiation on the<br />

Stainless Steel Seya beam line 021 at the<br />

<strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong> by measuring<br />

the displacement current from a gold surface<br />

after this radiation had reflected at a 45 o<br />

incident angle from another gold surface. A<br />

characteristic cos 2 () pattern was observed<br />

upon rotating these surfaces about the<br />

ionizing radiation propagation direction.<br />

After normalizing to the displacement<br />

current for the 45 o reflection surface and<br />

accounting for the reflection coefficients <strong>of</strong><br />

gold surfaces [1,2], we can place a minimum<br />

value for the linear<br />

polarization <strong>of</strong> 97%,<br />

independent <strong>of</strong> the ionizing<br />

photon energy from 24 eV<br />

down to zero order on the<br />

beam line monochromator<br />

grating.<br />

Propagation<br />

direction for<br />

synchrotron<br />

radiation<br />

A<br />

A<br />

We would like to thank the staff and administration <strong>of</strong> the <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong> for<br />

the opportunity and assistance in accomplishing this project. The kind assistance <strong>of</strong> Mark<br />

Bissen, Roger Hansen, Bob Julian, Chris Moore, Mary Severson and Dan Wallace is particularly<br />

noted.<br />

This work is supported by NSF Grant No. 9731869. The <strong>SRC</strong> is operated under Grant No.<br />

DMR-0084402.<br />

References:<br />

[1] J. A. R. Samson, Techniques <strong>of</strong> Vacuum Ultraviolet Spectroscopy (Wiley, 1967), pages 304-<br />

305.<br />

[2] D. W. Lynch and W. R. Hunter, "Optical Constants <strong>of</strong> Metals" in Handbook <strong>of</strong> Optical<br />

Constants <strong>of</strong> Solids, edited by E. D. Palik (Academic Press, New York, 1985), page 286.


TEMPERATURE DEPENDENCE OF N=1 BISMUTH-BASED<br />

HTSC AS SEEN IN ARPES<br />

L. Dudy 1 , B. Müller 1 , A. Krapf 1 , H. Dwelk 1 , H. Höchst 2 and R. Manzke 1<br />

1 Humboldt Universität zu Berlin 2 <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong> (<strong>SRC</strong>)<br />

High temperature superconductivity in the cuprates mainly results from the twodimensional<br />

electronic structure <strong>of</strong> the hole-doped CuO-planes. The maximum transition<br />

temperature <strong>of</strong> this kind <strong>of</strong> material is dependent <strong>of</strong> the number <strong>of</strong> CuO-planes per unit-cell<br />

and <strong>of</strong> the layers between the CuO-layers. These layers basically isolate the CuO-planes from<br />

each other and dope holes in them. The current status <strong>of</strong> our investigations focuses on the<br />

single-layer material Bi 2 Sr 2-x La x CuO 6+ (Bi-2201), which is available in high quality single<br />

crystals. Substituting divalent Sr by trivalent La varies the hole concentration <strong>of</strong> our samples<br />

almost continuously. Because <strong>of</strong> the low transition temperature this material provides in<br />

addition the advantage <strong>of</strong> investigating the normal state with extremely small temperature<br />

broadening. For photoemission measurements Bi-2201 also benefits <strong>of</strong> the absence <strong>of</strong> bilayer<br />

splitting. This effect occurs by the interaction <strong>of</strong> electrons between narrow CuO-planes<br />

as e.g. observed in Bi 2 Sr 2 CaCu 2 O 8+ (Bi-2212).<br />

In the past we presented high-resolution photoemission results performed at <strong>SRC</strong><br />

which below a certain temperature show along the direction <strong>of</strong> the CuO-bonds a<br />

characteristic two-peak structure [1]. It has been shown that the low energy sharp peak is<br />

strongly dependent <strong>of</strong> the applied polarization <strong>of</strong> the incident light (Fig.1a) [2]. To going<br />

further it seems like the structure isn’t visible in each M-direction (Fig. 1b). Another<br />

remarkable observation is that this structure is not only seen in the superconducting regime<br />

but also persists above the transition temperature and vanishes at higher temperatures<br />

(Fig.1c). The comparison <strong>of</strong> temperature dependence with a pseudo-1d-system [3] points to a<br />

pseudo-1d effect. In our interpretation these features are associated with one-dimensionality<br />

<strong>of</strong> the electronic structure leading to spin-charge separation [4]. The necessary condition for<br />

this state is the formation <strong>of</strong> an asymmetric electronic density, a phenomenon called stripes<br />

[5]. The present contribution follows this interpretation but pr<strong>of</strong>its from the raise <strong>of</strong> data<br />

collected at the <strong>SRC</strong> in order to explain the phase diagram more precisely and identifies the<br />

vanishing <strong>of</strong> the one-dimensionality with the closing <strong>of</strong> the pseudogap [6]. Recent results [7]<br />

<strong>of</strong> the universality <strong>of</strong> the pseudogap temperature T* for every CuO-plane in every cuprate<br />

will also be tried to be discussed.


Fig.1: Photoemission spectra <strong>of</strong> optimally doped Bi 2 Sr 2-x La x CuO 6+ (x=0.40, T c =29 K) in the normal state at the<br />

M-point <strong>of</strong> the Brillouin zone. Polarization dependence (a): Shown are two polarization geometries, where the<br />

sample was kept fixed and the electrical field vector E <strong>of</strong> the linearly polarized synchrotron radiation has been<br />

turned by 90°. T=35K, h=34eV, and E=30meV [2]. 90°-asymmetry (b): Shown are two spectra taken at<br />

otherwise identical conditions at the two equivalent M-points (,0) and (0,). T=35K, h=22eV, and<br />

E=16meV. Temperature dependence <strong>of</strong> weakly overdoped (BiPb) 2 Sr 2-x La x CuO 6+ (x=0.15, T c =25K) (c):<br />

Shown are spectra taken near M at (0.75,0) for two temperatures, at T=10K where the double structure is very<br />

pronounced and at T=90K it has vanished. The thick dotted curves in all three panels are the difference spectra.<br />

References:<br />

[1] C. Janowitz, R. Müller, L. Dudy, A. Krapf, R. Manzke, C. Ast, H. Höchst, Europhysics<br />

Letters 60 (2002) 615<br />

[2] R. Manzke, R. Müller, C. Janowitz, M. Schneider, A. Krapf, H. Dwelk, Phys. Rev. B<br />

(Rapid Commun.) 63 (2001) 100504<br />

[3] R. Claessen et al., Phys. Rev. Lett. 88, 096402 (2002)<br />

[4] J. Voit, Euro. Phys. J. B 5 (1998) 505<br />

[5] for an introduction see J. Orenstein, A.J. Millis, SCIENCE 288 (2000) 468 and references<br />

therein<br />

[6] for a review about the Pseudogap see T. Timusk und B. Statt Rep. Prog. Phys. 62, 61<br />

(1999)<br />

[7] T. Honma, P.H. Hor, H.H. Hsieh, M. Tanimoto, cond-mat/0309597, 2003


UPGRADING THE SAMPLE MANIPULATOR OF THE SCIENTA-<br />

2002 USER SYSTEM WITH AN ADDITIONAL ANGULAR DEGREE<br />

OF FREEDOM<br />

Chad T. Gundelach, Mike V. Fisher, Sergey Gorovikov and Hartmut Höchst<br />

<strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong>, <strong>University</strong> <strong>of</strong> Wisconsin-Madison,<br />

3731 Schneider Drive, Stoughton, WI 53589<br />

The <strong>SRC</strong> user community has access to two state <strong>of</strong> the art photoemission<br />

systems equipped with a 200 mm Scienta spherical electron analyzer. Both analyzers are<br />

capable <strong>of</strong> high energy and angular resolution. In the multidetection mode, the lens<br />

system has an angular dispersion <strong>of</strong> ~1mm/degree enabling the simultaneous detection <strong>of</strong><br />

about 10 degrees along the entrance slit <strong>of</strong> the analyzer with an angular resolution <strong>of</strong> ~0.2<br />

degree. By mounting the analyzer so that the angular dispersion plane is either horizontal<br />

or vertical, the 10-degree angular sampling window was used by various user groups to<br />

sample data in the azimuthal- or polar-direction relative to the sample normal. Depending<br />

on Users’ preferences, the orientation <strong>of</strong> the SES-200 system was switched several times<br />

during the past few years. Remounting the analyzer is a very elaborate and timeconsuming<br />

process. Frequent orientation changes not only increase the down time <strong>of</strong> the<br />

system but also can also seriously jeopardize the spectrometer performance or damage<br />

the fragile multi-channel plate detector.<br />

To minimize the needs for orientation changes on the SES-2002 system, we<br />

upgraded the sample manipulator with an additional degree <strong>of</strong> angular rotation. The<br />

poster describes the design <strong>of</strong> the new cryogenic sample manipulator and reports first test<br />

measurements utilizing the azimuthal and polar dispersion direction to map a large twodimensional<br />

section <strong>of</strong> the Brillouin zone. The design efforts were governed by space<br />

limitations, which led to coupling some mechanical components to the cold head’s outer<br />

cryo-shield without significantly diminishing the thermal performance. The polar<br />

rotation axis is established by capturing a ball in detents on both sides <strong>of</strong> the lower inner<br />

stage using set-screws from the outer stage. An additional benefit <strong>of</strong> supporting the lower<br />

inner stage with the outer stage is the improved mechanical and thermal positional<br />

stability. As the inner stage is heated, the thermal expansion <strong>of</strong> the upper inner stage is<br />

absorbed by bendable cooling straps while the sample itself remains in the fixed outer<br />

shield position. The mechanism for actuating the polar sample angle is located beneath<br />

the lower inner stage. It is based on a worm gear system consisting <strong>of</strong> a segment <strong>of</strong> a<br />

titanium gear that is driven by a silicon bronze worm. The self-locking characteristic <strong>of</strong><br />

the worm gear mechanism prevents unintentional rotation. Two 0.012” thick by 2 “ wide<br />

OFHC copper straps provide thermal coupling to the inner cryo-stage and allow the<br />

sample holder to rotate by ±30 degree around the polar axis. Measurements verified that<br />

the temperature drop along the flexible Cu bands is less than 1K.<br />

Two GaAlAs diode temperature sensors are mounted to the cold finger. The<br />

permanent stationary diode (PSD) is mounted on the upper inner stage near the heater,<br />

while the permanent moving diode (PMD) is mounted on the lower inner stage just above<br />

the sample area. The heater and diodes allow for closed loop control <strong>of</strong> the sample<br />

temperature using a Lakeshore temperature controller.


The sample manipulator has a cool<br />

down time <strong>of</strong> six hours. This long time<br />

constant is a result <strong>of</strong> the large thermal mass<br />

<strong>of</strong> the cold finger. The cryo head itself<br />

reaches a minimum temperature <strong>of</strong> ~10K in<br />

less than an hour under no load conditions.<br />

Temperature measurements with a calibrated<br />

GaAlAs diode mounted on the Ti-sample<br />

slug indicate a minimum sample temperature<br />

<strong>of</strong> ~18.5K. Another test investigated the<br />

impact radiative heating had on the minimum<br />

temperature, the difference between having<br />

the sample area shielded and unshielded was<br />

~ 4K.<br />

Extrapolating the temperature versus<br />

heater power curve for the two permanent<br />

diodes to 10K provides an estimate <strong>of</strong> the<br />

heat load at the inner stage to be ~0.5 W.<br />

This heat loss includes all sources <strong>of</strong> head<br />

load on the inner stage such as radiative heat<br />

and heat leak from the outer to inner stage<br />

(titanium balls and worm gear interface).<br />

By engaging the worm shaft with a<br />

non-magnetic ball driver, the polar angle can<br />

be actuated while the sample is in measuring<br />

position at an azimuthal angle <strong>of</strong> ~0 degree.<br />

The ball driver is attached to a retractable<br />

externally mounted rotary-linear feed<br />

through. The additional thermal losses<br />

imposed by the ball driver are minimal.<br />

While engaged for longer time periods, the<br />

temperature increase at the sample was only<br />

~0.2K.<br />

The functionality <strong>of</strong> the polar angle<br />

scan and the reproducibility <strong>of</strong> the<br />

mechanism were also tested in the<br />

photoemission chamber under UHV<br />

Figure 1 - Fermi Surface <strong>of</strong> BiSb alloy single<br />

crystal measured at a photon energy <strong>of</strong> 18 eV<br />

conditions. The improved thermal positional stability allowed photoemission<br />

measurements to be taken on a small sample (< 1 mm diameter) as it was heated from 18<br />

to 280 K without the need to reposition the sample. The figure on the right shows a twodimensional<br />

photoemission intensity map <strong>of</strong> a Bi0.86Sb0.14 crystal near the Fermi<br />

energy which was measured at 50K utilizing the newly added polar angular axis.<br />

Acknowledgements<br />

The <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong> (<strong>SRC</strong>) is funded by the National Science Foundation<br />

(NSF) under Grant No. DMR-0884402.


ANGLE-RESOLVED PHOTOEMISSION OF UAsSe AND USb 2<br />

E. Guziewicz 1 , T. Durakiewicz 1 , M.T. Butterfield 1 , C.G. Olson 2 , J.J. Joyce 1 , A.J. Arko 1 ,<br />

J.L. Sarrao 1 , A. Wojakowski 3 , T. Cichorek 3,4<br />

1 Los Alamos National Laboratory, Los Alamos, New Mexico, 87545<br />

2 Ames Laboratory, Iowa State <strong>University</strong>, Ames IA, USA<br />

3 Institute <strong>of</strong> Low Temperature and Structure Research, Polish Academy <strong>of</strong> Sciences,<br />

Wroclaw, Poland<br />

4 Max Planck Institute for Chemical Physics <strong>of</strong> Solids, D-01187 Dresden, Germany<br />

We have performed ARPES study on single crystals <strong>of</strong> UAsSe and USb 2 . Both compounds<br />

have tetragonal layered crystal structures and order magnetically along the c axis at low<br />

temperature. UAsSe shows ferromagnetic behavior below 116K, and USb 2 is an<br />

antiferromagnet below 203K.<br />

Magnetic behavior gives<br />

evidence for localization <strong>of</strong> the<br />

U5f states. However, the<br />

enhanced Sommerfeld<br />

coefficients (41 mJ/mol·K 2 for<br />

UAsSe and 26 mJ/mol·K 2<br />

Figure 1: Angle-resolved normal photoemission spectra <strong>of</strong> USb 2<br />

and UAsSe showing 3D character <strong>of</strong> these compounds.<br />

for<br />

USb 2 ) indicate some degree <strong>of</strong><br />

hybridization between the U5f<br />

and conduction band electrons.<br />

We have investigated this<br />

problem <strong>of</strong> correlation between<br />

the U5f electrons and their<br />

hybridization with the<br />

conduction band using ARPES.<br />

Photoemission spectra were<br />

taken on the PGM beamline at<br />

the <strong>SRC</strong> in Wisconsin with an<br />

overall energy resolution for<br />

the lowest photon energies<br />

(h=20 eV) <strong>of</strong> 24 meV. Both<br />

uranium compounds display a similar electronic structure within 100 meV <strong>of</strong> the Fermi<br />

edge, where a very narrow 5f character peak is found. The intensity <strong>of</strong> this peak does not<br />

follow a simple photoionization cross-section dependence for the U5f electrons, indicating<br />

a considerable degree <strong>of</strong> hybridization with the conduction band. In both UAsSe and USb 2<br />

we have found a dispersion <strong>of</strong> the peak near the Fermi edge <strong>of</strong> about 10 meV and 20 meV,<br />

respectively, which is additional evidence <strong>of</strong> hybridization. The dispersion in normal<br />

incidence spectra (Fig. 1) provides evidence that neither UAsSe nor USb 2 have purely 2D<br />

electronic structure and thus require treatment as 3D or quasi-2D materials. The narrow<br />

bands in these U-based magnetic materials are reminiscent <strong>of</strong> the band magnetism found in


Cr and Fe but for these U systems the band widths and dispersions are two orders <strong>of</strong><br />

magnitude smaller.<br />

Work Supported by the US Department <strong>of</strong> Energy, Office <strong>of</strong> Science. The <strong>SRC</strong> is<br />

operated under Grant No. DMR-0084402.


Hidden One Dimensionality and Non-Fermi Liquid ARPES Lineshapes<br />

<strong>of</strong> the Electronic Structure <strong>of</strong> η-Mo 4 O 11<br />

G. –H. Gweon 1* , S. –K. Mo 1 , J. W. Allen 1 , H. Höchst 2 , J. L. Sarrao 3 , and Z. Fisk 3<br />

1. Randall Laboratory <strong>of</strong> Physics, <strong>University</strong> <strong>of</strong> Michigan, Ann Arbor, MI 48109<br />

2. <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong>, <strong>University</strong> <strong>of</strong> Wisconsin, Stoughton, WI 53589<br />

3. National High Magnetic Field Lab., Florida State <strong>University</strong>, Tallahassee, FL 32306<br />

η-Mo 4 O 11 is a layered metal that undergoes two charge density wave (CDW) transitions at 109 K<br />

and 30 K, and is unique in showing a bulk quantum Hall effect [1]. Research so far indicates that<br />

this material has a “hidden one-dimensional” (hidden-1d) Fermi surface (FS) in the normal state<br />

(T > 109 K), whose nesting property drives the 109 K CDW formation [2]. Here, we directly<br />

confirm this picture by angle resolved photoemission spectroscopy (ARPES). Figure 1 shows the<br />

Fermi energy intensity map measured at T=150K and with hν=17eV at the 4m-NIM line <strong>of</strong> the<br />

<strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong>. The geometry <strong>of</strong> the FS is in good general agreement with that <strong>of</strong><br />

the band calculation, and can be seen as two vertical lines <strong>of</strong> hidden-1d FS coming from chains<br />

along the crystal b axis and double oblique hidden 1-d lines coming from chains along the crystal<br />

(b±c) axes. The latter are characterized by a single nesting vector Q CDW , similar to the situation<br />

<strong>of</strong> other 2-d hidden-1d materials like NaMo 6 O 17 and KMo 6 O 17 [3]. Fig 2 shows the temperature<br />

dependent change <strong>of</strong> the valence band spectrum at point A and B respectively. We have observed<br />

that there is a small gap opening <strong>of</strong> size ~15meV only at the point A accompanied with a change<br />

in the line shape, while at the point B, which is a part <strong>of</strong> the remnant FS that is not nested by<br />

Q CDW, the spectrum does not show any change as the temperature decreases. Even more<br />

interesting, this material also shows the same ARPES line shape anomalies and lack <strong>of</strong> Fermi<br />

edge in the angle integrated spectrum, that we have identified in other low dimensional metals<br />

and that are most easily rationalized within an electron fractionalization scenario that includes<br />

for the quasi-2d systems the idea <strong>of</strong> a “melted holon” part <strong>of</strong> the lineshape, arising from disorder<br />

[4]. This lineshape is neatly confined to the electronic bandwidth but is essentially featureless in<br />

energy and k. It is best seen in a region <strong>of</strong> k-space where all dispersing peaks lie above the<br />

Fermi energy. Fig. 3 shows the “melted holon” lineshape <strong>of</strong> η-Mo 4 O 11 obtained at such a point<br />

in k-space, along the yellow line marked in the Fig. 1. Disorder that could be responsible is<br />

known in this material [5]. More detailed studies on the lineshapes and also <strong>of</strong> the 30 K CDW<br />

transition are in progress.<br />

The work at <strong>University</strong> <strong>of</strong> Michigan is supported by the US NSF grant No. DMR-0302825. The<br />

<strong>SRC</strong> is supported by the US NSF grant No. DMR-00-84402.<br />

References<br />

[1] S. Hill et al., Phys. Rev. B 58, 10778 (1998).<br />

[2] E. Canadell et al., Inorganic Chemistry 28, 1466 (1989)<br />

[3] G. –H. Gweon et al., Phys. Rev. B 55, R14453 (1997).<br />

* current address: Advanced Light Source, Lawrence Berkeley National Laboratory, Berkeley, CA 94720


[4] G. –H. Gweon, J. W. Allen, and J. D. Denlinger, cond-mat/0303122.<br />

[5] Mallet et al, Phys. Rev. B 63, 165428 (2001).<br />

k (Å -1 )<br />

0 y 0.2 0.4<br />

A<br />

B<br />

Z<br />

Γ<br />

Q CDW1<br />

M<br />

Y<br />

-0.5 0 0.5<br />

k x (Å -1 )<br />

Figure 1. Normal state Fermi energy<br />

intensitiy map <strong>of</strong> η-Mo4O11. White arrow<br />

shows the nesting vector for T=109K CDW<br />

transition.<br />

Intensity (arb. units)<br />

(a)<br />

110 K<br />

50 K<br />

-0.2 -0.1 0 0.1<br />

(b)<br />

-0.2 -0.1 0 0.1<br />

E-E F (eV)<br />

-1 -0.5 0<br />

E-E F<br />

Figure 3. “Melted holon”<br />

lineshape at the Y point <strong>of</strong> the<br />

k-space.<br />

Figure 2. Temperature dependent<br />

change <strong>of</strong> the valence band spectrum at<br />

point A and B respectively.


THE FERMI SURFACES OF THIN Sb(111) FILMS<br />

Hartmut Höchst and Christian R. Ast<br />

<strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong>, <strong>University</strong> <strong>of</strong> Wisconsin-Madison<br />

3731 Schneider Drive, Stoughton, WI, 53589<br />

The unique electronic properties <strong>of</strong> the group-V elements Sb and Bi, their small<br />

Fermi energy combined with highly anisotropic electron and hole masses and carrier<br />

concentrations several orders <strong>of</strong> magnitude lower than in normal metals makes them<br />

prime candidates for potential thermoelectric converters. Devices based on nano-scale<br />

structures <strong>of</strong> Sb and Bi or alloys there<strong>of</strong> are among those with the highest conversion<br />

efficiency.[1,2]<br />

The rhombohedral (A7) crystal structure <strong>of</strong> the group V-elements causes a<br />

negative band gap. The conduction band minimum at the L-point is lower than the<br />

valence band maxima, which occur at the H-point for Sb and at the T-point for Bi. As a<br />

result, the Fermi surfaces (FS) consist <strong>of</strong> six pockets centered at the H-points or <strong>of</strong> two<br />

half-pockets centered at the T-points. Band structure calculations using different<br />

theoretical approaches agree in some basic overall Fermi surface features.[3-6].The<br />

calculated dimensions <strong>of</strong> the Fermi surface however have been far from experimental<br />

results based on magnetotransport and resonance measurements. These data were<br />

historically the most accurate even though indirect sources to judge the quality <strong>of</strong> the<br />

various theoretical approaches. More recently, photoemission spectroscopy (PES) was<br />

recognized to be a potentially useful additional tool to investigate the electronic<br />

properties near the FS. Compared to the classical FS measurements, PES has the<br />

advantage that information can be gathered from very small samples, at elevated<br />

temperatures and from materials, which exhibit strong impurity, compositional or<br />

structurally related scattering effects otherwise detrimental to most classical resonance<br />

based FS measurements.<br />

We report the first photoemission based FS data <strong>of</strong> Sb. Combining ARPES with<br />

the tuneability <strong>of</strong> synchrotron radiation allows determining the parallel components <strong>of</strong> the<br />

Fermi momentum k x and k y . Using a simple final state model provides also access to the<br />

perpendicular component k enables us to separate the three dimensional hole- pocket<br />

typical for bulk Sb from additional FS features which are <strong>of</strong> 2D-nature and most likely<br />

related to the presence <strong>of</strong> a Sb bilayer on the (111)-surface. Compared to Bi[7-9], which<br />

has extremely small bulk hole-pockets, preventing a detailed k-mapping due to<br />

experimental limitations in the momentum resolution, the carrier concentration in Sb is<br />

~200 times larger and thus opening up the possibility <strong>of</strong> tracing the contours <strong>of</strong> the bulk<br />

hole-FS. The location and cross section <strong>of</strong> the hole pocket in the mirror plane compares<br />

well with theoretical predictions based upon pseudopotential calculations by Falicov and<br />

Lin [3] while other more recent and advanced model calculations [4-6] seem not to<br />

comply with our data.<br />

Acknowledgements<br />

The <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong> (<strong>SRC</strong>) is funded by the National Science Foundation<br />

(NSF) under Grant No. DMR-0884402.


References<br />

[1] S. Cho, Y. Kim, A. Di Venere, et al., Journal <strong>of</strong> Vacuum Science & technology A:<br />

Vacuum Surfaces and Films 17, 2987 (1999).<br />

[2] X. Sun, Z. Zhang, and M. S. Dresselhaus, Applied . Physics Letters 74, 4005 (1999).<br />

[3] L. M. Falicov and P. J. Lin, Physical Review 141, 562 (1966).<br />

[4] J. Rose and R. Schuchardt, Physica Status Solidi (b) 117, 213 (1983).<br />

[5] Y. Liu and R. E. Allen, Physical Review B 52, 1566 (1995).<br />

[6] X. Gonze, J. P. Michenaud, and J. P. Vigneron, Physical Review B 41, 11827 (1990).<br />

[7] C. R. Ast and H. Höchst, Physical Review Letters 77, 177602 (2001)<br />

[8] C. R. Ast and H. Höchst, Physical Review B 66, 125103 (2002).<br />

[9] C.R. Ast and H. Höchst, Physical Review Letters 90, 016403 (2003)<br />

Fig.1 : Fermi level intensity map I(k || ) <strong>of</strong> Sb(111) measured at hv=25 eV<br />

Fig.2: Fermi level intensity map I(k ,k || ) <strong>of</strong> Sb(111). The perpendicular electron<br />

momentum k was changed by scanning the photon energy from 14-28 eV.


MOMENTUM DEPENDENT LOW ENERGY LOSSES IN CORE<br />

LEVEL PHOTOEMISSION SPECTRA OF POORLY CONDUCTING<br />

METALS?<br />

Hartmut Höchst and Christian R. Ast<br />

<strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong>, <strong>University</strong> <strong>of</strong> Wisconsin-Madison<br />

3731 Schneider Drive, Stoughton, WI, 53589<br />

The availability <strong>of</strong> high-quality samples having low impurity levels and good<br />

structural quality combined with instrumental improvements in synchrotron radiation<br />

beam lines and electron energy analyzers <strong>of</strong> high energy resolution led recently to<br />

photoemission core-level spectra exhibiting additional features which were previously<br />

not seen. Photoemission line-shape changes or multiple core-level components can occur<br />

through surface modifications, crystal-field splitting, vibrational contributions and<br />

excitations <strong>of</strong> electrons or phonons. [1] The theory <strong>of</strong> core-level photoemission and<br />

aspects <strong>of</strong> intrinsic and extrinsic excitations associated with the photo-hole, or<br />

photoelectron, were recently restated in a series <strong>of</strong> papers by Hedin and coworkers.[2]<br />

The relative strength <strong>of</strong> low energy excitations accompanying the “main line”<br />

photoemission spectrum is directly related to the dc resistivity and can be quite<br />

significant in materials <strong>of</strong> poor conductivity such as small-gap superconductors and<br />

semimetals.[3]<br />

We report angle resolved core-level photoemission spectra <strong>of</strong> the semimetals Sb<br />

and Bi where we observe a multiple peak structure separated by ~170-300 meV towards<br />

higher energy from the main 4d and 5d components, respectively.[4,5] Utilizing photons<br />

ranging from 30-150 eV, angular scans along the k-direction normal to the (111) surface<br />

and k || scans along high symmetry directions <strong>of</strong> the surface Brillouin-zone (SBZ), we find<br />

dispersion in the split-<strong>of</strong>f components commensurate with the bulk and surface Brillouinzone<br />

periodicity. Additionally, the peak width and intensity <strong>of</strong> the split-<strong>of</strong>f peak oscillates<br />

between final state symmetry points.<br />

From the wealth <strong>of</strong> accumulated information, we dismiss an earlier claim by<br />

Patthey et al. assigning the additional feature in the Bi 5d spectra to a surface- shifted<br />

component.[6] Since the d 5/2 and d 3/2 components show the same features symmetry<br />

arguments rule out crystal-field effects as the source <strong>of</strong> the observed peak splitting. The<br />

data can be explained by a k-dependent loss function consisting <strong>of</strong> several characteristic<br />

direct and indirect low energy interband transitions excited in various parts <strong>of</strong> the bulk<br />

Brillouin zone.<br />

Acknowledgements<br />

The <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong> (<strong>SRC</strong>) is funded by the National Science Foundation<br />

(NSF) under Grant No. DMR-0884402.<br />

References<br />

[1] J. N. Andersen, T. Balasubramanian, C.-O. Almbladh, et al., Physical Review Letters<br />

86, 4398 (2001).<br />

[2] L. Hedin and J. D. Lee, Physical Review B 64, 115109 (2001).


[3] D. L. Mills, Physical Review B 62, 11197 (2000).<br />

[4] H. Höchst, and C.R. Ast,, Journal <strong>of</strong> Electron Spectr. and Relat. Phenomena, (in<br />

press) 2003<br />

[5] C. R. Ast and H. Höchst, Physical Review Letter, in press (2003).<br />

[6] F. Patthey, W. D. Schneider, and H. Micklitz, Physical Review B 49, 11293 (1994).<br />

Fig. 1: Energy difference E 1 (filled dots) and relative peak intensity I 1 /I 0 (open dots) <strong>of</strong> the Bi 5d 5/2<br />

component measured at h=60 eV as a function <strong>of</strong> parallel momentum along and .<br />

Fig. 2: Fermi level region <strong>of</strong> the projected bulk band structure <strong>of</strong> Bi(111). Arrows indicate indirect (I) and<br />

direct (D) interband transitions related to the loss features observed by LEELS.


THE P(1s) AND P(2p) XAFS SPECTRA OF ELEMENTAL<br />

PHOSPHORUS, THEORY AND EXPERIMENT<br />

Astrid Jürgensen<br />

Canadian <strong>Synchrotron</strong> <strong>Radiation</strong> Facility, <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong>,<br />

3731 Schneider Drive, Stoughton, WI, USA 53589-3097<br />

X-ray absorption spectroscopy was used to study the local structure and chemical bonding <strong>of</strong><br />

elemental phosphorus. Three allotropes <strong>of</strong> this element exist in the solid state. Crystalline black-<br />

P, a semiconducting material with an orthorhombic puckered layer structure, is the most stable<br />

allotrope. White-P, the least stable allotrope, is composed <strong>of</strong> tetrahedral P 4 molecules. At room<br />

temperature it is a plastic crystal. Its structure is similar to -Mn with the P 4 molecules in the<br />

positions <strong>of</strong> the Mn atoms. The most common allotrope is the amorphous red-P. Like white-P it<br />

is composed <strong>of</strong> P 4 molecules. In the gas phase tetrahedral P 4 molecules are stable up to ~ 800 ºC.<br />

At higher temperatures they decompose into P 2 molecules that have a shorter P-P bond length<br />

and, like N 2 , a formal triple bond.<br />

Theoretical P(1s) and P(2p) XAFS spectra <strong>of</strong> gaseous and solid state phosphorus were<br />

calculated by ab initio FEFF and GSCF3 methods. These were compared and the spectral<br />

features were related to structural and electronic properties <strong>of</strong> the different allotropes <strong>of</strong><br />

elemental phosphorus. The calculation results were then used to explain the features observed in<br />

the experimental P(1s) and P(2p) spectra <strong>of</strong> red-P, which were obtained at the DCM and<br />

Grasshopper beamlines <strong>of</strong> CSRF, respectively. They were measured simultaneously by total<br />

electron yield (TEY) and fluorescence yield (FY) with the incident photon beam at normal<br />

incidence to the sample surface.<br />

The results <strong>of</strong> this study were presented at the XAFS 12 conference, held in Malmö, Sweden in<br />

June <strong>of</strong> 2003. A paper has been submitted to Physica Scripta to be published in the conference<br />

proceedings. A temperature dependent P(1s) EXAFS study <strong>of</strong> red-P is in progress to determine<br />

the Debye-Waller factor (the mean square deviation <strong>of</strong> the interatomic distance) <strong>of</strong> the first<br />

coordination shell in red-P.<br />

0.1<br />

0.05<br />

0<br />

-0.05<br />

-0.1<br />

3 6 9 12 15<br />

P4 molecule<br />

theory<br />

0 1 2 3 4 5 6<br />

0.35<br />

P4 molecule<br />

0.28<br />

theory<br />

0.21<br />

0.14<br />

0.07<br />

(k)<br />

0.1<br />

0.05<br />

0<br />

-0.05<br />

-0.1<br />

0.1<br />

0.05<br />

0<br />

-0.05<br />

-0.1<br />

white-P<br />

theory<br />

black-P<br />

theory<br />

FT(k 3 (k)) (Å -4 )<br />

0<br />

0.28<br />

0.21<br />

0.14<br />

0.07<br />

0<br />

0.28<br />

0.21<br />

0.14<br />

0.07<br />

white-P<br />

theory<br />

black-P<br />

theory<br />

0.1<br />

0.05<br />

0<br />

-0.05<br />

red-P<br />

experiment<br />

0<br />

0.12<br />

0.09<br />

0.06<br />

red-P<br />

experiment<br />

-0.1<br />

0.03<br />

3 6 9 12 15<br />

k (Å -1 )<br />

0<br />

0 1 2 3 4 5 6<br />

R (Å)<br />

The experimental P(1s) (top) and P(2p)<br />

(bottom) XANES spectra <strong>of</strong> red-P<br />

The P(1s) EXAFS spectrum (left) and its Fourier Transform (right). The first<br />

peak in the FT represents the first coordination shell with an interatomic<br />

distance <strong>of</strong> 2.212 Å.


CHARACTERIZATION OF SULPHATE INTERACTIONS WITH<br />

HEMATITE MINERALS USING XANES<br />

T.Kotzer 1 , J.Cutler 1 , A.Pratt 2 , J.Dutrizac 2<br />

1 Canadian Light Source, <strong>University</strong> <strong>of</strong> Saskatchewan, Saskatoon, Canada<br />

2 Natural Resources Canada - CANMET, Ottawa, Canada<br />

Sulphur bearing hematite is a product <strong>of</strong> some novel hydrometallurgical sphalerite<br />

treatments. Leaching tests aimed at investigating the stability <strong>of</strong> the hematitic leach residue have<br />

shown that only a fraction <strong>of</strong> the sulphur present is removed. In this study sulphur bearing<br />

hematites were leached in a variety <strong>of</strong> media and examined using XANES Sulphur K-edge and<br />

L-edge spectroscopy at the Canadian <strong>Synchrotron</strong> <strong>Radiation</strong> Facility, <strong>SRC</strong>-Madison, WI, to<br />

address questions regarding the chemical environment, oxidation state and physical distribution<br />

(surficial versus intercrystalline) <strong>of</strong> the sulphur within the hematite minerals. XANES S K-edge<br />

FY and TEY and L-edge TEY spectra indicate that the sulphur predominantly occurs as sulphate<br />

having a formal oxidation state <strong>of</strong> S 6+ and is largely distributed within the hematite matrix at<br />

concentrations between approximately 0.04 and 4.7 %. The XANES S K- and L-edge absorption<br />

spectra have also been used to assess the effectiveness <strong>of</strong> leach treatments employing HNO 3 and<br />

NH 4 OH solutions for the removal <strong>of</strong> sulphate from the hematite minerals. Here, the spectra<br />

indicate that a 0.05M HNO 3 solution was largely ineffective at removal <strong>of</strong> surficial and<br />

intercrystalline sulphate whereas 1 to 4M NH 4 OH leach solutions appear to have variably<br />

removed sulphate from the uppermost regions (5 to 30 nm depth) <strong>of</strong> the hematite minerals.<br />

Overall integration <strong>of</strong> the XANES S K-edge and L-edge TEY and FY spectra provides both<br />

unique information regarding the distribution <strong>of</strong> sulphate within the hematite matrix and a means<br />

to evaluate the effectiveness <strong>of</strong> leach processes used to remove sulphate from hematite minerals.


Exploratory Experiments: Photodetachment <strong>of</strong> Negative Ions<br />

by Energetic Photons<br />

Thomas Kvale 1 , Song Cheng 1 , David Seely 2 , and Jeffrey Thompson 3<br />

1<br />

The <strong>University</strong> <strong>of</strong> Toledo, Toledo, OH 43606<br />

2 Albion College, Albion, MI 49224<br />

3<br />

The <strong>University</strong> <strong>of</strong> Nevada, Reno, NV 89557<br />

Abstract<br />

One <strong>of</strong> the current investigations in Atomic Physics is the electron-electron interaction in<br />

atoms by placing atoms in "exotic" states. A unique opportunity exists to perform a series <strong>of</strong><br />

experiments which will probe this interaction by examining continuum states in various atomic<br />

species by photodetaching negative ions. <strong>Synchrotron</strong> radiation is useful in providing a<br />

continuously tunable source <strong>of</strong> photons over the energy range <strong>of</strong> interest. Various theories [1, 2]<br />

predict the photodetachment cross sections are rich in features (both minima and resonances) in this<br />

photon energy range. This situation makes these studies ideally suited for providing stringent tests<br />

<strong>of</strong> our current understanding <strong>of</strong> atomic structure. The specific experiments listed in this abstract are<br />

representative <strong>of</strong> the types <strong>of</strong> experiments possible and are only the first in an extended series <strong>of</strong><br />

experiments that are currently <strong>of</strong> great interest in atomic physics.<br />

The first series <strong>of</strong> experiments that are proposed will probe the photon energy region from<br />

4eV to 20 eV in a variety <strong>of</strong> relatively light elements. Gribakin, et al. [1] predicted the<br />

photodetachment cross sections for the outer np and nearby ns subshells <strong>of</strong> C - 2p 34 S, Si - 3p 34 S, and<br />

Ge - 4p 34 S. The np photodetachment cross sections have minima around 5eV above the negative ion<br />

ground states and then rise to maxima at around 8 eV before gradually falling again. The minima<br />

have widths <strong>of</strong> order <strong>of</strong> 0.5 eV, whereas the maxima have widths <strong>of</strong> several eV. By contrast, most<br />

<strong>of</strong> the predicted [2] resonances<br />

in Be - ( 4 P o , 4 D o , and 4 S o ) have<br />

widths <strong>of</strong> order <strong>of</strong> 10-50 meV.<br />

These proposed experiments<br />

complement previous<br />

photodetachment experiments <strong>of</strong><br />

the authors at lower photon<br />

energies. With an estimated<br />

photon flux <strong>of</strong> 10 12 photon/s, an<br />

ion current <strong>of</strong> 100 nA, and the<br />

predicted photodetachment cross<br />

sections, we expect a<br />

photoelectron current <strong>of</strong><br />

approximately 10-1000 e - /s. An<br />

apparatus (PHOTO-2) has been<br />

constructed for conducting this<br />

series <strong>of</strong> photodetachment<br />

experiments (see Figure).<br />

Briefly, it consists <strong>of</strong>: a 0 - 50kV negative ion accelerator; a dodecapole chamber for merging the<br />

ion and photon beams; an interaction chamber <strong>of</strong> approximately 30 cm long; and data acquisition<br />

chamber. Photodetached electrons are energy analyzed by a hemispherical energy analyzer located


in the data acquisition chamber. The various residual ion, neutral, and photon beams exit through<br />

an aperture in the outer hemisphere <strong>of</strong> the analyzer and are detected by a Faraday cup detector. The<br />

interaction region has magnetic field coils to null the earth's magnetic field throughout this region.<br />

The predicted continuum state structure (energy locations, widths, and magnitudes) are very<br />

sensitive to specific atomic structure theory approximations, and experimental measurements such<br />

as the ones currently proposed will provide stringent tests <strong>of</strong> our understanding <strong>of</strong> the electronelectron<br />

and electron-nucleus interactions that occur in atoms and ions.<br />

References<br />

[1] G.F. Gribakin, A.A. Gribakina, B.V. Gul'tsev, and V.K. Ivanov, J. Phys. B: At., Mol., Opt.<br />

Phys. 25, 1757 (1992).<br />

[2] Jose' Luis Sanz-Vicario and Eva Lindroth, Phys. Rev. A 68, 012702 (2003).<br />

[3] D. Calabrese, A.M. Covington, D. Carpenter, J.S. Thompson, T.J. Kvale, and R. Collier, J.<br />

Phys. B: At., Mol., Opt. Phys. 30, 4791 (1997).<br />

[4] A.M. Covington, D. Calabrese, W.W. Williams, J.S. Thompson, and T.J. Kvale, Phys. Rev.<br />

A 56, 4746 (1997).<br />

[5] W.W. Williams, D.L. Carpenter, A.M. Covington, J.S. Thompson, T.J. Kvale, and D.G.<br />

Seely, Phys. Rev. A 58, 3582 (1998).


CHEMOMECHANICAL PROPERTIES OF ANTIWEAR FILMS<br />

USING X-RAY ABSORPTION MICROSCOPY AND<br />

NANOINDENTATION TECHNIQUES<br />

Mark A. Nicholls 1 , G. Michael Bancr<strong>of</strong>t 1 , Peter R. Norton 1 *, Masoud Kasrai 1 , Gelsomina<br />

De Stasio 2 , Bradley H. Frazer 2 , and Lisa M. Wiese 3<br />

1 Department <strong>of</strong> Chemistry, <strong>University</strong> <strong>of</strong> Western Ontario,<br />

London, Ontario, Canada N6A 2B7<br />

2 Department <strong>of</strong> Physics, <strong>University</strong> <strong>of</strong> Wisconsin-Madison, Madison, WI 53706, USA<br />

3 Syncrotron <strong>Radiation</strong> <strong>Center</strong>, Univ. <strong>of</strong> Wisconsin-Madison, Stoughton, WI 53589, USA<br />

ABSTRACT<br />

The first chemomechanical comparison between an antiwear film formed from a solution<br />

containing zinc dialkyl-dithiophophates (ZDDPs) to a solution containing ZDDP plus a<br />

detergent (ZDDPdet) has been performed. X-ray absorption near edge structure<br />

(XANES) analysis has shown a difference in the type <strong>of</strong> polyphosphate, between each<br />

film. The ZDDPdet film has been found<br />

to contain short-chain polyphosphates<br />

throughout and contain a large amount <strong>of</strong><br />

CaCO 3 . X-ray photoelectron emission<br />

microscopy (X-PEEM) has provided<br />

detailed spatially resolved<br />

microchemistry <strong>of</strong> the films. The large<br />

Intensity (arb. units)<br />

[A]<br />

[B]<br />

[C]<br />

[D]<br />

[E]<br />

[F]<br />

1<br />

2<br />

s'<br />

3<br />

b<br />

a<br />

c<br />

s 1<br />

unreacted<br />

ZDDP<br />

Zn 10<br />

P 18<br />

O 55<br />

Ca 3<br />

(PO 4<br />

) 2<br />

ZDDP + Det.<br />

X-PEEM<br />

long-chain<br />

internal model*<br />

shorter-chain<br />

internal model*<br />

map-ave.<br />

long-chain<br />

pads in the ZDDP antiwear film have<br />

[G]<br />

map-ave.<br />

shorter-chain<br />

long-chain polyphosphates at the surface<br />

[H]<br />

and shorter-chain polyphosphates are<br />

130 140 150 160<br />

Photon Energy (eV)<br />

found in the lower lying regions. The


spatially resolved chemistry <strong>of</strong> the ZDDPdet film was found to be short-chain calcium<br />

phosphate throughout.<br />

Fiducial marks allowed for<br />

the re-location <strong>of</strong> the same<br />

areas with an imaging<br />

nanoindenter. This allowed<br />

the nanoscale mechanical<br />

properties, <strong>of</strong> selected<br />

antiwear pads, to be<br />

measured on the same length scale. The indentation modulus <strong>of</strong> the ZDDP antiwear pads<br />

were found to be heterogeneous, ~120 GPa at the center and ~90 GPa at the edges. The<br />

ZDDPdet antiwear pads were found to be more uniform and have a similar indentation<br />

modulus <strong>of</strong> ~90 GPa. A theory explaining this measured similarity, which is based on the<br />

probing depths <strong>of</strong> all techniques used, sheds new insight into the structure and<br />

mechanical response <strong>of</strong> ZDDP antiwear films.


ELECTRONIC PROPERTIES OF MN-SILICIDE ON SI(111)<br />

J. J. Paggel, K. Schwinge, G. Ctistis, U. Deffke, and P. Fumagalli<br />

Institut für Experimentalphysik, Freie Universität Berlin, Germany<br />

T. Miller and T.-C. Chiang<br />

Department <strong>of</strong> Physics, <strong>University</strong> <strong>of</strong> Illinois at Urbana-Champaign, IL<br />

and<br />

Frederick-Seitz Materials Research Laboratory, <strong>University</strong> <strong>of</strong> Illinois at Urbana-Champaign, IL<br />

The electron spin as new degree <strong>of</strong> freedom in electronics is widely discussed. One<br />

central problem is spin-injection into the semiconductor material, i.e. the transfer <strong>of</strong> the spinpolarization<br />

from the magnetic electrode into the semiconductor. It is generally believed that<br />

structural disorder at the interface will lower the injection efficiency. In order to achieve high<br />

spin-polarization in the semiconductor, epitaxial, magnetic films on semiconductors are therefore<br />

sought after.<br />

Mn – in its conventional antiferromagnetic -phase – might be a promising candidate for<br />

magnetic films on silicon. Its fcc-phase is lattice matched to Si. There is even a large probability<br />

for Mn to be ferromagnetic at RT in this phase. Growing Mn-films on Si(111) yields epitaxial<br />

films with six-fold symmetry. Auger electron spectroscopy indicates a large Si-content <strong>of</strong> the<br />

films, maybe even a Si surface layer. RHEED and STM indicate closed films <strong>of</strong> good quality.<br />

The nature <strong>of</strong> these films could not be revealed using our laboratory experiments. Photoelectron<br />

spectroscopy experiments, reported here, identify the film as metallic Mn-silicide. The bonding<br />

character between Mn and Si is largely covalent. The Si2p core-level from the silicide is very<br />

simple and shows a striking similarity to the core-level <strong>of</strong> the famous Si(111)-(77)<br />

reconstruction, turning the interest from potential application <strong>of</strong> the films to basic research, as the<br />

electronic properties <strong>of</strong> the Si(111)-(77) are still under debate. The Si(111)-(77) reconstruction<br />

shows at least two nearly dispersionless surface states, associated with adatoms, which are also<br />

identified as origin for the prominent surface core-level <strong>of</strong> the Si(111)-(77) reconstruction. The<br />

Si2p core-level from the silicide shows a similar surface component, such that an adatom surface<br />

state was expected. The latter has not been found yet.<br />

This work was funded by the Deutsche Forschungsgemeinschaft through SFB 290 and<br />

grant Pa661/5-1, the U.S. National Science Foundation (grant DMR-02-03003), the U.S.<br />

Department <strong>of</strong> Energy, Division <strong>of</strong> Materials Sciences (grant DEFG02-91ER45439), and the<br />

Petroleum Research Fund, administered by the American Chemical Society. The <strong>Synchrotron</strong><br />

<strong>Radiation</strong> <strong>Center</strong> is supported by the National Science Foundation under grant DMR-00-84402.


THE RELATIONSHIP BETWEEN PHOSPHATE SPECIATION IN SOILS<br />

AND BIOSOLIDS/MANURE MANAGEMENT PRACTICES: A P K-EDGE<br />

XANES SPECTROSCOPIC STUDY<br />

Derek Peak,Gurpal Toor, and James T Sims<br />

1 Dept. <strong>of</strong> Soil Science, 51 Campus Dr, Univ. <strong>of</strong> Saskatchewan. Saskatoon SK S7N 5A8<br />

2 Dept Plant and Soil Sciences, Univ. Delaware. 149 Townsend Hall Newark DE 19717<br />

The long-term application <strong>of</strong> animal manures and biosolids has resulted in serious<br />

environmental issues in many locations throughout the world. For example, the state <strong>of</strong><br />

Delaware has a large amount <strong>of</strong> farmland characterized as having “excessive P levels” due to<br />

heavy application <strong>of</strong> poultry waste for the last three decades. These soils serve as a significant<br />

source for phosphate movement into waters, including the Chesapeake Bay. Elevated phosphate<br />

levels in natural waters are a serious concern because it is typically the limiting nutrient for algae<br />

and other photosynthetic organisms. Increased phosphate can therefore result in algal blooms<br />

and ultimately fish kills and eutrophication.<br />

The tendency <strong>of</strong> phosphorus to move from a soil to surface waters is highly dependent<br />

upon its chemical form. In soils, phosphate has a tendency to form insoluble metal phosphate<br />

solids and to form strong sorption complexes with iron and aluminum oxides. These reactions<br />

result in phosphate typically being held very strongly by soils. In biosolids and manures,<br />

however, high levels <strong>of</strong> soluble phosphate and readily degradable organic phosphates are found<br />

along with more stable organic phosphates and calcium phosphate minerals. To attempt to<br />

reduce the environmental risk <strong>of</strong> manure applications, scientists have proposed that (a) the<br />

manures be chemically amended prior to application to reduce P availability or (b) the diet <strong>of</strong> the<br />

animals be modified to reduce P in manures. The presented research will perform direct<br />

chemical samples from a variety <strong>of</strong> animal types, diets, and chemical amendments using P K-<br />

Edge XANES spectroscopy coupled with linear combination fitting. This allows us to directly<br />

assess the success <strong>of</strong> different strategies to reduce manage phosphorus.<br />

We found that changes in animal diet do indeed play a large role in the resulting<br />

phosphate speciation. Adding supplemental P <strong>of</strong>ten served to increase the fraction <strong>of</strong> calcium<br />

phosphate minerals present. Take for example the samples in Figure 1. The DM01 sample on<br />

the left is a dairy manure sample from an animal with 29%P in its diet. The sample on the right<br />

is a dairy manure sample from an animal fed 58% P. There is a marked increase in calcium<br />

phosphate precipitation in the higher P diets, which can affect the solubility <strong>of</strong> P compared to<br />

organic and aqueous forms. XANES has similarly been useful in determining the chemistry <strong>of</strong><br />

phosphate in manures amended with aluminum coagulants.<br />

This research was supported by an NSERC Discovery grant and the Saskatchewan<br />

<strong>Synchrotron</strong> Institute.


Figure 1. Comparison <strong>of</strong> phosphate speciation in dairy manures from cattle with different diets.<br />

Sample DM01 is from an animal with a low P diet while DM41 is from an animal with a diet<br />

high in P.


PHOTOEMISSION STUDIES OF QUANTUM WELL SYSTEMS WITH<br />

VICINAL INTERFACES<br />

D. Ricci, T. Miller, and T.-C. Chiang<br />

Department <strong>of</strong> Physics, <strong>University</strong> <strong>of</strong> Illinois at Urbana-Champaign, 1110 West Green Street,<br />

Urbana, Illinois 61801-3080<br />

Frederick Seitz Materials Research Laboratory, <strong>University</strong> <strong>of</strong> Illinois at Urbana-Champaign,<br />

104 South Goodwin Avenue, Urbana, Illinois 61801-2902<br />

The electronic structures <strong>of</strong> thin metallic films on semiconductor surfaces have been<br />

shown to be pr<strong>of</strong>oundly influenced by the films’ interfaces with the substrate and with the<br />

vacuum. As seen using synchrotron-based angle-resolved photoemission spectroscopy, wellordered<br />

systems grown on low-index faces may exhibit discrete electronic spectra characteristic<br />

<strong>of</strong> quantum wells [1]. In turn, the quantized electronic structure has been observed to impact<br />

various physical properties <strong>of</strong> the film [2]. Vicinal surfaces possess periodic arrangements <strong>of</strong><br />

steps which influence the growth and development <strong>of</strong> adsorbate structures [3], and thus may<br />

produce systems with electronic properties that differ from those formed with substrates oriented<br />

along a high symmetry axis [4]. Currently, we are probing the surface and quantum well states<br />

arising in Ag/vicinal Si(111) and Ag/vicinal Si(001) systems, with an eye towards developing an<br />

understanding <strong>of</strong> their electronic and physical properties. In this regard, comparison to the<br />

existing results from the simpler on-axis systems should prove useful.<br />

This work is supported by the U.S. Department <strong>of</strong> Energy, Division <strong>of</strong> Materials Sciences<br />

(grant DEFG02-91ER45439). We acknowledge the Petroleum Research Fund, administered by<br />

the American Chemical Society, and the U.S. National Science Foundation (grant DMR-02-<br />

03003), for partial support <strong>of</strong> the synchrotron beamline operation and the central facilities <strong>of</strong> the<br />

Frederick Seitz Materials Research Laboratory. The <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong> <strong>of</strong> the<br />

<strong>University</strong> <strong>of</strong> Wisconsin-Madison is supported by the U.S. National Science Foundation (grant<br />

DMR-00-84402.).<br />

References:<br />

[1] M. Upton, T. Miller, and T.-C. Chiang, Phys. Rev. B (in press); J. J. Paggel, T. Miller, and<br />

T.-C. Chiang, Science 283, 1709 (1999); T.-C. Chiang, Surf. Sci. Rep. 39, 181 (2000), and<br />

references therein.<br />

[2] J. J. Paggel, C. M. Wei, M. Y. Chou, D.-A. Luh, T. Miller, and T.-C. Chiang, Phys. Rev. B66<br />

233403 (2002); D.-A. Luh, T. Miller, J. J. Paggel, and T.-C. Chiang, Phys. Rev. Lett. 88,<br />

256802 (2002); D.-A. Luh, T. Miller, J. J. Paggel, M. Y. Chou, and T.-C. Chiang, Science,<br />

292 1131 (2001).<br />

[3] E. Hoque, A. Petkova, M. Henzler, Surf. Sci. 515, 312 (2002)<br />

[4] A.P. Shapiro, T. Miller, and T.-C. Chiang, Phys. Rev. B, 38, 1779 (1988)


GROWTH OF THIN FILMS ON RECONSTRUCTED SURFACES:<br />

AG ON AU/SI(111)- 3 3<br />

S. –J. Tang, T. Miller, and T.-C. Chiang<br />

Department <strong>of</strong> Physics, <strong>University</strong> <strong>of</strong> Illinois at Urbana-Champaign, 1110 West Green Street,<br />

Urbana, Illinois 61801-3080<br />

Frederick Seitz Materials Research Laboratory, <strong>University</strong> <strong>of</strong> Illinois at Urbana-Champaign,<br />

104 South Goodwin Avenue, Urbana, Illinois 61801-2902<br />

Both for fundamental studies and for potential applications, it is <strong>of</strong>tentimes desireable to<br />

be able to grow uniform films in a layer-by-layer fashion on a substrate <strong>of</strong> different material. The<br />

interface with the substrate is an important factor controlling the nature <strong>of</strong> film growth.<br />

Generally, epitaxial growth <strong>of</strong> thin films is facilitated by a lattice-matched starting substrate, as<br />

there is a tendency for the substrate to serve as a template for further crystalline growth <strong>of</strong> the<br />

film. In typical cases there exists lattice mismatch which must be accommodated by strain or by<br />

the development <strong>of</strong> crystal imperfections in the substrate, film, or at the interface; consequently it<br />

may be more difficult to achieve layer-by-layer growth in these systems. Similar problems may<br />

be caused by reconstructions <strong>of</strong> the substrate surface, which may be quite complex. In these<br />

cases, it may happen that the presence <strong>of</strong> a third species at the interface can facilitate a desired<br />

growth pattern.<br />

Ag films grown on the Si (111) surface provide examples <strong>of</strong> these issues. The<br />

equilibrium (111) Si surface has the well-known, but complicated, (7x7) reconstruction. Ag<br />

and Si both have fcc as their basic lattice but have large lattice mismatch with each other ( Si: a=<br />

5.43 Å, Ag: a= 4.09 Å). Depending on growth conditions, Ag on this surface can form flat or<br />

three-dimensional islands, or strained epitaxial layers. In the current work, we use Au to produce<br />

a commensurate Au/Si(111)- 3 3 reconstructed surface as a seed layer before depositing Ag<br />

films. This has the effect <strong>of</strong> simplifying the surface reconstruction as well as modifying its<br />

chemical properties [1]. Subsequent growth <strong>of</strong> Ag films <strong>of</strong> various coverages was carried out at<br />

-165C, and the films were annealed to RT. The electronic structure <strong>of</strong> the Ag thin film was<br />

probed by angle resolved photoemission using the high-resolution low energy synchrotron beam<br />

provided by U1-NIM beamline at <strong>SRC</strong>.<br />

Figure 1 shows a series <strong>of</strong> energy distribution curves at normal emission from Ag films<br />

on Au/Si - 3 3 at coverages ranging from 6 to 22 monolayers (ML). Each spectrum shows a<br />

series <strong>of</strong> discrete peaks representing the states <strong>of</strong> the quantum well formed by the thin metallic<br />

film. As is evident, the quantum well states peaks become narrower and shift to lower binding<br />

energies as the Ag film coverage increases. This coverage dependent behavior <strong>of</strong> the quantum<br />

well states can be simply explained through the phase accumulation model. The appearance <strong>of</strong><br />

the intense and sharp Ag surface state near Fermi level is significant as it not only reflects good<br />

crystallinity <strong>of</strong> the film but also the important role <strong>of</strong> the intermediate 3 3 seed layer in<br />

reducing the strain in the Ag film. This is because in films grown without the seed layer, the<br />

strain due to the Ag/Si lattice mismatch is known to significantly shift the initial state energy <strong>of</strong>


Ag surface state such that it moves above the Fermi level and becomes depopulated and<br />

significantly less intense [2].<br />

Figure 2 shows the in-plane dispersion <strong>of</strong> the quantum well states and the surface state<br />

along the - K direction for a 22 ML thick Ag film. The bottom <strong>of</strong> the Ag surface state band is<br />

26 meV above the Fermi level. A fit<br />

Figure1<br />

Figure2<br />

7.5<br />

hv=22eV<br />

Ag on Au/Si- root3<br />

0.0<br />

Ag 22Ml Au/Si<br />

n=1<br />

7.0<br />

-0.2<br />

Photoemission Intensity (arb.units)<br />

6.5<br />

6.0<br />

5.5<br />

Ag ML<br />

22<br />

19<br />

16<br />

14<br />

12<br />

Initial state energy (eV)<br />

-0.4<br />

-0.6<br />

-0.8<br />

-1.0<br />

n=2<br />

n=3<br />

5.0<br />

10<br />

8<br />

-1.2<br />

n=4<br />

-3 -2 -1 0<br />

Initial state energy (eV)<br />

6<br />

-1.4<br />

-0.3 -0.2 -0.1 0.0 0.1 0.2 0.3<br />

k || (A -1 )<br />

to the data (solid curve) reveals the effective mass m*/m <strong>of</strong> the surface state band to be 0.404<br />

which is very close to the value 0.397 measured by F. Reinert et al [3] from the (111) surface <strong>of</strong><br />

bulk Ag. The dispersions <strong>of</strong> the quantum well states (crosses) can be compared with those <strong>of</strong> Ag<br />

bulk bands calculated at constant k (triangles) [4]. The match between the measurement and<br />

calculation is good for the quantum-well states with quantum numbers n from n = 2 to n = 4. For<br />

the quantum-well state with n = 1, the dispersion is significantly less than that <strong>of</strong> the other<br />

quantum well states and corresponding calculated bulk value. We note that the deeper quantumwell<br />

states overlap the bulk bands in the substrate and so are really resonances; for such states<br />

the reflection phase shift at the interface is relatively slowly-varying and so the dispersions in<br />

Fig. 2 are approximately at constant k . In contrast, the energy <strong>of</strong> the n = 1 state is within the<br />

energy gap <strong>of</strong> the substrate, where the interface phase shift is rapidly varying. In this case, the<br />

dispersion in Fig. 2 includes an implicit variation in k so that the curvature may not reflect the<br />

true effective mass. Further analysis <strong>of</strong> the dispersions and effective masses <strong>of</strong> quantum well<br />

states from different other coverages will be done.<br />

This work is supported by the U.S. National Science Foundation (grant DMR-02-03003).<br />

We acknowledge the Petroleum Research Fund, administered by the American Chemical<br />

Society, and the U.S. Department <strong>of</strong> Energy, Division <strong>of</strong> Materials Sciences (grant DEFG02-<br />

91ER45439), for partial support <strong>of</strong> the synchrotron beamline operation and the central facilities<br />

<strong>of</strong> the Frederick Seitz Materials Research Laboratory. The <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong> <strong>of</strong> the


<strong>University</strong> <strong>of</strong> Wisconsin-Madison is supported by the U.S. National Science Foundation (grant<br />

DMR-00-84402.).<br />

References:<br />

1. H. M. Zhang, K. Sakamoto, and R. I. G. Uhrberg, Phys. Rev. B64, 245421 (2001).<br />

2. G. Neuhold and K. Horn, Phys. Rev. Lett. 78, 1327 (1997).<br />

3. F. Reinert, G. Nicolay, S. Schmidt, D. Ehm, and S. Hüfner, Phys. Rev. B63, 115415<br />

(2001).<br />

4. N. V. Smith and L. F. Matheiss, Phys. Rev. B 9, 1341 (1974); N. V. Smith, Phys. Rev. B<br />

9, 1365 (1974)


HIGH-RESOLUTION STUDY OF THE LITHIUM INNER-SHELL<br />

RESONANCES<br />

R. Wehlitz 1 , J.B. Bluett 1 , and S.B.Whitfield 2<br />

1 <strong>SRC</strong>, UW-Madison, 3731 Schneider Dr., Stoughton, WI 53589<br />

2 Department <strong>of</strong> Physics and Astronomy, UW-Eau Claire, WI 54702<br />

The resonance region <strong>of</strong> lithium (58 – 81 eV) was investigated using monochromatized<br />

synchrotron radiation and ion time-<strong>of</strong>-flight (TOF) spectrometry. The photon beam <strong>of</strong><br />

the PGM undulator beam line passed through the experimental chamber containing<br />

vaporized lithium atoms. The lithium ions created were accelerated by a pulsed electric<br />

field into the TOF flight tube and detected by a Z-stack <strong>of</strong> microchannel plates [1].<br />

We have observed the autoionizing resonances to higher principal quantum numbers than<br />

in previous experiments [2,3]. Depending on the energy region we have used an energy<br />

resolution <strong>of</strong> 10 meV and 4.5 meV with step sizes as small as 0.5 meV.<br />

Several spectra were<br />

taken over short<br />

energy intervals as<br />

indicated by the<br />

different colors <strong>of</strong> the<br />

composed spectrum<br />

shown in Fig. 1. All<br />

spectra were<br />

normalized with<br />

respect to the photon<br />

flux and background<br />

corrected. The fit<br />

results along with the<br />

data <strong>of</strong> other authors<br />

will be presented for<br />

the various Rydberg<br />

series on the poster.<br />

This work was supported by NSF Grant No. 9987638. The <strong>SRC</strong> is operated under NSF<br />

Grant No. DMR-0084402.<br />

References:<br />

[1] R. Wehlitz, D. Luki, C. Koncz, and I.A. Sellin, Rev. Sci. Instrum. 73, 1671 (2002).<br />

[2] G. Mehlman, J.W. Cooper and E.B. Saloman, Phy. Rev. A 25, 2113 (1982).<br />

[3] L.M. Kiernan et al., J. Phys. B 29, L181 (1996).


RESONANCE PARAMETERS OF AUTOIONIZING Be 2pnl STATES<br />

R. Wehlitz 1 , D. Luki 2 , and J.B. Bluett 1<br />

1 <strong>SRC</strong>, UW-Madison, 3731 Schneider Dr., Stoughton, WI 53589<br />

2 Institute <strong>of</strong> Physics, 11001 Belgrade, Serbia and Montenegro<br />

While photoionization <strong>of</strong> helium has been studied thoroughly, beryllium (1s 2 2s 2 ), the<br />

next helium-like atom in the periodic table, has been investigated only marginally by<br />

comparison. Double excitations in the Be valence-shell region have been studied<br />

experimentally as well as theoretically in the past. In these experiments [1,2] vacuum<br />

sparks were used to photo-excite and –ionize Be atoms and absorption spectra were<br />

recorded on high-sensitive film.<br />

In general, for atomic photoexcitation resonances above the first ionization limit,<br />

autoionization becomes possible by interaction with one or more single-ionization<br />

continua. This leads to an<br />

asymmetric resonance pr<strong>of</strong>ile in the<br />

single-ionization cross section. A<br />

theoretical description <strong>of</strong> this process<br />

was introduced by Fano [3] and<br />

refined later by Shore and Starace.<br />

Since autoionization is a<br />

consequence <strong>of</strong> electron correlation,<br />

a measurement <strong>of</strong> the resonance<br />

pr<strong>of</strong>ile for comparison with theory<br />

can provide important information<br />

towards our understanding <strong>of</strong> how<br />

electron correlations affect a simple<br />

system. We will present our<br />

autoionization-pr<strong>of</strong>ile measurements<br />

<strong>of</strong> the Be 2pns (n=3-8) and 2pnd (n=3-5) double excitations [4]. Since the 1s electrons do<br />

not actively participate in the autoionization process, Be appears to be a system only<br />

slightly more complicated than He. However it is very different from He ins<strong>of</strong>ar as the<br />

series <strong>of</strong> autoionization resonances starts immediately above the first ionization threshold<br />

(see Fig. 1). Another difference to He is that the Be 2pns resonances are much broader<br />

due to a strong coupling to a rather weak continuum.<br />

This work was supported by NSF Grant No. 9987638. The <strong>SRC</strong> is operated under NSF<br />

Grant No. DMR-0084402.<br />

References:<br />

[1] G. Mehlman-Ball<strong>of</strong>fet and J.M. Esteva, Astrophys. J. 157, 945 (1969).<br />

[2] J.M. Esteva, G. Mehlman-Ball<strong>of</strong>fet, and J. Romand, J. Quant. Spectroc. Radiat.<br />

Transfer 12, 1291 (1972).<br />

[3] U. Fano, Phys. Rev. 124, 1866 (1961).<br />

[4] R. Wehlitz, D. Lukic, and J.B. Bluett, Phys. Rev. A, in press (2003).


RELATIVE PARTIAL CROSS SECTIONS AND ANGULAR<br />

DISTRIBUTION PARAMETERS OF THE<br />

LI 1S2S( 3,1 S) MAIN AND 1S2P( 3,1 P) CONJUGATE SHAKE-UP LINES IN<br />

THE REGION OF THE 1S3LN'L' AUTOIONIZING RESONANCES<br />

Scott B. Whitfield 1 and Ralf Wehlitz 2<br />

1 Department <strong>of</strong> Physics and Astronomy, <strong>University</strong> <strong>of</strong> Wisconsin, Eau Claire, WI 54701<br />

2 <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong>, <strong>University</strong> <strong>of</strong> Wisconsin, Stoughton, WI 53589<br />

Atomic Li is the simplest open-subshell atom, and is therefore an ideal candidate for<br />

detailed experimental and theoretical studies. Despite the extensive experimental work which has<br />

been carried out on atomic Li, there is only one rather old measurement [1] <strong>of</strong> the 1s2s( 1,3 S) main<br />

and 1s2p( 1,3 P) conjugate shake-up photolines in the region <strong>of</strong> the strong 1s3ln'l autoionizing<br />

resonances. We will present recent measurements <strong>of</strong> both the absolute partial cross sections, ,<br />

and the angular distribution parameters, , <strong>of</strong> these photoelectron lines in this resonance region.<br />

Our high resolution constant-ionic state spectra completely separate the dynamic behavior <strong>of</strong> all<br />

four photolines allowing a detailed comparison to a state-<strong>of</strong>-the-art R-matrix calculation [2]. The<br />

strong variations <strong>of</strong> both and <strong>of</strong> these lines across the 1s3ln'l autoionizing resonances are in<br />

generally excellent agreement with theory. This is illustrated in the figures above for the partial<br />

cross section <strong>of</strong> the 1s2s( 3 S) mainline and the 1s2s( 3 P) conjugate shakeup line.<br />

This work was supported by a Research Corporation College Cottrell Grant No. CC5243.<br />

The <strong>SRC</strong> is operated under Grant No. DMR-0084402.<br />

References:<br />

[1] T.A. Ferrett, D.W. Lindle, P.A. Heimann, W.D. Brewer, U. Becker, H.G. Kerh<strong>of</strong>f, and D. A.<br />

Shirley, Phys. Rev. A, 36, 3172 (1987).<br />

[2] L. Vo Ky, P. Faucher, A. Hibbert, J.M. Li, Y.Z. Qu, J. Yan, J.C. Chang, and F. Bely-Dubau,<br />

Phys. Rev. A, 57, 1045 (1998).


P<br />

A<br />

R<br />

T<br />

I<br />

C<br />

I<br />

P<br />

A<br />

N<br />

T<br />

S


Christian Ast<br />

Max-Planck Institute fhr Festk`rperforschung<br />

Heisenbergstra8e 1<br />

70569 Stuttgart<br />

Germany<br />

Phone: 011-49-711-689-1728<br />

FAX: 011-49-711-689-1662<br />

Email: c.ast@fkf.mpg.de<br />

Franco Cerrina<br />

<strong>Center</strong> for NanoTechnology and<br />

Department <strong>of</strong> Elec. & Comp. Engineering<br />

Room 2442 Engineering Hall<br />

Madison, WI 53706<br />

Phone: (608) 263-4955; FAX: (608) 262-0971<br />

Email: cerrina@nanotech.wisc.edu<br />

Matt Daniels<br />

<strong>University</strong> <strong>of</strong> Wisconsin-Madison, Physics<br />

C/o <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong><br />

3731 Schneider Drive<br />

Stoughton, WI 53589<br />

Phone: (608) 877-2382; FAX: (608) 877-2001<br />

Email: mjdaniels@wisc.edu<br />

Blake M. Dirksen<br />

Loras College<br />

1450 Alta Vista<br />

Dubuque, IA 52001<br />

Phone: (563) 588-7581; FAX: (563) 557-4070<br />

Email: bmdirksen@wisc.edu<br />

Tomasz Durakiewicz<br />

Los Alamos National Laboratory<br />

MST-10 Group, Mailstop K764<br />

Los Alamos, NM 87545<br />

Phone: (505) 259-4767; FAX: (505) 665-7652<br />

Email: tomasz@lanl.gov<br />

Rob Erhardt<br />

<strong>University</strong> <strong>of</strong> Wisconsin-Madison, Physics<br />

C/o <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong><br />

3731 Schneider Drive<br />

Stoughton, WI 53589<br />

Phone: (608) 877-2382; FAX: (608) 877-2001<br />

Email: roberhardt@yahoo.com<br />

<strong>SRC</strong> Users’ Group <strong>Meeting</strong><br />

October 25, 2003<br />

PARTICIPANTS


Cherice M. Evans<br />

Department <strong>of</strong> Chemistry<br />

Queens College – CUNY<br />

Flushing, NY 111367<br />

Phone: (718) 997-4216; FAX: (718) 997-5531<br />

Email: cevans@forbin.qc.edu<br />

Brad Frazer<br />

Department <strong>of</strong> Physics<br />

<strong>University</strong> <strong>of</strong> Wisconsin-Madison<br />

Madison, WI 53706<br />

Phone: (608) 877-2366; FAX: (608) 877-2001<br />

Email: bhfrazer@wisc.edu<br />

Viktoriya Golovkina<br />

<strong>Center</strong> for NanoTechnology<br />

<strong>University</strong> <strong>of</strong> Wisconsin-Madison<br />

3731 Schneider Drive<br />

Stoughton, WI 53589<br />

Phone: (608) 877-2429; FAX: (608) 877-2401<br />

Email: golovkina@nanotech.wisc.edu<br />

Franz Himpsel<br />

<strong>University</strong> <strong>of</strong> Wisconsin-Madison<br />

Department <strong>of</strong> Physics<br />

Room B317 Sterling Hall<br />

Madison, WI 53706-1390<br />

Phone: (608) 263-5590; FAX: (608) 265-2334<br />

Email: fhimpsel@facstaff.wisc.edu<br />

Carol Hirschmugl<br />

<strong>University</strong> <strong>of</strong> Wisconsin-Milwaukee<br />

Department <strong>of</strong> Physics<br />

P.O. Box 413<br />

Milwaukee, WI 53201<br />

Phone: (414) 229-5748; FAX: (414) 229-5589<br />

Email: cjhirsch@csd.uwm.edu<br />

Anphong Ho<br />

<strong>Center</strong> for NanoTechnology<br />

<strong>University</strong> <strong>of</strong> Wisconsin-Madison<br />

3731 Schneider Drive<br />

Stoughton, WI 53589<br />

Phone: (608) 877-2421; FAX: (608) 877-2401<br />

Email: ho@nanotech.wisc.edu<br />

Yongfeng Hu<br />

Canadian <strong>Synchrotron</strong> <strong>Radiation</strong> Facility<br />

C/o <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong><br />

3731 Schneider Drive<br />

Stoughton, WI 53589<br />

Phone: (608) 877-2229; FAX: (608) 877-2001<br />

Email: yfhu@facstaff.wisc.edu


David L. Huber<br />

Physical Sciences Laboratory<br />

<strong>University</strong> <strong>of</strong> Wisconsin-Madison<br />

3725 Schneider Drive<br />

Stoughton, WI 53589<br />

Phone: (608) 877-2250; FAX: (608) 877-2001<br />

Email: dhuber@psl.wisc.edu<br />

Barbara Illman<br />

<strong>University</strong> <strong>of</strong> Wisconsin-Madison<br />

John Joyce<br />

Los Alamos National Laboratory<br />

Mailstop K764<br />

Los Alamos, NM 87545<br />

Phone: (505) 667-6431; FAX: (505) 665-7652<br />

Email: jjoyce@lanl.gov<br />

Astrid Jurgensen<br />

Canadian <strong>Synchrotron</strong> <strong>Radiation</strong> Facility<br />

C/o <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong><br />

3731 Schneider Drive<br />

Stoughton, WI 53589<br />

Phone: (608) 877-2331; FAX: (608) 877-2001<br />

Email: ajurgensen@src.wisc.edu<br />

Adam Kaminski<br />

Department <strong>of</strong> Physics<br />

<strong>University</strong> <strong>of</strong> Wales Swansea<br />

Singleton Park<br />

Swansea, SA2 8PP United Kingdom<br />

Phone: (011) 44 (0) 1792-295016<br />

FAX: (011) 44 (0) 1792-295324<br />

Email: a.Kaminski@swansea.ac.uk<br />

Thomas Kotzer<br />

Canadian Light Source<br />

<strong>University</strong> <strong>of</strong> Saskatchewan<br />

101 Perimeter Road<br />

Saskatoon, SK S7N OX4 CANADA<br />

Phone: (306) 657-3594; FAX: (306) 657-3535<br />

Email: tom.kotzer@lightsource.ca<br />

Brian Krosschell<br />

Dept. <strong>of</strong> Physics and Astronomy<br />

<strong>University</strong> <strong>of</strong> Wisconsin-Eau Claire<br />

459 Sutherland Hall<br />

620 Hilltop Circle Drive<br />

Eau Claire, WI 54701<br />

Phone: (715) 836-4075; FAX: (715) 836-3955<br />

Email: krosscbd@uwec.edu


Thomas Kvale<br />

Department <strong>of</strong> Physics & Astronomy, M/S 111<br />

<strong>University</strong> <strong>of</strong> Toledo<br />

Toledo, OH 43606<br />

Phone: (419) 530-2980; FAX: (419) 530-2723<br />

Email: tjk@physics.utoledo.edu<br />

Uday Lanke<br />

Saskatchewan <strong>University</strong><br />

C/o <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong><br />

3731 Schneider Drive<br />

Stoughton, WI 53589<br />

Phone: (608) 877-2344; FAX: (608) 877-2001<br />

Email: ulanke@src.wisc.edu<br />

Gerald J. Lapeyre<br />

Department <strong>of</strong> Physics, EPS 264 Building<br />

Montana State <strong>University</strong><br />

Bozeman, MT 50717-0350<br />

Phone: (406) 994-6153; FAX: (406) 994-4452<br />

Email: lapeyre@physics.montana.edu<br />

Zhiwei Li<br />

Materials Science<br />

<strong>University</strong> <strong>of</strong> Wisconsin-Madison<br />

Madison, WI 53706<br />

Phone: (608) 877-2348<br />

Email: lizw@cae.wisc.edu<br />

Jessica McChesney<br />

Department <strong>of</strong> Physics<br />

<strong>University</strong> <strong>of</strong> Wisconsin-Madison<br />

1150 <strong>University</strong> Avenue<br />

Madison, Wi 53706<br />

Phone: (608) 262-5047; FAX: (608) 265-2334<br />

Email: jlmcchesney@wisc.edu<br />

David McIlroy<br />

Department <strong>of</strong> Physics<br />

Engineering and Physics Building<br />

<strong>University</strong> <strong>of</strong> Idaho<br />

Moscow, ID 83844-0903<br />

Phone: (208) 885-6809; FAX: (208) 885-4055<br />

Email: dmcilroy@uidaho.edu<br />

Kenneth McLaughlin, Ph.D.<br />

Department <strong>of</strong> Physics and Engineering<br />

Loras College<br />

1450 Alta Vista<br />

Dubuque, IA 52001<br />

Phone: (563) 588-7581; FAX: (563) 557-4070<br />

Email: Kenneth.mclaughlin@loras.edu


Tom Miller<br />

<strong>University</strong> <strong>of</strong> Illinois at Urbana-Champaign<br />

C/o <strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong><br />

3731 Schneider Drive<br />

Stoughton, WI 53589<br />

Phone: (608) 877-2342; FAX: (608) 877-2001<br />

Email: tamille1@facstaff.wisc.edu<br />

Paul Nealey<br />

Dept. <strong>of</strong> Chemical Engineering<br />

<strong>University</strong> <strong>of</strong> Wisconsin-Madison<br />

Room 3020 Engineering Hall<br />

Madison, WI 53706<br />

Phone: (608) 265-8171<br />

Email: nealey@engr.wisc.edu<br />

Cliff Olson<br />

Ames Laboratory<br />

C/o Physical Sciences Laboratory<br />

3725 Schneider Drive<br />

Stoughton, WI 53589<br />

Phone: (608) 877-2224; FAX: (608) 877-2001<br />

Email: cgolson@facstaff.wisc.edu<br />

Marshall Onellion<br />

Department <strong>of</strong> Physics<br />

<strong>University</strong> <strong>of</strong> Wisconsin-Madison<br />

475 N. Charter Strett<br />

Madison, WI 53705<br />

Phone: (608) 263-6829; FAX: (608) 265-2334<br />

Email: onellion@facstaff.wisc.edu<br />

Derek Peak<br />

Department <strong>of</strong> Soil Science<br />

<strong>University</strong> <strong>of</strong> Saskatchewan<br />

51 Campus Drive<br />

Saskatoon SK S7N 5A8 CANADA<br />

Phone: (306) 966-6806; FAX: (306) 966-6881<br />

Email: derek.peak@usask.ca<br />

George Sawatzky<br />

Department <strong>of</strong> Physics<br />

<strong>University</strong> <strong>of</strong> British Columbia<br />

Vancouver<br />

Email: sawatzky@physics.ubc.ca<br />

Steven Sahyun<br />

Department <strong>of</strong> Physics<br />

<strong>University</strong> <strong>of</strong> Wisconsin-Whitewater<br />

800 W. Main Street<br />

Whitewater, WI 53190<br />

Phone: (262) 472-5113<br />

Email: sahyuns@uww.edu


David Seely<br />

Department <strong>of</strong> Physics<br />

Albion College<br />

Albion, MI 492242<br />

Phone: (517) 629-0267; FAX: (517) 629-0264<br />

Email: dseely@albion.edu<br />

Steve Southworth<br />

Building 203<br />

Argonne National Laboratory<br />

Argonne, IL 60439<br />

Phone: (630) 252-3894; FAX: (630) 252-6210<br />

Email: southworth@anl.gov<br />

Kim Tan<br />

Canadian <strong>Synchrotron</strong> <strong>Radiation</strong> Facility<br />

C/o Physical Sciences Laboratory<br />

3725 Schneider Drive<br />

Stoughton, WI 53589<br />

Phone: (608) 877-2229; FAX: (608) 877-2001<br />

Email: ktan2@facstaff.wisc.edu<br />

Don Thielman<br />

<strong>Center</strong> for NanoTechnology<br />

<strong>University</strong> <strong>of</strong> Wisconsin-Madison<br />

3731 Schneider Drive<br />

Stoughton, WI 53589<br />

Phone: (608) 877-2412; FAX: (608) 877-2401<br />

Email: thielman@nanotech.wisc.edu<br />

Josh Thomas<br />

<strong>University</strong> <strong>of</strong> Toledo<br />

2732 Kenwood Blvd.<br />

Toledo, OH 43606<br />

Phone: (419) 277-2753<br />

Email: jthomas5@pop3.utoledo.edu<br />

James G. Tobin<br />

Lawrence Livermore National Laboratory<br />

7000 East Avenue, P. O. Box 808, L-356<br />

Livermore, CA 94550<br />

Phone: (925) 422-7247; FAX: (925) 423-7040<br />

Email: tobin1@llnl.gov<br />

Scott Whitfield<br />

Dept. <strong>of</strong> Physics and Astronomy<br />

<strong>University</strong> <strong>of</strong> Wisconsin-Eau Claire<br />

105 Garfield Street<br />

Eau Claire, WI 54702<br />

Phone: (715) 836-2199; FAX: (715) 836-3955<br />

Email: whitfisb@uwec.edu


Eric Wiedemann<br />

<strong>University</strong> <strong>of</strong> Wisconsin-Madison<br />

927 Chandler Street<br />

Madison, WI 53715<br />

Phone: (608) 347-4033<br />

Email: ewiedemann@wisc.edu<br />

David Williams<br />

Advanced Photon Source<br />

9700 S. Cass Avenue<br />

Argonne, IL 60439<br />

Fan Zheng<br />

Department <strong>of</strong> Physics<br />

<strong>University</strong> <strong>of</strong> Wisconsin-Madison<br />

1150 <strong>University</strong> Avenue<br />

Madison, WI 53706<br />

Phone: (608) 262-5047; FAX: (608) 265-2334<br />

Email: fzheng@wisc.edu<br />

PROGRAM ADVISORY COMMITTEE MEMBERS<br />

David L. Ederer<br />

Department <strong>of</strong> Physics<br />

Tulane <strong>University</strong><br />

2001 P. Stern Science <strong>Center</strong><br />

New Orleans, LA 70118<br />

Phone: (504) 865-5520; FAX: (504) 862-8702<br />

Email: dlederer@mailhost.tcs.tulane.edu<br />

Carolyn A. Larabell<br />

Lawrence Berkeley Laboratory<br />

Box 0452, S 1327<br />

Department <strong>of</strong> Anatomy<br />

One Cyclotron Road<br />

Berkeley, CA 94720<br />

Phone: (415) 514-0423; FAX: (415) 476-4856<br />

Email: larabel@itsa.ucsf.edu<br />

Roy F. Willis<br />

Department <strong>of</strong> Physics<br />

Davey Laboratory<br />

Pennsylvania State <strong>University</strong><br />

<strong>University</strong> Park, PA 16802<br />

Phone: (814) 865-6101; FAX: (814) 865-3604<br />

Email: rfw4@psu.edu


Linda Young<br />

Argonne National Laboratory<br />

Chemistry Division<br />

Building 203, Room F125<br />

9700 South Cass Avenue<br />

Argonne, IL 60439<br />

Phone: (630) 252-8878; FAX: (630) 252-6210<br />

Email: young@anlphy.phy.anl.gov<br />

<strong>SRC</strong> STAFF AND STUDENTS<br />

<strong>Synchrotron</strong> <strong>Radiation</strong> <strong>Center</strong><br />

<strong>University</strong> <strong>of</strong> Wisconsin-Madison<br />

3731 Schneider Drive<br />

Stoughton, WI 53589<br />

Phone: (608) 877-2000; FAX: (608) 877-2001<br />

Joe Bisognano<br />

Phone: (608) 877-2163<br />

Email: jbisognano@src.wisc.edu<br />

Mark Bissen<br />

Phone: (608) 877-2146<br />

Email: mbissen@src.wisc.edu<br />

Jacques Bluett<br />

Phone: (608) 658-3646<br />

Email: jbluett@src.wisc.edu<br />

Juan Carlos Campuzano<br />

Phone: (608) 877-2382<br />

Email: jcc@uic.edu<br />

Gelsomina De Stasio<br />

Phone: (608) 877-2381<br />

Email: pupa@src.wisc.edu<br />

Mike Fisher<br />

Phone: (608) 877-2148<br />

Email: mfisher@src.wisc.edu<br />

Sergey Gorovikov<br />

Phone: (608) 877-2339<br />

Email: sgorovikov@src.wisc.edu<br />

Mike Green<br />

Phone: (608) 877-2159<br />

Email: mgreen@src.wisc.edu<br />

Chad Gundelach<br />

Phone: (608) 877-2141<br />

Email: cgundelach@src.wisc.edu


Pete Hagen<br />

Phone: (608) 877-2154<br />

Email: phagen@src.wisc.edu<br />

Roger Hansen<br />

Phone: (608) 877-2143<br />

Email: rhansen@src.wisc.edu<br />

Richard Hatch<br />

Phone: (608) 877-2369<br />

Email: rhatch@src.wisc.edu<br />

Ken Jacobs<br />

Phone: (608) 877-2142<br />

Email: kjacobs@src.wisc.edu<br />

Pavle Jaranic<br />

Phone: (608) 877-2000<br />

Email: pjaranic@src.wisc.edu<br />

Bob Julian<br />

Phone: (608) 877-2158<br />

Email: rjulian@src.wisc.edu<br />

Michelle Kirch<br />

Phone: (608) 877-2135<br />

Email: mkirch@src.wisc.edu<br />

Pamela Layton<br />

Phone: (608) 877-2134<br />

Email: playton@src.wisc.edu<br />

Bob Legg<br />

Phone: (608) 877-2162<br />

Email: rlegg@src.wisc.edu<br />

Chris Moore<br />

Phone: (608) 877-2137<br />

Email: cmoore@src.wisc.edu<br />

Bruce Neumann<br />

Phone: (608) 877-2157<br />

Email: bneumann@src.wisc.edu<br />

Esther Olson<br />

Phone: (608) 877-2295<br />

Email: eolson@pslsrc.wisc.edu<br />

John Stott<br />

Phone: (608) 877-2151<br />

Email: jstott@src.wisc.edu<br />

James Taylor<br />

Phone: (608) 877-2403<br />

Email: jwtaylor@src.wisc.edu


Dan Wallace<br />

Phone: (608) 877-2139<br />

Email: dwallace@src.wisc.edu<br />

Ralf Wehlitz<br />

Phone: (608) 877-2164<br />

Email: rwehlitz@src.wisc.edu


NOTES

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!