04.08.2013 Views

Impurities in a Homogeneous Electron Gas - Physics at Oregon ...

Impurities in a Homogeneous Electron Gas - Physics at Oregon ...

Impurities in a Homogeneous Electron Gas - Physics at Oregon ...

SHOW MORE
SHOW LESS

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

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

<strong>Impurities</strong> <strong>in</strong> a <strong>Homogeneous</strong> <strong>Electron</strong> <strong>Gas</strong><br />

by<br />

Jung-Hwan Song<br />

A THESIS<br />

submitted to<br />

<strong>Oregon</strong> St<strong>at</strong>e University<br />

<strong>in</strong> partial fulfillment of<br />

the requirements for the<br />

degree of<br />

Doctor of Philosophy<br />

Presented December 7, 2004<br />

Commencement June 2005


AN ABSTRACT OF THE THESIS OF<br />

Jung-Hwan Song for the degree of Doctor of Philosophy <strong>in</strong> <strong>Physics</strong> presented on<br />

December 7, 2004.<br />

Title: <strong>Impurities</strong> <strong>in</strong> a <strong>Homogeneous</strong> <strong>Electron</strong> <strong>Gas</strong><br />

Abstract approved:<br />

Henri J. F. Jansen<br />

Immersion energies for an impurity <strong>in</strong> a homogeneous electron gas with a uni-<br />

form positive background charge density have been calcul<strong>at</strong>ed numerically us<strong>in</strong>g<br />

density functional theory. The numerical aspects of this problem are very de-<br />

mand<strong>in</strong>g and have not been properly discussed <strong>in</strong> previous work. The numerical<br />

problems are rel<strong>at</strong>ed to approxim<strong>at</strong>ions of <strong>in</strong>f<strong>in</strong>ity and cont<strong>in</strong>uity, and they have<br />

been corrected us<strong>in</strong>g physics based on the Friedel sum rule and Friedel oscill<strong>at</strong>ions.<br />

The numerical precision is tested extensively. Immersion energies are obta<strong>in</strong>ed for<br />

non-sp<strong>in</strong>-polarized systems, and are compared with published d<strong>at</strong>a. Numerical<br />

results, such as phase shifts, density of st<strong>at</strong>es, dielectric constants, and compress-<br />

ibilities, are obta<strong>in</strong>ed and compared with analytical theories. Immersion energies<br />

for excited systems are obta<strong>in</strong>ed by vary<strong>in</strong>g the number of electrons <strong>in</strong> the bound<br />

st<strong>at</strong>es of an impurity. The model is extended to sp<strong>in</strong>-polarized systems and is<br />

tested <strong>in</strong> detail for a carbon impurity. The sp<strong>in</strong>-coupl<strong>in</strong>g with an external mag-<br />

netic field is considered ma<strong>in</strong>ly for a hydrogen impurity. These new results show<br />

very <strong>in</strong>terest<strong>in</strong>g behavior <strong>at</strong> low densities.


c○ Copyright by Jung-Hwan Song<br />

December 7, 2004<br />

All Rights Reserved


<strong>Impurities</strong> <strong>in</strong> a <strong>Homogeneous</strong> <strong>Electron</strong> <strong>Gas</strong><br />

by<br />

Jung-Hwan Song<br />

A THESIS<br />

submitted to<br />

<strong>Oregon</strong> St<strong>at</strong>e University<br />

<strong>in</strong> partial fulfillment of<br />

the requirements for the<br />

degree of<br />

Doctor of Philosophy<br />

Presented December 7, 2004<br />

Commencement June 2005


Doctor of Philosophy thesis of Jung-Hwan Song presented on December 7, 2004<br />

APPROVED:<br />

Major Professor, represent<strong>in</strong>g <strong>Physics</strong><br />

Chair of the Department of <strong>Physics</strong><br />

Dean of the Gradu<strong>at</strong>e School<br />

I understand th<strong>at</strong> my thesis will become part of the permanent collection of<br />

<strong>Oregon</strong> St<strong>at</strong>e University libraries. My sign<strong>at</strong>ure below authorizes release of my<br />

thesis to any reader upon request.<br />

Jung-Hwan Song, Author


ACKNOWLEDGMENT<br />

I would like to thank the people who helped me come to Corvallis and over-<br />

come culture differences.<br />

I cannot fully express my gr<strong>at</strong>itude to my Major Professor, Henri J. F. Jansen,<br />

who has been a most important contributor to this thesis, and gave me guidel<strong>in</strong>es<br />

and helps not only for physics problems but also for my personal problems. He has<br />

been very k<strong>in</strong>d, generous, and has answered all my questions, <strong>in</strong>clud<strong>in</strong>g the trivial<br />

ones. I really appreci<strong>at</strong>e his expertise and experience, and look forward to further<br />

collabor<strong>at</strong>ions.<br />

I would like to express my appreci<strong>at</strong>ion to Professors William W. Warren,<br />

Rub<strong>in</strong> H. Landau, Tom Giebultowicz, and Solomon C. S. Yim for serv<strong>in</strong>g on my<br />

committee.<br />

I am thankful to my friends and office-m<strong>at</strong>es, especially Jae-Hyuk Lee, Porn-<br />

r<strong>at</strong> W<strong>at</strong>tanakasiwich, and David M<strong>at</strong>usevitsch, for mak<strong>in</strong>g my stay a very pleasant<br />

one.<br />

And f<strong>in</strong>ally, I am deeply gr<strong>at</strong>eful to my parents and my wife for their uncon-<br />

ditional love and support. This thesis might have taken much more time to be<br />

f<strong>in</strong>ished without helps from my wife, JungIn.<br />

F<strong>in</strong>ancial support was partially provided by ONR.


TABLE OF CONTENTS<br />

Page<br />

1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1<br />

2 DENSITY FUNCTIONAL THEORY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4<br />

2.1 The Hohenberg-Kohn theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4<br />

2.1.1 Uniqueness of the external potential <strong>in</strong> terms of the density . 5<br />

2.1.2 Vari<strong>at</strong>ional pr<strong>in</strong>ciple . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7<br />

2.1.3 Universal functional FHK[n] . . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

2.2 The Kohn-Sham equ<strong>at</strong>ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

2.2.1 Non-<strong>in</strong>teract<strong>in</strong>g particle system <strong>in</strong> an external potential vNI(r) 9<br />

2.2.2 Interact<strong>in</strong>g particle system <strong>in</strong> an external potential v(r) . . . . 10<br />

2.2.3 Ground st<strong>at</strong>e energy of an <strong>in</strong>teract<strong>in</strong>g system . . . . . . . . . . . . 12<br />

2.2.4 The chemical potential µ and Fermi surface . . . . . . . . . . . . . 12<br />

2.2.5 Excited st<strong>at</strong>es; Eigenvalues <strong>in</strong> the Kohn-Sham equ<strong>at</strong>ions . . . 14<br />

2.2.6 Sp<strong>in</strong>-polarized systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15<br />

2.2.7 Exchange-correl<strong>at</strong>ion; LDA . . . . . . . . . . . . . . . . . . . . . . . . . . 17<br />

2.2.7.1 A functional . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17<br />

2.2.7.2 Exchange-correl<strong>at</strong>ion energy . . . . . . . . . . . . . . . . . . . . . . 17<br />

2.2.7.3 Interpol<strong>at</strong>ion scheme; Sp<strong>in</strong>-<strong>in</strong>dependent . . . . . . . . . . . . . 21<br />

2.2.7.4 Interpol<strong>at</strong>ion scheme; Sp<strong>in</strong>-dependent . . . . . . . . . . . . . . . 23<br />

3 A MODEL AND ITS PROPERTIES: AN IMPURITY IN A HOMOGE-<br />

NEOUS ELECTRON GAS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30<br />

3.1 An impurity <strong>in</strong> a homogeneous electron gas. . . . . . . . . . . . . . . . . . . . . . . . . 30<br />

3.2 Energy calcul<strong>at</strong>ion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32<br />

3.2.1 Symmetry <strong>in</strong> potentials and phase shifts . . . . . . . . . . . . . . . . 33<br />

3.2.2 Density of <strong>in</strong>duced st<strong>at</strong>es <strong>in</strong> terms of phase shifts . . . . . . . . . 39<br />

3.2.3 Immersion energies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39<br />

3.3 Friedel sum rule and Friedel oscill<strong>at</strong>ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43<br />

3.4 Dielectric functions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44


TABLE OF CONTENTS (Cont<strong>in</strong>ued)<br />

Page<br />

3.5 Virtual bound st<strong>at</strong>e resonance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47<br />

3.6 Sc<strong>at</strong>ter<strong>in</strong>g length and bound st<strong>at</strong>e energy . . . . . . . . . . . . . . . . . . . . . . . . . . 51<br />

4 IMPLEMENTATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55<br />

4.1 Perspective scheme for self-consistent solutions . . . . . . . . . . . . . . . . . . . . . 55<br />

4.1.1 Non-sp<strong>in</strong> polarized system . . . . . . . . . . . . . . . . . . . . . . . . . . . 55<br />

4.1.2 Sp<strong>in</strong> polarized system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57<br />

4.2 Mix<strong>in</strong>g scheme . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57<br />

4.3 Search for eigenst<strong>at</strong>es and eigenvalues of the <strong>in</strong>f<strong>in</strong>ite system . . . . . . . . 60<br />

4.3.1 Bound st<strong>at</strong>es . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60<br />

4.3.2 Sc<strong>at</strong>tered st<strong>at</strong>es . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63<br />

4.4 Numerical problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65<br />

4.4.1 Convergence of phase shifts . . . . . . . . . . . . . . . . . . . . . . . . . . 65<br />

4.4.2 Correction for potentials . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67<br />

4.4.3 Calcul<strong>at</strong>ions for <strong>in</strong>duced electrons and energies . . . . . . . . . . . 70<br />

4.4.4 Mesh for r-space and k-space . . . . . . . . . . . . . . . . . . . . . . . . 71<br />

4.4.5 Cancel<strong>at</strong>ion of numerical errors <strong>in</strong> immersion energies: Number<br />

of r po<strong>in</strong>ts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76<br />

4.4.6 Further numerical precision tests . . . . . . . . . . . . . . . . . . . . . . 79<br />

4.5 Self-consistent solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83<br />

4.6 Sp<strong>in</strong> paramagnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

5 RESULTS AND DISCUSSION. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88<br />

5.1 Non-sp<strong>in</strong>-polarized system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89<br />

5.1.1 Behavior of a system dur<strong>in</strong>g the numerical calcul<strong>at</strong>ions . . . . 89<br />

5.1.2 Charge densities and potentials . . . . . . . . . . . . . . . . . . . . . . . 89<br />

5.1.3 Phase shift, Sc<strong>at</strong>ter<strong>in</strong>g length, and Density of <strong>in</strong>duced st<strong>at</strong>es 93<br />

5.1.4 Dielectric constant and compressibility . . . . . . . . . . . . . . . . . 101<br />

5.1.5 Immersion energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106


5.1.6 Excited system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109<br />

5.2 Sp<strong>in</strong> polarized system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113<br />

5.2.1 Zero external magnetic field . . . . . . . . . . . . . . . . . . . . . . . . . 113<br />

5.2.1.1 Carbon impurity <strong>at</strong> low background densities . . . . . . . . . 113<br />

5.2.1.2 Electrical Resistivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119<br />

5.2.1.3 Excited system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121<br />

5.2.2 Non-zero external magnetic field . . . . . . . . . . . . . . . . . . . . . . 148<br />

6 CONCLUSIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166<br />

BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169<br />

APPENDIX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171<br />

APPENDIX A Atomic Rydberg units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172<br />

APPENDIX B Comparison with published d<strong>at</strong>a. . . . . . . . . . . . . . . . . . . . . . . . . . 174<br />

APPENDIX C Boundary conditions for non-spherical potentials <strong>at</strong> small r176<br />

APPENDIX D S<strong>in</strong>gular value decomposition(SVD) . . . . . . . . . . . . . . . . . . . . . . 179<br />

APPENDIX E Behavior of phase shifts for E → ∞ . . . . . . . . . . . . . . . . . . . . . . 182<br />

APPENDIX F Sc<strong>at</strong>ter<strong>in</strong>g length and bound st<strong>at</strong>e energy . . . . . . . . . . . . . . . . . 184


LIST OF FIGURES<br />

Figure Page<br />

2.1 The exchange ɛx , correl<strong>at</strong>ion ɛc, and Thomas-Fermi k<strong>in</strong>etic energies<br />

TT F = 2.2099/r 2 s per electron from the parametriz<strong>at</strong>ion of Hed<strong>in</strong>-<br />

Lundqvist. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20<br />

2.2 The exchange ɛx , correl<strong>at</strong>ion ɛc, and Thomas-Fermi k<strong>in</strong>etic energies<br />

TT F = 2.2099/r 2 s multiplied by the electron density n. . . . . . . . . . . . 21<br />

2.3 The exchange vx and correl<strong>at</strong>ion vc potentials. . . . . . . . . . . . . . . . . . 22<br />

2.4 The exchange-correl<strong>at</strong>ion energy per electron ɛxc from the parametriz<strong>at</strong>ion<br />

of von Barth and Hed<strong>in</strong>. m = n ζ where n is the total<br />

density. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25<br />

2.5 The exchange-correl<strong>at</strong>ion energy nɛxc. . . . . . . . . . . . . . . . . . . . . . . 26<br />

2.6 The exchange-correl<strong>at</strong>ion ɛxc and Thomas-Fermi k<strong>in</strong>etic energies<br />

TT F = 2.2099/r 2 s multiplied by the electron density n. . . . . . . . . . . . 28<br />

2.7 The exchange-correl<strong>at</strong>ion and Thomas-Fermi k<strong>in</strong>etic energies for extremely<br />

low densities. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28<br />

2.8 The exchange-correl<strong>at</strong>ion potential vxc for the sp<strong>in</strong>-up density. . . . . 29<br />

2.9 The exchange-correl<strong>at</strong>ion potential vxc for the sp<strong>in</strong>-down density. . . 29<br />

3.1 An impurity system consists of an impurity <strong>at</strong>om and a homogeneous<br />

electron gas. The electron density fluctu<strong>at</strong>es due to the impurity<br />

<strong>at</strong>om. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30<br />

3.2 The density of st<strong>at</strong>es ρ0(ɛ) <strong>in</strong> the conduction band and the correspond<strong>in</strong>g<br />

phase shift. ɛd = −0.3 and ∆(ɛ) = (−0.13) 2 πρ0(ɛ) are<br />

used. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50<br />

3.3 There is a virtual bound st<strong>at</strong>e resonance <strong>at</strong> the solution of ɛ − ɛd −<br />

Λ(ɛ) = 0. The density of st<strong>at</strong>es ∆ρ(ɛ) <strong>in</strong>duced by the impurity.<br />

Calcul<strong>at</strong>ed from Fig. 3.2 <strong>in</strong> M<strong>at</strong>hem<strong>at</strong>ica. . . . . . . . . . . . . . . . . . . . . 50<br />

3.4 Shows how the density of st<strong>at</strong>es ∆ρ(ɛ) varies <strong>in</strong> the conduction band<br />

for the different strength of the potential. Calcul<strong>at</strong>ed from Fig. 3.2<br />

<strong>in</strong> M<strong>at</strong>hem<strong>at</strong>ica. The number of electrons <strong>in</strong> the conduction band<br />

is obta<strong>in</strong>ed numerically. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51


LIST OF FIGURES (Cont<strong>in</strong>ued)<br />

Figure Page<br />

4.1 Flow chart for non-polarized systems. . . . . . . . . . . . . . . . . . . . . . . . 56<br />

4.2 Phase shift values of l = 0 for a zero potential. . . . . . . . . . . . . . . . . 65<br />

4.3 Phase shift values for an <strong>at</strong>tractive potential of Yukawa type and a<br />

repulsive potential. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66<br />

4.4 Typical example show<strong>in</strong>g defects <strong>in</strong> the output potential dur<strong>in</strong>g the<br />

numerical calcul<strong>at</strong>ions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68<br />

4.5 Typical examples to show how the effective potentials behave before<br />

and after the correction. These potentials are calcul<strong>at</strong>ed for a<br />

hydrogen impurity and n0 = 0.0025. . . . . . . . . . . . . . . . . . . . . . . . . . 70<br />

4.6 Calcul<strong>at</strong>ed for a hydrogen impurity and n0 = 0.0025. This example<br />

shows the oscill<strong>at</strong>ions of <strong>in</strong>tegral values of the <strong>in</strong>duced electronic<br />

density as a function of r. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71<br />

4.7 Typical example for the effective potential. The logarithmic mesh<br />

is used for the rapid vari<strong>at</strong>ion of the potential <strong>in</strong> the vic<strong>in</strong>ity of the<br />

orig<strong>in</strong>. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

4.8 Example for phase shifts. The phase shift varies rapidly <strong>at</strong> the<br />

bottom of the conduction band if the energy of a bound st<strong>at</strong>e is<br />

close to the bottom of the conduction band. . . . . . . . . . . . . . . . . . . 73<br />

4.9 Integrand <strong>in</strong> (3.22). Calcul<strong>at</strong>ed for a Hydrogen impurity <strong>at</strong> 0.04<br />

background density. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74<br />

4.10 The density of a free carbon <strong>at</strong>om and the density <strong>in</strong>duced by a<br />

carbon impurity <strong>in</strong> a homogeneous electron gas. The 1s and 2s core<br />

st<strong>at</strong>es rema<strong>in</strong> almost the same. These calcul<strong>at</strong>ions are performed<br />

for a sp<strong>in</strong> polarized system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76<br />

4.11 Comparison between immersion energies calcul<strong>at</strong>ed with a constant<br />

free <strong>at</strong>om energy and the free <strong>at</strong>om energy calcul<strong>at</strong>ed with the same<br />

r-mesh as the impurity system. The immersion energies calcul<strong>at</strong>ed<br />

with the same free <strong>at</strong>om energy show rel<strong>at</strong>ively large vari<strong>at</strong>ion with<br />

the different number of r-po<strong>in</strong>ts due to the numerical errors of core<br />

st<strong>at</strong>es. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78


LIST OF FIGURES (Cont<strong>in</strong>ued)<br />

Figure Page<br />

4.12 Fitt<strong>in</strong>g by <strong>in</strong>verse power law to f<strong>in</strong>d the correct immersion energy.<br />

D<strong>at</strong>a from Table 4.1 and Table 4.2 . . . . . . . . . . . . . . . . . . . . . . . . . . 79<br />

4.13 Extra densities <strong>in</strong>duced by an impurity. . . . . . . . . . . . . . . . . . . . . . . 81<br />

4.14 Immersion energy versus the <strong>in</strong>verse of the number of r po<strong>in</strong>ts <strong>in</strong><br />

the l<strong>in</strong>ear r-mesh. D<strong>at</strong>a from Table 4.4. . . . . . . . . . . . . . . . . . . . . . 82<br />

4.15 The density of st<strong>at</strong>es <strong>at</strong> absolute zero temper<strong>at</strong>ure. With an external<br />

magnetic field, the energy of the sp<strong>in</strong>s (magnetic-moment-up)<br />

parallel to the field is lowered by µB, while the energy of the sp<strong>in</strong>s<br />

(magnetic-moment-down) antiparallel to the field is raised by µB.<br />

The chemical potential of the sp<strong>in</strong>-up band is equal to th<strong>at</strong> of the<br />

sp<strong>in</strong>-down band. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

5.1 Vari<strong>at</strong>ion of the number of extra electrons for a system with a hydrogen<br />

impurity dur<strong>in</strong>g the iter<strong>at</strong>ive loops of the numerical calcul<strong>at</strong>ions.<br />

The background electronic density is 0.06. . . . . . . . . . . . . . . . . . . . 90<br />

5.2 The electronic density <strong>in</strong>duced by a hydrogen impurity. It is plotted<br />

as a function of 2kF r. Friedel oscill<strong>at</strong>ions are <strong>in</strong>dependent of the<br />

background density. The range of the background density is from<br />

0.0005 to 0.06. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91<br />

5.3 The electronic density <strong>in</strong>duced by a hydrogen impurity. . . . . . . . . . . 92<br />

5.4 The electronic density <strong>in</strong>duced by a hydrogen impurity <strong>in</strong> the bound<br />

st<strong>at</strong>es. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93<br />

5.5 The electronic density <strong>in</strong>duced by a hydrogen impurity <strong>in</strong> the conduction<br />

band. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94<br />

5.6 Phase shifts (l = 0) for a hydrogen impurity. Plotted for low background<br />

densities only. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95<br />

5.7 Inverse of sc<strong>at</strong>ter<strong>in</strong>g length versus background density for a hydrogen<br />

impurity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96<br />

5.8 Phase shifts for a hydrogen impurity for two different background<br />

densities (0.0025 and 0.06). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97


LIST OF FIGURES (Cont<strong>in</strong>ued)<br />

Figure Page<br />

5.9 Phase shifts (l = 0) for a hydrogen impurity. . . . . . . . . . . . . . . . . . . 98<br />

5.10 The density of st<strong>at</strong>es for a hydrogen impurity for two different background<br />

densities (0.0025 and 0.06). . . . . . . . . . . . . . . . . . . . . . . . . . 98<br />

5.11 Phase shifts of l = 0 and l = 1 for a lithium impurity. . . . . . . . . . . . 100<br />

5.12 Dielectric constants which are numerically calcul<strong>at</strong>ed for a hydrogen<br />

impurity. The Thomas-Fermi dielectric constants are also plotted<br />

for comparison. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102<br />

5.13 Compressibility calcul<strong>at</strong>ed numerically for a hydrogen impurity. A<br />

dotted l<strong>in</strong>e which is taken from [6] is for the result predicted by<br />

theories such as Hubbard, VS, and SS. Both l<strong>in</strong>es are go<strong>in</strong>g through<br />

zero around rs = 5.0. The electronic density which corresponds to<br />

rs of the bottom x-axis is given <strong>in</strong> the top x-axis. . . . . . . . . . . . . . . 103<br />

5.14 Comparison of the compressibility calcul<strong>at</strong>ed numerically for a hydrogen<br />

impurity with sc<strong>at</strong>ter<strong>in</strong>g length. . . . . . . . . . . . . . . . . . . . . . 104<br />

5.15 Immersion energies for several different impurities. 1.0 <strong>in</strong> the legend<br />

<strong>in</strong>dic<strong>at</strong>es th<strong>at</strong> the number of electrons <strong>in</strong> the bound st<strong>at</strong>es is limited<br />

to 1.0. 465 po<strong>in</strong>ts for the logarithmic r-space and 0.81205 × 10 −5<br />

for the first value of r-mesh are used <strong>in</strong> these calcul<strong>at</strong>ions. . . . . . . . 106<br />

5.16 Vari<strong>at</strong>ion <strong>in</strong> the k<strong>in</strong>etic, Coulomb, and exchange-correl<strong>at</strong>ion energies<br />

vs the background density for a hydrogen impurity. 465 po<strong>in</strong>ts for<br />

the logarithmic r-space and 0.81205×10 −5 for the first vale of r-mesh<br />

are used <strong>in</strong> these calcul<strong>at</strong>ions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107<br />

5.17 Immersion energies for the different number of electrons <strong>in</strong> the<br />

bound st<strong>at</strong>es. 465 po<strong>in</strong>ts for the logarithmic r-space and 0.81205 ×<br />

10 −5 for the first value of r-mesh are used <strong>in</strong> these calcul<strong>at</strong>ions. . . . 109<br />

5.18 Vari<strong>at</strong>ion <strong>in</strong> the bound st<strong>at</strong>e energy for the different number of electrons<br />

<strong>in</strong> the bound st<strong>at</strong>e. The second plot is obta<strong>in</strong>ed us<strong>in</strong>g (5.7)<br />

and slopes between two adjacent immersion energies <strong>in</strong> Fig. 5.17.<br />

465 po<strong>in</strong>ts for the logarithmic r-space and 0.81205 × 10 −5 for the<br />

first value of r-mesh are used <strong>in</strong> these calcul<strong>at</strong>ions. . . . . . . . . . . . . . 110


LIST OF FIGURES (Cont<strong>in</strong>ued)<br />

Figure Page<br />

5.19 Comparison of immersion energies for a carbon impurity between a<br />

non-sp<strong>in</strong>-polarized system and a sp<strong>in</strong>-polarized system. The same<br />

reference energy for a free carbon <strong>at</strong>om is used for both calcul<strong>at</strong>ions. 113<br />

5.20 Plot of ∆N vs a background density where ∆N is the difference <strong>in</strong><br />

the number of electrons between sp<strong>in</strong>-up and sp<strong>in</strong>-down electrons <strong>in</strong><br />

a system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114<br />

5.21 The electronic density and the sp<strong>in</strong> density <strong>in</strong>duced by a carbon<br />

impurity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115<br />

5.22 The electronic density (sp<strong>in</strong>-up and sp<strong>in</strong>-down) <strong>in</strong>duced by a carbon<br />

impurity <strong>in</strong> the conduction band. . . . . . . . . . . . . . . . . . . . . . . . . . . . 116<br />

5.23 The density of st<strong>at</strong>es <strong>in</strong>duced by a carbon impurity. . . . . . . . . . . . . 117<br />

5.24 The values of Q versus Z for the background density of 0.01. The<br />

calcul<strong>at</strong>ions are performed only for Z = 1, 3, 6. Note th<strong>at</strong>, <strong>at</strong> this<br />

background density, there is no difference between sp<strong>in</strong>-polarized<br />

and non-sp<strong>in</strong>-polarized systems for these impurities. . . . . . . . . . . . . 120<br />

5.25 Immersion energies and ∆N for a carbon impurity vs the number<br />

of sp<strong>in</strong>-up electrons <strong>in</strong> a 2p bound st<strong>at</strong>e. ∆N is the difference <strong>in</strong> the<br />

number of electrons between sp<strong>in</strong>-up and sp<strong>in</strong>-down electrons <strong>in</strong> a<br />

system. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122<br />

5.26 Bound st<strong>at</strong>e energies vs the number of sp<strong>in</strong>-up electrons <strong>in</strong> a 2p<br />

bound st<strong>at</strong>e. The second plot shows the vari<strong>at</strong>ion of the curv<strong>at</strong>ure,<br />

the energy m<strong>in</strong>imum, and the number of electrons <strong>in</strong> a 2p bound<br />

st<strong>at</strong>e correspond<strong>in</strong>g to the energy m<strong>in</strong>imum <strong>in</strong> the first plot. . . . . . . 123<br />

5.27 The density and the sp<strong>in</strong> density <strong>in</strong>duced by a carbon impurity <strong>at</strong><br />

0.0005 background density. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124<br />

5.28 The density <strong>in</strong> the bound st<strong>at</strong>es of sp<strong>in</strong>-up and sp<strong>in</strong>-down electrons<br />

<strong>in</strong>duced by a carbon impurity <strong>at</strong> 0.0005 background density. . . . . . . 125<br />

5.29 The density and the sp<strong>in</strong> density <strong>in</strong> the conduction band <strong>in</strong>duced<br />

by a carbon impurity <strong>at</strong> 0.0005 background density. . . . . . . . . . . . . 126


LIST OF FIGURES (Cont<strong>in</strong>ued)<br />

Figure Page<br />

5.30 The density and the sp<strong>in</strong> density <strong>in</strong>duced by a carbon impurity <strong>at</strong><br />

0.00075 background density. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127<br />

5.31 The density <strong>in</strong> the bound st<strong>at</strong>es of sp<strong>in</strong>-up and sp<strong>in</strong>-down electrons<br />

<strong>in</strong>duced by a carbon impurity <strong>at</strong> 0.00075 background density. . . . . . 128<br />

5.32 The density and the sp<strong>in</strong> density <strong>in</strong> the conduction band <strong>in</strong>duced<br />

by a carbon impurity <strong>at</strong> 0.00075 background density. . . . . . . . . . . . 129<br />

5.33 The density and the sp<strong>in</strong> density <strong>in</strong>duced by a carbon impurity <strong>at</strong><br />

0.001 background density. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130<br />

5.34 The density <strong>in</strong> the bound st<strong>at</strong>es of sp<strong>in</strong>-up and sp<strong>in</strong>-down electrons<br />

<strong>in</strong>duced by a carbon impurity <strong>at</strong> 0.001 background density. . . . . . . . 131<br />

5.35 The density and the sp<strong>in</strong> density <strong>in</strong> the conduction band <strong>in</strong>duced<br />

by a carbon impurity <strong>at</strong> 0.001 background density. . . . . . . . . . . . . . 132<br />

5.36 The density and the sp<strong>in</strong> density <strong>in</strong>duced by a carbon impurity <strong>at</strong><br />

0.0011 background density. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133<br />

5.37 The density <strong>in</strong> the bound st<strong>at</strong>es of sp<strong>in</strong>-up and sp<strong>in</strong>-down electrons<br />

<strong>in</strong>duced by a carbon impurity <strong>at</strong> 0.0011 background density. . . . . . . 134<br />

5.38 The density and the sp<strong>in</strong> density <strong>in</strong> the conduction band <strong>in</strong>duced<br />

by a carbon impurity <strong>at</strong> 0.0011 background density. . . . . . . . . . . . . 135<br />

5.39 The effective potential of the sp<strong>in</strong>-up band vs the number of sp<strong>in</strong>-up<br />

electrons <strong>in</strong> a 2p bound st<strong>at</strong>e for 0.0005 background density. . . . . . . 137<br />

5.40 The Coulomb potentials calcul<strong>at</strong>ed by <strong>in</strong>duced sp<strong>in</strong>-up and sp<strong>in</strong>down<br />

electrons vs the number of sp<strong>in</strong>-up electrons <strong>in</strong> a 2p bound<br />

st<strong>at</strong>e for 0.0005 background density. . . . . . . . . . . . . . . . . . . . . . . . . . 138<br />

5.41 The Coulomb potentials vs the number of sp<strong>in</strong>-up electrons <strong>in</strong> a 2p<br />

bound st<strong>at</strong>e for 0.0005 background density. The second plot shows<br />

the area which corresponds to the box <strong>in</strong> the first plot. . . . . . . . . . . 139<br />

5.42 The exchange-correl<strong>at</strong>ion potentials of <strong>in</strong>duced sp<strong>in</strong>-up and sp<strong>in</strong>down<br />

electrons vs the number of sp<strong>in</strong>-up electrons <strong>in</strong> a 2p bound<br />

st<strong>at</strong>e for 0.0005 background density. . . . . . . . . . . . . . . . . . . . . . . . . . 140


LIST OF FIGURES (Cont<strong>in</strong>ued)<br />

Figure Page<br />

5.43 Phase shifts of l = 0 and l = 1 for sp<strong>in</strong>-up and sp<strong>in</strong>-down electrons. 142<br />

5.44 The density of <strong>in</strong>duced st<strong>at</strong>es for sp<strong>in</strong>-up and sp<strong>in</strong>-down electrons<br />

<strong>at</strong> 0.0005 background density. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143<br />

5.45 The density of <strong>in</strong>duced st<strong>at</strong>es for sp<strong>in</strong>-up and sp<strong>in</strong>-down electrons<br />

<strong>at</strong> 0.001 background density. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144<br />

5.46 The exchange-correl<strong>at</strong>ion potential with different <strong>in</strong>duced densities<br />

<strong>in</strong> the sp<strong>in</strong>-up band. Background density is 0.0005. The <strong>in</strong>duced<br />

density <strong>in</strong> the sp<strong>in</strong>-down band is assumed to be 0.047. . . . . . . . . . . 145<br />

5.47 Partial wave decompositions of the Friedel sum rule. These plots<br />

show the change <strong>in</strong> the number of electrons <strong>in</strong>duced by a carbon<br />

impurity for each band. 0.0005 is used for a background density.<br />

Bound st<strong>at</strong>es are not <strong>in</strong>cluded. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146<br />

5.48 Bound st<strong>at</strong>e energies, the conduction band m<strong>in</strong>imums, the sp<strong>in</strong>moment,<br />

and the immersion energies vs the magnetic field for a<br />

hydrogen impurity with 0.0025 background density. . . . . . . . . . . . . . 148<br />

5.49 The exchange-correl<strong>at</strong>ion potential with different background densities<br />

of the sp<strong>in</strong>-up band. The total background density is 0.0025 and<br />

the <strong>in</strong>duced density is assumed to be 0.03 for each sp<strong>in</strong>-up and sp<strong>in</strong>down<br />

band. This plot shows how the exchange-correl<strong>at</strong>ion potential<br />

for each band may vary with the <strong>in</strong>creas<strong>in</strong>g magnetic field. . . . . . . . 149<br />

5.50 The effective potentials for the different magnetic field 0.01 and 0.04.<br />

A hydrogen impurity with 0.0025 background density is used. . . . . . 151<br />

5.51 The effective potential for a hydrogen impurity. No external magnetic<br />

field is applied. The potential of the sp<strong>in</strong>-up band is identical<br />

to th<strong>at</strong> of the sp<strong>in</strong>-down band. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152<br />

5.52 Partial wave decompositions of the Friedel sum rule. These plots<br />

show the change <strong>in</strong> the number of electrons <strong>in</strong>duced by a hydrogen<br />

impurity for each band. 0.0025 is used for a background density. . . 153<br />

5.53 Partial wave decompositions of the Friedel sum rule. no external<br />

magnetic field is applied. This plot for the sp<strong>in</strong>-up band is identical<br />

with th<strong>at</strong> for the sp<strong>in</strong>-down band. . . . . . . . . . . . . . . . . . . . . . . . . . . 154


LIST OF FIGURES (Cont<strong>in</strong>ued)<br />

Figure Page<br />

5.54 The <strong>in</strong>duced electronic densities <strong>in</strong> the conduction band for different<br />

magnetic fields. A hydrogen impurity with 0.0025 background<br />

density is used. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155<br />

5.55 The <strong>in</strong>duced electronic densities <strong>in</strong> the conduction band of sp<strong>in</strong>-up<br />

electrons for each angular momentum st<strong>at</strong>e. These plots correspond<br />

to the magnetic field 0.08 <strong>in</strong> Fig. 5.48 and Fig. 5.54. A hydrogen<br />

impurity with 0.0025 background density is used. . . . . . . . . . . . . . . 156<br />

5.56 The density of st<strong>at</strong>es of the sp<strong>in</strong>-up and sp<strong>in</strong>-down bands for different<br />

magnetic fields. The arrows <strong>in</strong> the first plot <strong>in</strong>dic<strong>at</strong>e the<br />

loc<strong>at</strong>ions of the sc<strong>at</strong>ter<strong>in</strong>g resonances. A hydrogen impurity with<br />

0.0025 background density is used. . . . . . . . . . . . . . . . . . . . . . . . . . . 157<br />

5.57 Comparison of kF of the sp<strong>in</strong>-down band with k values which correspond<br />

to the sc<strong>at</strong>ter<strong>in</strong>g resonances <strong>in</strong> Fig. 5.56. . . . . . . . . . . . . . . 158<br />

5.58 Bound st<strong>at</strong>e energies, the conduction band m<strong>in</strong>imums, the sp<strong>in</strong>moment,<br />

and the immersion energies vs the magnetic field for a<br />

hydrogen impurity with two different background densities, 0.01 and<br />

0.04. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160<br />

5.59 The exchange-correl<strong>at</strong>ion potential with different background densities<br />

of the sp<strong>in</strong>-up band. The total background density is 0.04 and<br />

the <strong>in</strong>duced density is assumed to be 0.03 for each sp<strong>in</strong>-up and sp<strong>in</strong>down<br />

band. This plot shows how the exchange-correl<strong>at</strong>ion potential<br />

for each band may vary with the <strong>in</strong>creas<strong>in</strong>g magnetic field. . . . . . . . 161<br />

5.60 The <strong>in</strong>tegr<strong>at</strong>ed values of the density of st<strong>at</strong>es of the sp<strong>in</strong>-up band<br />

for different magnetic fields to f<strong>in</strong>d the total <strong>in</strong>duced number of<br />

electrons <strong>in</strong> the sp<strong>in</strong>-up conduction band. A hydrogen impurity<br />

with 0.04 background density is used. The band m<strong>in</strong>imums are<br />

shifted <strong>in</strong> such a way th<strong>at</strong> they have the same m<strong>in</strong>imum. . . . . . . . . 162<br />

5.61 Partial wave decompositions of the Friedel sum rule. These plots<br />

show the change <strong>in</strong> the number of electrons <strong>in</strong>duced by a hydrogen<br />

impurity for each band. 0.04 is used for a background density. . . . . 163<br />

5.62 The density <strong>in</strong>duced by a hydrogen impurity for each conduction<br />

band. The background density is 0.04 and the magnetic field is 0.3. 164


LIST OF FIGURES (Cont<strong>in</strong>ued)<br />

Figure Page<br />

6.1 Behavior of phase shifts for k → ∞. . . . . . . . . . . . . . . . . . . . . . . . . . 182<br />

6.2 Sc<strong>at</strong>ter<strong>in</strong>g length versus 1/ |EB|, where EB is the bound st<strong>at</strong>e<br />

energy rel<strong>at</strong>ive to the conduction band m<strong>in</strong>imum. . . . . . . . . . . . . . . 184


LIST OF TABLES<br />

Table Page<br />

3.1 Energies of free <strong>at</strong>oms (<strong>in</strong> Rydbergs). . . . . . . . . . . . . . . . . . . . . . . . . 31<br />

3.2 Ioniz<strong>at</strong>ion energies of one electron from a neutral <strong>at</strong>om (<strong>in</strong> Rydbergs).<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31<br />

4.1 Energies of a carbon impurity system ∆E and energies of a free<br />

carbon <strong>at</strong>om E<strong>at</strong>om (<strong>in</strong> Rydbergs). A carbon impurity <strong>in</strong> a sp<strong>in</strong><br />

polarized system and a background density of 0.002 are used for<br />

calcul<strong>at</strong>ions. The r-mesh <strong>in</strong> the table represents the number of rpo<strong>in</strong>ts<br />

used <strong>in</strong> the range from 0 to 8 a.u., which corresponds to the<br />

logarithmic r-mesh and <strong>in</strong>cludes most core st<strong>at</strong>es. (See Sec. 4.4.4.) . 77<br />

4.2 Example for a hydrogen impurity. (Non-sp<strong>in</strong> polarized system.) A<br />

background density of 0.0025 is used for these calcul<strong>at</strong>ions. . . . . . . . 78<br />

4.3 Immersion energies for a hydrogen impurity. 70 k po<strong>in</strong>ts, 550 r<br />

po<strong>in</strong>ts for the logarithmic r-mesh, and 0.1 for the <strong>in</strong>terval of the<br />

l<strong>in</strong>ear r-mesh are used for these calcul<strong>at</strong>ions. . . . . . . . . . . . . . . . . . . 80<br />

4.4 Immersion energies for a hydrogen impurity. 70 k po<strong>in</strong>ts, 80 for<br />

Rmax, and 550 r po<strong>in</strong>ts for the logarithmic r-mesh are used for<br />

calcul<strong>at</strong>ions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82<br />

4.5 Immersion energies for a hydrogen impurity. There is no difference<br />

with<strong>in</strong> 10 −5 Rydbergs for k po<strong>in</strong>ts more than 70. 80 for Rmax and<br />

550 r po<strong>in</strong>ts for the logarithmic r-mesh are used for these calcul<strong>at</strong>ions. 83


IMPURITIES IN A HOMOGENEOUS ELECTRON GAS<br />

1. INTRODUCTION<br />

The electronic structure of many-electron systems such as <strong>at</strong>oms, molecules,<br />

metals, and semiconductors is of gre<strong>at</strong> importance for analyz<strong>in</strong>g and develop<strong>in</strong>g<br />

m<strong>at</strong>erials. Hartree-Fock or second quantiz<strong>at</strong>ion calcul<strong>at</strong>ions for many-electron sys-<br />

tems, especially for solids, are, however, almost impossible or comput<strong>at</strong>ionally very<br />

demand<strong>in</strong>g. One of the most successful theoretical and practical approaches to such<br />

complex systems is density functional theory, <strong>in</strong> which systems are described by<br />

the electron density r<strong>at</strong>her than wave functions. Density functional theory has now<br />

been generalized for molecular dynamics, sp<strong>in</strong>-polarized systems, superconductors,<br />

excited systems, time-dependent systems, and so on.<br />

In 1964, Hohenberg and Kohn published density functional theory, <strong>in</strong> which<br />

the ground st<strong>at</strong>e properties of a many-electron system can be calcul<strong>at</strong>ed <strong>in</strong> terms<br />

of the electron density by m<strong>in</strong>imiz<strong>in</strong>g energy functionals. [1] The Hohenberg-Kohn<br />

theorem, however, has a critical shortcom<strong>in</strong>g <strong>in</strong> th<strong>at</strong> the energy functionals are<br />

impossible to determ<strong>in</strong>e or can only be approxim<strong>at</strong>ed for slowly vary<strong>in</strong>g charge dis-<br />

tributions, as <strong>in</strong> the Thomas-Fermi theory. The Kohn-Sham equ<strong>at</strong>ions form an ef-<br />

fective implement<strong>at</strong>ion of the Hohenberg-Kohn theorem. They were derived by <strong>in</strong>-<br />

troduc<strong>in</strong>g a non-<strong>in</strong>teract<strong>in</strong>g k<strong>in</strong>etic energy functional and the exchange-correl<strong>at</strong>ion<br />

energy functional and utiliz<strong>in</strong>g the energy vari<strong>at</strong>ional pr<strong>in</strong>ciple of the Hohenberg-<br />

Kohn theorem. [2] S<strong>in</strong>ce the exchange-correl<strong>at</strong>ion term <strong>in</strong>cludes every <strong>in</strong>teraction<br />

other than Coulomb <strong>in</strong>teraction and depends on the entire density distribution, it<br />

is, <strong>at</strong> present, not possible to calcul<strong>at</strong>e the exchange-correl<strong>at</strong>ion <strong>in</strong>teraction exactly.


Hence approxim<strong>at</strong>ions are needed, for <strong>in</strong>stance, LDA(Local Density Approxima-<br />

tion) and GGA(Generalized Gradient Approxim<strong>at</strong>ion). 1 S<strong>in</strong>ce the Kohn-Sham<br />

equ<strong>at</strong>ions are a self-consistent set of one-particle equ<strong>at</strong>ions, they have been used<br />

widely for computer calcul<strong>at</strong>ions such as the band structure. [3] [4]<br />

Computer calcul<strong>at</strong>ions for a model of an impurity <strong>at</strong>om <strong>in</strong> a homogeneous<br />

electron gas were performed <strong>in</strong> 1981 by Puska, Niem<strong>in</strong>en, and Man<strong>in</strong>en. [5] This<br />

model has been developed orig<strong>in</strong>ally to study metal vacancies and has been ap-<br />

plied to chemisorbed <strong>at</strong>oms on metal surfaces and to molecular bond<strong>in</strong>g. [6] [7]<br />

The density of st<strong>at</strong>es <strong>in</strong>duced by impurities were calcul<strong>at</strong>ed, as well as the im-<br />

mersion energies. Recently, numerical calcul<strong>at</strong>ions have also been performed for<br />

non-spherical systems, such as an impurity mov<strong>in</strong>g through a uniform electronic<br />

gas, and di<strong>at</strong>omic molecules <strong>in</strong> an electronic gas, with<strong>in</strong> the Kohn-Sham scheme.<br />

[8] [9]<br />

The important properties of this model were already discussed by J. Friedel<br />

<strong>in</strong> 1950s <strong>in</strong> the context of metallic alloys. [10] [11] [12] For a given perturb<strong>at</strong>ion<br />

potential due to impurity <strong>at</strong>oms, a free-electron gas shows two important fe<strong>at</strong>ures,<br />

the Friedel sum rule and Friedel oscill<strong>at</strong>ions, which are rel<strong>at</strong>ed to the total <strong>in</strong>duced<br />

charge <strong>in</strong> the vic<strong>in</strong>ity of an impurity and to the long-range oscill<strong>at</strong>ions of the elec-<br />

tronic density <strong>at</strong> large distances. In Friedel’s papers, transition metal impurities <strong>in</strong><br />

metals such as Al and Cu were discussed. He considered the d shell of a transition<br />

metal impurity <strong>at</strong> the Fermi energy and used the concept of virtual bound st<strong>at</strong>es.<br />

1 The exchange-correl<strong>at</strong>ion <strong>in</strong>teraction <strong>in</strong> LDA depends only on the local electron density.<br />

In GGA it depends not only on the local density but also on the gradient of the local<br />

density.<br />

2


The density of st<strong>at</strong>es <strong>at</strong> the Fermi energy <strong>in</strong>duced by an impurity is used <strong>in</strong> the<br />

discussion of a variety of physical properties such as Pauli paramagnetism, orbital<br />

diamagnetism, electrical resistivity, and thermal properties.<br />

This model can be used to study the behavior of a s<strong>in</strong>gle impurity, for <strong>in</strong>-<br />

stance, to analyze the viol<strong>at</strong>ion of Hund’s second rule for a s<strong>in</strong>gle iron <strong>at</strong>om <strong>in</strong><br />

metals. This is important for properties like crystall<strong>in</strong>e magnetic anisotropy and<br />

quench<strong>in</strong>g of the orbital moment. It can also be considered as a first approxima-<br />

tion to analyze any system, consist<strong>in</strong>g of impurities and an electronic gas, when the<br />

impurities hardly <strong>in</strong>teract and l<strong>at</strong>tice vibr<strong>at</strong>ions can be neglected. This model can<br />

be used to <strong>in</strong>vestig<strong>at</strong>e the magnetic behavior of an electron gas due to an impurity.<br />

Analysis of the sp<strong>at</strong>ial distribution of an electron gas and the concentr<strong>at</strong>ion of<br />

impurities may also shed light on phenomena like diffusion and <strong>in</strong>direct exchange<br />

<strong>in</strong>teractions via conduction electrons.<br />

The calcul<strong>at</strong>ions <strong>in</strong> this thesis first follow the scheme used <strong>in</strong> the Puska, Niem-<br />

<strong>in</strong>en, and Man<strong>in</strong>en paper. LDA is used for the exchange-correl<strong>at</strong>ion <strong>in</strong>teraction <strong>in</strong><br />

this work. Their model is extended for excited systems and sp<strong>in</strong>-polarized systems<br />

with or without an external magnetic field. The sp<strong>in</strong>-moment for the model system<br />

is also calcul<strong>at</strong>ed, consider<strong>in</strong>g sp<strong>in</strong>-coupl<strong>in</strong>g only, with an external magnetic field.<br />

A variety of results such as immersion energies, phase shifts, potentials, dielec-<br />

tric functions, partial-wave decompositions of the Friedel sum rule, and electronic<br />

densities are obta<strong>in</strong>ed.<br />

Density functional theory and the Kohn-Sham equ<strong>at</strong>ion are briefly expla<strong>in</strong>ed<br />

<strong>in</strong> Chapter 2. The model used <strong>in</strong> this work is <strong>in</strong>troduced, and theoretical calcula-<br />

tions and general properties of an electronic gas are discussed <strong>in</strong> Chapter 3. The<br />

numerical methods developed <strong>in</strong> this work are expla<strong>in</strong>ed <strong>in</strong> Chapter 4. The results<br />

and discussion follow <strong>in</strong> Chapter 5. Chapter 6 is dedic<strong>at</strong>ed to conclusions.<br />

3


2. DENSITY FUNCTIONAL THEORY<br />

The Hohenberg-Kohn theorem provides the theoretical found<strong>at</strong>ion for func-<br />

tional calcul<strong>at</strong>ions of the electronic structure for <strong>in</strong>teract<strong>in</strong>g particle systems and<br />

suggests a universal functional FHK[n] which is <strong>in</strong>dependent of the n<strong>at</strong>ure of the<br />

external potential <strong>in</strong> a system. The Kohn-Sham equ<strong>at</strong>ions can be obta<strong>in</strong>ed us<strong>in</strong>g a<br />

vari<strong>at</strong>ion pr<strong>in</strong>ciple by compar<strong>in</strong>g with a Schröd<strong>in</strong>ger equ<strong>at</strong>ion and they <strong>in</strong>corpor<strong>at</strong>e<br />

FHK[n] via the exchange-correl<strong>at</strong>ion energy Exc[n]. Exc[n] can be approxim<strong>at</strong>ed as<br />

<strong>in</strong> LDA or GGA, s<strong>in</strong>ce FHK[n] is impossible to f<strong>in</strong>d. The Kohn-Sham equ<strong>at</strong>ions are<br />

a set of self-consistent one-particle equ<strong>at</strong>ions and have a variety of applic<strong>at</strong>ions,<br />

especially <strong>in</strong> a numerical way. In this chapter we will review the Hohenberg-Kohn<br />

theorem and the Kohn-Sham equ<strong>at</strong>ions briefly. The Kohn-Sham equ<strong>at</strong>ions will<br />

be extended for a sp<strong>in</strong>-polarized system. The exchange-correl<strong>at</strong>ion energies and<br />

potentials will be discussed and expressed explicitly with<strong>in</strong> the LDA scheme.<br />

2.1. The Hohenberg-Kohn theorem<br />

In 1964, Hohenberg and Kohn po<strong>in</strong>ted out th<strong>at</strong> it is possible to derive the<br />

ground st<strong>at</strong>e properties of a many-electron system <strong>in</strong> terms of the electron density.<br />

[1] The Hamiltonian for <strong>in</strong>teract<strong>in</strong>g particles under the <strong>in</strong>fluence of an external<br />

potential v(r) is of the form [13]<br />

where, <strong>in</strong> the second quantized not<strong>at</strong>ion,<br />

ˆH = ˆ T + ˆ V + ˆ W (2.1)<br />

4


ˆT = <br />

<br />

σ<br />

ˆV = <br />

<br />

σ<br />

ˆW = 1<br />

2<br />

<br />

<br />

σ,σ ′<br />

ψ † <br />

σ(r) − 2<br />

2m ∇2 <br />

(r) ψσ(r) dr (2.2)<br />

v σ (r)ψ † σ(r)ψσ(r) dr (2.3)<br />

ψ † σ(r)ψ †<br />

σ ′(r ′ )w(r, r ′ )ψσ ′(r′ )ψσ(r) drdr ′<br />

5<br />

(2.4)<br />

where σ is for sp<strong>in</strong> degeneracy and m is the mass of particles. Note th<strong>at</strong> ˆ W<br />

<strong>in</strong>dic<strong>at</strong>es a two-particle <strong>in</strong>teraction such as the Coulomb <strong>in</strong>teraction.<br />

We assume for simplicity th<strong>at</strong> the ground st<strong>at</strong>e is non-degener<strong>at</strong>e. (The<br />

theory can be, however, easily extended to the degener<strong>at</strong>e case [13].) S<strong>in</strong>ce a non-<br />

degener<strong>at</strong>e ground st<strong>at</strong>e φ is determ<strong>in</strong>ed by solv<strong>in</strong>g the Schröd<strong>in</strong>ger equ<strong>at</strong>ion with<br />

an external potential v(r)<br />

ˆH|φ〉 = ( ˆ T + ˆ V + ˆ W )|φ〉 = Egs|φ〉 , (2.5)<br />

a map C : V ↦→ Φ is def<strong>in</strong>ed from the set of external potentials V to the set of<br />

ground st<strong>at</strong>es Φ. Now, for any ground st<strong>at</strong>e, the electronic density <strong>in</strong> the ground<br />

st<strong>at</strong>e φ can be calcul<strong>at</strong>ed by<br />

n(r) = <br />

〈φ|ψ † σ(r)ψσ(r)|φ〉 , (2.6)<br />

σ<br />

which leads to a second map D : Φ ↦→ N from the set of ground st<strong>at</strong>es to the set<br />

of ground st<strong>at</strong>e densities. The map CD : V ↦→ N clearly tells us th<strong>at</strong> n(r) is a<br />

functional of v(r).<br />

2.1.1. Uniqueness of the external potential <strong>in</strong> terms of the density<br />

The Hohenberg - Kohn theorem st<strong>at</strong>es th<strong>at</strong> the map C and D are both<br />

<strong>in</strong>vertible uniquely, th<strong>at</strong> is, the external potential is uniquely determ<strong>in</strong>ed up to<br />

an additive constant by the ground st<strong>at</strong>e density. One can show this simply by


educio ad absurdum as <strong>in</strong> [1]. For the map C, assume th<strong>at</strong> two different external<br />

potentials give rise to the same ground st<strong>at</strong>e. Suppose<br />

( ˆ T + ˆ V1 + ˆ W )|φ〉 = Egs,1|φ〉 (2.7)<br />

and ( ˆ T + ˆ V2 + ˆ W )|φ〉 = Egs,2|φ〉 (2.8)<br />

where ˆ V1 = ˆ V2 + c by assumption. Note th<strong>at</strong> two potentials with only a constant<br />

difference are equivalent. By subtraction, one has, due to non-degeneracy,<br />

( ˆ V1 − ˆ V2)|φ〉 = (Egs,1 − Egs,2)|φ〉 (2.9)<br />

S<strong>in</strong>ce ˆ V1 and ˆ V2 are multiplic<strong>at</strong>ive oper<strong>at</strong>ors, one must have ˆ V1 − ˆ V2 = Egs,1 − Egs,2<br />

which is <strong>in</strong> contradiction with the assumption <strong>at</strong> the beg<strong>in</strong>n<strong>in</strong>g. More clearly, the<br />

st<strong>at</strong>ement can be proved as follows. Assume, for simplicity, th<strong>at</strong> the system has a<br />

s<strong>in</strong>gle particle with a sp<strong>in</strong> σ <strong>at</strong> position r. Then,<br />

〈rσ|( ˆ V1 − ˆ V2)|φ〉 = 〈rσ|(Egs,1 − Egs,2)|φ〉<br />

(v σ 1 (r) − v σ 2 (r))φσ(r) = (Egs,1 − Egs,2)φσ(r)<br />

(v σ 1 (r) − v σ 2 (r)) = (Egs,1 − Egs,2)<br />

assum<strong>in</strong>g th<strong>at</strong> |φ〉 does not vanish. F<strong>in</strong>ally,<br />

ˆV1 − ˆ V2 = <br />

<br />

dr (v σ 1 (r) − v σ 2 (r)))ψ † σ(r)ψσ(r)<br />

σ<br />

= (Egs,1 − Egs,2) <br />

<br />

= Egs,1 − Egs,2<br />

σ<br />

dr ψ † σ(r)ψσ(r)<br />

Two different potentials, hence, lead to the different ground st<strong>at</strong>es and the<br />

map C is <strong>in</strong>vertible.<br />

The proof for the map D follows by the same procedure. Assume th<strong>at</strong> two<br />

different ground st<strong>at</strong>es |φ1〉 and |φ2〉 yield the same ground st<strong>at</strong>e density, n1(r) =<br />

n2(r). Then one has<br />

6


Similarly<br />

Egs,1 = 〈φ1| ˆ H1|φ1〉 < 〈φ2| ˆ H1|φ2〉 = 〈φ2| ˆ H2 + ˆ V1 − ˆ V2|φ2〉<br />

<br />

= Egs,2 + dr n(r)(v1(r) − v2(r))<br />

Egs,2 = 〈φ2| ˆ H2|φ2〉 < 〈φ1| ˆ <br />

H2|φ1〉 = Egs,1 +<br />

Now, add<strong>in</strong>g of both <strong>in</strong>equalities leads to a contradiction aga<strong>in</strong><br />

Egs,1 + Egs,2 < Egs,1 + Egs,2<br />

dr n(r)(v2(r) − v1(r))<br />

7<br />

(2.10)<br />

and two different ground st<strong>at</strong>es |φ1〉 and |φ2〉, hence, yield different ground st<strong>at</strong>e<br />

densities. The map D must be <strong>in</strong>vertible. Due to this <strong>in</strong>vertibility, the ground st<strong>at</strong>e<br />

expect<strong>at</strong>ion value of any observable Ô (e.g. ˆ H, the Hamiltonian) is determ<strong>in</strong>ed<br />

by the ground st<strong>at</strong>e density and is thus a unique functional of the ground st<strong>at</strong>e<br />

density.<br />

〈φ[n]| Ô|φ[n]〉 = O[n] (2.11)<br />

The map CD is, therefore, <strong>in</strong>vertible and is thus the ground st<strong>at</strong>e density uniquely<br />

determ<strong>in</strong>es the external potential up to an additive constant.<br />

n,<br />

2.1.2. Vari<strong>at</strong>ional pr<strong>in</strong>ciple<br />

Now, the energy functional with the external potential ˆ V is, for any density<br />

E[n] = 〈φ[n]| ˆ T + ˆ V + ˆ W |φ[n]〉 . (2.12)<br />

S<strong>in</strong>ce there exist a non-degener<strong>at</strong>e ground st<strong>at</strong>e density ngs and ground st<strong>at</strong>e en-<br />

ergy Egs for this specific external potential via the map CD, we derive from the<br />

vari<strong>at</strong>ional pr<strong>in</strong>ciple


Egs = E[ngs] = 〈φ[ngs]| ˆ T + ˆ V + ˆ W |φ[ngs]〉<br />

E[n] = 〈φ[n]| ˆ T + ˆ V + ˆ W |φ[n]〉<br />

Egs < E[n] if n = ngs .<br />

The exact ground st<strong>at</strong>e density is, therefore, determ<strong>in</strong>ed by m<strong>in</strong>imiz<strong>in</strong>g the energy<br />

functional E[n]: Egs = m<strong>in</strong><br />

n∈N E[n].<br />

Therefore, one can conclude th<strong>at</strong> there exists a unique energy functional for<br />

a given external potential from the Hohenberg and Kohn theorem and the exact<br />

ground st<strong>at</strong>e density is determ<strong>in</strong>ed by the vari<strong>at</strong>ional equ<strong>at</strong>ion δE[n] = 0.<br />

2.1.3. Universal functional FHK[n]<br />

S<strong>in</strong>ce the map D is <strong>in</strong>dependent of the external potential of the particular<br />

system, one can write the energy functional E[n] as<br />

E[n] = 〈φ[n]| ˆ T + ˆ V + ˆ W |φ[n]〉<br />

<br />

(2.13)<br />

= FHK[n] + dr v0(r)n(r)<br />

where FHK[n] = 〈φ[n]| ˆ T + ˆ W |φ[n]〉. As one can see FHK[n] does not depend on the<br />

external potential (e.g. the <strong>in</strong>teraction between electrons and nuclei), and it is <strong>in</strong><br />

pr<strong>in</strong>ciple possible to calcul<strong>at</strong>e the universal functional FHK[n], which is the same for<br />

all electronic systems such as <strong>at</strong>oms, molecules and solids s<strong>in</strong>ce, for these systems,<br />

ˆW is the Coulomb repulsion between electrons. Now, the functional FHK[n] has<br />

a very important mean<strong>in</strong>g, s<strong>in</strong>ce once FHK[n] is known, one can apply it for any<br />

systems (any number of particles and any external potential).<br />

8


2.2. The Kohn-Sham equ<strong>at</strong>ions<br />

At first sight, density functional theory seems to lead to a new way for the<br />

physics of many-body systems. The universal functional FHK[n] is, however, <strong>in</strong><br />

most cases not possible to obta<strong>in</strong>. One can make some progress to f<strong>in</strong>d approxi-<br />

m<strong>at</strong>e expressions for FHK[n] for a gas of slowly vary<strong>in</strong>g density or almost constant<br />

density. [1] Kohn and Sham, <strong>in</strong> 1965, suggested an altern<strong>at</strong>ive way to solve the<br />

problem, us<strong>in</strong>g a set of self-consistent equ<strong>at</strong>ions for s<strong>in</strong>gle particles <strong>in</strong>clud<strong>in</strong>g the<br />

exchange-correl<strong>at</strong>ion potentials. [2] This s<strong>in</strong>gle particle picture, the Kohn-Sham<br />

equ<strong>at</strong>ions, of a many-body system is quite accessible especially <strong>in</strong> a numerical way<br />

and, thus, provides a new comput<strong>at</strong>ional way us<strong>in</strong>g s<strong>in</strong>gle particle aspects of solid<br />

st<strong>at</strong>e physics.<br />

The key po<strong>in</strong>t <strong>in</strong> deriv<strong>in</strong>g the Kohn-Sham equ<strong>at</strong>ions is th<strong>at</strong>, if n is the ground<br />

st<strong>at</strong>e density of an <strong>in</strong>teract<strong>in</strong>g particle system, one can assume there exists a unique<br />

external potential vNI of a non-<strong>in</strong>teract<strong>in</strong>g particle system correspond<strong>in</strong>g to the<br />

ground st<strong>at</strong>e density n. (Uniqueness of vNI is proven from the Hohenberg-Kohn<br />

theorem, but the existence of vNI must be assumed.) It is possible, now, to de-<br />

rive one-particle equ<strong>at</strong>ions of the <strong>in</strong>teract<strong>in</strong>g system by comparison with the non-<br />

<strong>in</strong>teract<strong>in</strong>g system.<br />

2.2.1. Non-<strong>in</strong>teract<strong>in</strong>g particle system <strong>in</strong> an external potential vNI(r)<br />

The Schröd<strong>in</strong>ger equ<strong>at</strong>ion for a non-<strong>in</strong>teract<strong>in</strong>g system <strong>in</strong> an external poten-<br />

tial vNI(r) is<br />

<br />

− 2<br />

2m ∇2 <br />

+ vNI(r) φi(r) = ɛiφi(r) . (2.14)<br />

The ground st<strong>at</strong>e is aga<strong>in</strong> assumed to be non-degener<strong>at</strong>e for simplicity.<br />

9


The ground st<strong>at</strong>e density nNI is<br />

nNI(r) =<br />

where φi(r) occupies the lowest eigenst<strong>at</strong>es.<br />

10<br />

N<br />

|φi(r)| 2 . (2.15)<br />

i<br />

If the ground st<strong>at</strong>e density nNI is given, the external potential vNI(r) is<br />

determ<strong>in</strong>ed up to a trivial constant from the Hohenberg-Kohn theorem. S<strong>in</strong>ce the<br />

wave functions φi(r) of the Schröd<strong>in</strong>ger equ<strong>at</strong>ion are, thus, unique functionals of<br />

the density nNI, φi(r) = φi([n]; r), the k<strong>in</strong>etic energy TNI[n] of the non-<strong>in</strong>teract<strong>in</strong>g<br />

system is a unique functional and hence so is the total energy. The k<strong>in</strong>etic energy<br />

and total energy functional with the ground st<strong>at</strong>e density nNI are<br />

TNI[n] = <br />

<br />

i,σ<br />

dr φ ∗ <br />

i,σ(r) − 2<br />

2m ∇2 <br />

(r) φi,σ(r)<br />

<br />

(2.16)<br />

and ENI[n] = TNI[n] + dr vNI(r)n(r) . (2.17)<br />

The ground st<strong>at</strong>e density, equivalently, can be obta<strong>in</strong>ed from the vari<strong>at</strong>ional<br />

pr<strong>in</strong>ciple δE = 0 [13] of the Hohenberg-Kohn theorem or the Euler equ<strong>at</strong>ions:<br />

<br />

<br />

δ<br />

ENI[n] − µNI<br />

δn(r)<br />

<br />

n(r) dr = δTNI[n]<br />

δn(r) + vNI(r) − µNI = 0 (2.18)<br />

where µNI is a Lagrange multiplier to keep the total number of particles constant,<br />

n(r) dr = N.<br />

2.2.2. Interact<strong>in</strong>g particle system <strong>in</strong> an external potential v(r)<br />

The energy functional for an <strong>in</strong>teract<strong>in</strong>g particle system is, from the<br />

Hohenberg-Kohn theorem,<br />

<br />

E[n] = FHK[n] + dr vI(r)n(r) (2.19)


If we def<strong>in</strong>e Exc[n] as<br />

Exc[n] = FHK[n] − 1<br />

<br />

2<br />

11<br />

dr dr ′ n(r)w(r, r ′ )n(r ′ ) − TNI[n] (2.20)<br />

where TNI[n] is the non-<strong>in</strong>teract<strong>in</strong>g k<strong>in</strong>etic energy, the energy functional follows<br />

the form<br />

E[n] = TNI[n] + 1<br />

<br />

2<br />

dr dr ′ n(r)w(r, r ′ )n(r ′ <br />

) + Exc[n] +<br />

dr vI(r)n(r) (2.21)<br />

The ground st<strong>at</strong>e density nI(r) of the <strong>in</strong>teract<strong>in</strong>g system can be obta<strong>in</strong>ed<br />

from the Euler equ<strong>at</strong>ion, [14]<br />

where<br />

δ<br />

δn(r) [EI[n]<br />

<br />

− µI dr n(r)] (2.22)<br />

= δTNI[n]<br />

δn(r) + vI(r)<br />

<br />

+ dr ′ w(r, r ′ )n(r ′ ) + vxc([n]; r) − µI = 0 (2.23)<br />

vxc([n]; r) = δExc[n]<br />

δn(r)<br />

As mentioned before, if there exists the unique external potential vNI(r) for<br />

a non-<strong>in</strong>teract<strong>in</strong>g system correspond<strong>in</strong>g to the ground st<strong>at</strong>e density nI(r) of the<br />

<strong>in</strong>teract<strong>in</strong>g system, one can f<strong>in</strong>d from (2.18) and (2.22) th<strong>at</strong><br />

<br />

vNI(r) = vI(r) +<br />

<br />

= vI(r) +<br />

dr ′ w(r, r ′ )n(r ′ ) + vxc([n]; r) − (µI − µNI)<br />

.<br />

dr ′ w(r, r ′ )n(r ′ ) + vxc([n]; r) (2.24)<br />

where the arbitrary constant <strong>in</strong> vNI(r) is chosen to cancel out the term (µI −<br />

µNI) and, of course, the potentials on both sides of the equ<strong>at</strong>ion s<strong>at</strong>isfy the same<br />

boundary conditions.<br />

One can f<strong>in</strong>ally write down, the Kohn-Sham equ<strong>at</strong>ions, a set of the self-<br />

consistent equ<strong>at</strong>ions of a non-<strong>in</strong>teract<strong>in</strong>g system with vNI(r) correspond<strong>in</strong>g to the<br />

<strong>in</strong>teract<strong>in</strong>g system.


[−<br />

12<br />

2<br />

2m ∇2 <br />

+ vI(r) + dr ′ w(r, r ′ )n(r ′ ) + vxc([n]; r)]φi(r) = ɛiφi(r) (2.25)<br />

N<br />

n(r) = |φi(r)| 2<br />

for the lowest eigenst<strong>at</strong>es i = 1, · · · , N (2.26)<br />

i<br />

These equ<strong>at</strong>ions are the self-consistent equ<strong>at</strong>ions <strong>in</strong> a sense th<strong>at</strong> one can solve<br />

(2.25) to obta<strong>in</strong> the ground st<strong>at</strong>e density and the effective s<strong>in</strong>gle particle potential<br />

depends on the density.<br />

2.2.3. Ground st<strong>at</strong>e energy of an <strong>in</strong>teract<strong>in</strong>g system<br />

The ground st<strong>at</strong>e energy for the Kohn-Sham equ<strong>at</strong>ions can be easily derived.<br />

From (2.25),<br />

<br />

ɛi =<br />

T [n] =<br />

N<br />

i<br />

dr φ ∗ i (r)(− 2<br />

ɛi −<br />

where vNI follows (2.24).<br />

Eg =<br />

<br />

2m ∇2 <br />

)φi(r) +<br />

dr |φi(r)| 2 vNI(r) (2.27)<br />

dr n(r)vNI(r) (2.28)<br />

By <strong>in</strong>sert<strong>in</strong>g (2.28) <strong>in</strong>to (2.21), the ground st<strong>at</strong>e energy Eg is<br />

N<br />

i<br />

ɛi − 1<br />

<br />

2<br />

dr dr ′ n(r)w(r, r ′ )n(r ′ <br />

) + Exc[n] −<br />

2.2.4. The chemical potential µ and Fermi surface<br />

The follow<strong>in</strong>g proof is taken from [13].<br />

dr vxc(r)n(r) . (2.29)<br />

With the def<strong>in</strong>ition of the energy functional (2.21), equ<strong>at</strong>ion (2.22) leads to<br />

δEI[n]<br />

δn(r)<br />

= µI<br />

(2.30)


If EN is the ground st<strong>at</strong>e energy and nN is the ground st<strong>at</strong>e density of a N-particle<br />

system,<br />

Then,<br />

∂EN<br />

∂N<br />

EN+ɛ − EN<br />

= lim<br />

ɛ→0 ɛ<br />

δEI[n]<br />

δn(r)<br />

<br />

<br />

<br />

nN<br />

<br />

1<br />

= lim<br />

ɛ→0 ɛ<br />

<br />

1<br />

= lim<br />

ɛ→0 ɛ<br />

= µI(N)<br />

13<br />

= µI[N] (2.31)<br />

dr δEI[n]<br />

δn(r)<br />

<br />

<br />

<br />

nN<br />

(nN+ɛ(r) − nN(r))<br />

dr µI(N)(nN+ɛ(r) − nN(r))<br />

The Lagrangian multiplier, therefore, equals the exact chemical potential of the<br />

<strong>in</strong>teract<strong>in</strong>g system.<br />

The eigenvalues of the Kohn-Sham equ<strong>at</strong>ion are less than or equal to the<br />

chemical potential µ. The Kohn-Sham Fermi surface <strong>in</strong> k space is, thus, def<strong>in</strong>ed<br />

by ɛFermi ≡ µ. The Kohn-Sham Fermi surface is, <strong>in</strong> general, not the same as<br />

the physical Fermi surface, which can be found from the Dyson equ<strong>at</strong>ion, s<strong>in</strong>ce<br />

veff(r) <strong>in</strong> the Kohn-Sham equ<strong>at</strong>ion is not the same as the self-energy Σ(r, r ′ ; ɛi)<br />

<strong>in</strong> the Dyson equ<strong>at</strong>ion. They are, however, identical <strong>in</strong> the case of a uniform<br />

system of <strong>in</strong>teract<strong>in</strong>g particles s<strong>in</strong>ce the Fermi surface must be spherical no m<strong>at</strong>ter<br />

wh<strong>at</strong> equ<strong>at</strong>ions are used. Even <strong>in</strong> the case of a non-uniform system, it turns<br />

out th<strong>at</strong> the difference between two Fermi surfaces is small s<strong>in</strong>ce the electrost<strong>at</strong>ic<br />

potential(external and Coulomb potentials) which is the primary term to cause the<br />

anisotropy of the Fermi surface is the same for both cases. (See Chapter 2, [14]<br />

and section 6.4, [13] for more details.)


2.2.5. Excited st<strong>at</strong>es; Eigenvalues <strong>in</strong> the Kohn-Sham equ<strong>at</strong>ions<br />

Koopmans’ theorem st<strong>at</strong>es th<strong>at</strong> the eigenvalue of the eigenst<strong>at</strong>e m <strong>in</strong> the<br />

Hartree-Fock equ<strong>at</strong>ion is the same as the neg<strong>at</strong>ive of the energy required to remove<br />

the electron <strong>in</strong> the st<strong>at</strong>e m. [15] (This theorem does not apply to small systems. See<br />

[15].) It is n<strong>at</strong>ural, thus, to th<strong>in</strong>k of higher eigenst<strong>at</strong>es <strong>in</strong> the Kohn-Sham equ<strong>at</strong>ion<br />

as those <strong>in</strong> the Hartree-Fock equ<strong>at</strong>ion and obta<strong>in</strong> excit<strong>at</strong>ion energies from the<br />

Kohn-Sham equ<strong>at</strong>ion.<br />

Density functional theory is, however, essentially a ground st<strong>at</strong>e and st<strong>at</strong>ic<br />

(frequency-<strong>in</strong>dependent) theory, which means th<strong>at</strong> eigenvalues or energy levels of<br />

the Kohn-Sham equ<strong>at</strong>ion do not represent excit<strong>at</strong>ion energies which are required<br />

to excite electrons. The primary reason is th<strong>at</strong> the effective exchange-correl<strong>at</strong>ion<br />

potentials must vary <strong>in</strong> the excited system. (See Section 4, [13].) One can also<br />

compare the real eigenvalues of the Kohn-Sham equ<strong>at</strong>ion with the excit<strong>at</strong>ion en-<br />

ergies of the Dyson equ<strong>at</strong>ion. The excit<strong>at</strong>ion energies of the Dyson equ<strong>at</strong>ion are,<br />

however, <strong>in</strong> general complex, which means the f<strong>in</strong>ite lifetime, and the real parts<br />

are <strong>in</strong> general different from the eigenvalues of the Kohn-Sham equ<strong>at</strong>ion as well.<br />

[14]<br />

In order to f<strong>in</strong>d the excit<strong>at</strong>ion energy <strong>in</strong> the Kohn-Sham scheme, one needs<br />

to solve two Kohn-Sham equ<strong>at</strong>ions and f<strong>in</strong>d two self-consistent solutions for the<br />

ground st<strong>at</strong>e and excited st<strong>at</strong>e. In this thesis, the excited st<strong>at</strong>es for a hydrogen<br />

and a carbon impurity will be exam<strong>in</strong>ed by vary<strong>in</strong>g the number of electrons <strong>in</strong> the<br />

bound st<strong>at</strong>es. It is, however, possible to f<strong>in</strong>d the approxim<strong>at</strong>e excit<strong>at</strong>ion energy<br />

us<strong>in</strong>g Janak’s theorem and the Sl<strong>at</strong>er transition st<strong>at</strong>e [13] and also, for an <strong>in</strong>f<strong>in</strong>ite<br />

system, derive a theorem of self-consistent equ<strong>at</strong>ions similar to Koopmans’ theorem<br />

(P. 142, [14]).<br />

14


2.2.6. Sp<strong>in</strong>-polarized systems<br />

If there’s an external magnetic field, one generally has to consider the elec-<br />

tronic orbital coupl<strong>in</strong>g as well as sp<strong>in</strong> coupl<strong>in</strong>g. In this thesis, only sp<strong>in</strong> coupl<strong>in</strong>g<br />

is discussed and the Kohn-Sham equ<strong>at</strong>ions are derived accord<strong>in</strong>gly.<br />

The external potential term, (2.1), is modified as follows. [13]<br />

ˆV = <br />

<br />

σ1σ2<br />

dr ψ † σ1 (r) [vσ1 (r)δσ1σ2 + µBB(r) · σσ1σ2(r)] ψσ2(r) (2.32)<br />

where µB = e/2mc is the Bohr magneton.<br />

The magnetic moment density oper<strong>at</strong>or is def<strong>in</strong>ed as<br />

<br />

ˆm(r) = −µB ψ † σ1 (r)σσ1σ2(r)ψσ2(r) (2.33)<br />

σ1σ2<br />

where σ denotes the vector of Pauli m<strong>at</strong>rices.<br />

With the density oper<strong>at</strong>or<br />

ˆn(r) = ˆn+(r) + ˆn−(r) = <br />

ˆnσ(r) = <br />

ψ † σ(r)ψσ(r) , (2.34)<br />

the external potential term can be rewritten as<br />

<br />

ˆV =<br />

σ<br />

σ<br />

dr [v(r)ˆn(r) − B(r) · ˆm(r)] . (2.35)<br />

Now the ground st<strong>at</strong>e density n(r) and magnetiz<strong>at</strong>ion density m(r) is<br />

n(r) = 〈φ|ˆn(r)|φ〉 (2.36)<br />

m(r) = 〈φ| ˆm(r)|φ〉 . (2.37)<br />

One can work either with (n(r), m(r)) or, equivalently, with the ground sp<strong>in</strong><br />

density m<strong>at</strong>rix nσ1σ2(r) = 〈φ|ψ † σ1 σσ1σ2(r)ψσ2(r)|φ〉. Note th<strong>at</strong> both cases have<br />

four <strong>in</strong>dependent real functions.<br />

15


For simplicity, it is assumed th<strong>at</strong> only z-components <strong>in</strong> the external magnetic<br />

field B(r) and the magnetiz<strong>at</strong>ion m(r) have non-vanish<strong>in</strong>g values. Note th<strong>at</strong>, for<br />

a z-component of m(r),<br />

m(r) = −µB(n+(r) − n−(r)) . (2.38)<br />

As <strong>in</strong> the case of the sp<strong>in</strong>-<strong>in</strong>dependent Hohenberg-Kohn theorem, two differ-<br />

ent non-degener<strong>at</strong>e ground st<strong>at</strong>es lead to two different sp<strong>in</strong> density m<strong>at</strong>rices nss ′(r)<br />

or, equivalently, to two different sets of (n(r), m(r)) or (n+(r), n−(r)).<br />

Now one can derive the Kohn-Sham equ<strong>at</strong>ions for a sp<strong>in</strong>-polarized system<br />

from the vari<strong>at</strong>ional pr<strong>in</strong>ciple of the Hohenberg-Kohn theorem as <strong>in</strong> the sp<strong>in</strong>-<br />

<strong>in</strong>dependent case, Sec. 2.2.<br />

[− 2<br />

2m ∇2 + vI(r) + σµBB(r) +<br />

nσ(r) =<br />

<br />

dr ′ w(r, r ′ )n(r ′ )<br />

+ v σ xc([n+, n−]; r)]φ σ i (r) = ɛ σ i φ σ i (r) (2.39)<br />

N<br />

|φ σ i (r)| 2 for the lowest eigenst<strong>at</strong>es i = 1, · · · , N (2.40)<br />

i<br />

where σ denotes + or − correspond<strong>in</strong>g to the sp<strong>in</strong> projection <strong>in</strong> the z direction.<br />

dent case,<br />

The exchange-correl<strong>at</strong>ion potential v σ xc([n+, n−]; r) is, as <strong>in</strong> the sp<strong>in</strong> <strong>in</strong>depen-<br />

v σ xc([n+, n−]; r) = δEσ xc[n+, n−]<br />

δnσ(r)<br />

16<br />

. (2.41)<br />

and the functional Eσ xc[n+, n−] is, as <strong>in</strong> the sp<strong>in</strong> <strong>in</strong>dependent case, given by<br />

E σ xc[n+, n−] = FHK[n+, n−] − 1<br />

<br />

2<br />

n(r)w(r, r ′ )n(r ′ ) − TNI[n+, n−] (2.42)<br />

The ground st<strong>at</strong>e energy with an analogy to (2.29) is given by<br />

N <br />

Eg = ɛ<br />

i σ<br />

σ i − 1<br />

<br />

dr n(r)w(r, r<br />

2<br />

′ )n(r ′ ) + Exc[n+, n−]<br />

− <br />

<br />

dr v σ xc([n+, n−]; r)nσ(r) . (2.43)<br />

σ


2.2.7. Exchange-correl<strong>at</strong>ion; LDA<br />

At this po<strong>in</strong>t the only problem <strong>in</strong> the Kohn-Sham scheme is to f<strong>in</strong>d a func-<br />

tional Exc[n] <strong>in</strong>stead of a functional FHK[n] <strong>in</strong> the Hohenberg-Kohn theorem.<br />

In this work, the local density approxim<strong>at</strong>ion(LDA) is used for the exchange-<br />

correl<strong>at</strong>ion functional Exc[n].<br />

2.2.7.1. A functional<br />

Suppose there is a functional E[n] for a system with a vary<strong>in</strong>g electron density.<br />

A functional E[n], <strong>in</strong> general, is nonlocal and, thus, E[n] depends on the whole<br />

distribution of the electron density. [16]<br />

If E[n] has the form E[n] = e[n] dr where e[n] is a density functional, for<br />

a very slowly vary<strong>in</strong>g density, e[n] can be expanded <strong>in</strong> Taylor series <strong>in</strong> terms of a<br />

density gradient.<br />

e[n] = e0(n(r)) + e1(n(r))∇n(r) + · · · (2.44)<br />

In LDA, us<strong>in</strong>g the first term only, the exchange-correl<strong>at</strong>ion functional becomes<br />

<br />

Exc[n] = n(r)ɛxc(n(r)) dr (2.45)<br />

where ɛxc(n(r)) is the exchange-correl<strong>at</strong>ion energy per electron for a uniform gas<br />

of <strong>in</strong>teract<strong>in</strong>g electrons with density n. [17] The problem is now reduced to f<strong>in</strong>d<strong>in</strong>g<br />

the exchange-correl<strong>at</strong>ion energy <strong>in</strong> a homogeneous electron gas.<br />

2.2.7.2. Exchange-correl<strong>at</strong>ion energy<br />

The energy of the Hartree-Fock model with plane-wave st<strong>at</strong>es can be easily<br />

obta<strong>in</strong>ed us<strong>in</strong>g the Sl<strong>at</strong>er determ<strong>in</strong>ant or us<strong>in</strong>g the field oper<strong>at</strong>ors <strong>in</strong> the second<br />

17


quantiz<strong>at</strong>ion not<strong>at</strong>ion (Ch. 5, [15]) or the exchange term of the self energy (Ch.<br />

5, [18]) s<strong>in</strong>ce the Hartree energy is zero. The exchange energy, subtract<strong>in</strong>g the<br />

k<strong>in</strong>etic energy, is given by<br />

Ex(k) = − e2 kF<br />

π<br />

<br />

2kF k ln<br />

<br />

kF<br />

<br />

+ k <br />

<br />

kF<br />

− k <br />

1 + k2 F − k2<br />

18<br />

(2.46)<br />

The average exchange energy contribution per particle to the ground st<strong>at</strong>e energy<br />

is easily obta<strong>in</strong>ed as well (See [15] and [18]). Allow<strong>in</strong>g two electrons (sp<strong>in</strong>-up and<br />

sp<strong>in</strong>-down) per st<strong>at</strong>e,<br />

ɛx = 2 1<br />

2N<br />

<br />

k<br />

nkEx(k) = − 3 e<br />

4<br />

2kF π<br />

(2.47)<br />

The exchange part (called the Kohn-Sham or Gáspár potential [17]) of the chemical<br />

potential is, us<strong>in</strong>g the Seitz’s theorem [18], given by<br />

µx = d<br />

dn (nɛx) = − e2<br />

π kF = 4<br />

3 ɛx<br />

(2.48)<br />

Collect<strong>in</strong>g the k<strong>in</strong>etic term and the exchange term gives the first two terms for the<br />

ground st<strong>at</strong>e energy per particle<br />

ɛgs = 2.2099<br />

r 2 s<br />

− 0.9163<br />

rs<br />

where rs is the Wigner-Seitz radius, def<strong>in</strong>ed by n 4π<br />

3 r3 s = 1.<br />

+ · · · Ryd. (2.49)<br />

The exclusion pr<strong>in</strong>ciple <strong>in</strong> the Hartree-Fock model <strong>in</strong>troduces the strong re-<br />

pulsion between the electrons of parallel sp<strong>in</strong> but the Hartree-Fock model fails to<br />

correl<strong>at</strong>e the motion of electrons of antiparallel sp<strong>in</strong> s<strong>in</strong>ce the electrons of antipar-<br />

allel sp<strong>in</strong> tends to separ<strong>at</strong>e due to the Coulomb repulsion. Now the difference<br />

between the exact energy and the Hartree-Fock energy is called the correl<strong>at</strong>ion<br />

energy and thus one can write<br />

ɛgs = 2.2099<br />

r 2 s<br />

− 0.9163<br />

rs<br />

+ ɛc Ryd. (2.50)


The high density limit of rs → 0 for ɛc was first given by Gell-Mann and Brueckner<br />

(1957) and the same result has been obta<strong>in</strong>ed by many other people. Qu<strong>in</strong>n and<br />

Ferrell (1958) used the Seitz’s theorem to f<strong>in</strong>d the correl<strong>at</strong>ion energy from the<br />

correl<strong>at</strong>ion chemical potential µc. They found µc by calcul<strong>at</strong><strong>in</strong>g the self-energy<br />

terms of an electron <strong>at</strong> the Fermi energy. The exchange-correl<strong>at</strong>ion terms <strong>in</strong> the<br />

self-energy can be c<strong>at</strong>egorized <strong>in</strong> the Feynman diagrams by the number of <strong>in</strong>ternal<br />

Coulomb l<strong>in</strong>es [18]. The exchange energy term has only one Coulomb l<strong>in</strong>e. The<br />

correl<strong>at</strong>ion terms have two or more Coulomb l<strong>in</strong>es. The correl<strong>at</strong>ion energy from<br />

one of self-energy diagrams was given by Onsager et al. (1966), which is 0.0436.<br />

The correl<strong>at</strong>ion energy <strong>in</strong> the RPA is<br />

ɛc = −0.142 + 0.0622 ln rs + · · · Ryd. (2.51)<br />

Collect<strong>in</strong>g these terms gives the result, with another term from Carr and<br />

Maradud<strong>in</strong> (1964),<br />

ɛc = −0.094 + 0.0622 ln rs + 0.018rs ln rs + · · · Ryd. (2.52)<br />

This result gives, however, positive values for low densities (rs >∼ 2.5), which<br />

is not correct s<strong>in</strong>ce the correl<strong>at</strong>ion energy must be neg<strong>at</strong>ive consider<strong>in</strong>g the pair<br />

distribution function. (The correl<strong>at</strong>ion effects lower the energy of an <strong>in</strong>teract<strong>in</strong>g<br />

electron system and <strong>in</strong>creases the magnitude of the b<strong>in</strong>d<strong>in</strong>g energy.) [18]<br />

Wigner (1934) found th<strong>at</strong> the system with the positive background charge<br />

ga<strong>in</strong>s energy by localiz<strong>at</strong>ion of electrons. The total energy for a simple system <strong>in</strong><br />

which there is an electron <strong>at</strong> the center of a sphere <strong>in</strong> the unit cell of the l<strong>at</strong>tice<br />

(Wigner-Seitz model) and the electric field is zero outside of each sphere due to<br />

the uniform positive charge is ∼ −1.8/rs. In the low density limit, subtract<strong>in</strong>g the<br />

exchange correl<strong>at</strong>ion energy −0.9163/rs,<br />

19


FIGURE 2.1. The exchange ɛx , correl<strong>at</strong>ion ɛc, and Thomas-Fermi k<strong>in</strong>etic energies<br />

TT F = 2.2099/r 2 s per electron from the parametriz<strong>at</strong>ion of Hed<strong>in</strong>-Lundqvist.<br />

ɛc = − 0.8757<br />

rs<br />

20<br />

Ryd. (2.53)<br />

Interpol<strong>at</strong>ion schemes between these two limits have been made by several<br />

people. For <strong>in</strong>stance, Nozières and P<strong>in</strong>es (1958) recommend, for the range of actual<br />

metallic densities, the <strong>in</strong>terpol<strong>at</strong>ion result ɛc −0.115 + 0.031 ln rs Ryd.<br />

The correl<strong>at</strong>ion energy also can be calcul<strong>at</strong>ed us<strong>in</strong>g dielectric functions for<br />

which there are several models such as Thomas-Fermi, RPA, Hubbard, S<strong>in</strong>gwi-<br />

Sjölander. The S<strong>in</strong>gwi-Sjölander correl<strong>at</strong>ion energy is considered to be the best<br />

due to its positive pair distribution function for most densities. [18]


FIGURE 2.2. The exchange ɛx , correl<strong>at</strong>ion ɛc, and Thomas-Fermi k<strong>in</strong>etic energies<br />

TT F = 2.2099/r 2 s multiplied by the electron density n.<br />

2.2.7.3. Interpol<strong>at</strong>ion scheme; Sp<strong>in</strong>-<strong>in</strong>dependent<br />

In this work, the <strong>in</strong>terpol<strong>at</strong>ion scheme by Hed<strong>in</strong> and Lundqvist (See [17] and<br />

[14]) is used. They used a model given by<br />

µxc(rs) = µx(rs) + µc(rs) = β(rs)µx(rs) . (2.54)<br />

where µxc(rs) is the exchange-correl<strong>at</strong>ion contribution to the effective potential<br />

veff and β(rs) is a correl<strong>at</strong>ion enhancement factor.<br />

With the expression<br />

one can obta<strong>in</strong>, us<strong>in</strong>g (2.48),<br />

<br />

β(rs) = 1 + Bx ln 1 + 1<br />

<br />

x<br />

21<br />

, (2.55)


FIGURE 2.3. The exchange vx and correl<strong>at</strong>ion vc potentials.<br />

<br />

µc(rs) = −C ln 1 + 1<br />

<br />

Ryd. (2.56)<br />

x<br />

<br />

ɛc(rs) = −C (1 + x 3 <br />

) ln 1 + 1<br />

<br />

+<br />

x<br />

x<br />

2 − x2 − 1<br />

<br />

Ryd. (2.57)<br />

3<br />

where x = rs/A and C = 2B((9π)/4) 1/3 /πA. The parameters<br />

22<br />

A = 21 and C = 0.045 (2.58)<br />

are chosen to reproduce the S<strong>in</strong>gwi et al. (1970) results for exc.<br />

S<strong>in</strong>ce β(rs) varies from 1.0 to 1.33 for 0 ≤ rs ≤ 6, the exchange effect of ɛxc<br />

dom<strong>in</strong><strong>at</strong>es the correl<strong>at</strong>ion contribution for the range of actual metallic densities.<br />

The behavior of the exchange and correl<strong>at</strong>ion energies is shown <strong>in</strong> Fig. 2.1 and<br />

Fig. 2.2. The correl<strong>at</strong>ion energies are always neg<strong>at</strong>ive as it should be. A system


of the <strong>in</strong>teract<strong>in</strong>g electrons reaches a lower exchange-correl<strong>at</strong>ion energy as the<br />

density <strong>in</strong>creases. A simple model, which <strong>in</strong>cludes the exchange-correl<strong>at</strong>ion and<br />

Thomas-Fermi k<strong>in</strong>etic energy only, as <strong>in</strong> Fig. 2.2, however, has a m<strong>in</strong>imum energy<br />

<strong>at</strong> n 0.015. One can argue th<strong>at</strong>, if Coulomb effects and the vari<strong>at</strong>ion of the<br />

wave functions are neglected, a system of <strong>in</strong>teract<strong>in</strong>g electrons favors the density<br />

n 0.015.<br />

2.2.7.4. Interpol<strong>at</strong>ion scheme; Sp<strong>in</strong>-dependent<br />

The LDA can be extended for the sp<strong>in</strong>-polarized case (called the local sp<strong>in</strong><br />

density approxim<strong>at</strong>ion, LSDA) and, by choos<strong>in</strong>g the z-axis along the local sp<strong>in</strong><br />

direction,<br />

<br />

Exc[n+, n−] =<br />

and the variables are now n+(r) and n−(r) or<br />

dr (n+(r) + n−(r))ɛxc(n+(r), n−(r)) (2.59)<br />

n(r) = n+(r) + n−(r)<br />

and ζ(r) = n+(r) − n−(r)<br />

n(r)<br />

where ζ describes the degree of local magnetiz<strong>at</strong>ion. For <strong>in</strong>stance, the exchange<br />

energy can be, us<strong>in</strong>g a simple superposition pr<strong>in</strong>ciple (Ch.5, [13]), written as<br />

Ex,pol[n, ζ] = 1<br />

2 Ex[(1 + ζ)n] + 1<br />

2 Ex[(1 − ζ)n] ,<br />

which can be expanded <strong>in</strong> a series us<strong>in</strong>g a gradient expansion. In lowest order, the<br />

exchange energy per particle is<br />

ɛ 0 x,pol(r) = 1<br />

2 [(1 + ζ(r))4/3 + (1 − ζ(r)) 4/3 ]ɛ 0 x<br />

where ɛ 0 x follows the form (2.47).<br />

23<br />

(2.60)


For the sp<strong>in</strong>-polarized case, there are several parametriz<strong>at</strong>ions, such as von<br />

Barth and Hed<strong>in</strong> (1972), Gunnarsson and Lundqvist (1976), and Monte Carlo<br />

results of Ceperley and Alder (1980). (See [13] or their papers for more details.)<br />

In this work, the parametriz<strong>at</strong>ion of von Barth and Hed<strong>in</strong> is used. They used a<br />

generalized random phase expression for ɛxc <strong>in</strong> terms of the polariz<strong>at</strong>ion propag<strong>at</strong>or<br />

and the two-bubble r<strong>in</strong>g approxim<strong>at</strong>ion for the irreducible propag<strong>at</strong>or. [19]<br />

The parametriz<strong>at</strong>ion is based on an <strong>in</strong>terpol<strong>at</strong>ion between the paramag-<br />

netic(unpolarized, ζ = 0) and the ferromagnetic(fully polarized, ζ = ±1) st<strong>at</strong>e:<br />

ɛxc(n(r), ζ(r)) = ɛx(n(r), ζ(r)) + ɛc(n(r), ζ(r)) (2.61)<br />

ɛi(n(r), ζ(r)) = ɛi(n, ζ = 0) + (ɛi(n, ζ = 1) − ɛi(n, ζ = 0))f(ζ(r)) (2.62)<br />

where i denotes the exchange (i = x) or correl<strong>at</strong>ion (i = c) contribution to the<br />

exchange-correl<strong>at</strong>ion energy per particle. The <strong>in</strong>terpol<strong>at</strong>ion function is:<br />

f(ζ(r)) = (1 + ζ(r))4/3 + (1 − ζ(r)) 4/3 − 2<br />

2 4/3 − 2<br />

24<br />

(2.63)<br />

from which one can obta<strong>in</strong> aga<strong>in</strong> the ζ-dependence of the exchange energy term<br />

(2.60). The exchange energy <strong>in</strong> the paramagnetic limit, ɛx(n, ζ = 0), is given by<br />

(2.47). In the ferromagnetic limit, ɛx(n, ζ = ±1) = 2 1/3 ɛx(n, ζ = 0).<br />

The correl<strong>at</strong>ion energy follows aga<strong>in</strong> the Hed<strong>in</strong>-Lundqvist form us<strong>in</strong>g the<br />

parameters<br />

(2.59).<br />

ζ = 0 : C P = 0.0504, A P = 30<br />

ζ = 1 : C F = 0.0254, A F = 75<br />

The exchange-correl<strong>at</strong>ion potential can be easily calcul<strong>at</strong>ed us<strong>in</strong>g (2.41) and<br />

v σ xc = ∂<br />

∂n σ ((n+ + n−)ɛxc(n+, n−)) (2.64)<br />

For <strong>in</strong>stance, v + xc = ɛxc(n+, n−)) + n ∂ɛxc<br />

∂n+<br />

(2.65)


FIGURE 2.4. The exchange-correl<strong>at</strong>ion energy per electron ɛxc from the parametriz<strong>at</strong>ion<br />

of von Barth and Hed<strong>in</strong>. m = n ζ where n is the total density.<br />

After some algebra,<br />

v + xc = ( 4<br />

3 ɛx(n, ζ = 0) + γ(ɛc(n, ζ = ±1) − ɛc(n, ζ = 0)))(1 + ζ) 1/3<br />

−C P <br />

ln 1 + AP<br />

<br />

− γ(ɛc(n, ζ = ±1) − ɛc(n, ζ = 0)) + τcf(ζ)<br />

rs<br />

where τc = −CF <br />

ln 1 + AF<br />

<br />

+ C rs<br />

P <br />

ln 1 + AP<br />

<br />

− rs<br />

4<br />

3 (ɛc(n, ζ = ±1) − ɛc(n, ζ = 0))<br />

and γ =<br />

4<br />

3(21/3−1) .<br />

S<strong>in</strong>ce the correl<strong>at</strong>ion energy for the sp<strong>in</strong> polarized case uses the Hed<strong>in</strong>-<br />

Lundqvist model which is already obta<strong>in</strong>ed for the non-sp<strong>in</strong> polarized case (al-<br />

though the coefficients are different) the behavior of ɛxc and vxc for the sp<strong>in</strong>-<br />

polarized case is not much different from those for the non-sp<strong>in</strong> polarized case.<br />

(See Fig. 2.4 along m = 0.) Note th<strong>at</strong> ɛxc is the exchange-correl<strong>at</strong>ion energy per<br />

electron.<br />

25


FIGURE 2.5. The exchange-correl<strong>at</strong>ion energy nɛxc.<br />

The behavior of the correl<strong>at</strong>ion effect <strong>in</strong> a polarized case (m = 0) is, however,<br />

different due to the degree of local magnetiz<strong>at</strong>ion. Notice th<strong>at</strong> the exchange-<br />

correl<strong>at</strong>ion potential for sp<strong>in</strong>-down has the mirror image of th<strong>at</strong> of sp<strong>in</strong>-up and<br />

s<strong>at</strong>isfies the symmetry rel<strong>at</strong>ion v + xc(ζ) = v − xc(−ζ). Along the l<strong>in</strong>e m = n (fully<br />

polarized limit) <strong>in</strong> Fig. 2.8 and Fig. 2.9, the exchange potential vanishes and the<br />

correl<strong>at</strong>ion potential opposes this effect. [13] As a result, the correl<strong>at</strong>ion effect<br />

reduces the polariz<strong>at</strong>ion dependence of vx, which is consistent with the behavior<br />

of the pair distribution function. [18]<br />

A system with a uniform density can reach a lower exchange-correl<strong>at</strong>ion<br />

energy as the magnitude of m <strong>in</strong>creases as shown <strong>in</strong> Fig. 2.5. The k<strong>in</strong>etic en-<br />

ergy n · TT F , however, makes the situ<strong>at</strong>ion opposite and dom<strong>in</strong><strong>at</strong>es the exchange-<br />

correl<strong>at</strong>ion energy especially for high densities. The system reaches a higher energy<br />

26


due to the term n · TT F as |m| <strong>in</strong>creases <strong>in</strong> most of the density range.(See Fig. 2.6)<br />

One can notice, however, <strong>in</strong> Fig. 2.7, th<strong>at</strong> for extremely low densities (n ≤ 0.001)<br />

the exchange-correl<strong>at</strong>ion energy dom<strong>in</strong><strong>at</strong>es the k<strong>in</strong>etic energy n · TT F as |m| <strong>in</strong>-<br />

creases and thus n · TT F does not keep the system away from a polarized st<strong>at</strong>e.<br />

The density n <strong>at</strong> which there’s no dom<strong>in</strong>ant term between n · TT F and n · exc and<br />

the system is <strong>in</strong> equilibrium is ∼ 0.00153.<br />

One can conclude th<strong>at</strong> the exchange energy favors a sp<strong>in</strong>-polarized solution<br />

and the k<strong>in</strong>etic energy opposes the exchange-correl<strong>at</strong>ion effects.<br />

27


FIGURE 2.6. The exchange-correl<strong>at</strong>ion ɛxc and Thomas-Fermi k<strong>in</strong>etic energies<br />

TT F = 2.2099/r 2 s multiplied by the electron density n.<br />

FIGURE 2.7. The exchange-correl<strong>at</strong>ion and Thomas-Fermi k<strong>in</strong>etic energies for<br />

extremely low densities.<br />

28


FIGURE 2.8. The exchange-correl<strong>at</strong>ion potential vxc for the sp<strong>in</strong>-up density.<br />

FIGURE 2.9. The exchange-correl<strong>at</strong>ion potential vxc for the sp<strong>in</strong>-down density.<br />

29


3. A MODEL AND ITS PROPERTIES: AN IMPURITY IN A<br />

HOMOGENEOUS ELECTRON GAS<br />

Immersion energy calcul<strong>at</strong>ions for an impurity <strong>in</strong> a homogeneous electron<br />

gas can be simplified us<strong>in</strong>g the symmetry of the potentials. Physical quantities,<br />

like density of st<strong>at</strong>es and phase shifts, are very important <strong>in</strong> the discussion of<br />

theoretical aspects, such as immersion energy calcul<strong>at</strong>ions and the behavior of an<br />

impurity, as well as <strong>in</strong> the discussion of numerical aspects, such as normaliz<strong>at</strong>ion<br />

conditions for sc<strong>at</strong>tered wave functions and criteria for the maximum number of<br />

angular momentum l for each momentum k. Various aspects of an impurity system<br />

can be exam<strong>in</strong>ed <strong>in</strong> terms of Friedel oscill<strong>at</strong>ions, dielectric functions, virtual bound<br />

st<strong>at</strong>e resonances, and sc<strong>at</strong>ter<strong>in</strong>g lengths.<br />

3.1. An impurity <strong>in</strong> a homogeneous electron gas<br />

The model system used <strong>in</strong> this work consists of an impurity <strong>at</strong>om immersed<br />

<strong>in</strong> a homogeneous electron gas with a uniform positive background charge density.<br />

(See Fig. 3.1.) This impurity system is neutral and also <strong>in</strong>f<strong>in</strong>ite. The immersion<br />

energy is now calcul<strong>at</strong>ed by subtract<strong>in</strong>g the energies of two isol<strong>at</strong>ed systems, an<br />

impurity <strong>at</strong>om and the background, from the impurity system.<br />

FIGURE 3.1. An impurity system consists of an impurity <strong>at</strong>om and a homogeneous<br />

electron gas. The electron density fluctu<strong>at</strong>es due to the impurity <strong>at</strong>om.<br />

30


Atoms V. B. H. Experiment 1 KS-KC 1 KS-X 1 HF 1<br />

Li −14.710 −14.956 −14.656 −14.349 −14.865<br />

Ne −256.747 −257.880 −256.349 −254.986 −257.094<br />

K −1195.943 −1199.97 −1196.215 −1193.387 −1198.329<br />

TABLE 3.1. Energies of free <strong>at</strong>oms (<strong>in</strong> Rydbergs).<br />

Atoms V. B. H. Experiment 1 KS-KC 1 KS-X 1 HF 1<br />

Li 0.417 0.396 0.396 0.331 0.393<br />

Ne 1.641 1.585 1.657 1.551 1.461<br />

K 0.347 0.319 0.330 0.271 0.295<br />

TABLE 3.2. Ioniz<strong>at</strong>ion energies of one electron from a neutral <strong>at</strong>om (<strong>in</strong> Rydbergs).<br />

Eimm = Eimp − Epure − E<strong>at</strong>om<br />

31<br />

(3.1)<br />

The energy of a free <strong>at</strong>om E<strong>at</strong>om is easily calcul<strong>at</strong>ed by solv<strong>in</strong>g the Kohn-Sham<br />

equ<strong>at</strong>ions for bound st<strong>at</strong>es. Some examples from this work (V. B. H.), <strong>in</strong> which<br />

sp<strong>in</strong>-polarized and spherically symmetric systems are used, are shown <strong>in</strong> Table 3.1<br />

and 3.2 and are compared with the results of Tong and Sham [20]. KS-KC denotes<br />

the exchange-correl<strong>at</strong>ion energy <strong>in</strong> which the correl<strong>at</strong>ion energy is simply <strong>in</strong>ter-<br />

pol<strong>at</strong>ed between Gell-Mann and Brueckner and Wigner schemes. KS-K and HF<br />

1 Ref. [20]


denote the exchange energy only (− 3 e<br />

4<br />

2kF ) and Hartree-Fock calcul<strong>at</strong>ions respec-<br />

π<br />

tively. (See [20] for more details.) Ioniz<strong>at</strong>ion energies can be found by evalu<strong>at</strong><strong>in</strong>g<br />

E[N] − E[N − 1], as mentioned <strong>in</strong> Sec. 2.2.5.<br />

Differences between numerical calcul<strong>at</strong>ions and experimental values for free<br />

<strong>at</strong>om energies are larger than those for ioniz<strong>at</strong>ion energies. LDA gives rel<strong>at</strong>ively<br />

large numerical errors for core st<strong>at</strong>es such as 1s and 2s st<strong>at</strong>es due to the rapid<br />

vari<strong>at</strong>ions <strong>in</strong> the density. Numerical errors are, however, reduced <strong>in</strong> ioniz<strong>at</strong>ion<br />

energy calcul<strong>at</strong>ions s<strong>in</strong>ce core st<strong>at</strong>es <strong>in</strong> E[N] and E[N − 1] do not vary much and<br />

thus errors are canceled out. The same reason<strong>in</strong>g applies to immersion energies,<br />

which is expla<strong>in</strong>ed <strong>in</strong> detail <strong>in</strong> Sec. 4.4.5.<br />

3.2. Energy calcul<strong>at</strong>ion<br />

S<strong>in</strong>ce this work focuses on spherically symmetric potentials, the Kohn-Sham<br />

equ<strong>at</strong>ions and density and energy calcul<strong>at</strong>ions can be gre<strong>at</strong>ly simplified and the<br />

comput<strong>at</strong>ional work can be reduced us<strong>in</strong>g this spherical symmetry. It is, however,<br />

important to extend this work to non-spherical systems, s<strong>in</strong>ce most systems such<br />

as molecules and solids have non-spherical potentials and even <strong>in</strong> the case of a<br />

s<strong>in</strong>gle impurity <strong>at</strong>om, the electronic structure and immersion energies depend on<br />

the non-spherical densities if the angular momentum shell of an impurity <strong>at</strong>om is<br />

not fully occupied and thus the spherical symmetry is broken. Phase shifts are<br />

very important quantities used to calcul<strong>at</strong>e density of st<strong>at</strong>es and energy and also<br />

provide boundary conditions for sc<strong>at</strong>tered st<strong>at</strong>es. In this section, the Kohn-Sham<br />

equ<strong>at</strong>ions for non-spherical potentials are briefly <strong>in</strong>troduced and phase shifts and<br />

energy calcul<strong>at</strong>ions will be reviewed.<br />

32


3.2.1. Symmetry <strong>in</strong> potentials and phase shifts<br />

S<strong>in</strong>ce angular momentum shells <strong>in</strong> general have cyl<strong>in</strong>drical symmetry (φ sym-<br />

metry), non-spherical potentials have the same symmetry. (Note th<strong>at</strong> vxc depends<br />

on only local density and is φ-symmetric. The readers may refer to [21], [9], and [8]<br />

for this section.) The Kohn-Sham equ<strong>at</strong>ions can be simplified accord<strong>in</strong>gly. There-<br />

fore, if the <strong>in</strong>put density used to calcul<strong>at</strong>e the potentials of the KS equ<strong>at</strong>ions is<br />

φ-symmetric, the output density from the KS equ<strong>at</strong>ions is also φ-symmetric s<strong>in</strong>ce<br />

[H, Lz] = 0 and the iter<strong>at</strong>ive procedure of the KS equ<strong>at</strong>ions does not break the<br />

symmetry. The Kohn-Sham equ<strong>at</strong>ions are, <strong>in</strong> Rydberg <strong>at</strong>omic units (APPENDIX<br />

B),<br />

−∇ 2 + Veff(r) ψi(r) = ɛiψi(r) . (3.2)<br />

The angular dependence of the Kohn-Sham st<strong>at</strong>es ψi(r) can be expanded <strong>in</strong> terms<br />

of spherical harmonics.<br />

ψi(r) = <br />

lm<br />

33<br />

uilm(r)<br />

Ylm(Ωr) . (3.3)<br />

r<br />

One can express the electronic density n(r) as a sum over bound st<strong>at</strong>es and sc<strong>at</strong>-<br />

tered st<strong>at</strong>es with a reference of zero-energy.<br />

The total density is, as <strong>in</strong> (2.15),<br />

n(r) = n bound (r) + n cond (r) =<br />

lim<br />

r→∞ Veff(r) = 0 . (3.4)<br />

N<br />

i<br />

|ψ bound<br />

i (r)| 2 +<br />

M<br />

j<br />

|ψ cond<br />

j (r)| 2<br />

(3.5)<br />

where i and j are order<strong>in</strong>g <strong>in</strong>dices for the lowest occupied st<strong>at</strong>es, and i <strong>in</strong> |ψbound i (r)|<br />

and j <strong>in</strong> |ψcond i (r)| denote the KS st<strong>at</strong>es <strong>in</strong> which ɛi < 0 and ɛj > 0, respectively.


Note th<strong>at</strong> m is a good quantum number s<strong>in</strong>ce [H, Lz] = 0 and the Kohn-Sham<br />

equ<strong>at</strong>ions can be written down for each ɛi and m value.<br />

Us<strong>in</strong>g (3.3), one obta<strong>in</strong>s coupled equ<strong>at</strong>ions, similar to radial equ<strong>at</strong>ions for a<br />

spherically symmetric potential,<br />

−∇ 2 + Veff(r) <br />

by projection onto Yl ′ m(Ωr),<br />

where<br />

d 2<br />

dr 2 + ɛi −<br />

V m<br />

<br />

ll ′ (r) =<br />

lm<br />

l(l + 1)<br />

r2 <br />

uilm(r)<br />

<br />

Ylm(Ωr) = ɛi<br />

r<br />

lm<br />

uilm(r) = <br />

l ′<br />

uilm(r)<br />

Ylm(Ωr) .<br />

r<br />

34<br />

V m<br />

ll ′ (r)uil ′ m(r) , (3.6)<br />

dΩ Y ∗<br />

lm(Ωr)Veff(r, θ)Yl ′ m(Ωr) . (3.7)<br />

Notice th<strong>at</strong> s<strong>in</strong>ce Veff(r, θ) does not depend on φ, V m<br />

ll ′ (r) and the solutions <strong>in</strong> 3.6<br />

are real.<br />

The l<strong>in</strong>early <strong>in</strong>dependent solutions u j<br />

lm (j is a simple order<strong>in</strong>g <strong>in</strong>dex2 vary<strong>in</strong>g<br />

from 0 to lmax) can be found from coupled equ<strong>at</strong>ions (3.6) and their behavior <strong>at</strong><br />

small r is<br />

lim<br />

r→0 uj<br />

lm (r) ∼ rj+|l−j|+1<br />

∞<br />

b l ir i−1 , (3.8)<br />

which provides the boundary conditions of <strong>in</strong>dependent solutions <strong>at</strong> small r. (See<br />

APPENDIX C.)<br />

One can now expand uilm <strong>in</strong> terms of l<strong>in</strong>early <strong>in</strong>dependent solutions.<br />

uilm(r) = <br />

j<br />

i=1<br />

α m j u j<br />

ilm (r) , (3.9)<br />

2 one has to use the cut-off maximum lmax <strong>in</strong> a computer program.


where the coefficients α m j can be determ<strong>in</strong>ed by the asymptotic behavior of the<br />

solutions and do not depend on l but on ɛi and m. One can determ<strong>in</strong>e α m j by<br />

boundary conditions, which uniquely determ<strong>in</strong>es the radial wave functions uilm(r)<br />

and Kohn-Sham eigenst<strong>at</strong>es ψi(r). For <strong>in</strong>stance, one has, for bound st<strong>at</strong>es, two sets<br />

of l<strong>in</strong>early <strong>in</strong>dependent of solutions which are u j<br />

ilm (r) from the outward <strong>in</strong>tegr<strong>at</strong>ion<br />

started from the orig<strong>in</strong> and v j<br />

ilm (r) from the <strong>in</strong>ward <strong>in</strong>tegr<strong>at</strong>ion started from the<br />

<strong>in</strong>f<strong>in</strong>ity. The asymptotic behavior of v j<br />

ilm (r) is<br />

lim<br />

r→∞ vj<br />

lm (r) ∼ δjl exp −√ −ɛr<br />

35<br />

. (3.10)<br />

The condition for the existence of bound st<strong>at</strong>es is th<strong>at</strong> the two sets of l<strong>in</strong>early<br />

<strong>in</strong>dependent solutions must m<strong>at</strong>ch <strong>at</strong> an <strong>in</strong>termedi<strong>at</strong>e r.<br />

<br />

j<br />

<br />

j<br />

α m j<br />

α m j u j<br />

ilm<br />

d<br />

dr uj<br />

ilm<br />

(r) = <br />

j<br />

(r) = <br />

j<br />

β m j v j<br />

ilm (r) (3.11)<br />

β m j<br />

d<br />

dr vj<br />

ilm (r) . (3.12)<br />

S<strong>in</strong>ce a nontrivial solution exist only if the determ<strong>in</strong>ant of (3.11) is zero, one needs<br />

to search for the zeroes of the determ<strong>in</strong>ant to f<strong>in</strong>d bound st<strong>at</strong>es. Notice the Eq.<br />

(3.11) becomes the standard Wronskian condition for the spherically symmetric<br />

potentials. Once the bound st<strong>at</strong>e energy has been determ<strong>in</strong>ed, one can use the<br />

s<strong>in</strong>gular value decomposition method (APPENDIX D) to f<strong>in</strong>d the eigenfunction<br />

ψi(r). For sc<strong>at</strong>tered st<strong>at</strong>es, one can use the asymptotic behavior of the sc<strong>at</strong>tered<br />

wave function for large r to determ<strong>in</strong>e the coefficients α m j . Notice th<strong>at</strong> the wave<br />

number vector k <strong>in</strong> uklm(r) also depends on the direction of the <strong>in</strong>cident plane<br />

wave, which thus leads to considerably more comput<strong>at</strong>ional work compared to<br />

spherical systems. (The readers refer to [9] and [8] for more details.)


In addition, it is known, <strong>in</strong> a case of a free <strong>at</strong>om calcul<strong>at</strong>ion, th<strong>at</strong> the energy<br />

difference between the non-spherical potential and the spherical potential is much<br />

smaller than th<strong>at</strong> between the non-sp<strong>in</strong> polarized and sp<strong>in</strong> polarized systems. [22]<br />

A spherical potential can, therefore, be considered the best for a first try<br />

and a sp<strong>in</strong>-polarized system with a spherical potential still has many <strong>in</strong>terest<strong>in</strong>g<br />

properties. If the potential is spherically symmetric, the angular momentum l<br />

is a constant of a motion and also a good quantum number s<strong>in</strong>ce [H, L 2 ] = 0.<br />

The eigenenergies do not depend on l and m and eigenst<strong>at</strong>es are degener<strong>at</strong>e. The<br />

coupled equ<strong>at</strong>ions (3.6) can be simplified to the radial equ<strong>at</strong>ions.<br />

2 d<br />

dr2 + ɛi,l<br />

l(l + 1)<br />

−<br />

r2 <br />

ui,l(r) = V (r)ui,l(r) . (3.13)<br />

For sc<strong>at</strong>tered st<strong>at</strong>es,<br />

d 2<br />

dr 2 + k2 −<br />

36<br />

l(l + 1)<br />

r2 <br />

uk,l(r) = V (r)uk,l(r) . (3.14)<br />

If the potential vanishes for the exterior region r > Rb, the solutions for the<br />

sc<strong>at</strong>tered st<strong>at</strong>es <strong>in</strong> (3.14) follow a form of free particle waves.<br />

Rk,l(r) = uk,l(r)<br />

r = Aljl(kr) + Blnl(kr) r > Rb (3.15)<br />

Us<strong>in</strong>g the asymptotic approxim<strong>at</strong>ions 3 ,<br />

uk,l(r)<br />

r<br />

s<strong>in</strong>(kr − l<br />

= Al<br />

π<br />

2 ) cos(kr − l<br />

− Bl<br />

kr<br />

π<br />

2 )<br />

kr<br />

for kr ≫ l . (3.16)<br />

If the potential vanishes <strong>at</strong> the orig<strong>in</strong> (r = 0), Bl must be zero as nl diverges <strong>at</strong><br />

the orig<strong>in</strong>. The <strong>in</strong>tensity of sc<strong>at</strong>ter<strong>in</strong>g by the potential, hence, can be compared<br />

with the r<strong>at</strong>io Bl/Al. The phase shift δl is now def<strong>in</strong>ed by<br />

Bl<br />

Al<br />

For large kr, (3.16) can be written as<br />

= − tan δl . (3.17)


uk,l(r)<br />

r<br />

π s<strong>in</strong>(kr − l + δl) 2 <br />

kr<br />

37<br />

for kr ≫ l , (3.18)<br />

from which one can deduce th<strong>at</strong> the quantity δl represents the phase difference<br />

of the radial wave function between the sc<strong>at</strong>ter<strong>in</strong>g (V = 0 and δl = 0) and the<br />

non-sc<strong>at</strong>ter<strong>in</strong>g (V = 0 and δl = 0) cases.<br />

The radial-wave solution of a free particle <strong>in</strong> the sc<strong>at</strong>ter<strong>in</strong>g case can be rewrit-<br />

ten, us<strong>in</strong>g (3.17), as<br />

Rk,l(r) = uk,l(r)<br />

r = eiδl [cos δk,ljl(kr) − s<strong>in</strong> δk,lnl(kr)] r > Rb , (3.19)<br />

where the normaliz<strong>at</strong>ion constant is chosen so th<strong>at</strong> Rk,l(r) is jl(kr) for V = 0.<br />

Note th<strong>at</strong> the wave function for r > Rb is [23]<br />

ψk(r) =<br />

1<br />

(2π) 3/2<br />

∞<br />

(2l + 1)Rk,l(r)Pl(cos θ) r > Rb (3.20)<br />

l=0<br />

and thus the electronic density is<br />

n(r) = 2<br />

(2π) 3<br />

kF<br />

dk (|ψk(r)|<br />

0<br />

2<br />

(3.21)<br />

= 8π<br />

(2π) 3<br />

<br />

kF<br />

(2l + 1) dk k 2 (Rk,l(r)) 2 . (3.22)<br />

l<br />

0<br />

The boundary condition for the <strong>in</strong>terior radial wave function (r < Rb) is the<br />

logarithmic deriv<strong>at</strong>ive <strong>at</strong> the boundary.<br />

<br />

1<br />

R<strong>in</strong>t dR<br />

k,l<br />

<strong>in</strong>t<br />

<br />

k,l<br />

1<br />

=<br />

dr<br />

Rext k,l<br />

r=Rb<br />

dR ext<br />

k,l<br />

dr<br />

where R<strong>in</strong>t k,l is a <strong>in</strong>terior radial wave function and Rext k,l<br />

tion which is given by (3.15).<br />

<br />

r=Rb<br />

One can easily obta<strong>in</strong>, us<strong>in</strong>g the phase shift def<strong>in</strong>ition (3.17),<br />

(3.23)<br />

a exterior radial wave func-<br />

tan δl = kRbj ′ l (kRb)uk,l(Rb) + jl(kRb)uk,l(Rb) − Rbjl(kRb)u ′ k,l (Rb)<br />

kRbn ′ l (kRb)uk,l(Rb) + nl(kRb)uk,l(Rb) − Rbnl(kRb)u ′ k,l (Rb)<br />

(3.24)


where uk,l(r) is a <strong>in</strong>terior solution (for r < Rb).<br />

Once the radial equ<strong>at</strong>ions (3.14) for r < Rb are solved, the phase shift can be<br />

obta<strong>in</strong>ed by (3.24) and normaliz<strong>at</strong>ion of the radial wave function can be performed<br />

by (3.19). For very small kr, one can use the asymptotic approxim<strong>at</strong>ions 3 aga<strong>in</strong><br />

and obta<strong>in</strong> the expression<br />

tan δl → alk 2l+1<br />

For s-waves, this equ<strong>at</strong>ion becomes<br />

38<br />

as k → 0 . (3.25)<br />

tan δ0 δ0 → a0k as k → 0 . (3.26)<br />

One sees th<strong>at</strong> the sc<strong>at</strong>ter<strong>in</strong>g length 4 a0 is a slope of δ0(k) for small k values, which<br />

can be used to determ<strong>in</strong>e whether the potential is <strong>at</strong>tractive or repulsive and also<br />

to determ<strong>in</strong>e the required <strong>in</strong>terval of a k-mesh 5 for small k values. If there’s a<br />

virtual bound st<strong>at</strong>e (sc<strong>at</strong>ter<strong>in</strong>g resonance) around the bottom of conduction band,<br />

the sc<strong>at</strong>ter<strong>in</strong>g length will be very large due to the large density of <strong>in</strong>duced st<strong>at</strong>es for<br />

small k values. In this case, one needs to use a correspond<strong>in</strong>g very small <strong>in</strong>terval<br />

<strong>in</strong> a k-mesh around k = 0.<br />

3 r → 0, krjl(kr) (kr) l+1 /(Π l i=0 2i + 1) and − krnl(kr) (Π l−1<br />

i=0 2i + 1)/(kr)l ,<br />

r → ∞, krjl(kr) s<strong>in</strong>(kr − lπ/2) and − krnl(kr) cos(kr − lπ/2).<br />

4 a1 is called the sc<strong>at</strong>ter<strong>in</strong>g volume.<br />

5 In the computer program, one has to use discrete values of k for the cont<strong>in</strong>uum st<strong>at</strong>es<br />

k = 0 · · · kF and the <strong>in</strong>terval between k values affects the numerical precision of <strong>in</strong>te-<br />

gr<strong>at</strong>ion <strong>in</strong> k-space. The larger the magnitude of a0 is , the smaller the <strong>in</strong>terval between<br />

small k values is.


3.2.2. Density of <strong>in</strong>duced st<strong>at</strong>es <strong>in</strong> terms of phase shifts<br />

The density of <strong>in</strong>duced st<strong>at</strong>es can easily derived from the phase shifts as<br />

follows. [18] If one can assume all charges are conta<strong>in</strong>ed <strong>in</strong> a very large sphere of<br />

radius R0 (V (R0) → ∞), the radial wave functions for sc<strong>at</strong>tered st<strong>at</strong>es must vanish<br />

<strong>at</strong> r = R0. The asymptotic approxim<strong>at</strong>ions 3 can be used aga<strong>in</strong> due to the large R0<br />

and one can obta<strong>in</strong>, us<strong>in</strong>g (3.16), boundary conditions for free particle st<strong>at</strong>es and<br />

for sc<strong>at</strong>tered st<strong>at</strong>es. For free particle st<strong>at</strong>es,<br />

and for sc<strong>at</strong>tered st<strong>at</strong>es,<br />

kR0 − lπ<br />

2<br />

= nfπ<br />

kR0 − lπ<br />

2 + δl,m(k) = nsπ .<br />

One can assume th<strong>at</strong> k is cont<strong>in</strong>uous due to a very large radius R0 (kR0 ≫ l) and<br />

obta<strong>in</strong> the number of <strong>in</strong>duced st<strong>at</strong>es per k, △D(k),<br />

where dδl,m(k)<br />

dk<br />

△D(k)dk = d(ns − nf)<br />

dk =<br />

dk<br />

1<br />

π<br />

<br />

l,m,σ<br />

dδ σ l,m (k)<br />

dk<br />

39<br />

dk , (3.27)<br />

is summed over l, m, and σ to <strong>in</strong>clude the degener<strong>at</strong>e st<strong>at</strong>es.<br />

3.2.3. Immersion energies<br />

If E<strong>at</strong>om is given <strong>in</strong> the immersion equ<strong>at</strong>ion (3.1), the only term which one<br />

needs to f<strong>in</strong>d is Eimp − Epure. This is composed of the <strong>in</strong>duced k<strong>in</strong>etic energy, the<br />

electrost<strong>at</strong>ic energy, and the exchange-correl<strong>at</strong>ion energy.<br />

Eimp − Epure = △T + △C + △Exc<br />

(3.28)


1. The k<strong>in</strong>etic energy △T<br />

The k<strong>in</strong>etic energy which is similar to (2.28) can be written as<br />

△T =<br />

N<br />

i<br />

ɛi −<br />

<br />

dr n(r)veff(r)<br />

= △T bound + △T cond<br />

N<br />

<br />

= ɛi − dr n bound <br />

(r)veff(r) +<br />

i<br />

dr n cond <br />

(r)veff(r)<br />

where i <strong>in</strong>cludes the sp<strong>in</strong> s and bound and cond denote bound st<strong>at</strong>es and<br />

sc<strong>at</strong>tered st<strong>at</strong>es respectively.<br />

For bound st<strong>at</strong>es,<br />

△T bound =<br />

N<br />

i<br />

ɛi −<br />

where i is only for the occupied bound st<strong>at</strong>es.<br />

Similarly, for sc<strong>at</strong>tered st<strong>at</strong>es, s<strong>in</strong>ce k is cont<strong>in</strong>uous,<br />

△T cond =<br />

kF<br />

0<br />

<br />

k 2 <br />

△D(k) dk −<br />

,<br />

40<br />

dr n bound (r)veff(r) , (3.29)<br />

dr n cond (r)veff(r) , (3.30)<br />

where △D(k) is the number of st<strong>at</strong>es per k <strong>in</strong>duced by the potential veff.<br />

The first term can be rewritten, us<strong>in</strong>g (3.27) and E = k 2 , as<br />

kF<br />

0<br />

k 2 △D(k) dk = 1<br />

kF <br />

2<br />

k<br />

π 0<br />

= 1<br />

π<br />

= 1<br />

π<br />

<br />

l,m,σ<br />

<br />

l,m,σ<br />

l,m,σ<br />

EF<br />

0<br />

dδ σ l,m (k)<br />

dk<br />

E dδσ l,m (E)<br />

dE<br />

EF δ σ l,m(EF ) − 1<br />

π<br />

dk (3.31)<br />

dE (3.32)<br />

<br />

l,m,σ<br />

EF<br />

0<br />

δ σ l,m(E) dE , (3.33)<br />

where EF is the Fermi energy which is k2 F . Add<strong>in</strong>g the k<strong>in</strong>etic energies for<br />

the bound and sc<strong>at</strong>tered st<strong>at</strong>es together,


△T =<br />

N<br />

i<br />

+ 1<br />

π<br />

ɛi −<br />

<br />

l,m,σ<br />

<br />

dr n(r)veff(r)<br />

EF δ σ l,m(EF ) − 1<br />

π<br />

<br />

l,m,σ<br />

EF<br />

0<br />

41<br />

δ σ l,m(E) dE . (3.34)<br />

Note th<strong>at</strong> the first term <strong>in</strong> (3.33) is the energy, EF Z, required to add the<br />

extra Z electrons to the system, which is easily verified by the Friedel sum<br />

rule (See (3.45).) and the second term is the k<strong>in</strong>etic energy caused by the<br />

<strong>in</strong>teraction between an impurity and an electron gas <strong>in</strong> a metal.<br />

2. The electrost<strong>at</strong>ic energy ∆C<br />

S<strong>in</strong>ce the Hartree energy of the background system which has a uniform<br />

density n0 is zero, one can write the Coulomb energy as<br />

△C = 1<br />

2 e2<br />

<br />

dr dr ′ ∆n(r)∆n(r′ )<br />

|r ′ <br />

− e dr ∆n(r)Vext (r) (3.35)<br />

− r|<br />

= e 2<br />

<br />

1<br />

dr dr<br />

2<br />

′ ∆n(r′ )<br />

|r ′ <br />

Z<br />

− ∆n(r) , (3.36)<br />

− r| r<br />

where ∆n(r) = n(r) − n0(r) and the second term ( eZ ) is the external po-<br />

r<br />

tential due to the nucleus of an impurity <strong>at</strong>om. However, this can be further<br />

simplified us<strong>in</strong>g [24]<br />

<br />

dr ′ 1<br />

|r − r ′ | =<br />

∞<br />

r<br />

l=0<br />

l <<br />

r l+1<br />

><br />

Rb<br />

= 2π<br />

r0<br />

<br />

dr ′ r ′2<br />

Pl(cos θ)r ′2 dr ′ s<strong>in</strong> θdθdφ (3.37)<br />

∞<br />

l=0<br />

rl <<br />

r l+1<br />

1<br />

Pl(x)dx , (3.38)<br />

> −1<br />

where θ is the angle between r and r ′ and Pl(x) are the Legendre polynomials.<br />

Us<strong>in</strong>g the fact th<strong>at</strong> P0(x) = 1 and 1<br />

−1 dx Pl(x)P0(x) = 2 · δl,0,<br />

F<strong>in</strong>ally,<br />

<br />

dr ′<br />

1<br />

|r − r ′ |<br />

r<br />

1<br />

= 4π r<br />

r r0<br />

′2 dr ′ +<br />

Rb<br />

r<br />

r ′ dr ′<br />

<br />

. (3.39)


Rb<br />

∆C = 4π<br />

× e 2<br />

r0<br />

<br />

2π<br />

r∆n(r)<br />

r<br />

3. The exchange-correl<strong>at</strong>ion energy ∆Exc<br />

r0<br />

42<br />

∆n(r ′ )r ′2 dr ′ Rb<br />

+ 2πr ∆n(r<br />

r<br />

′ )r ′ dr ′ <br />

− Z dr . (3.40)<br />

The exchange-correl<strong>at</strong>ion energy for non-sp<strong>in</strong>-polarized systems is simply<br />

Exc[n]−Exc[n0], <strong>in</strong> which n0 is the background electron density and Exc[n] =<br />

dr n(r)ɛxc(n(r)).<br />

The exchange-correl<strong>at</strong>ion energy for sp<strong>in</strong>-polarized systems can be written<br />

as<br />

∆Exc = Exc[n + , n − ] − Exc[n + 0 , n − 0 ] for sp<strong>in</strong>-up (3.41)<br />

+ Exc[n − , n + ] − Exc[n − 0 , n + 0 ] for sp<strong>in</strong>-down, (3.42)<br />

where n + is the sp<strong>in</strong>-up density and n − is the sp<strong>in</strong>-down density.<br />

Now, the immersion energy is<br />

N<br />

<br />

Eimm = ɛi − dr n(r)veff(r)<br />

i<br />

+ 1 <br />

EF δ<br />

π<br />

l,m,σ<br />

σ l,m(EF ) − 1 <br />

π<br />

l,m,σ<br />

+ 4πe 2<br />

Rb<br />

r<br />

r∆n(r) 2π<br />

r0<br />

+ Exc[n + , n − ] − Exc[n + 0 , n − 0 ]<br />

EF<br />

δ σ l,m(E) dE<br />

0<br />

∆n(r<br />

r0<br />

′ )r ′2 dr ′ Rb<br />

+ 2πr<br />

r<br />

∆n(r ′ )r ′ dr ′ <br />

− Z dr<br />

+ Exc[n − , n + ] − Exc[n − 0 , n + 0 ] − E<strong>at</strong>om . (3.43)<br />

Note th<strong>at</strong> the effective potential energy is, by (2.24) and (3.39),<br />

Veff(r) = 4π e2 r<br />

1<br />

∆n(r<br />

r<br />

′ )r ′2 dr ′ Rb<br />

+ ∆n(r ′ )r ′ dr ′<br />

<br />

r0<br />

r<br />

− e2 Z<br />

r + vxc(n(r)) − vxc(n0) . (3.44)


3.3. Friedel sum rule and Friedel oscill<strong>at</strong>ions<br />

Two important phenomena <strong>in</strong> many-body problems are the screen<strong>in</strong>g of an<br />

impurity <strong>at</strong>om <strong>in</strong> an electron gas and the oscill<strong>at</strong>ory behavior of the electron density<br />

<strong>in</strong> the vic<strong>in</strong>ity of an impurity <strong>at</strong>om. [10] [11] [12]<br />

The number of <strong>in</strong>duced electrons due to an impurity <strong>at</strong>om can be easily<br />

obta<strong>in</strong>ed us<strong>in</strong>g (3.27).<br />

∆N = Z =<br />

kF<br />

0<br />

dk △D(k) =<br />

kF<br />

0<br />

dk 1<br />

π<br />

<br />

l,m,σ<br />

δ σ l,m (k)<br />

dk<br />

= 2<br />

π<br />

43<br />

<br />

(2l + 1)δl(kF ) , (3.45)<br />

where Z is the positive valence on an impurity, the factor of 2 is the sp<strong>in</strong> degeneracy<br />

σ, and (2l + 1) is the orbital degeneracy rel<strong>at</strong>ed to m. This rel<strong>at</strong>ion is called the<br />

Friedel sum rule and st<strong>at</strong>es th<strong>at</strong> the displaced electronic charge ∆N · e exactly<br />

cancels the charge Z · |e| of the impurity and th<strong>at</strong> the Coulomb potential of 1/r<br />

becomes a screened potential with the density of the screen<strong>in</strong>g electrons determ<strong>in</strong>ed<br />

<strong>in</strong> a self-consistent way. It should be noted th<strong>at</strong> if an impurity <strong>at</strong>om has one<br />

electron <strong>in</strong> bound st<strong>at</strong>es and thus has an excess valency Z − 1, one has to remove<br />

the contribution δ/π of the s<strong>in</strong>gle bound st<strong>at</strong>e on the right-hand side of (3.45).<br />

The value of the phase shift δ for a bound st<strong>at</strong>e should be π to be consistent. [15]<br />

This is actually another st<strong>at</strong>ement of Lev<strong>in</strong>son’s theorem.<br />

In order to see how the density varies <strong>at</strong> large distances (Friedel oscill<strong>at</strong>ions),<br />

one needs to take the asymptotic limit. (See Sec. 3.2.1, [15], and [18].) S<strong>in</strong>ce the<br />

radial wave function of free particles depends on only jl(kr)) as mentioned <strong>in</strong> Sec.<br />

3.2.1,<br />

l


n(r) − n0 = ∆n(r) = 2<br />

(2π) 3<br />

kF<br />

0<br />

= 8π<br />

(2π) 3<br />

1<br />

lim ∆n(r) =<br />

r→∞ π2r2 = 1<br />

π 2 r 2<br />

<br />

(2l + 1)<br />

l<br />

<br />

(2l + 1)<br />

l<br />

<br />

(2l + 1)<br />

l<br />

kF<br />

0<br />

kF<br />

0<br />

kF<br />

0<br />

44<br />

dk (|ψk(r)| 2 − |φk(r)| 2 ) (3.46)<br />

dk k 2 (uk,l(r)) 2<br />

(<br />

r2 − (jl(kr)) 2 ) (3.47)<br />

dk (s<strong>in</strong> 2 (kr + δl(k) − lπ/2) − (s<strong>in</strong> 2 (kr − lπ/2) (3.48)<br />

dk 1<br />

2 (cos(2kr − lπ) − (cos(2kr + 2δl(k) − lπ) , (3.49)<br />

where the normaliz<strong>at</strong>ion constant for radial solutions uk,l(r)<br />

r<br />

is chosen as <strong>in</strong> (3.19).<br />

Notice th<strong>at</strong> one can use (3.47) to f<strong>in</strong>d the density difference for sc<strong>at</strong>tered<br />

st<strong>at</strong>es numerically.<br />

Approxim<strong>at</strong>ely us<strong>in</strong>g δl(k) = δl(kF ) + (k − kF )(dδ/dk), [18]<br />

(−1)l<br />

lim ∆n(r) <br />

r→∞ 2π2r2 (−1)l<br />

4π 2 r 3<br />

− (−1)l<br />

4π 2 r 3<br />

<br />

(2l + 1)<br />

l<br />

kF<br />

× dk cos(2kr) − cos(2kr + 2δl(kF ) + 2(k − kF )<br />

0<br />

dδ<br />

dk )<br />

<br />

(3.50)<br />

<br />

(2l + 1) [s<strong>in</strong>(2kF r) − (s<strong>in</strong>(2kF r + 2δl(kF )))] (3.51)<br />

l<br />

<br />

l<br />

2(2l + 1) s<strong>in</strong>(δl(kF )) cos(2kF r + δl(kF )) + O( 1<br />

) (3.52)<br />

r4 The electronic density difference <strong>at</strong> large distances oscill<strong>at</strong>es with the wave number<br />

2kF and decreases as r −3 . The Friedel’s theorem is <strong>in</strong>dependent of the n<strong>at</strong>ure of<br />

the impurity <strong>at</strong>om and the impurity determ<strong>in</strong>es only the values of phase shifts. [18]<br />

The wave length λ ∼ π/kF is thus frequently used to verify the numerical results<br />

as <strong>in</strong> Ch. 4 and 5.<br />

3.4. Dielectric functions<br />

The screen<strong>in</strong>g of the Coulomb potential of an impurity and the behavior of<br />

the <strong>in</strong>duced electronic density can be derived us<strong>in</strong>g dielectric functions as well. [15]


[18] It is convenient to def<strong>in</strong>e the dielectric functions us<strong>in</strong>g potentials s<strong>in</strong>ce one has<br />

to use potentials <strong>in</strong> the Kohn-Sham equ<strong>at</strong>ions and the Fourier components of the<br />

potential are easily calcul<strong>at</strong>ed numerically. The dielectric function can be found<br />

<strong>in</strong> the usual def<strong>in</strong>ition,<br />

Eq + 4πPq = Dq = ɛ(w, q)Eq , (3.53)<br />

where Eq is the longitud<strong>in</strong>al electric field. If one writes the effective potential V eff<br />

as the external potential V ext plus the <strong>in</strong>duced potential V <strong>in</strong>d , V eff = V ext + V <strong>in</strong>d ,<br />

us<strong>in</strong>g eEq = −iqVq, one can obta<strong>in</strong> from (3.53)<br />

V eff<br />

q<br />

− V <strong>in</strong>d<br />

q<br />

ɛ(w, q) =<br />

= V ext<br />

q<br />

V ext<br />

q<br />

V eff<br />

q<br />

= ɛ(w, q)V eff<br />

q<br />

45<br />

(3.54)<br />

. (3.55)<br />

Note th<strong>at</strong> Pq = −E <strong>in</strong>d<br />

q /4π. S<strong>in</strong>ce the dielectric function is calcul<strong>at</strong>ed as the r<strong>at</strong>io<br />

of the external potential to the effective potential, one can estim<strong>at</strong>e the response<br />

of an electron gas due to an impurity <strong>at</strong>om, which is very important <strong>in</strong> a sense th<strong>at</strong><br />

one needs to determ<strong>in</strong>e the stability of the self-consistent procedure for a particular<br />

impurity <strong>at</strong>om. (See Sec. 4.2.)<br />

If the charge of the impurity is Z, the external potential becomes<br />

Z<br />

r<br />

1<br />

=<br />

(2π) 3<br />

<br />

The effective potential <strong>in</strong> the l<strong>in</strong>ear response region is<br />

V eff (r, w) = 1<br />

(2π) 3<br />

<br />

= 1<br />

(2π) 3<br />

= 1<br />

(2π) 3<br />

<br />

<br />

iq·r Z4π<br />

dq e<br />

q2 . (3.56)<br />

dq e iq·r V eff (w, q) (3.57)<br />

dq e iq·r V ext (w, q)<br />

ɛ(w, q)<br />

(3.58)<br />

dq e iq·r Z4π<br />

q2ɛ(w, q)<br />

. (3.59)


Density functional theory is essentially a st<strong>at</strong>ic theory and further more if the<br />

impurity is st<strong>at</strong>ionary, one can use w = 0. The Poisson equ<strong>at</strong>ion is<br />

∇ 2 V eff (r) = −4π[∆n(r) + Zδ(r)] , (3.60)<br />

where ∆n(r) is the charge <strong>in</strong>duced by Zδ(r). The screen<strong>in</strong>g electronic density is<br />

now<br />

∆n(r) = Z<br />

(2π) 3<br />

<br />

dq e iq·r<br />

<br />

1<br />

− 1<br />

ɛ(0, q)<br />

46<br />

. (3.61)<br />

One notices th<strong>at</strong> if there is a s<strong>in</strong>gularity <strong>in</strong> the dielectric constant, the screen<strong>in</strong>g<br />

electronic density will be oscill<strong>at</strong><strong>in</strong>g as the Friedel oscill<strong>at</strong>ions.<br />

The total <strong>in</strong>duced charge is<br />

<br />

∆N =<br />

dr ∆n(r) = Z<br />

(2π) 3<br />

<br />

dq dr e iq·r<br />

<br />

1<br />

ɛ(0, q)<br />

<br />

1<br />

= Z dq δ(q) − 1<br />

ɛ(0, q)<br />

<br />

1<br />

= Z − 1<br />

ɛ(0, 0)<br />

<br />

− 1<br />

(3.62)<br />

(3.63)<br />

(3.64)<br />

If one use the Thomas-Fermi dielectric constant for ɛ(0, q) ( [15], [18], and [25]),<br />

where k 2 s = 4e2 mkF<br />

2 π<br />

ɛT F (0, q) = q2 + k2 s<br />

q2 <br />

1<br />

∆N = Z<br />

ɛT F (0, 0)<br />

(3.65)<br />

<br />

− 1 = −Z , (3.66)<br />

= 4.6/rs and rs is the Wigner-Seitz radius. The Thomas-Fermi<br />

dielectric constant shows the correct Friedel sum rule result but does not <strong>in</strong>clude<br />

the Friedel oscill<strong>at</strong>ions. The effective potential screened by the electron gas is<br />

V eff (r, w) = 1<br />

(2π) 3<br />

<br />

= Z<br />

r e−ksr<br />

iq·r Z4π<br />

dq e<br />

q2 q 2<br />

q 2 + k 2 s<br />

(3.67)<br />

. (3.68)


Note th<strong>at</strong> this potential is also exactly the same as the Yukawa potential. The<br />

screen<strong>in</strong>g length is def<strong>in</strong>ed by (3.67) as ls = 1/ks ∼ r 1/2<br />

s .<br />

The L<strong>in</strong>dhard dielectric constant (<strong>at</strong> T = 0) can be derived by the self-<br />

consistent field [15] or RPA approxim<strong>at</strong>ions [18] or straightforward calcul<strong>at</strong>ions<br />

[25]. The dielectric constant <strong>in</strong> the L<strong>in</strong>dhard theory is<br />

ɛL<strong>in</strong>(q) = 1 + 4π<br />

q2 e2mkF 22π2 <br />

1 + 4k2 <br />

F − q2 <br />

ln <br />

q + 2kF <br />

<br />

4qkF q − 2kF<br />

47<br />

(3.69)<br />

Note th<strong>at</strong> the L<strong>in</strong>dhard dielectric constant is not analytic <strong>at</strong> q = 2kF , which results<br />

<strong>in</strong> th<strong>at</strong> the <strong>in</strong>duced density and the screened effective potential of an impurity<br />

charge oscill<strong>at</strong>e as 1<br />

r 3 cos 2kF r. The Friedel’s theorem has been verified aga<strong>in</strong> with<br />

dielectric constants. The limit of the L<strong>in</strong>dhard dielectric constant as q → 0 is just<br />

the Thomas-Fermi dielectric constant. As one can notice, the dielectric constant is<br />

<strong>in</strong> general gett<strong>in</strong>g close to 1 as q <strong>in</strong>creases, which means the effective potential <strong>at</strong><br />

high q’s will be almost same as the external potential and thus the <strong>in</strong>duced density<br />

<strong>at</strong> high q’s does not contribute much. However, if there is a virtual bound st<strong>at</strong>e <strong>at</strong><br />

high q values, one can see a small peak <strong>in</strong> the dielectric function <strong>at</strong> correspond<strong>in</strong>g<br />

q values s<strong>in</strong>ce there’s a contribution of the <strong>in</strong>duced density <strong>at</strong> the q values to the<br />

dielectric constant. (see Ch. 5.)<br />

3.5. Virtual bound st<strong>at</strong>e resonance<br />

If an impurity <strong>at</strong>om has bound st<strong>at</strong>es, the electrons <strong>in</strong> bound st<strong>at</strong>es will be<br />

localized <strong>in</strong> the vic<strong>in</strong>ity of the impurity ion. If the <strong>at</strong>tractive effective potential of<br />

the impurity is reduced or not sufficiently strong enough to have a bound st<strong>at</strong>e, the<br />

bound st<strong>at</strong>e will be merged <strong>in</strong>to the conduction band <strong>in</strong> a metal. In this case, the<br />

electrons with positive energy rel<strong>at</strong>ive to the bottom of the conduction band still<br />

tend to localize around the impurity ion and <strong>in</strong>duce a narrow peak <strong>in</strong> the density


of st<strong>at</strong>es of the conduction band, which is called a virtual bound st<strong>at</strong>e resonance.<br />

A virtual bound st<strong>at</strong>e can be seen very well <strong>in</strong> the non-<strong>in</strong>teract<strong>in</strong>g Anderson model<br />

by calcul<strong>at</strong>ions of density of st<strong>at</strong>es us<strong>in</strong>g Green’s functions. [26] The Hamiltonian<br />

takes the form, <strong>in</strong> the non-<strong>in</strong>teract<strong>in</strong>g Anderson model,<br />

H = <br />

σ<br />

ɛdc †<br />

d,σ cd,σ + <br />

k,σ<br />

ɛkc †<br />

k,σ<br />

48<br />

k,σck,σ + <br />

(Vkc †<br />

d,σck,σ + V ∗<br />

k c †<br />

k,σcd,σ) , (3.70)<br />

where Vk is an overlap m<strong>at</strong>rix element between the <strong>at</strong>omic d level of the impurity<br />

ion and the conduction electrons and ɛd is the energy of the d level of the impurity.<br />

By tak<strong>in</strong>g appropri<strong>at</strong>e m<strong>at</strong>rix elements of (ɛ ± is − H)G ± (ɛ) = I, one can f<strong>in</strong>d<br />

the coupled equ<strong>at</strong>ions of Green’s functions, which lead to the T m<strong>at</strong>rix element by<br />

compar<strong>in</strong>g with the expression G + = G + 0 + G + 0 T (ɛ + )G + 0 . Note th<strong>at</strong> ɛ ± = ɛ ± is.<br />

where<br />

〈k|T (ɛ)|k ′ 〉 = V ∗<br />

k G +<br />

d,d (ɛ)V ′ k , (3.71)<br />

G +<br />

d,d (ɛ) =<br />

1<br />

ɛ + is − ɛd − <br />

k<br />

|Vk| 2<br />

ɛ+is−ɛk<br />

. (3.72)<br />

One can f<strong>in</strong>d the phase shift from T m<strong>at</strong>rix element us<strong>in</strong>g the prescription<br />

lims→+0<br />

where<br />

1<br />

x−is<br />

1 = P + iπδ(x).<br />

x<br />

δ(ɛ) = arg det T (ɛ + ) (3.73)<br />

= π<br />

<br />

−ɛ + ɛd + Λ(ɛ)<br />

− tan−1<br />

(3.74)<br />

2 ∆(ɛ)<br />

Λ(ɛ) = P |Vk| 2<br />

, ∆(ɛ) = π<br />

ɛ − ɛk<br />

<br />

|Vk| 2 δ(ɛ − ɛk) . (3.75)<br />

k<br />

One can derive the density of st<strong>at</strong>es ρ(ɛ) from the expression of the Green’s function<br />

ρ(ɛ) = ∓ 1<br />

π ImTrG± (ɛ) [26] or from the Friedel sum rule (3.45). The density of st<strong>at</strong>e<br />

<strong>in</strong>duced by the impurity is given by<br />

k


∆ρ(ɛ) = 1 dδ(ɛ)<br />

π dɛ<br />

49<br />

, (3.76)<br />

where δ(ɛ) = 2 <br />

l (2l + 1)δl(ɛ) for spherically symmetric potentials. If one tries to<br />

f<strong>in</strong>d Λ(ɛ) and ∆(ɛ) with a simple density of st<strong>at</strong>es,<br />

ρ0(ɛ) = ρ0 for − D < ɛ < D , (3.77)<br />

by neglect<strong>in</strong>g the k dependence of Vk, one can obta<strong>in</strong><br />

∆(ɛ) = πρ0|V | 2 = C for − D < ɛ < D , (3.78)<br />

= 0 for ɛ ≤ −D or ɛ ≥ D (3.79)<br />

Λ(ɛ) = ρ0|V | 2 <br />

<br />

ln <br />

D + ɛ<br />

<br />

D<br />

− ɛ<br />

. (3.80)<br />

Note th<strong>at</strong> Λ(ɛ) can be easily calcul<strong>at</strong>ed by the Kramers-Kronig rel<strong>at</strong>ion.<br />

Λ(ɛ) = 1<br />

D<br />

dɛ<br />

π −D<br />

′ ∆(ɛ′ )<br />

ɛ − ɛ ′<br />

(3.81)<br />

It can be shown th<strong>at</strong> the density of <strong>in</strong>duced st<strong>at</strong>es approxim<strong>at</strong>ely follows a<br />

Lorentzian form and thus there’s a virtual bound st<strong>at</strong>e <strong>at</strong> ɛ ′ d .<br />

δ(ɛ) = π<br />

2 − tan−1 −ɛ + ɛ′ d<br />

C<br />

∆ρ(ɛ) = 1<br />

π<br />

dδ(ɛ)<br />

dɛ<br />

= 1<br />

π<br />

(3.82)<br />

C<br />

(ɛ − ɛ ′ d )2 + C 2 , (3.83)<br />

where ɛ ′ d − ɛd − Λ(ɛ ′ d ) = 0. Fig. 3.3 shows the density of st<strong>at</strong>es <strong>in</strong>duced by the<br />

impurity with the density of st<strong>at</strong>es (Fig. 3.2) <strong>in</strong> the conduction band. S<strong>in</strong>ce the<br />

potential (V = −0.13) is weak and ɛd = −0.3 is well with<strong>in</strong> the conduction band,<br />

there is a resonance around the energy of −0.3. This virtual bound st<strong>at</strong>e becomes<br />

a real bound st<strong>at</strong>e below the bottom of the conduction band for V < −0.6. (Fig.<br />

3.4) S<strong>in</strong>ce there is one electron <strong>in</strong> a bound st<strong>at</strong>e for V < −0.6, the number of<br />

electrons <strong>in</strong> the conduction band is zero as shown <strong>in</strong> Fig. 3.4. Note th<strong>at</strong> <strong>in</strong> this


FIGURE 3.2. The density of st<strong>at</strong>es ρ0(ɛ) <strong>in</strong> the conduction band and the correspond<strong>in</strong>g<br />

phase shift. ɛd = −0.3 and ∆(ɛ) = (−0.13) 2 πρ0(ɛ) are used.<br />

FIGURE 3.3. There is a virtual bound st<strong>at</strong>e resonance <strong>at</strong> the solution of<br />

ɛ − ɛd − Λ(ɛ) = 0. The density of st<strong>at</strong>es ∆ρ(ɛ) <strong>in</strong>duced by the impurity. Calcul<strong>at</strong>ed<br />

from Fig. 3.2 <strong>in</strong> M<strong>at</strong>hem<strong>at</strong>ica.<br />

case the system has always one extra electron due to the impurity, regardless of<br />

the number of virtual bound st<strong>at</strong>es. Actually, the transition between bound and<br />

unbound st<strong>at</strong>es is smooth and the properties such as the total energy are smooth<br />

functions of the strength of the potential as well. [27]<br />

50


FIGURE 3.4. Shows how the density of st<strong>at</strong>es ∆ρ(ɛ) varies <strong>in</strong> the conduction band<br />

for the different strength of the potential. Calcul<strong>at</strong>ed from Fig. 3.2 <strong>in</strong> M<strong>at</strong>hem<strong>at</strong>ica.<br />

The number of electrons <strong>in</strong> the conduction band is obta<strong>in</strong>ed numerically.<br />

3.6. Sc<strong>at</strong>ter<strong>in</strong>g length and bound st<strong>at</strong>e energy<br />

The purpose of this section is to give emphasis to the rel<strong>at</strong>ionship between<br />

the sc<strong>at</strong>ter<strong>in</strong>g length and the bound st<strong>at</strong>e energy, which easily can be derived for<br />

51


sc<strong>at</strong>ter<strong>in</strong>g by a spherically symmetric square-well potential <strong>at</strong> low energies (such<br />

as a deuteron system). [28] [29] Notice th<strong>at</strong> it is sufficient to consider only the<br />

S-wave for low energies. Suppose the potential has the form,<br />

V (r) = −V0<br />

for r < r0<br />

52<br />

= 0 for r > r0 . (3.84)<br />

One can immedi<strong>at</strong>ely write the correspond<strong>in</strong>g radial equ<strong>at</strong>ions and solutions for<br />

bound st<strong>at</strong>es of the energy EB,<br />

d 2 u(r)<br />

dr 2 + k2 u(r) = 0 and u(r) = A s<strong>in</strong>(kr) for r < r0 (3.85)<br />

d 2 u(r)<br />

dr 2 − k2 Bu(r) = 0 and u(r) = B exp(−kBr) for r > r0 , (3.86)<br />

with k 2 = V0 − |EB| and k 2 B = |EB|. The boundary condition <strong>at</strong> r = r0 gives rise<br />

to<br />

k cot(ka) = −kB . (3.87)<br />

S<strong>in</strong>ce the deuteron system has a small b<strong>in</strong>d<strong>in</strong>g energy |E|, us<strong>in</strong>g k0 k(k0 = V 1/2<br />

0 ),<br />

If k0r0 π/2, one obta<strong>in</strong>s<br />

k0 tan( π<br />

2 − k0r0) = −kB . (3.88)<br />

k0r0 = π kB<br />

+<br />

2 k0<br />

For low-energy sc<strong>at</strong>ter<strong>in</strong>g, similarly,<br />

. (3.89)<br />

u(r) = A s<strong>in</strong>(kr) for r < r0 (3.90)<br />

u(r) = B s<strong>in</strong>(k ′ r + δ) for r > r0 , (3.91)<br />

with k 2 = V0 + E and k ′2 = E. Us<strong>in</strong>g the boundary condition <strong>at</strong> r = r0,


k cot kr0 = k ′ cot(k ′ r0 + δ) (3.92)<br />

tan δ =<br />

k ′<br />

k tan kr0 − tan k ′ r0<br />

1 + k′<br />

k tan kr0 tan k ′ r0<br />

For k ′ → 0, one can use k → k0, which leads to<br />

δ tan δ k ′ <br />

tan k0r0<br />

r0 − 1<br />

k0r0<br />

Note th<strong>at</strong> the total cross section is given by<br />

lim<br />

k ′ 4π<br />

=<br />

→0 k ′2 s<strong>in</strong>2 δ0 4π( δ0<br />

k<br />

<br />

= 4π<br />

′ )2<br />

where the sc<strong>at</strong>ter<strong>in</strong>g length a is def<strong>in</strong>ed by<br />

r0<br />

tan k0r0<br />

k0r0<br />

53<br />

. (3.93)<br />

. (3.94)<br />

2 − 1<br />

(3.95)<br />

(3.96)<br />

= 4πa 2 , (3.97)<br />

δ0 → −ak ′<br />

and k0 cot k0r0 = 1<br />

as k ′ → 0 (3.98)<br />

r0 − a<br />

The quantity 4πa 2 only depends on the potential.<br />

One can also rewrite (3.93) to obta<strong>in</strong><br />

k ′ cot δ = k cot kr0 + k ′ tan k ′ r0<br />

1 − k<br />

k ′ cot kr0 tan k ′ r0<br />

. (3.99)<br />

(3.100)<br />

One f<strong>in</strong>ally can expand (3.100) <strong>in</strong> terms of k ′ keep<strong>in</strong>g terms up to k ′2 , with (3.99),<br />

k = k ′2 + k 2 0 k0 + k′2<br />

2k0 , and tan k′ r0 k ′ r0 + 1<br />

3 (k′ r0) 3 ,<br />

where reff = r0<br />

k ′ cot δ − 1 1<br />

+<br />

a<br />

<br />

1 − 1<br />

(k0r0) 2<br />

r0<br />

a<br />

1 r0 − ( 3 a )2<br />

expansion for k ′ cot δ for low-energy sc<strong>at</strong>ter<strong>in</strong>g.<br />

2 k′2reff + · · ·<br />

<br />

, (3.101)<br />

. This result is known as the effective range<br />

Now, <strong>in</strong> order to rel<strong>at</strong>e the bound st<strong>at</strong>e energy to the sc<strong>at</strong>ter<strong>in</strong>g length, one<br />

needs to write from (3.89), s<strong>in</strong>ce kB/k0 ≪ 1,


tan k0r0 = tan(π/2 + kB/k0) = −k0/kB . (3.102)<br />

One can now rewrite (3.100) with tan k ′ r0 k ′ r0 as<br />

k ′ cot δ − kB<br />

1 + r0kB<br />

r0<br />

54<br />

+ k<br />

1 + r0kB<br />

′2 . (3.103)<br />

By compar<strong>in</strong>g this equ<strong>at</strong>ion with the effective range expansion (3.101), one can<br />

obta<strong>in</strong><br />

a = r0 + 1<br />

kB<br />

. (3.104)<br />

This result provides a way to see whether the self-consistent solutions from the<br />

numerical calcul<strong>at</strong>ions are correct or not by compar<strong>in</strong>g the sc<strong>at</strong>ter<strong>in</strong>g length a and<br />

the b<strong>in</strong>d<strong>in</strong>g energy. The potential range r0 also can be estim<strong>at</strong>ed from the equ<strong>at</strong>ion<br />

above.


4. IMPLEMENTATION<br />

In general, the procedure to implement a theory and transl<strong>at</strong>e it <strong>in</strong>to com-<br />

puter codes is not easy due to the limit<strong>at</strong>ions of a computer. Two numerical<br />

problems which directly affect the numerical precision orig<strong>in</strong><strong>at</strong>e from approxima-<br />

tions to <strong>in</strong>f<strong>in</strong>ity and cont<strong>in</strong>uity, which frequently become the largest problems and<br />

can be expressed only approxim<strong>at</strong>ely <strong>in</strong> numerical calcul<strong>at</strong>ions. In most cases, one<br />

needs to use a large cut-off value for <strong>in</strong>f<strong>in</strong>ity and very small <strong>in</strong>terval for cont<strong>in</strong>uity<br />

<strong>in</strong> numerical calcul<strong>at</strong>ions to obta<strong>in</strong> a reasonable answer. It is important to f<strong>in</strong>d<br />

out how those parameters can be determ<strong>in</strong>ed. In this chapter, the iter<strong>at</strong>ive scheme<br />

to accomplish the self-consistent solutions and the numerical methods to f<strong>in</strong>d the<br />

eigenst<strong>at</strong>es will be <strong>in</strong>troduced. The technique to solve the numerical problems, <strong>in</strong><br />

which physical understand<strong>in</strong>g is <strong>in</strong>dispensable, will be discussed. It needs to be<br />

emphasized th<strong>at</strong> spherically symmetric potentials are used throughout this work.<br />

4.1. Perspective scheme for self-consistent solutions<br />

4.1.1. Non-sp<strong>in</strong> polarized system<br />

Fig. 4.1 shows one possible flow chart to f<strong>in</strong>d a self-consistent solution for<br />

the KS equ<strong>at</strong>ions. One can also start with the <strong>in</strong>itial electronic density, but <strong>in</strong><br />

this case the effective potential is more convenient to deal with s<strong>in</strong>ce the rel<strong>at</strong>ive<br />

numerical response of the effective potential <strong>in</strong> r-space is much larger than th<strong>at</strong> of<br />

the density. With an <strong>in</strong>put effective potential, one can solve the KS equ<strong>at</strong>ions to<br />

f<strong>in</strong>d the density, from which one can construct the output effective potential. A<br />

self-consistent solution means th<strong>at</strong> the <strong>in</strong>put effective potential must be the same<br />

as the output effective potential. In practice, one can stop the calcul<strong>at</strong>ions if the<br />

55


FIGURE 4.1. Flow chart for non-polarized systems.<br />

<strong>in</strong>put potential is very close to the output potential, when numerical precision<br />

required for the immersion energy can be achieved and thus further calcul<strong>at</strong>ions<br />

do not affect the f<strong>in</strong>al answers. In order to compare the <strong>in</strong>put potential with the<br />

output potential, one can use<br />

(∆V ) 2 = 1<br />

<br />

Ω<br />

56<br />

dr [V <strong>in</strong>put<br />

eff (r) − V output<br />

eff (r)] 2 , (4.1)<br />

where Ω is the total volume of the system. The potential difference ∆V can be<br />

used to f<strong>in</strong>d the physical quantities such as immersion energy, bound st<strong>at</strong>e energy,


and magnetiz<strong>at</strong>ion s<strong>in</strong>ce the self-consistent solution means th<strong>at</strong> ∆V = 0. One can<br />

<strong>at</strong>tempt to determ<strong>in</strong>e the physical quantities correspond<strong>in</strong>g to the zero value of<br />

∆V by us<strong>in</strong>g an appropri<strong>at</strong>e fitt<strong>in</strong>g method for the d<strong>at</strong>a po<strong>in</strong>ts.<br />

In order to make progress, one needs a numerical method to determ<strong>in</strong>e the<br />

<strong>in</strong>put potential from the output potential for the next iter<strong>at</strong>ion, which will be<br />

expla<strong>in</strong>ed <strong>in</strong> Sec. 4.2.<br />

4.1.2. Sp<strong>in</strong> polarized system<br />

One big difference between sp<strong>in</strong> polarized systems and non-sp<strong>in</strong> polarized sys-<br />

tems is the need to solve two different sets of KS equ<strong>at</strong>ions, for n+(r) and n−(r),<br />

which doubles the calcul<strong>at</strong>ion time. Only the exchange-correl<strong>at</strong>ion energies and po-<br />

tentials are, however, explicitly dependent upon the sp<strong>in</strong> density (n+(r) − n−(r))<br />

as <strong>in</strong> (2.41), (2.42), and Sec. 2.2.7.4. The Coulomb potentials and energies do not<br />

depend on the sp<strong>in</strong> density, but only on the total density (n+(r) + n−(r)).<br />

4.2. Mix<strong>in</strong>g scheme<br />

S<strong>in</strong>ce the KS equ<strong>at</strong>ions are solved self-consistently <strong>in</strong> this work and the time<br />

to f<strong>in</strong>d self-consistent solutions is directly proportional to the number of iter<strong>at</strong>ions<br />

of the self-consistent loop <strong>in</strong> Fig. 4.1, it is important to f<strong>in</strong>d an efficient method<br />

to construct the next <strong>in</strong>put potential from the output potential. The simplest<br />

mix<strong>in</strong>g scheme is V <strong>in</strong>itial,i+1<br />

eff<br />

57<br />

(r) = V f<strong>in</strong>al,i<br />

eff (r). This simple scheme, however, does<br />

not work s<strong>in</strong>ce the response <strong>in</strong> the output potential is usually enormous compared<br />

to the change <strong>in</strong> the <strong>in</strong>put potential, even for the high densities of an electron


gas. 1 The ma<strong>in</strong> purpose of a mix<strong>in</strong>g scheme is to make the convergence faster,<br />

th<strong>at</strong> is, to reduce the program-runn<strong>in</strong>g time to reach self-consistency while the<br />

solutions ma<strong>in</strong>ta<strong>in</strong> the stability such th<strong>at</strong> they don’t diverge. Another straight<br />

mix<strong>in</strong>g scheme is<br />

V <strong>in</strong>,i+1<br />

eff<br />

(r) = V <strong>in</strong>,i<br />

eff<br />

58<br />

out,i <strong>in</strong>,i<br />

(r) + a (V (r) − V (r)) (4.2)<br />

where i is the number of the iter<strong>at</strong>ion. The mix<strong>in</strong>g r<strong>at</strong>io a ensures the stability<br />

of the convergence dur<strong>in</strong>g the self-consistent calcul<strong>at</strong>ions. One can keep reduc<strong>in</strong>g<br />

a until stability is acquired. One way to speed up the convergence is Broyden’s<br />

method which is<br />

where F i (r) = V out,i<br />

eff<br />

eff<br />

eff<br />

V <strong>in</strong>,i+1<br />

eff (r) = V <strong>in</strong>,i<br />

eff (r) + a(r) F i (r)<br />

<strong>in</strong>,i <strong>in</strong>,i−1<br />

Veff (r) − Veff (r)<br />

a(r) =<br />

F i (r) − F i−1 (r)<br />

<strong>in</strong>,i<br />

(r) − V (r). One can obta<strong>in</strong> Broyden’s method by try-<br />

eff<br />

<strong>in</strong>g to f<strong>in</strong>d the po<strong>in</strong>t on the l<strong>in</strong>e of V <strong>in</strong><br />

out<br />

eff (r) = Veff (r) between the two po<strong>in</strong>ts,<br />

(V <strong>in</strong>,i−1<br />

eff<br />

(r), V out,i−1<br />

(r)) and (V <strong>in</strong>,i out,i<br />

(r), V (r)). S<strong>in</strong>ce the mix<strong>in</strong>g r<strong>at</strong>io <strong>in</strong> Broy-<br />

eff<br />

eff<br />

eff<br />

den’s method depends on the <strong>in</strong>put and output potentials of last two iter<strong>at</strong>ions,<br />

discont<strong>in</strong>uities <strong>in</strong> the potential of r-space may show up dur<strong>in</strong>g the calcul<strong>at</strong>ions and<br />

the <strong>in</strong>tegr<strong>at</strong>ion for wave functions based on the potential may go wrong as well.<br />

Another problem <strong>in</strong> this case is the long range tail of Friedel oscill<strong>at</strong>ion, which is<br />

usually strongly coupled to the electronic density <strong>in</strong> the vic<strong>in</strong>ity of an impurity<br />

due to the Coulomb potential. The idea to avoid this problem and speed up the<br />

1 Results for dielectric constants suggest the response of electron gases of high densities<br />

is small compared to those of low densities. See chapter 5.<br />

,


convergence <strong>at</strong> the same time is to use the straight mix<strong>in</strong>g scheme and the mix<strong>in</strong>g<br />

r<strong>at</strong>io of an exponential-like function.<br />

The mix<strong>in</strong>g scheme is now<br />

a(r) = a exp(−α r 2 ) . (4.3)<br />

V <strong>in</strong>,i+1<br />

eff (r) = V <strong>in</strong>,i<br />

eff (r) + a exp(−α r2 ) (V out,i<br />

eff<br />

59<br />

<strong>in</strong>,i<br />

(r) − V (r)) . (4.4)<br />

There are two parameters a and α <strong>in</strong> this mix<strong>in</strong>g scheme. a can be determ<strong>in</strong>ed<br />

by consider<strong>in</strong>g the response of a region <strong>in</strong> the vic<strong>in</strong>ity of an impurity and α is<br />

a parameter to suppress the strong response of the tail of Friedel oscill<strong>at</strong>ions.<br />

Now one can allow as large a change as possible <strong>in</strong> the vic<strong>in</strong>ity of an impurity<br />

by <strong>in</strong>creas<strong>in</strong>g the a value while one suppresses the response of the tail by us<strong>in</strong>g<br />

large value of α. In most cases, 0.02 for a and 0.002 for α have been found<br />

appropri<strong>at</strong>e. Another idea to make convergence even faster is to utilize the last<br />

two <strong>in</strong>put potentials and add the response of those to the mix<strong>in</strong>g scheme.<br />

V <strong>in</strong>,i+1<br />

eff<br />

(r) = V <strong>in</strong>,i<br />

eff<br />

= V <strong>in</strong>,i<br />

eff<br />

One can make two vectors V <strong>in</strong>,i<br />

eff<br />

out,i <strong>in</strong>,i <strong>in</strong>,i <strong>in</strong>,i−1<br />

(r) + a(r) (V (r) − V (r) + V (r) − V (r))<br />

eff<br />

eff<br />

out,i <strong>in</strong>,i−1<br />

(r) + a(r) (Veff (r) − Veff (r)) . (4.5)<br />

out,i<br />

<strong>in</strong>,i−1<br />

(r) → V (r) and V (r) → V <strong>in</strong>,i<br />

(r) for each r.<br />

eff<br />

If those vectors have the same directions, the mix<strong>in</strong>g scheme (4.5) will acceler<strong>at</strong>e<br />

the convergence (large mix<strong>in</strong>g r<strong>at</strong>io) compared to the previous mix<strong>in</strong>g scheme (4.4)<br />

and if they have opposite directions, (4.5) will deceler<strong>at</strong>e the convergence (small<br />

mix<strong>in</strong>g r<strong>at</strong>io) to avoid the possible oscill<strong>at</strong>ions of potentials between two iter<strong>at</strong>ions. 2<br />

2 The divergence can be noticed usually by fluctu<strong>at</strong>ions of potentials between iter<strong>at</strong>ions<br />

dur<strong>in</strong>g the numerical calcul<strong>at</strong>ions.<br />

eff<br />

eff<br />

eff<br />

eff<br />

eff


Once s<strong>at</strong>isfactory self-consistency is obta<strong>in</strong>ed <strong>in</strong> the vic<strong>in</strong>ity of an impurity,<br />

one can reduce a to avoid the possible large vari<strong>at</strong>ion of the density around the<br />

impurity and reduce α to allow now larger vari<strong>at</strong>ion of the tail of potentials to<br />

make convergence faster. One can implement this <strong>in</strong> the program by calcul<strong>at</strong><strong>in</strong>g<br />

the potential difference ∆V of (4.1) and have the program autom<strong>at</strong>ically vary<br />

the parameters a and α. If ∆V is small enough, especially <strong>in</strong> the vic<strong>in</strong>ity of an<br />

impurity, and the convergence is too slow, a and α can be reduced by a small<br />

amount dur<strong>in</strong>g calcul<strong>at</strong>ions. One idea is to calcul<strong>at</strong>e the rel<strong>at</strong>ive r<strong>at</strong>io of ∆V<br />

between two iter<strong>at</strong>ions.<br />

R<strong>at</strong>io =<br />

(∆Vi) 2 − (∆Vi−1) 2<br />

((∆Vi) 2 + (∆Vi−1) 2 )/2.0<br />

60<br />

. (4.6)<br />

If the r<strong>at</strong>io is too small, one can try to reduce a and α values. However one needs<br />

to f<strong>in</strong>d the appropri<strong>at</strong>e iter<strong>at</strong>ions <strong>at</strong> which the parameters can be reduced and the<br />

values of parameters by trial and error. It has been found <strong>in</strong> this work th<strong>at</strong> if the<br />

∆V 2 is smaller than 10 −8 ∼ 10 −9 and the r<strong>at</strong>io is smaller than 0.002 ∼ 0.005, a<br />

and α can be reduced by 10% ∼ 20%. A m<strong>in</strong>imum number of iter<strong>at</strong>ions between<br />

the iter<strong>at</strong>ions <strong>at</strong> which the parameters are reduced, however, often needs to be set<br />

to ensure the convergence of the calcul<strong>at</strong>ions.<br />

4.3. Search for eigenst<strong>at</strong>es and eigenvalues of the <strong>in</strong>f<strong>in</strong>ite system<br />

4.3.1. Bound st<strong>at</strong>es<br />

The method for the search for bound st<strong>at</strong>es is well established and calcu-<br />

l<strong>at</strong>ions are quite simple for spherically symmetric potentials. [30] S<strong>in</strong>ce only the<br />

pr<strong>in</strong>ciple quantum number n and the angular momentum number l for spherically<br />

symmetric potentials are necessary to dist<strong>in</strong>guish eigenst<strong>at</strong>es, each eigenst<strong>at</strong>e can


e c<strong>at</strong>egorized by n and l or the number of nodes nnodes and l. 3 One can then<br />

search for bound st<strong>at</strong>es from the m<strong>in</strong>imum energy to the maximum energy us<strong>in</strong>g<br />

nnodes and l. Note th<strong>at</strong> the possible lowest energy for the bound st<strong>at</strong>es can be set<br />

to − Z2<br />

n 2 Ryd., which is bound st<strong>at</strong>e energy for a pure Coulomb potential, s<strong>in</strong>ce once<br />

the pure Coulomb potential is screened, the energy eigenvalues are always higher<br />

than those for the pure Coulomb potential. An upper limit for bound st<strong>at</strong>es is<br />

zero with a reference of zero energy limr→∞ Veff(r) = 0.<br />

The method used <strong>in</strong> this work is known as the bisectional algorithm. [30]<br />

There is a better and faster method than the bisectional algorithm, which has<br />

never been tried <strong>in</strong> this work s<strong>in</strong>ce the time required to search for bound st<strong>at</strong>es is<br />

rel<strong>at</strong>ively quite small compared to th<strong>at</strong> for sc<strong>at</strong>tered st<strong>at</strong>es and is usually less than<br />

a few seconds for today’s computers. For the bisectional algorithm, one takes half<br />

the m<strong>in</strong>imum energy as the <strong>in</strong>itial energy for the energy eigenvalue one tries to<br />

achieve and <strong>in</strong>tegr<strong>at</strong>es the KS equ<strong>at</strong>ion with this energy. If the current energy is<br />

higher or smaller than the eigenenergy, one can use the current energy as an upper<br />

limit or a lower limit for the next iter<strong>at</strong>ion and <strong>in</strong>tegr<strong>at</strong>e the KS equ<strong>at</strong>ion with the<br />

new middle energy between the upper and lower limits. One can try to repe<strong>at</strong> the<br />

process until the bound st<strong>at</strong>e energy is obta<strong>in</strong>ed with<strong>in</strong> the required tolerance for<br />

the numerical error.<br />

In order to decide whether the current energy is higher or smaller than the<br />

eigenenergy, one uses two conditions, which both must be s<strong>at</strong>isfied for correct wave<br />

functions. The first one is th<strong>at</strong> the number of nodes <strong>in</strong> the wave functions must<br />

be the same as required <strong>in</strong> nnodes = n − l − 1. The second one is th<strong>at</strong> the first<br />

3 Sp<strong>in</strong>-orbit coupl<strong>in</strong>g is neglected <strong>in</strong> this work, and hence n = nnodes + l + 1.<br />

61


deriv<strong>at</strong>ive of one wave function <strong>in</strong>tegr<strong>at</strong>ed from the orig<strong>in</strong> outwards must m<strong>at</strong>ch <strong>at</strong><br />

a m<strong>at</strong>ch<strong>in</strong>g po<strong>in</strong>t with the first deriv<strong>at</strong>ive of the other wave function <strong>in</strong>tegr<strong>at</strong>ed from<br />

the maximum radius <strong>in</strong>wards. The numerical best po<strong>in</strong>t for a m<strong>at</strong>ch<strong>in</strong>g po<strong>in</strong>t is an<br />

<strong>in</strong>nermost classical turn<strong>in</strong>g po<strong>in</strong>t, s<strong>in</strong>ce one can expect th<strong>at</strong> the oscill<strong>at</strong>ory behavior<br />

of a solution becomes a decreas<strong>in</strong>g exponential-like behavior <strong>at</strong> this turn<strong>in</strong>g po<strong>in</strong>t.<br />

Note th<strong>at</strong> Friedel oscill<strong>at</strong>ions <strong>in</strong> the tail of potentials may give rise to a number of<br />

turn<strong>in</strong>g po<strong>in</strong>ts. First, one can narrow the range of the energy us<strong>in</strong>g the number<br />

of nodes. If the number of nodes <strong>in</strong> a wave function is larger or smaller than th<strong>at</strong><br />

required, the current energy is higher or lower than the eigenenergy. If the number<br />

of nodes is correct, one can check if first deriv<strong>at</strong>ives of wave functions m<strong>at</strong>ch <strong>at</strong><br />

the m<strong>at</strong>ch<strong>in</strong>g po<strong>in</strong>t. 4 If the magnitude of a first deriv<strong>at</strong>ive of an outward solution<br />

is larger or smaller than th<strong>at</strong> of an <strong>in</strong>ward solution, the current energy is higher<br />

or lower than the eigenenergy. If the energy is needed to be higher than the <strong>in</strong>itial<br />

maximum energy (zero as <strong>in</strong> (3.4)), one can stop the search s<strong>in</strong>ce there is no bound<br />

st<strong>at</strong>e.<br />

The KS equ<strong>at</strong>ions can be <strong>in</strong>tegr<strong>at</strong>ed only with appropri<strong>at</strong>e boundary condi-<br />

tions. For bound st<strong>at</strong>es,<br />

as r → 0<br />

un,l(r) → r l+1<br />

as r → ∞<br />

and u ′ n,l(r) → (l + 1)r l<br />

62<br />

, (4.7)<br />

un,l(r) → exp(− √ −Er) and u ′ n,l(r) → − √ −E exp(− √ −Er) . (4.8)<br />

4 S<strong>in</strong>ce the numerical solutions are not normalized, one has to ensure they both have<br />

the same numerical scale to compare first deriv<strong>at</strong>ives, for <strong>in</strong>stance, by us<strong>in</strong>g logarithmic<br />

deriv<strong>at</strong>ives.


The normaliz<strong>at</strong>ion condition for bound st<strong>at</strong>es is<br />

<br />

63<br />

dr |u(r)| 2 = 1 . (4.9)<br />

The density difference for spherically symmetric potentials due to the bound st<strong>at</strong>es<br />

of an impurity is given by<br />

n(r) − n0 = (2l + 1) u<br />

2<br />

4π<br />

2 n,l (r)<br />

r2 , (4.10)<br />

where the factor 2 is for sp<strong>in</strong> degeneracy 5<br />

as follows.<br />

4.3.2. Sc<strong>at</strong>tered st<strong>at</strong>es<br />

n,l<br />

The Fermi wave vector kF of a homogeneous electron gas can be calcul<strong>at</strong>ed<br />

2 Ω<br />

(2π) 3<br />

4π<br />

3 k3 F = N<br />

where the left hand side is the number of st<strong>at</strong>es available <strong>in</strong> a system of the volume<br />

Ω, the factor 2 <strong>in</strong> the left hand side is for sp<strong>in</strong> degeneracy, and N is the total number<br />

of electrons <strong>in</strong> a system. In the thermodynamic limit, the above equ<strong>at</strong>ion leads to<br />

kF = (3π 2 n0) 1/3 , (4.11)<br />

where n0 is the electronic density of the system. S<strong>in</strong>ce the system used <strong>in</strong> this<br />

work is extended to <strong>in</strong>f<strong>in</strong>ity, and thus there are no boundary conditions, there are<br />

5 The factor (2l + 1)/4π is obta<strong>in</strong>ed by us<strong>in</strong>g the addition theorem:<br />

l<br />

m=−l<br />

|Yl,m(θ, φ)| 2 =<br />

2l + 1<br />

4π<br />

.


no quantized k values and all k values up to kF are available. In addition, the<br />

chemical potential is assumed to be a constant and the same as the Fermi energy<br />

of a homogeneous electron gas <strong>at</strong> T = 0 s<strong>in</strong>ce the dimension of the system is much<br />

larger than th<strong>at</strong> of an impurity. Now one needs to <strong>in</strong>tegr<strong>at</strong>e the KS equ<strong>at</strong>ions<br />

<strong>in</strong> the k-mesh for the cont<strong>in</strong>uum st<strong>at</strong>es k = 0 · · · kF . The wave functions can be<br />

normalized by (3.24) and the density difference for sc<strong>at</strong>tered st<strong>at</strong>es is given by<br />

(3.47).<br />

The ma<strong>in</strong> numerical problem is th<strong>at</strong> the maximum number of angular mo-<br />

mentum values should be determ<strong>in</strong>ed for each k. Theoretically kR is a good value<br />

for the maximum l where R is the radius of the potential. [23] On the other hand,<br />

only 10 l values were used <strong>in</strong> the work of Puska, Niem<strong>in</strong>en, and Man<strong>in</strong>en.(1981)<br />

None of these are, however, good criteria for the maximum l value. The kR tends<br />

to neglect the low-energy sc<strong>at</strong>ter<strong>in</strong>g due to very small k values. Ten for the max-<br />

imum l value is more than enough for small k values and is too small for large k<br />

values. It is found th<strong>at</strong> us<strong>in</strong>g too small a number for the maximum l value gives<br />

an error <strong>in</strong> the energy and an <strong>in</strong>stability <strong>in</strong> terms of convergence. A better way<br />

is us<strong>in</strong>g the values of the phase shift to see how close the sc<strong>at</strong>tered st<strong>at</strong>e <strong>at</strong> the<br />

current l value is to the free particle st<strong>at</strong>e. In other words, if the phase shift is<br />

sufficiently small and thus l values larger than the current l value are not impor-<br />

tant, the calcul<strong>at</strong>ion for the current k value can be term<strong>in</strong><strong>at</strong>ed and one beg<strong>in</strong>s the<br />

calcul<strong>at</strong>ion for the next k value <strong>in</strong> the k-mesh. A good value for the smallest phase<br />

shift is 10 −8 , which turns out to give a good convergence.<br />

64


FIGURE 4.2. Phase shift values of l = 0 for a zero potential.<br />

4.4. Numerical problems<br />

4.4.1. Convergence of phase shifts<br />

The accuracy of the phase shifts needs to be tested to see if 10 −8 suggested<br />

for the smallest phase shift <strong>in</strong> the previous section is well with<strong>in</strong> the precision of<br />

the numerical error. One idea taken from [21] is to test the phase shift values for<br />

zero potential s<strong>in</strong>ce the phase shift is zero for the zero potential. One can see the<br />

peak <strong>in</strong> Fig. 4.2 around k = 0.6 and th<strong>at</strong> the phase shift devi<strong>at</strong>es more strongly<br />

from zero after this value. The test has been performed for the electronic density<br />

65


FIGURE 4.3. Phase shift values for an <strong>at</strong>tractive potential of Yukawa type and a<br />

repulsive potential.<br />

of n0 = 0.6. 6 One can conclude th<strong>at</strong> the numerical error <strong>in</strong> the phase shifts is<br />

less than 3 · 10 −11 for the range of actual metallic densities. It is found th<strong>at</strong> the<br />

numerical error of higher l values is smaller than th<strong>at</strong> of l = 0. This numerical<br />

behavior may be, however, dependent on the numerical libraries one uses. Fig.<br />

4.3 shows the rel<strong>at</strong>ionship between the maximum radius Rm used <strong>in</strong> the numerical<br />

calcul<strong>at</strong>ion and the phase shift values. As seen <strong>in</strong> Fig. 4.3, if Rm is too small and<br />

thus the potential does not go to zero fast enough, the phase shift calcul<strong>at</strong>ions<br />

fail. If the potential has such a long range, phase shift values <strong>at</strong> very low l show a<br />

6 Parameters for the test are R0 = 8.1250520·10 −6 , Rlog = 8.0, Rm = 31.02, Nlog = 465,<br />

and Nl<strong>in</strong> = 230, where Nlog is the number of po<strong>in</strong>ts used <strong>in</strong> the logarithmic r-space<br />

between R0 and Rlog and Nl<strong>in</strong> is the number of po<strong>in</strong>ts used <strong>in</strong> the l<strong>in</strong>ear r-space between<br />

Rlog and Rm.<br />

66


zig-zag behavior, which is similar to th<strong>at</strong> of hard sphere sc<strong>at</strong>ter<strong>in</strong>g for high energy<br />

approxim<strong>at</strong>ions. [23] They are, however, eventually approach<strong>in</strong>g zero as l <strong>in</strong>creases.<br />

Then one can stop the calcul<strong>at</strong>ion for each k if phase shift values are small enough,<br />

for <strong>in</strong>stance, less than 10 −8 as argued <strong>in</strong> the previous section.<br />

4.4.2. Correction for potentials<br />

The ma<strong>in</strong> problem <strong>in</strong> solv<strong>in</strong>g the KS equ<strong>at</strong>ions numerically is th<strong>at</strong> the effective<br />

potentials <strong>in</strong> the KS equ<strong>at</strong>ions have physically and numerically essential defects,<br />

especially <strong>at</strong> the beg<strong>in</strong>n<strong>in</strong>g of calcul<strong>at</strong>ions. These defects may not only slow down<br />

the convergence but also lead the calcul<strong>at</strong>ions to <strong>in</strong>correct self-consistent solutions.<br />

The first problem is th<strong>at</strong>, dur<strong>in</strong>g the numerical calcul<strong>at</strong>ions, s<strong>in</strong>ce the effective<br />

potential has not reached self-consistency yet, especially <strong>at</strong> the beg<strong>in</strong>n<strong>in</strong>g, there<br />

may be a surplus or deficiency of charge to screen the impurity and the system does<br />

not s<strong>at</strong>isfy the Friedel sum rule. The effective potentials <strong>in</strong> this case are oscill<strong>at</strong><strong>in</strong>g<br />

around some constant value due to the extra charge. This extra charge appears as<br />

a constant (<strong>in</strong>tersect<strong>in</strong>g po<strong>in</strong>t with r = 0) <strong>in</strong> the rV plot of Fig. 4.4. The second<br />

problem is th<strong>at</strong>, s<strong>in</strong>ce the system is extended to <strong>in</strong>f<strong>in</strong>ity, the upper limit of the<br />

<strong>in</strong>tegral <strong>in</strong> the Coulomb potential must be <strong>in</strong>f<strong>in</strong>ity as well. No m<strong>at</strong>ter how large<br />

a radius is used for the numerical calcul<strong>at</strong>ions, the <strong>in</strong>tegral from the maximum<br />

radius to <strong>in</strong>f<strong>in</strong>ity <strong>in</strong> the Coulomb potential is always miss<strong>in</strong>g, which appears as a<br />

slope <strong>in</strong> the rV plot <strong>in</strong> Fig. 4.4. If one uses naively this <strong>in</strong>correct output potential<br />

to mix with the <strong>in</strong>put potential for the next iter<strong>at</strong>ion, the phase shift calcul<strong>at</strong>ions<br />

cannot be performed s<strong>in</strong>ce the phase shift calcul<strong>at</strong>ions and thus the normaliz<strong>at</strong>ion<br />

for sc<strong>at</strong>tered wave functions are based on the fact th<strong>at</strong> the potentials are zero or<br />

67


FIGURE 4.4. Typical example show<strong>in</strong>g defects <strong>in</strong> the output potential dur<strong>in</strong>g the<br />

numerical calcul<strong>at</strong>ions.<br />

almost zero <strong>at</strong> the end of the maximum radius. 7 In addition, special phase shifts<br />

have to be used for Coulomb potentials. 8<br />

7 At the maximum radius, it is assumed th<strong>at</strong> an impurity is completely screened and<br />

the system s<strong>at</strong>isfies the Friedel sum rule.<br />

8 This may be much more difficult than zero Coulomb potentials. Besides, allow<strong>in</strong>g non-<br />

zero Coulomb potential <strong>at</strong> the maximum radius may be a wrong guide for a self-consistent<br />

solution of the system used <strong>in</strong> this work.<br />

68


One can simply make the potential zero <strong>at</strong> the maximum radius and run<br />

the program for many iter<strong>at</strong>ions until the extra charge (Constant <strong>in</strong> Fig. 4.4)<br />

disappears. This is, however, not only time consum<strong>in</strong>g, but also the potential is<br />

not likely to go to the zero <strong>at</strong> the maximum radius due to the Friedel oscill<strong>at</strong>ions.<br />

One can also try to remove the extra slope <strong>in</strong> the rV plot us<strong>in</strong>g an analytic or<br />

a numerical model for the miss<strong>in</strong>g <strong>in</strong>tegral, but this turns out to be not good<br />

enough. 9 The best way is simply to determ<strong>in</strong>e the extra constant and slope <strong>in</strong> the<br />

rV plot and subtract them from rV . One can not, however, simply subtract the<br />

constant from rV because the potential <strong>in</strong> the vic<strong>in</strong>ity of an impurity is affected by<br />

the <strong>in</strong>teraction between the impurity and an electron gas as well as by the Friedel<br />

oscill<strong>at</strong>ions and thus one has to allow for a vari<strong>at</strong>ion of the electronic charge around<br />

an impurity. S<strong>in</strong>ce <strong>at</strong> the same time the potential must be oscill<strong>at</strong><strong>in</strong>g around zero<br />

<strong>at</strong> the maximum radius, the formula to correct the potential can be written as<br />

r × V = r × V − r × (Slope + Constant/Rm) , (4.12)<br />

where Rm is the maximum radius r <strong>in</strong> the r-mesh for the numerical calcul<strong>at</strong>ions.<br />

One can try other functions <strong>in</strong>stead of a l<strong>in</strong>ear r <strong>in</strong> r × Constant/Rm. However<br />

other functions may suppress the vari<strong>at</strong>ion around an impurity too much or too<br />

little, either of which makes the convergence slower. The slope and constant can<br />

be easily calcul<strong>at</strong>ed by four po<strong>in</strong>ts of maximum and m<strong>in</strong>imum peaks near the end<br />

of the potential. Then one can set up two l<strong>in</strong>e equ<strong>at</strong>ions, from which one can f<strong>in</strong>d<br />

the average slope and constant <strong>at</strong> r = 0. Fig. 4.5 shows the typical examples of<br />

the effective potentials before and after the correction and the right figure <strong>in</strong> Fig.<br />

4.5 shows th<strong>at</strong> the potential is oscill<strong>at</strong><strong>in</strong>g around zero <strong>at</strong> large r values after the<br />

9 It has been found th<strong>at</strong> there is still a noticeable slope after correct<strong>in</strong>g the potential.<br />

69


FIGURE 4.5. Typical examples to show how the effective potentials behave before<br />

and after the correction. These potentials are calcul<strong>at</strong>ed for a hydrogen impurity<br />

and n0 = 0.0025.<br />

correction, which means an impurity is completely screened and Friedel sum rule<br />

is s<strong>at</strong>isfied as it should be.<br />

4.4.3. Calcul<strong>at</strong>ions for <strong>in</strong>duced electrons and energies<br />

One detail previous papers [5] have overlooked is th<strong>at</strong> after <strong>in</strong>clud<strong>in</strong>g all of<br />

the screen<strong>in</strong>g charge around impurity the electronic density is simply oscill<strong>at</strong><strong>in</strong>g<br />

due to the Friedel oscill<strong>at</strong>ions and thus the <strong>in</strong>tegr<strong>at</strong>ed values of the electronic<br />

density, needed to f<strong>in</strong>d the number of electrons <strong>in</strong>duced due to an impurity, are<br />

also oscill<strong>at</strong><strong>in</strong>g, which is shown <strong>in</strong> Fig. 4.6. In addition, the number of <strong>in</strong>duced<br />

electrons <strong>in</strong> the system should not depend on the maximum value of r s<strong>in</strong>ce the<br />

system goes to <strong>in</strong>f<strong>in</strong>ity and thus this number can not be found by simply <strong>in</strong>tegr<strong>at</strong><strong>in</strong>g<br />

from zero to the maximum of r. In order to f<strong>in</strong>d the charge <strong>in</strong>duced <strong>in</strong> the system,<br />

<strong>in</strong>tegr<strong>at</strong>ed values of the charge density as a function of r should be exam<strong>in</strong>ed as <strong>in</strong><br />

70


FIGURE 4.6. Calcul<strong>at</strong>ed for a hydrogen impurity and n0 = 0.0025. This example<br />

shows the oscill<strong>at</strong>ions of <strong>in</strong>tegral values of the <strong>in</strong>duced electronic density as a<br />

function of r.<br />

Fig. 4.6 and the average charge from the oscill<strong>at</strong>ion should be calcul<strong>at</strong>ed to f<strong>in</strong>d the<br />

total number of electrons <strong>in</strong>duced due to an impurity. It is important to note th<strong>at</strong><br />

the r dependence of energies such as the exchange-correl<strong>at</strong>ion and the Coulomb<br />

energy can be removed <strong>in</strong> the same way.<br />

4.4.4. Mesh for r-space and k-space<br />

In order to <strong>in</strong>tegr<strong>at</strong>e functions numerically, one needs to use small <strong>in</strong>terval to<br />

approxim<strong>at</strong>e the cont<strong>in</strong>uous space over which the functions are to be <strong>in</strong>tegr<strong>at</strong>ed.<br />

71


FIGURE 4.7. Typical example for the effective potential. The logarithmic mesh<br />

is used for the rapid vari<strong>at</strong>ion of the potential <strong>in</strong> the vic<strong>in</strong>ity of the orig<strong>in</strong>.<br />

The term, mesh, is def<strong>in</strong>ed <strong>in</strong> this work as how <strong>in</strong>tervals vary between two adjacent<br />

po<strong>in</strong>ts <strong>in</strong> a cont<strong>in</strong>uous space. If the <strong>in</strong>terval does not vary and is simply a constant,<br />

the <strong>in</strong>tegral space is called a l<strong>in</strong>ear mesh. S<strong>in</strong>ce the numerical precision directly<br />

depends on the mesh used for the <strong>in</strong>tegral <strong>in</strong> r or k-space, the behavior of functions<br />

which are to be <strong>in</strong>tegr<strong>at</strong>ed needs to be exam<strong>in</strong>ed to choose the appropri<strong>at</strong>e mesh for<br />

each space. The r-space can be c<strong>at</strong>egorized <strong>in</strong>to two regions, based on the density<br />

vari<strong>at</strong>ion. The first region covers the core st<strong>at</strong>es <strong>in</strong> which the electronic density<br />

usually varies rapidly. In the second region, the net electronic charge <strong>in</strong>duced<br />

by a electron gas is almost zero and the Friedel oscill<strong>at</strong>ions dom<strong>in</strong><strong>at</strong>e the density<br />

vari<strong>at</strong>ion. Therefore a logarithmic r-mesh is appropri<strong>at</strong>e for the first region which<br />

<strong>in</strong>cludes the core st<strong>at</strong>es <strong>in</strong> the vic<strong>in</strong>ity of an impurity and after th<strong>at</strong>, a simple l<strong>in</strong>ear<br />

72


FIGURE 4.8. Example for phase shifts. The phase shift varies rapidly <strong>at</strong> the<br />

bottom of the conduction band if the energy of a bound st<strong>at</strong>e is close to the<br />

bottom of the conduction band.<br />

r-mesh can be used for the second region as shown <strong>in</strong> Fig. 4.7. In this work, the<br />

logarithmic r-mesh always corresponds to the first region and the l<strong>in</strong>ear r-mesh<br />

corresponds to the second region.<br />

Phase shifts <strong>at</strong> l = 0 need to be exam<strong>in</strong>ed <strong>in</strong> order to see how the wave<br />

functions behave <strong>in</strong> k-space. Based on this <strong>in</strong>form<strong>at</strong>ion, one can choose the mesh <strong>in</strong><br />

k-space. The appropri<strong>at</strong>e mesh <strong>in</strong> k-space is, however, quite difficult to determ<strong>in</strong>e<br />

before calcul<strong>at</strong>ions s<strong>in</strong>ce <strong>at</strong> least the energies of bound st<strong>at</strong>es, or the sc<strong>at</strong>ter<strong>in</strong>g<br />

lengths are necessary to estim<strong>at</strong>e how large the slopes of phase shifts <strong>at</strong> l = 0 are<br />

<strong>in</strong> the vic<strong>in</strong>ity of the bottom of the conduction band. Note th<strong>at</strong> the sc<strong>at</strong>ter<strong>in</strong>g<br />

length is <strong>in</strong>versely proportional to the energy of bound st<strong>at</strong>es . (See Sec. 3.6.) One<br />

73


FIGURE 4.9. Integrand <strong>in</strong> (3.22). Calcul<strong>at</strong>ed for a Hydrogen impurity <strong>at</strong> 0.04<br />

background density.<br />

may try very low or high background densities of an electron gas first and try to<br />

determ<strong>in</strong>e the range of the background densities where the energies of bound st<strong>at</strong>es<br />

are close to the bottom of the conduction band. (See Fig. 4.8.) If the calcul<strong>at</strong>ions<br />

need to be performed for this range of the background density, the phase shifts <strong>in</strong><br />

k-space also need to be exam<strong>in</strong>ed <strong>in</strong> order to determ<strong>in</strong>e the range of k-space where<br />

the phase shifts vary fast. In th<strong>at</strong> case, a different k-mesh is necessary due to the<br />

different vari<strong>at</strong>ion of the phase shifts. One can try the logarithmic mesh for the<br />

fast vari<strong>at</strong>ion of phase shifts <strong>at</strong> the bottom of the conduction band as <strong>in</strong> r-space.<br />

However, it turns out th<strong>at</strong> a logarithmic mesh causes more numerical errors than<br />

the simple l<strong>in</strong>ear mesh. Fig. 4.9 shows how the wave functions <strong>in</strong> k-space vary <strong>at</strong><br />

different r values. A logarithmic mesh may work well <strong>at</strong> low r values but causes<br />

74


more numerical errors for the ord<strong>in</strong>ary oscill<strong>at</strong><strong>in</strong>g wave functions <strong>in</strong> k-space. A<br />

better way is simply to use two different l<strong>in</strong>ear k-meshes with different <strong>in</strong>tervals<br />

to deal with the rapid vari<strong>at</strong>ion <strong>at</strong> the bottom of the conduction band. The very<br />

small <strong>in</strong>terval should be used for the range of k <strong>in</strong> which the phase shifts vary fast.<br />

Because of the smooth behavior of the phase shifts, a large <strong>in</strong>terval can be used<br />

for large k values <strong>in</strong> order to reduce the time to <strong>in</strong>tegr<strong>at</strong>e <strong>in</strong> k-space.<br />

75


FIGURE 4.10. The density of a free carbon <strong>at</strong>om and the density <strong>in</strong>duced by a<br />

carbon impurity <strong>in</strong> a homogeneous electron gas. The 1s and 2s core st<strong>at</strong>es rema<strong>in</strong><br />

almost the same. These calcul<strong>at</strong>ions are performed for a sp<strong>in</strong> polarized system.<br />

4.4.5. Cancel<strong>at</strong>ion of numerical errors <strong>in</strong> immersion energies: Number of r<br />

po<strong>in</strong>ts<br />

In this section, the calcul<strong>at</strong>ions are based on the logarithmic r-mesh which is<br />

used for the range from 0 to 8 a.u. and on the l<strong>in</strong>ear r-mesh which is used after 8<br />

a.u. Due to the approxim<strong>at</strong>ions <strong>in</strong> LDA, the rapid vari<strong>at</strong>ion of the density of the<br />

core st<strong>at</strong>es <strong>in</strong> the vic<strong>in</strong>ity of an impurity may cause significant numerical errors.<br />

However, s<strong>in</strong>ce the core st<strong>at</strong>es hardly <strong>in</strong>teract with the environment, numerical<br />

errors of the core st<strong>at</strong>es of an impurity system can be reduced by those of a free<br />

<strong>at</strong>om <strong>in</strong> the immersion energy calcul<strong>at</strong>ions. Th<strong>at</strong> is,<br />

76<br />

Oimpuritysystem − Ofree<strong>at</strong>om 0 , (4.13)


where O denotes the numerical errors <strong>in</strong> the core st<strong>at</strong>es. Fig. 4.10 shows th<strong>at</strong> the<br />

core st<strong>at</strong>es of an impurity actually do not <strong>in</strong>teract with a homogeneous electron gas<br />

and hence do not contribute to the immersion energy significantly. Table 4.1 and<br />

Fig. 4.11 illustr<strong>at</strong>e how the r-mesh <strong>in</strong> the vic<strong>in</strong>ity of an impurity affects the energies<br />

and how numerical errors <strong>in</strong> the immersion energies can be reduced with<strong>in</strong> a few<br />

milli-Rydbergs by calcul<strong>at</strong><strong>in</strong>g the free <strong>at</strong>om energy with the same r-mesh as for<br />

the impurity system. These results may depend on the range of the logarithmic r-<br />

mesh which covers the core st<strong>at</strong>es. The hydrogen impurity for a non-sp<strong>in</strong> polarized<br />

system <strong>in</strong> Table 4.2 shows the same numerical behavior. Therefore it is important<br />

to use the same r-mesh <strong>in</strong> calcul<strong>at</strong><strong>in</strong>g the free <strong>at</strong>om energy and the energy of an<br />

impurity system to avoid large numerical errors due to the core st<strong>at</strong>es.<br />

In order to m<strong>in</strong>imize the rema<strong>in</strong>der of the numerical errors due to the <strong>in</strong>terval<br />

<strong>in</strong> the r-mesh and f<strong>in</strong>d the correct immersion energy, one can <strong>at</strong>tempt to use the<br />

l<strong>in</strong>ear rel<strong>at</strong>ionship between the immersion energies and the number of r-po<strong>in</strong>ts as<br />

r-mesh ∆E E<strong>at</strong>om Eimm = ∆E − E<strong>at</strong>om<br />

525 -73.43094 -73.29770 -0.13324<br />

550 -73.51473 -73.38030 -0.13443<br />

600 -73.64172 -73.50606 -0.13566<br />

650 -73.75003 -73.61365 -0.13638<br />

TABLE 4.1. Energies of a carbon impurity system ∆E and energies of a free<br />

carbon <strong>at</strong>om E<strong>at</strong>om (<strong>in</strong> Rydbergs). A carbon impurity <strong>in</strong> a sp<strong>in</strong> polarized system<br />

and a background density of 0.002 are used for calcul<strong>at</strong>ions. The r-mesh <strong>in</strong> the<br />

table represents the number of r- po<strong>in</strong>ts used <strong>in</strong> the range from 0 to 8 a.u., which<br />

corresponds to the logarithmic r-mesh and <strong>in</strong>cludes most core st<strong>at</strong>es. (See Sec.<br />

4.4.4.)<br />

77


-mesh ∆E E<strong>at</strong>om Eimm = ∆E − E<strong>at</strong>om<br />

450 -1.06535 -0.87278 -0.19257<br />

550 -1.07140 -0.87725 -0.19415<br />

650 -1.07576 -0.88037 -0.19539<br />

750 -1.07884 -0.88285 -0.19599<br />

850 -1.08129 -0.88458 -0.19671<br />

1050 -1.08471 -0.88707 -0.19764<br />

TABLE 4.2. Example for a hydrogen impurity. (Non-sp<strong>in</strong> polarized system.) A<br />

background density of 0.0025 is used for these calcul<strong>at</strong>ions.<br />

FIGURE 4.11. Comparison between immersion energies calcul<strong>at</strong>ed with a constant<br />

free <strong>at</strong>om energy and the free <strong>at</strong>om energy calcul<strong>at</strong>ed with the same r-mesh as the<br />

impurity system. The immersion energies calcul<strong>at</strong>ed with the same free <strong>at</strong>om<br />

energy show rel<strong>at</strong>ively large vari<strong>at</strong>ion with the different number of r-po<strong>in</strong>ts due to<br />

the numerical errors of core st<strong>at</strong>es.<br />

78


FIGURE 4.12. Fitt<strong>in</strong>g by <strong>in</strong>verse power law to f<strong>in</strong>d the correct immersion energy.<br />

D<strong>at</strong>a from Table 4.1 and Table 4.2 .<br />

<strong>in</strong> Fig. 4.12. Now s<strong>in</strong>ce the number of r-po<strong>in</strong>ts is <strong>in</strong> the denom<strong>in</strong><strong>at</strong>or, the number of<br />

r-po<strong>in</strong>ts can be easily extrapol<strong>at</strong>ed to <strong>in</strong>f<strong>in</strong>ity to make the <strong>in</strong>terval of r-mesh zero.<br />

Us<strong>in</strong>g this method, the immersion energy can be determ<strong>in</strong>ed up to 1 milli-rydberg.<br />

4.4.6. Further numerical precision tests<br />

The results of numerical calcul<strong>at</strong>ions may depend on other numerical param-<br />

eters such as the maximum radius Rmax and the number of po<strong>in</strong>ts <strong>in</strong> k and r-space.<br />

The effect of the logarithmic r-mesh <strong>in</strong> the vic<strong>in</strong>ity of an impurity is already tested<br />

<strong>in</strong> Sec. 4.4.5 and it is shown th<strong>at</strong> the immersion energy can be obta<strong>in</strong>ed with<strong>in</strong> 1<br />

milli-rydberg. Other important parameters are the maximum radius Rmax and the<br />

number of po<strong>in</strong>ts <strong>in</strong> k and r-space. The tests <strong>in</strong> this section have been performed<br />

for a Hydrogen impurity <strong>in</strong> a non-sp<strong>in</strong> polarized system of a homogeneous electron<br />

gas. The electronic density is 0.0025. The logarithmic r-mesh is used for the range<br />

from 0 to 8 a.u. and the l<strong>in</strong>ear r-mesh is used after 8 a.u. The maximum radius<br />

79


Rmax Immersion energy(Ry.)<br />

40 Not converged<br />

50 -0.19392<br />

60 -0.19449<br />

70 -0.19395<br />

80 -0.19402<br />

90 -0.19415<br />

120 -0.19411<br />

TABLE 4.3. Immersion energies for a hydrogen impurity. 70 k po<strong>in</strong>ts, 550 r po<strong>in</strong>ts<br />

for the logarithmic r-mesh, and 0.1 for the <strong>in</strong>terval of the l<strong>in</strong>ear r-mesh are used<br />

for these calcul<strong>at</strong>ions.<br />

is a very important parameter for convergence and stability of the calcul<strong>at</strong>ions.<br />

S<strong>in</strong>ce the phase shift calcul<strong>at</strong>ions are based on a zero potential <strong>at</strong> the maximum<br />

radius Rmax, the Coulomb potential due to the extra charge must vanish <strong>at</strong> least<br />

<strong>at</strong> Rmax. Therefore it is important to <strong>in</strong>clude enough Friedel oscill<strong>at</strong>ions and also<br />

not to loose any significant charge of the bound st<strong>at</strong>es. For <strong>in</strong>stance, the conver-<br />

gence required to obta<strong>in</strong> an immersion energy with<strong>in</strong> a error bar of 1 milli-Rydberg<br />

((∆V ) 2 10 −11 , See Sec. 4.1.1.) can not be accomplished for Rmax = 40 <strong>in</strong> Table<br />

4.3. S<strong>in</strong>ce the bound st<strong>at</strong>e has a very short range <strong>at</strong> this background density, the<br />

ma<strong>in</strong> reason is the long tail of Friedel oscill<strong>at</strong>ions which prevent the self-consistent<br />

calcul<strong>at</strong>ions from converg<strong>in</strong>g. (See Fig. 4.13.) One good barometer for the Friedel<br />

oscill<strong>at</strong>ions is to <strong>in</strong>clude <strong>at</strong> least 10 λ’s where λ = π/kF . As the background density<br />

<strong>in</strong>creases, the bound st<strong>at</strong>e plays a more important role and the extended bound<br />

st<strong>at</strong>e should be considered to determ<strong>in</strong>e the maximum radius Rmax.<br />

80


FIGURE 4.13. Extra densities <strong>in</strong>duced by an impurity.<br />

The immersion energies <strong>in</strong> Table 4.3 show small fluctu<strong>at</strong>ion, less than 1 milli-<br />

Rydberg, especially for Rmax larger than 70. Note th<strong>at</strong> the convergence process<br />

is smooth for Rmax larger than 70. The results for the number of r po<strong>in</strong>ts <strong>in</strong><br />

the l<strong>in</strong>ear r-mesh show a very regular p<strong>at</strong>tern <strong>in</strong> Table 4.4, similar to th<strong>at</strong> for the<br />

number of r po<strong>in</strong>ts <strong>in</strong> the logarithmic r-mesh. Fig. 4.14 shows th<strong>at</strong> the immersion<br />

energy decreases as the number of r po<strong>in</strong>ts for the l<strong>in</strong>ear r-mesh <strong>in</strong>creases. The<br />

difference is, however, not larger than 1 milli-Rydberg. One can use simply 0.1 for<br />

the <strong>in</strong>terval of the l<strong>in</strong>ear r-mesh to obta<strong>in</strong> the immersion energy with<strong>in</strong> an error<br />

bar of 1 milli-Rydberg. Another important parameter is the number of k po<strong>in</strong>ts<br />

<strong>in</strong> k-mesh. S<strong>in</strong>ce the k-mesh is a simple l<strong>in</strong>ear mesh, one can expect simply th<strong>at</strong><br />

a larger number of k po<strong>in</strong>ts give rise to the better results. However, for smooth<br />

81


Number of r po<strong>in</strong>ts Immersion energy(Ry.)<br />

320 -0.19353<br />

520 -0.19387<br />

720 -0.19402<br />

920 -0.19411<br />

1120 -0.19417<br />

TABLE 4.4. Immersion energies for a hydrogen impurity. 70 k po<strong>in</strong>ts, 80 for Rmax,<br />

and 550 r po<strong>in</strong>ts for the logarithmic r-mesh are used for calcul<strong>at</strong>ions.<br />

FIGURE 4.14. Immersion energy versus the <strong>in</strong>verse of the number of r po<strong>in</strong>ts <strong>in</strong><br />

the l<strong>in</strong>ear r-mesh. D<strong>at</strong>a from Table 4.4.<br />

82


Number of k po<strong>in</strong>ts Immersion energy(Ry.)<br />

60 -0.19403<br />

70 -0.19402<br />

80 -0.19402<br />

90 -0.19402<br />

100 -0.19402<br />

TABLE 4.5. Immersion energies for a hydrogen impurity. There is no difference<br />

with<strong>in</strong> 10 −5 Rydbergs for k po<strong>in</strong>ts more than 70. 80 for Rmax and 550 r po<strong>in</strong>ts for<br />

the logarithmic r-mesh are used for these calcul<strong>at</strong>ions.<br />

behavior of phase shifts, <strong>in</strong>creas<strong>in</strong>g the number of k po<strong>in</strong>ts does not improve the<br />

results significantly. The results for different number of k po<strong>in</strong>ts are given <strong>in</strong><br />

Table 4.5, which show there is no noticeable improvements with<strong>in</strong> 10 −5 Ry. <strong>in</strong><br />

the immersion energies for more than 70 k po<strong>in</strong>ts. It should be noted th<strong>at</strong> this<br />

result may be quite different for systems which have a large sc<strong>at</strong>ter<strong>in</strong>g length as<br />

mentioned <strong>in</strong> Sec. 4.4.4. Increas<strong>in</strong>g the number of k po<strong>in</strong>ts for those systems can<br />

adjust the significant charge amount correctly and yield better immersion energies.<br />

4.5. Self-consistent solutions<br />

The self-consistent solutions may depend on the methods to solve numerical<br />

problems as well as the boundary conditions. Incorrect numerical parameters or<br />

methods may lead to <strong>in</strong>correct self-consistent solutions. Th<strong>at</strong> is, self-consistency<br />

is not a sufficient condition for a correct solution. The only way to obta<strong>in</strong> the<br />

correct self-consistent solutions is to compare the solutions with the results of<br />

83


other theories or experiments. There are several ways to see if the iter<strong>at</strong>ive process<br />

converges correctly. The extra charge <strong>in</strong>duced by an impurity can be calcul<strong>at</strong>ed by<br />

the Friedel sum rule, <strong>in</strong>tegr<strong>at</strong><strong>in</strong>g the electronic density <strong>in</strong> r-space, or <strong>in</strong>tegr<strong>at</strong><strong>in</strong>g<br />

the density of st<strong>at</strong>es <strong>in</strong> energy space.<br />

<br />

Z = 4π<br />

dr r 2 ∆n(r) =<br />

ɛF<br />

−∞<br />

dɛ ∆ρ(ɛ) = Zb + 2<br />

π<br />

<br />

(2l + 1)δl(kF ) ,<br />

where Zb is the number of electrons <strong>in</strong> the bound st<strong>at</strong>es and the density of st<strong>at</strong>es<br />

∆ρ(ɛ) can be calcul<strong>at</strong>ed from the phase shift values as <strong>in</strong> (3.27):<br />

∆ρ(ɛ) = 2<br />

π<br />

<br />

l<br />

(2l + 1) dδl(ɛ)<br />

dɛ<br />

l<br />

84<br />

. (4.14)<br />

All three results for the extra charge must be very close to each other after the<br />

required convergence is achieved. The electronic density far away from an impurity<br />

should follow the form of the Friedel oscill<strong>at</strong>ions as mentioned <strong>in</strong> Ch. 3.<br />

∼ CF<br />

cos(2kF r + δ)<br />

r 3<br />

<strong>at</strong> large r ⇒ λ ∼ 2π/2kF<br />

Therefore the wave length of oscill<strong>at</strong>ions of the electronic density, λ, should be<br />

about 2π/2kF <strong>at</strong> large distances. There could be a virtual bound st<strong>at</strong>e if the<br />

number of bound st<strong>at</strong>e electrons is smaller than th<strong>at</strong> of the positive charge of the<br />

impurity. At low background densities, s<strong>in</strong>ce only low energies are allowed for the<br />

conduction band, only s-waves can contribute to the virtual bound st<strong>at</strong>e resonance<br />

<strong>in</strong> the density of st<strong>at</strong>es. The sc<strong>at</strong>ter<strong>in</strong>g length can be very important as <strong>in</strong> the case<br />

of a hydrogen impurity when the virtual bound st<strong>at</strong>e resonance is loc<strong>at</strong>ed near<br />

the conduction band m<strong>in</strong>imum and the non-zero angular momentum st<strong>at</strong>es can be<br />

neglected. In th<strong>at</strong> case, the sign of the sc<strong>at</strong>ter<strong>in</strong>g length can also be used to verify<br />

the behavior of the conduction band dur<strong>in</strong>g the iter<strong>at</strong>ive loops of the numerical<br />

calcul<strong>at</strong>ions.


FIGURE 4.15. The density of st<strong>at</strong>es <strong>at</strong> absolute zero temper<strong>at</strong>ure. With an<br />

external magnetic field, the energy of the sp<strong>in</strong>s (magnetic-moment-up) parallel to<br />

the field is lowered by µB, while the energy of the sp<strong>in</strong>s (magnetic-moment-down)<br />

antiparallel to the field is raised by µB. The chemical potential of the sp<strong>in</strong>-up<br />

band is equal to th<strong>at</strong> of the sp<strong>in</strong>-down band.<br />

4.6. Sp<strong>in</strong> paramagnetism<br />

A system with an external magnetic field is discussed <strong>in</strong> Sec. 2.2.6. In this<br />

section, details of the numerical implement<strong>at</strong>ion will be discussed. Note th<strong>at</strong> the<br />

sp<strong>in</strong>-up band is assumed to be parallel to an external magnetic field <strong>in</strong> this work. 10<br />

The energy shift by an external magnetic field <strong>in</strong> Fig. 4.15 is ±µB where − and<br />

+ should be taken for the sp<strong>in</strong>-up and sp<strong>in</strong>-down band respectively and µ is the<br />

Bohr magneton. Some electrons near the Fermi energy transfer from the sp<strong>in</strong>-down<br />

band to the sp<strong>in</strong>-up band such th<strong>at</strong> the chemical potential (Fermi energy) of the<br />

10 The term, sp<strong>in</strong>-up, <strong>in</strong> this work denotes sp<strong>in</strong>-magnetic-moment-up s<strong>in</strong>ce the magnetic<br />

moment is opposite <strong>in</strong> sign to the sp<strong>in</strong> S for an electron.<br />

85


sp<strong>in</strong>-up band equals th<strong>at</strong> of the sp<strong>in</strong>-down band. The range of k correspond<strong>in</strong>g to<br />

the range of the energy from the conduction band m<strong>in</strong>imum to the Fermi energy<br />

adjusts as well.<br />

sp<strong>in</strong>-up: 0 ↔ √ ɛF + u<br />

sp<strong>in</strong>-down: 0 ↔ √ ɛF − u ,<br />

where u = µB. The number of electrons is:<br />

N+ = V<br />

(2π) 3<br />

4<br />

3 π(ɛF + u) 3/2 for sp<strong>in</strong>-up (4.15)<br />

N− = V<br />

(2π) 3<br />

86<br />

4<br />

3 π(ɛF − u) 3/2 for sp<strong>in</strong>-down. (4.16)<br />

Now one can easily obta<strong>in</strong> the equ<strong>at</strong>ion for the Fermi energy ɛF .<br />

6π 2 n = (ɛF + u) 3/2 + (ɛF − u) 3/2 . (4.17)<br />

Note th<strong>at</strong> the Bohr magneton µ <strong>in</strong> <strong>at</strong>omic Rydberg unit is √ 2 and one <strong>at</strong>omic<br />

Rydberg unit for the magnetic <strong>in</strong>duction B corresponds to 3.324 × 10 5 T <strong>in</strong> SI.<br />

S<strong>in</strong>ce the Fermi wave vector kF of the sp<strong>in</strong>-up band is different from th<strong>at</strong><br />

of the sp<strong>in</strong>-down band with an external magnetic field, one can expect from the<br />

Friedel oscill<strong>at</strong>ions th<strong>at</strong> the electronic density for each band behaves differently<br />

even <strong>at</strong> large distances. Now, s<strong>in</strong>ce there are two different frequencies <strong>in</strong> the Friedel<br />

oscill<strong>at</strong>ions of the density, there should be p<strong>at</strong>terns of be<strong>at</strong>s 11 <strong>in</strong> the total electronic<br />

density <strong>in</strong>duced by an impurity. The Coulomb potential calcul<strong>at</strong>ed from the total<br />

11 The term, be<strong>at</strong>, is orig<strong>in</strong>ally used to describe fluctu<strong>at</strong>ions <strong>in</strong> the amplitude vari<strong>at</strong>ion<br />

caused by two sound waves of different frequency. In this work, be<strong>at</strong>s are used to<br />

emphasize the altern<strong>at</strong><strong>in</strong>g constructive and destructive <strong>in</strong>terference of two electronic<br />

density oscill<strong>at</strong>ions of different frequency.


electronic density, hence, has also p<strong>at</strong>terns of be<strong>at</strong>s as well. In th<strong>at</strong> case one can<br />

not use the method for the potential correction, which was <strong>in</strong>troduced <strong>in</strong> Sec.<br />

4.4.2 because the method takes advantage of the shape of the regular oscill<strong>at</strong>ions.<br />

However, one can avoid the be<strong>at</strong> p<strong>at</strong>terns by calcul<strong>at</strong><strong>in</strong>g the Coulomb potential<br />

separ<strong>at</strong>ely from the density of the sp<strong>in</strong>-up and sp<strong>in</strong>-down band and apply<strong>in</strong>g the<br />

method for each part of the Coulomb potential. After correct<strong>in</strong>g each Coulomb<br />

potential, one can f<strong>in</strong>d the total Coulomb potential by add<strong>in</strong>g both contributions<br />

to Coulomb potential and f<strong>in</strong>ally the effective potential for each band by add<strong>in</strong>g the<br />

exchange-correl<strong>at</strong>ion potential of each band to the total Coulomb potential. The<br />

Coulomb potential separ<strong>at</strong>ely calcul<strong>at</strong>ed from the density of each band, however,<br />

still may have irregular p<strong>at</strong>terns but not as strong as the total Coulomb potential.<br />

This is a second order effect s<strong>in</strong>ce the calcul<strong>at</strong>ions for the density are essentially<br />

based on an effective potential which has these be<strong>at</strong> p<strong>at</strong>terns. Therefore it is<br />

important to use a very large maximum radius to reduce the numerical error <strong>in</strong><br />

case th<strong>at</strong> the irregular p<strong>at</strong>tern causes a large numerical error. One often needs<br />

to exam<strong>in</strong>e the Coulomb potential and the effective potential dur<strong>in</strong>g the iter<strong>at</strong>ive<br />

loops of numerical calcul<strong>at</strong>ions and f<strong>in</strong>d the appropri<strong>at</strong>e maximum radius. Note<br />

th<strong>at</strong> the be<strong>at</strong> frequency kbe<strong>at</strong> is the difference of the two frequencies.<br />

kbe<strong>at</strong> = 2k +<br />

F<br />

− 2k−<br />

F<br />

87<br />

. (4.18)<br />

Thus one can estim<strong>at</strong>e how many be<strong>at</strong>s exist for a given range of r <strong>in</strong> the electronic<br />

density.


5. RESULTS AND DISCUSSION<br />

Two important quantities for the model used <strong>in</strong> this work are the positive<br />

charge of an impurity and the electronic background density. The Coulomb poten-<br />

tial is determ<strong>in</strong>ed by the positive charge of an impurity and the <strong>in</strong>duced electronic<br />

density while the exchange-correl<strong>at</strong>ion potential is determ<strong>in</strong>ed by the <strong>in</strong>duced elec-<br />

tronic density and the electronic background density. If the <strong>in</strong>duced electronic den-<br />

sity rema<strong>in</strong>s the same for the same impurity, 1 the exchange-correl<strong>at</strong>ion <strong>in</strong>teraction<br />

makes a difference between two different background densities, which, for a differ-<br />

ent background density, leads to a different self-consistent solution with different<br />

<strong>in</strong>duced charge densities, potentials, bound st<strong>at</strong>es, phase shifts, densities of st<strong>at</strong>es,<br />

and dielectric functions. The system still obeys the Friedel sum rule and exhibits<br />

Friedel oscill<strong>at</strong>ions. The results <strong>in</strong> this chapter are c<strong>at</strong>egorized by sp<strong>in</strong> polariza-<br />

tion. For a sp<strong>in</strong>-polarized system, the total sp<strong>in</strong>-moment will also be an important<br />

physical quantity, which may dist<strong>in</strong>guish between self-consistent solutions.<br />

1 This is not true. As the difference between background densities becomes larger, the<br />

response of an electron gas becomes more important due to the difference <strong>in</strong> Fermi<br />

energies and <strong>in</strong> Friedel oscill<strong>at</strong>ions. This effect is rel<strong>at</strong>ed to the amount of <strong>in</strong>duced<br />

charge <strong>in</strong> the vic<strong>in</strong>ity of an impurity as well as to the oscill<strong>at</strong>ions of the density <strong>at</strong> large<br />

distances.<br />

88


5.1. Non-sp<strong>in</strong>-polarized system<br />

The non-sp<strong>in</strong>-polarized system is tested ma<strong>in</strong>ly with a hydrogen impurity.<br />

The discussion <strong>in</strong> this section will proceed us<strong>in</strong>g a hydrogen impurity unless noted<br />

differently.<br />

5.1.1. Behavior of a system dur<strong>in</strong>g the numerical calcul<strong>at</strong>ions<br />

The number of electrons <strong>in</strong>duced by a hydrogen impurity should always be<br />

equal to one <strong>in</strong> order to screen the hydrogen impurity <strong>in</strong> a self-consistent manner,<br />

as mentioned <strong>in</strong> Ch. 3. Fig. 5.1 shows one example how the system evolves to<br />

s<strong>at</strong>isfy self-consistency. S<strong>in</strong>ce the <strong>in</strong>put effective potential for the first iter<strong>at</strong>ion is<br />

calcul<strong>at</strong>ed from the electronic density of a free hydrogen <strong>at</strong>om plus a background<br />

density of 0.06, the effective potential is sufficiently <strong>at</strong>tractive <strong>at</strong> the beg<strong>in</strong>n<strong>in</strong>g and<br />

thus there are two extra electrons <strong>in</strong> the bound st<strong>at</strong>es. The background density<br />

is, however, very high. S<strong>in</strong>ce the system with two electrons <strong>in</strong> the bound st<strong>at</strong>es<br />

of a hydrogen impurity can not be the self-consistent solutions for such a high<br />

background density, the system goes through a transition from hav<strong>in</strong>g a bound<br />

st<strong>at</strong>e to not hav<strong>in</strong>g a bound st<strong>at</strong>e and adds one extra electron <strong>in</strong> the conduction<br />

band. Notice <strong>in</strong> Fig. 5.1 th<strong>at</strong> the Friedel sum dur<strong>in</strong>g the transition responds to<br />

the change <strong>in</strong> the number of electrons <strong>in</strong> the bound st<strong>at</strong>es, while the conduction<br />

band cont<strong>in</strong>uously adds electrons to make one extra electron.<br />

5.1.2. Charge densities and potentials<br />

As already mentioned <strong>in</strong> Ch. 3, the Friedel oscill<strong>at</strong>ions are <strong>in</strong>dependent of<br />

the n<strong>at</strong>ure of an impurity. The <strong>in</strong>duced electronic densities <strong>at</strong> large distances<br />

89


FIGURE 5.1. Vari<strong>at</strong>ion of the number of extra electrons for a system with a<br />

hydrogen impurity dur<strong>in</strong>g the iter<strong>at</strong>ive loops of the numerical calcul<strong>at</strong>ions. The<br />

background electronic density is 0.06.<br />

90


FIGURE 5.2. The electronic density <strong>in</strong>duced by a hydrogen impurity. It is plotted<br />

as a function of 2kF r. Friedel oscill<strong>at</strong>ions are <strong>in</strong>dependent of the background<br />

density. The range of the background density is from 0.0005 to 0.06.<br />

show Friedel oscill<strong>at</strong>ions and they all have the same frequency 2kF , as illustr<strong>at</strong>ed<br />

<strong>in</strong> Fig. 5.2. The <strong>in</strong>duced electronic densities as a function of r are shown <strong>in</strong><br />

Fig. 5.3. Note th<strong>at</strong> although there is always one extra electron <strong>in</strong> the system<br />

of a hydrogen impurity, regardless of the background density, the amount of the<br />

electronic charge <strong>in</strong> the vic<strong>in</strong>ity of the impurity (r : 0 ∼ 4) is dependent upon<br />

the background density. There is rel<strong>at</strong>ively a large amount of electrons near the<br />

impurity for low background densities as a result of the effect of bound st<strong>at</strong>es<br />

and the Friedel oscill<strong>at</strong>ions <strong>in</strong> the conduction band. (Fig. 5.4 and Fig. 5.5) These<br />

phenomena are rel<strong>at</strong>ed to the behavior of the dielectric functions, different from<br />

91


FIGURE 5.3. The electronic density <strong>in</strong>duced by a hydrogen impurity.<br />

the Thomas-Fermi dielectric functions, which will be discussed <strong>in</strong> Sec. 5.1.4. The<br />

electronic density <strong>in</strong> the bound st<strong>at</strong>es becomes more extended and the maximum<br />

value of the density decreases as the background density <strong>in</strong>creases. (Fig. 5.4) The<br />

bound st<strong>at</strong>es disappear for the background density higher than about 0.03. The<br />

numerical calcul<strong>at</strong>ions for the background density between 0.02 and 0.03 are very<br />

difficult due to the gre<strong>at</strong>ly extended bound st<strong>at</strong>es. These results are not <strong>in</strong>cluded<br />

<strong>in</strong> Fig. 5.4 and Fig. 5.5 s<strong>in</strong>ce the required convergence could not be obta<strong>in</strong>ed. As<br />

discussed <strong>in</strong> Sec. 3.5, however, the transition from bound st<strong>at</strong>es to resonance st<strong>at</strong>es<br />

should not result <strong>in</strong> any non-analytic behavior of the properties of the system such<br />

as the immersion energy. The maximum value of the <strong>in</strong>duced electronic density <strong>in</strong><br />

92


FIGURE 5.4. The electronic density <strong>in</strong>duced by a hydrogen impurity <strong>in</strong> the bound<br />

st<strong>at</strong>es.<br />

the conduction band <strong>in</strong>creases <strong>in</strong> the vic<strong>in</strong>ity of the impurity as the background<br />

density <strong>in</strong>creases. (Fig. 5.5) As a result, the impurity is always screened sufficiently<br />

regardless of the existence of bound st<strong>at</strong>es.<br />

5.1.3. Phase shift, Sc<strong>at</strong>ter<strong>in</strong>g length, and Density of <strong>in</strong>duced st<strong>at</strong>es<br />

Phase shifts for l = 0 <strong>at</strong> zero energy have the value of π for low background<br />

densities, due to the presence of a bound st<strong>at</strong>e of a hydrogen impurity. The result<br />

is shown <strong>in</strong> Fig. 5.6. Note th<strong>at</strong> there are two electrons <strong>in</strong> the bound st<strong>at</strong>e due<br />

to the sp<strong>in</strong> degeneracy. The vari<strong>at</strong>ions of the phase shifts of p-waves <strong>in</strong> the case<br />

93


FIGURE 5.5. The electronic density <strong>in</strong>duced by a hydrogen impurity <strong>in</strong> the conduction<br />

band.<br />

of a hydrogen impurity are rel<strong>at</strong>ively very small compared to those of s-waves 2<br />

and thus the lead<strong>in</strong>g contribution to the change <strong>in</strong> the density of st<strong>at</strong>es arises<br />

from the phase shifts for l = 0. The slope of the phase shift (or the sc<strong>at</strong>ter<strong>in</strong>g<br />

length) for low background densities is neg<strong>at</strong>ive, from which one sees th<strong>at</strong> the<br />

2 This is only true if there are two electrons <strong>in</strong> the bound st<strong>at</strong>es. If the number of elec-<br />

trons is smaller than two, phase shifts of p-waves also become important and contribute<br />

significantly to the change <strong>in</strong> the number of electrons <strong>in</strong> the conduction band. See Sec.<br />

5.1.6.<br />

94


FIGURE 5.6. Phase shifts (l = 0) for a hydrogen impurity. Plotted for low<br />

background densities only.<br />

potential is <strong>at</strong>tractive enough to have a first bound st<strong>at</strong>e and, consider<strong>in</strong>g the<br />

density of st<strong>at</strong>es, the conduction band loses one electron to keep the number of<br />

total extra electrons one. The magnitude of the slope of the phase shifts <strong>in</strong> general<br />

becomes smaller <strong>in</strong> Fig. 5.6 as the background density decreases, which one can<br />

expect consider<strong>in</strong>g the rel<strong>at</strong>ionship between the strength of the potential and the<br />

sc<strong>at</strong>ter<strong>in</strong>g length. 3 However, as the background density decreases more, the phase<br />

3 One may expect th<strong>at</strong> the potentials become more <strong>at</strong>tractive as the background density<br />

decreases, which means th<strong>at</strong> the sc<strong>at</strong>ter<strong>in</strong>g length decreases <strong>in</strong> magnitude. [31]<br />

95


FIGURE 5.7. Inverse of sc<strong>at</strong>ter<strong>in</strong>g length versus background density for a hydrogen<br />

impurity.<br />

shifts near the conduction band maximum tend to have a larger neg<strong>at</strong>ive slope than<br />

those of the higher background densities. At extremely low background densities<br />

such as 0.0005, even the slope near the conduction band m<strong>in</strong>imum becomes larger.<br />

(See Fig. 5.7.) The reason for this is th<strong>at</strong> a smaller neg<strong>at</strong>ive slope <strong>in</strong> the phase<br />

shifts can not yield a correct neg<strong>at</strong>ive excess charge <strong>at</strong> such a low background<br />

density to s<strong>at</strong>isfy the Friedel sum rule as the conduction band becomes narrower. 4<br />

4 Moreover, as the background density decreases, the contribution of s-waves to the<br />

number of <strong>in</strong>duced electrons becomes larger, and thus the phase shifts of s-waves <strong>at</strong> the<br />

Fermi energy should respond accord<strong>in</strong>gly.<br />

96


FIGURE 5.8. Phase shifts for a hydrogen impurity for two different background<br />

densities (0.0025 and 0.06).<br />

As a result, the system behaves such th<strong>at</strong> the potentials are less <strong>at</strong>tractive <strong>at</strong> the<br />

extremely low background densities. For high background densities (≥ 0.03) for a<br />

hydrogen impurity, s<strong>in</strong>ce there are no electrons <strong>in</strong> the bound st<strong>at</strong>es, there should<br />

be one extra electron <strong>in</strong> the conduction band, for which the contribution of s-waves<br />

to a virtual bound st<strong>at</strong>e resonance is significantly larger than the higher angular<br />

momentum st<strong>at</strong>es. (Fig. 5.8) The sc<strong>at</strong>ter<strong>in</strong>g lengths correspond<strong>in</strong>g to the phase<br />

shifts <strong>in</strong> Fig. 5.9 are shown <strong>in</strong> Fig. 5.7. The transition from a bound st<strong>at</strong>e to a<br />

resonance st<strong>at</strong>e occurs <strong>at</strong> about 0.03 of the background density, close to when the<br />

sign of the sc<strong>at</strong>ter<strong>in</strong>g length changes. The <strong>in</strong>verse of the sc<strong>at</strong>ter<strong>in</strong>g length <strong>in</strong>creases<br />

as the background density <strong>in</strong>creases from 0.0015. The magnitude of the sc<strong>at</strong>ter<strong>in</strong>g<br />

length, however, rapidly <strong>in</strong>creases as the background density decreases from 0.0015<br />

and becomes large <strong>at</strong> extremely low background densities (≤ 0.001), which cre<strong>at</strong>es<br />

a m<strong>in</strong>imum po<strong>in</strong>t <strong>at</strong> about 0.0015 <strong>in</strong> Fig. 5.7. This behavior is already discussed<br />

with the phase shifts above.<br />

97


FIGURE 5.9. Phase shifts (l = 0) for a hydrogen impurity.<br />

FIGURE 5.10. The density of st<strong>at</strong>es for a hydrogen impurity for two different<br />

background densities (0.0025 and 0.06).<br />

98


Typical examples for the density of st<strong>at</strong>es are shown <strong>in</strong> Fig. 5.10. At the<br />

background density of 0.0025, there are two electrons <strong>in</strong> the bound st<strong>at</strong>e, and the<br />

lead<strong>in</strong>g contribution to the conduction density of st<strong>at</strong>es comes from s-st<strong>at</strong>es, which<br />

show a neg<strong>at</strong>ive sc<strong>at</strong>ter<strong>in</strong>g length and a sharp neg<strong>at</strong>ive peak <strong>at</strong> the conduction<br />

band m<strong>in</strong>imum <strong>in</strong> the density of st<strong>at</strong>es. The conduction band loses one electron,<br />

which removes the excess charge of the bound st<strong>at</strong>e electrons. At the background<br />

density of 0.06, there are no electrons <strong>in</strong> the bound st<strong>at</strong>es. The effective potential<br />

is, however, <strong>at</strong>tractive enough to <strong>in</strong>duce an extra electron <strong>in</strong> the conduction band,<br />

which appears as a narrow peak <strong>in</strong> the conduction density of st<strong>at</strong>es due to the<br />

rapid <strong>in</strong>crease of the phase shift of l = 0 <strong>in</strong> the region of the conduction band<br />

m<strong>in</strong>imum. Therefore one sees th<strong>at</strong> the Friedel sum rule is always s<strong>at</strong>isfied and can<br />

be achieved <strong>at</strong> self-consistency dur<strong>in</strong>g the numerical calcul<strong>at</strong>ions. The phase shifts<br />

for a lithium impurity are shown <strong>in</strong> Fig. 5.11. At the background density of 0.0005,<br />

the electron configur<strong>at</strong>ion of a lithium impurity is 1s 2 2s 2 and the phase shift of<br />

the zero energy must be 2π (third branch 5 of tan δ0) by Lev<strong>in</strong>son’s theorem. The<br />

phase shift <strong>at</strong> 0.0005 shows a large neg<strong>at</strong>ive slope, which gives rise to the neg<strong>at</strong>ive<br />

density of st<strong>at</strong>es <strong>in</strong> the conduction band to remove the excess charge due to the<br />

bound st<strong>at</strong>es. As the background density <strong>in</strong>creases, the effective potential becomes<br />

less <strong>at</strong>tractive and the system loses a bound st<strong>at</strong>e (second branch of tanδ0) and<br />

the sc<strong>at</strong>ter<strong>in</strong>g length becomes positive. (phase shifts <strong>at</strong> 0.0025 <strong>in</strong> Fig. 5.11) If the<br />

background density further <strong>in</strong>creases, the sc<strong>at</strong>ter<strong>in</strong>g length decreases (phase shifts<br />

<strong>at</strong> 0.03 <strong>in</strong> Fig. 5.11) and becomes neg<strong>at</strong>ive. At the higher background density the<br />

5 See (3.26) and [31].<br />

99


FIGURE 5.11. Phase shifts of l = 0 and l = 1 for a lithium impurity.<br />

100


same cycle of the phase shifts repe<strong>at</strong>s, after loos<strong>in</strong>g one more bound st<strong>at</strong>e. (first<br />

branch of tanδ0)<br />

5.1.4. Dielectric constant and compressibility<br />

101<br />

The dielectric constants can be easily calcul<strong>at</strong>ed us<strong>in</strong>g the def<strong>in</strong>ition <strong>in</strong> (3.55)<br />

and the Fourier transform.<br />

V ext<br />

q<br />

ɛ(w, q) =<br />

V eff<br />

q<br />

<br />

−Ze dr exp(−iqr)<br />

=<br />

2<br />

r<br />

dr exp(−iqr)V (r)<br />

=<br />

−e 2 Z<br />

q dr V (r)r s<strong>in</strong>(qr)<br />

(5.1)<br />

(5.2)<br />

. (5.3)<br />

The Thomas-Fermi dielectric constant, already <strong>in</strong>troduced <strong>in</strong> Sec. 3.4, is given by<br />

where q 2 T F = 4e2 mkF<br />

2 π<br />

ɛT F (0, q) = q2 + q 2 T F<br />

q 2 , (5.4)<br />

. The Thomas-Fermi dielectric constants <strong>in</strong> Fig. 5.12 show a<br />

simple behavior which gives a screen<strong>in</strong>g length 6 proportional to n −1/3 . One can ex-<br />

pect from the Thomas-Fermi dielectric constants th<strong>at</strong> a higher background density<br />

has a shorter screen<strong>in</strong>g length. The numerical results of the dielectric constants<br />

for a hydrogen impurity and the Thomas-Fermi dielectric constants both approach<br />

the value of one as q → ∞. The density functional calcul<strong>at</strong>ions, however, show<br />

quite different results for the dielectric constants, especially <strong>at</strong> small q values. Note<br />

th<strong>at</strong> the presence of electrons <strong>in</strong> the bound st<strong>at</strong>es and the long wave length of the<br />

6 The screen<strong>in</strong>g length for the Thomas-Fermi dielectric constants can be easily obta<strong>in</strong>ed<br />

from the Yukawa potential, l = 1/qT F . See (3.67).


102<br />

FIGURE 5.12. Dielectric constants which are numerically calcul<strong>at</strong>ed for a hydrogen<br />

impurity. The Thomas-Fermi dielectric constants are also plotted for comparison.


103<br />

FIGURE 5.13. Compressibility calcul<strong>at</strong>ed numerically for a hydrogen impurity.<br />

A dotted l<strong>in</strong>e which is taken from [6] is for the result predicted by theories such<br />

as Hubbard, VS, and SS. Both l<strong>in</strong>es are go<strong>in</strong>g through zero around rs = 5.0. The<br />

electronic density which corresponds to rs of the bottom x-axis is given <strong>in</strong> the top<br />

x-axis.<br />

Friedel oscill<strong>at</strong>ions result <strong>in</strong> a larger amount of screen<strong>in</strong>g charge near the impu-<br />

rity for low background densities than for high background densities, as already<br />

mentioned <strong>in</strong> Sec. 5.1.2. One sees th<strong>at</strong> dielectric constants may have very large<br />

neg<strong>at</strong>ive or positive values for a certa<strong>in</strong> q while the <strong>in</strong>tegral term <strong>in</strong> (5.3) <strong>in</strong>creases<br />

from neg<strong>at</strong>ive to positive values near zero. Actually, it turns out th<strong>at</strong> the <strong>in</strong>tegral<br />

term <strong>in</strong> (5.3) changes its sign from neg<strong>at</strong>ive to positive as the low background<br />

density (n 0.00191 or rs 5.0) decreases further. At extremely low background<br />

densities (≤ 0.0015), therefore, the dielectric constants for q → 0 are neg<strong>at</strong>ive, as


FIGURE 5.14. Comparison of the compressibility calcul<strong>at</strong>ed numerically for a<br />

hydrogen impurity with sc<strong>at</strong>ter<strong>in</strong>g length.<br />

illustr<strong>at</strong>ed by Fig. 5.12. This result can be verified by discuss<strong>in</strong>g compressibilities.<br />

The compressibility can be obta<strong>in</strong>ed <strong>in</strong> terms of the dielectric function by [13]<br />

lim<br />

q→0 ɛ(q, 0) = q2 + q2 T F<br />

q2 K<br />

Kf<br />

104<br />

, (5.5)<br />

where qT F = 6πne 2 /ɛF is the Thomas-Fermi wave number and the quantity Kf =<br />

3(3π 2 ) −2/3 n −5/3 [1/e 2 a0] is the isothermal compressibility of the non-<strong>in</strong>teract<strong>in</strong>g<br />

electron gas <strong>at</strong> zero temper<strong>at</strong>ure. The compressibility is already obta<strong>in</strong>ed by sev-<br />

eral theories. 7 The RPA gives the constant value 1.0 for Kf/K while many other<br />

7 Readers should refer to [18] for more details.


theories such as Hubbard, S<strong>in</strong>gwi-Sjölander, and Vashishta-S<strong>in</strong>gwi give results for<br />

Kf/K, which all go through 0 around rs = 5.0. [18] The density functional cal-<br />

cul<strong>at</strong>ions for the compressibility give rise to nearly the same result for Kf/K as<br />

those theories for low background densities but to much larger values for high<br />

background densities, which results <strong>in</strong> the curved l<strong>in</strong>e of a hyperbolic type <strong>in</strong> Fig.<br />

5.13. One sees from (5.5) and Fig. 5.13 th<strong>at</strong> the dielectric constant for the high<br />

background density as q → 0 has a smaller value than expected by theories 8 while<br />

the dielectric constant for the low background density has a larger value. 9 In other<br />

words, if the screen<strong>in</strong>g length can be def<strong>in</strong>ed for any potential 10 , it can be st<strong>at</strong>ed<br />

th<strong>at</strong> the screen<strong>in</strong>g length for the high background density has a larger value than<br />

expected by analytical theories while the screen<strong>in</strong>g length for a low background<br />

density has a smaller value.<br />

105<br />

As shown <strong>in</strong> Fig. 5.14, one sees th<strong>at</strong> the neg<strong>at</strong>ive m<strong>in</strong>imum (<strong>at</strong> rs 5.0)<br />

of a value of the sc<strong>at</strong>ter<strong>in</strong>g length corresponds to <strong>in</strong>f<strong>in</strong>ity of the compressibility<br />

and <strong>in</strong>f<strong>in</strong>ity (<strong>at</strong> rs 2.0) of a value of the sc<strong>at</strong>ter<strong>in</strong>g length corresponds to the<br />

compressibility close to Kf. Note th<strong>at</strong> <strong>in</strong>f<strong>in</strong>ity of a sc<strong>at</strong>ter<strong>in</strong>g length <strong>in</strong>dic<strong>at</strong>es the<br />

transition from a bound st<strong>at</strong>e to a resonance st<strong>at</strong>e. One may th<strong>in</strong>k th<strong>at</strong> the most<br />

<strong>at</strong>tractive potential corresponds to the maximum of the compressibility, which<br />

8 See the dotted l<strong>in</strong>e <strong>in</strong> Fig. 5.13. As rs → 0, the expected dielectric constant is the<br />

Thomas-Fermi dielectric constant.<br />

9 Note th<strong>at</strong>, <strong>in</strong> Fig. 5.13, the dielectric constants calcul<strong>at</strong>ed by theories are always larger<br />

than or equal to the Thomas-Fermi dielectric constants.<br />

10 One can try to f<strong>in</strong>d the screen<strong>in</strong>g length for a particular potential by compar<strong>in</strong>g with<br />

the Thomas-Fermi dielectric constant. l = 1/qT F = ((ɛ(q) − 1)q 2 ) −1/2 as q → 0


106<br />

FIGURE 5.15. Immersion energies for several different impurities. 1.0 <strong>in</strong> the<br />

legend <strong>in</strong>dic<strong>at</strong>es th<strong>at</strong> the number of electrons <strong>in</strong> the bound st<strong>at</strong>es is limited to 1.0.<br />

465 po<strong>in</strong>ts for the logarithmic r-space and 0.81205 × 10 −5 for the first value of<br />

r-mesh are used <strong>in</strong> these calcul<strong>at</strong>ions.<br />

may be rel<strong>at</strong>ed to the localiz<strong>at</strong>ion of electrons 11 . The rel<strong>at</strong>ionship between the<br />

sc<strong>at</strong>ter<strong>in</strong>g length and the compressibility, however, needs further <strong>in</strong>vestig<strong>at</strong>ion.<br />

5.1.5. Immersion energy<br />

Immersion energies for H, Li, and C are shown <strong>in</strong> Fig. 5.15. A m<strong>in</strong>imum <strong>in</strong><br />

the immersion energy exists for H and C, while calcul<strong>at</strong>ions for Li can not verify<br />

11 Wigner l<strong>at</strong>tice [18]


107<br />

FIGURE 5.16. Vari<strong>at</strong>ion <strong>in</strong> the k<strong>in</strong>etic, Coulomb, and exchange-correl<strong>at</strong>ion energies<br />

vs the background density for a hydrogen impurity. 465 po<strong>in</strong>ts for the<br />

logarithmic r-space and 0.81205 × 10 −5 for the first vale of r-mesh are used <strong>in</strong><br />

these calcul<strong>at</strong>ions.<br />

the existence of a m<strong>in</strong>imum immersion energy. 12 The background densities <strong>at</strong><br />

which the m<strong>in</strong>imum immersion energies exist are 0.0024 for H and 0.0033 for C.<br />

The immersion energy after the m<strong>in</strong>imum value <strong>in</strong>creases almost l<strong>in</strong>early as the<br />

background density <strong>in</strong>creases, which is due to the repulsion rel<strong>at</strong>ed to the <strong>in</strong>crease<br />

12 To verify the existence of the m<strong>in</strong>imum immersion energy for Li, it is necessary to<br />

perform the calcul<strong>at</strong>ions for very low background densities as shown <strong>in</strong> Fig. 5.15, which<br />

is, however, very difficult due to the long range wave length of the Friedel oscill<strong>at</strong>ions as<br />

mentioned <strong>in</strong> Ch. 4.


<strong>in</strong> the k<strong>in</strong>etic energy. [5] The Coulomb and exchange-correl<strong>at</strong>ion energies actually<br />

become lower as the background density <strong>in</strong>creases. The r<strong>at</strong>e of the <strong>in</strong>crease <strong>in</strong> the<br />

k<strong>in</strong>etic energy is, however, large enough to compens<strong>at</strong>e this decrease, as illustr<strong>at</strong>ed<br />

<strong>in</strong> Fig. 5.16.<br />

108<br />

The zero background density limit <strong>in</strong> this <strong>in</strong>f<strong>in</strong>ite system may be different<br />

from a free <strong>at</strong>om without an electron gas. S<strong>in</strong>ce the system, <strong>at</strong> constant chem-<br />

ical potential, is extended to <strong>in</strong>f<strong>in</strong>ity, the impurity will be screened, no m<strong>at</strong>ter<br />

how small the background density, (but not zero) and, for this non-sp<strong>in</strong>-polarized<br />

<strong>in</strong>f<strong>in</strong>ite system, the impurity is not allowed to have partially occupied orbitals. 13<br />

Calcul<strong>at</strong>ions show th<strong>at</strong> a hydrogen impurity has no electrons <strong>in</strong> the bound st<strong>at</strong>es<br />

for high background densities( 0.03) and two electrons <strong>in</strong> the bound st<strong>at</strong>es for<br />

low background densities but no electrons aga<strong>in</strong> for extremely low background<br />

densities( 0.0001), which can be seen from the fact th<strong>at</strong> a free hydrogen <strong>at</strong>om<br />

can not have two electrons <strong>in</strong> the bound st<strong>at</strong>es <strong>in</strong> LDA. The <strong>in</strong>terest<strong>in</strong>g result<br />

is th<strong>at</strong> the immersion energy for a hydrogen impurity shows a l<strong>in</strong>ear behavior as<br />

illustr<strong>at</strong>ed <strong>in</strong> Fig. 5.15 when the hydrogen is forced to have only one electron <strong>in</strong><br />

the bound st<strong>at</strong>es regardless of the background density. Immersion energy <strong>in</strong> this<br />

case approaches zero as the background density goes to zero, which one can easily<br />

expect from the fact th<strong>at</strong> a free hydrogen <strong>at</strong>om has only one electron. A helium<br />

impurity which has two bound st<strong>at</strong>e electrons <strong>in</strong> a homogeneous electron gas shows<br />

the same l<strong>in</strong>ear behavior <strong>in</strong> Fig. 5.15.


109<br />

FIGURE 5.17. Immersion energies for the different number of electrons <strong>in</strong> the<br />

bound st<strong>at</strong>es. 465 po<strong>in</strong>ts for the logarithmic r-space and 0.81205 × 10 −5 for the<br />

first value of r-mesh are used <strong>in</strong> these calcul<strong>at</strong>ions.<br />

5.1.6. Excited system<br />

One can easily calcul<strong>at</strong>e the immersion energies of excited systems and obta<strong>in</strong><br />

the excit<strong>at</strong>ion energies by vary<strong>in</strong>g the number of electrons <strong>in</strong> the bound st<strong>at</strong>es as<br />

mentioned <strong>in</strong> Sec. 2.2.5. The bound st<strong>at</strong>e energy decreases as the number of<br />

electrons <strong>in</strong> an 1s st<strong>at</strong>e decreases while the immersion energy <strong>in</strong>creases. (See Fig.<br />

5.17.) This can be verified by Janak’s theorem:<br />

∂Eimm<br />

∂ni<br />

= ɛi(ni) , (5.6)<br />

13 For <strong>in</strong>stance, only two or zero electrons are allowed <strong>in</strong> an 1s st<strong>at</strong>e.


110<br />

FIGURE 5.18. Vari<strong>at</strong>ion <strong>in</strong> the bound st<strong>at</strong>e energy for the different number of<br />

electrons <strong>in</strong> the bound st<strong>at</strong>e. The second plot is obta<strong>in</strong>ed us<strong>in</strong>g (5.7) and slopes<br />

between two adjacent immersion energies <strong>in</strong> Fig. 5.17. 465 po<strong>in</strong>ts for the logarithmic<br />

r-space and 0.81205 × 10 −5 for the first value of r-mesh are used <strong>in</strong> these<br />

calcul<strong>at</strong>ions.


where ni denotes the occup<strong>at</strong>ion number of the level i. However, one needs to<br />

consider the change <strong>in</strong> the number of electrons <strong>at</strong> the Fermi energy as well as the<br />

bound st<strong>at</strong>es <strong>in</strong> order to compare the bound st<strong>at</strong>e energy with the vari<strong>at</strong>ion of<br />

the immersion energy us<strong>in</strong>g Janak’s theorem. Janak’s theorem (5.6) should be<br />

modified as<br />

∂Eimm<br />

∂ni<br />

111<br />

= −ɛF + ɛi(ni) , (5.7)<br />

where ɛF is the Fermi energy. Fig. 5.18 shows the result of (5.7) with the bound<br />

st<strong>at</strong>e energy plot. Bound st<strong>at</strong>e energies obta<strong>in</strong>ed from (5.7) are <strong>in</strong> error with<br />

respect to actually calcul<strong>at</strong>ed bound st<strong>at</strong>e energies by an energy difference smaller<br />

than 0.005 Rydbergs. 14 Bound st<strong>at</strong>e energies of a free hydrogen <strong>at</strong>om are <strong>in</strong>cluded<br />

<strong>in</strong> Fig. 5.18. One should, however, recall th<strong>at</strong> the zero background density limit<br />

may be different from a free <strong>at</strong>om as discussed <strong>in</strong> Sec. 5.1.5. One sees from the<br />

Friedel sum rule th<strong>at</strong> the screen<strong>in</strong>g <strong>in</strong> the model used <strong>in</strong> this work will always occur<br />

as long as there is a background density. The extra electrons <strong>in</strong> the conduction<br />

band will respond to the excess core charge, which makes the zero background<br />

density limit different from a free <strong>at</strong>om. A hydrogen impurity which has only<br />

one electron <strong>in</strong> a homogeneous electronic gas is, however, expected to reach a free<br />

hydrogen <strong>at</strong>om as the background density goes to zero. S<strong>in</strong>ce the extra electron<br />

<strong>in</strong> the conduction band <strong>in</strong> this case must be zero, as the background density goes<br />

to zero, the only response of the conduction band will be th<strong>at</strong> the wave length<br />

of the Friedel oscill<strong>at</strong>ion gets longer and it is found th<strong>at</strong> the amplitude of the<br />

Friedel oscill<strong>at</strong>ion near an impurity gets smaller. This st<strong>at</strong>ement is, however, not<br />

14 These differences are ma<strong>in</strong>ly due to <strong>in</strong>accur<strong>at</strong>e numerical deriv<strong>at</strong>ives.


verified numerically for a background density lower than 0.001 due to the numerical<br />

problems discussed <strong>in</strong> Ch. 4.<br />

112


113<br />

FIGURE 5.19. Comparison of immersion energies for a carbon impurity between a<br />

non-sp<strong>in</strong>-polarized system and a sp<strong>in</strong>-polarized system. The same reference energy<br />

for a free carbon <strong>at</strong>om is used for both calcul<strong>at</strong>ions.<br />

5.2. Sp<strong>in</strong> polarized system<br />

5.2.1. Zero external magnetic field<br />

5.2.1.1. Carbon impurity <strong>at</strong> low background densities<br />

A carbon impurity is ma<strong>in</strong>ly used for a sp<strong>in</strong>-polarized system <strong>in</strong> this work<br />

and <strong>in</strong>terest<strong>in</strong>g results are obta<strong>in</strong>ed <strong>at</strong> low background densities. Fig. 5.19 shows<br />

immersion energies for a sp<strong>in</strong>-polarized system and a non-sp<strong>in</strong>-polarized system of a<br />

carbon impurity. Overall, immersion energies of a sp<strong>in</strong>-polarized system are slightly<br />

lower than those of a non-sp<strong>in</strong>-polarized system. Note th<strong>at</strong> these calcul<strong>at</strong>ions are<br />

based on the same energy of a free <strong>at</strong>om, th<strong>at</strong> is, the same reference and thus only


FIGURE 5.20. Plot of ∆N vs a background density where ∆N is the difference<br />

<strong>in</strong> the number of electrons between sp<strong>in</strong>-up and sp<strong>in</strong>-down electrons <strong>in</strong> a system.<br />

rel<strong>at</strong>ive energies between two systems are mean<strong>in</strong>gful. One important difference<br />

between the two systems is the response of the conduction band to the number of<br />

extra electrons <strong>in</strong> the bound st<strong>at</strong>es <strong>at</strong> low background densities(≤ 0.005). For both<br />

systems, there are only four electrons <strong>in</strong> the bound st<strong>at</strong>es with the configur<strong>at</strong>ion<br />

1s 2 2s 2 , for a background density higher than 0.002. Note th<strong>at</strong> there should be<br />

two electrons <strong>in</strong> the conduction band due to the Friedel sum rule if there are<br />

four electrons <strong>in</strong> the bound st<strong>at</strong>es of a carbon impurity, which is true for both<br />

systems. However, a sp<strong>in</strong>-polarized system shows a non-zero magnetiz<strong>at</strong>ion for a<br />

background density lower than 0.005, which makes a difference <strong>in</strong> the immersion<br />

114<br />

energy compared to the non-sp<strong>in</strong>-polarized system. One sees th<strong>at</strong>, <strong>in</strong> Fig. 5.20,


FIGURE 5.21. The electronic density and the sp<strong>in</strong> density <strong>in</strong>duced by a carbon<br />

impurity.<br />

115


FIGURE 5.22. The electronic density (sp<strong>in</strong>-up and sp<strong>in</strong>-down) <strong>in</strong>duced by a carbon<br />

impurity <strong>in</strong> the conduction band.<br />

116


FIGURE 5.23. The density of st<strong>at</strong>es <strong>in</strong>duced by a carbon impurity.<br />

117


∆N (which is the difference <strong>in</strong> the number of electrons between sp<strong>in</strong>-up and sp<strong>in</strong>-<br />

down electrons <strong>in</strong> a system) becomes larger as the background density decreases<br />

and thus a sp<strong>in</strong>-polarized system favors a sp<strong>in</strong>-polarized solution for a background<br />

density lower than 0.005. Fig. 5.21 shows th<strong>at</strong> the sp<strong>in</strong> density becomes larger<br />

<strong>in</strong> the vic<strong>in</strong>ity of an impurity as the background density decreases while the total<br />

density <strong>in</strong>duced by an impurity rema<strong>in</strong>s almost the same. This sp<strong>in</strong> density also<br />

can be seen <strong>in</strong> Fig. 5.22. The maximum peak of the density <strong>in</strong>duced by an impurity<br />

<strong>in</strong> the conduction band of sp<strong>in</strong>-up electrons <strong>in</strong>creases <strong>in</strong> the vic<strong>in</strong>ity of an impurity<br />

between 0.002 and 0.005 background densities and decreases <strong>in</strong> the conduction<br />

band of sp<strong>in</strong>-down electrons. The density of st<strong>at</strong>es <strong>in</strong> the conduction bands of<br />

sp<strong>in</strong>-up and sp<strong>in</strong>-down electrons <strong>in</strong> Fig. 5.23 clearly show th<strong>at</strong> there is a resonance<br />

<strong>in</strong> the conduction band of sp<strong>in</strong>-up electrons between 0.002 and 0.005 background<br />

densities while there is none <strong>in</strong> the conduction band of sp<strong>in</strong>-down electrons. One<br />

other difference between a sp<strong>in</strong>-polarized system and a non-sp<strong>in</strong>-polarized system<br />

is th<strong>at</strong> a sp<strong>in</strong>-polarized system has a 2p bound st<strong>at</strong>e <strong>in</strong> the sp<strong>in</strong>-up band for a<br />

background density smaller than 0.002 and hence there are seven bound st<strong>at</strong>e<br />

electrons <strong>in</strong> a system while a non-sp<strong>in</strong>-polarized system can not have six electrons<br />

<strong>in</strong> a 2p bound st<strong>at</strong>e. 15 One sees now <strong>in</strong> Fig. 5.22 th<strong>at</strong>, due to a 2p bound st<strong>at</strong>e,<br />

the <strong>in</strong>duced density <strong>in</strong> the conduction band of sp<strong>in</strong>-up electrons sharply drops<br />

<strong>in</strong> the vic<strong>in</strong>ity of an impurity as the background density decreases from 0.002<br />

15 A non-sp<strong>in</strong>-polarized system is not tested for a background density lower than 0.0005.<br />

It is, however, found by consider<strong>in</strong>g the r<strong>at</strong>e of the vari<strong>at</strong>ion of the bound st<strong>at</strong>e energy<br />

as a background density decreases th<strong>at</strong> a non-sp<strong>in</strong>-polarized system does not have a 2p<br />

bound st<strong>at</strong>e <strong>at</strong> any background density.<br />

118


to 0.001. Note th<strong>at</strong> the <strong>in</strong>duced density <strong>in</strong> the conduction band of sp<strong>in</strong>-down<br />

electrons <strong>at</strong> 0.0004 background density has a rel<strong>at</strong>ively large hole (radius around<br />

10) compared to th<strong>at</strong> <strong>at</strong> 0.0005 background density, while there is a small difference<br />

of the <strong>in</strong>duced density <strong>in</strong> the conduction band of sp<strong>in</strong>-up electrons between 0.0004<br />

and 0.0005 background densities. One can expect th<strong>at</strong> ∆N <strong>in</strong> Fig. 5.20 is even<br />

larger than 3.0 <strong>at</strong> 0.0004 background density. 16 One possible explan<strong>at</strong>ion for this<br />

is th<strong>at</strong>, due to the strong effect of the exchange-correl<strong>at</strong>ion <strong>in</strong>teraction, a system<br />

may get a lower immersion energy by mak<strong>in</strong>g the magnetiz<strong>at</strong>ion larger <strong>at</strong> these<br />

extremely low background densities. (See Fig. 2.7, Fig. 2.8, and Fig. 2.9.)<br />

5.2.1.2. Electrical Resistivity<br />

The general idea is taken from [15] and [32].<br />

With<strong>in</strong> the scheme of density functional theory, the s<strong>in</strong>gle particle wave func-<br />

tions are used to calcul<strong>at</strong>e the ground st<strong>at</strong>e densities. The use of these wave func-<br />

tions for other purposes is not formally justified. However, the phase shifts (which<br />

are rel<strong>at</strong>ed to the wave functions) give reasonable agreement with experiment as<br />

discussed <strong>in</strong> [32].<br />

119<br />

The contribution of sc<strong>at</strong>ter<strong>in</strong>g centers to the electrical resistivity can be de-<br />

rived <strong>in</strong> terms of the phase shifts by<br />

∆ρ = 4πni<br />

e 2 kF Zh<br />

∞<br />

(l + 1) s<strong>in</strong> 2 [δl(ɛF ) − δl+1(ɛF )] , (5.8)<br />

l=0<br />

16 At 0.0004 background density, it is difficult to obta<strong>in</strong> a self-consistent solution with<br />

the required convergence due to the strong tendency toward a sp<strong>in</strong>-polarized system.


120<br />

FIGURE 5.24. The values of Q versus Z for the background density of 0.01. The<br />

calcul<strong>at</strong>ions are performed only for Z = 1, 3, 6. Note th<strong>at</strong>, <strong>at</strong> this background density,<br />

there is no difference between sp<strong>in</strong>-polarized and non-sp<strong>in</strong>-polarized systems<br />

for these impurities.<br />

where ni is the impurity concentr<strong>at</strong>ion and Zh is the number of valence electrons<br />

per host ion. 17 (See [32] and [15].) A useful expression is<br />

Q = 3<br />

∞<br />

(l + 1) s<strong>in</strong> 2 [δl(ɛF ) − δl+1(ɛF )] , (5.9)<br />

kF r 3 s<br />

l=0<br />

where rs is the Wigner-Seitz radius. The electronic stopp<strong>in</strong>g power for an ion<br />

slowly mov<strong>in</strong>g through the electron gas can be obta<strong>in</strong>ed from this equ<strong>at</strong>ion by<br />

multiply<strong>in</strong>g Q by vI, where vI is the velocity of an ion. Fig. 5.24 shows the Q<br />

values for H, Li, and C. The result agrees with Ref. [32]. The one important fe<strong>at</strong>ure<br />

17 CGS units are used <strong>in</strong> this section.


of such results is the oscill<strong>at</strong>ions of Q values as a function of Z. 18 These oscill<strong>at</strong>ions<br />

were observed experimentally <strong>in</strong> impurity residual resistivity measurements and <strong>in</strong><br />

the stopp<strong>in</strong>g power d<strong>at</strong>a for well-channeled slow ions. [33] The reader should refer<br />

to [32] for more details.<br />

5.2.1.3. Excited system<br />

121<br />

As discussed <strong>in</strong> Sec. 5.2.1.1, a sp<strong>in</strong>-polarized system with a carbon impurity<br />

has a very strong tendency toward a large sp<strong>in</strong>-moment solution(parallel to sp<strong>in</strong>-<br />

up) <strong>at</strong> low background densities(< 0.005). (See Fig. 5.20.) However, a strong<br />

tendency toward an anti-sp<strong>in</strong>-moment solution(parallel to sp<strong>in</strong>-down) <strong>in</strong> the con-<br />

duction band has been observed for excited systems as well. Fig. 5.25 shows the<br />

vari<strong>at</strong>ion of immersion energies and ∆N as a function of the number of sp<strong>in</strong>-up<br />

electrons <strong>in</strong> a 2p bound st<strong>at</strong>e. 19 The small vari<strong>at</strong>ion <strong>in</strong> the slope of the immersion<br />

energies <strong>in</strong> Fig. 5.25 can be verified with the bound st<strong>at</strong>e energies (Fig. 5.26) of a<br />

2p st<strong>at</strong>e us<strong>in</strong>g the Janak’s theorem as <strong>in</strong> Sec. 5.1.6. <strong>Electron</strong>s <strong>in</strong> the conduction<br />

band behave <strong>in</strong> such a way th<strong>at</strong> the difference <strong>in</strong> the total <strong>in</strong>duced density due to<br />

the change <strong>in</strong> a 2p bound st<strong>at</strong>e is as small as possible as illustr<strong>at</strong>ed <strong>in</strong> Fig. 5.27 ∼<br />

Fig. 5.38. However, one sees immedi<strong>at</strong>ely <strong>in</strong> Fig. 5.25 th<strong>at</strong> decreas<strong>in</strong>g the number<br />

of sp<strong>in</strong>-up electrons <strong>in</strong> a 2p bound st<strong>at</strong>e results <strong>in</strong> reduc<strong>in</strong>g the sp<strong>in</strong>-moment, th<strong>at</strong><br />

is, <strong>in</strong> a large anti-sp<strong>in</strong>-moment <strong>in</strong> the conduction band.<br />

18 In order to observe the oscill<strong>at</strong>ions of Q values, one needs to obta<strong>in</strong> the results for Z<br />

<strong>at</strong> least up to Z = 20. See [32].<br />

19 Note th<strong>at</strong> there is no 2p bound st<strong>at</strong>e for sp<strong>in</strong>-down electrons.


122<br />

FIGURE 5.25. Immersion energies and ∆N for a carbon impurity vs the number<br />

of sp<strong>in</strong>-up electrons <strong>in</strong> a 2p bound st<strong>at</strong>e. ∆N is the difference <strong>in</strong> the number of<br />

electrons between sp<strong>in</strong>-up and sp<strong>in</strong>-down electrons <strong>in</strong> a system.


123<br />

FIGURE 5.26. Bound st<strong>at</strong>e energies vs the number of sp<strong>in</strong>-up electrons <strong>in</strong> a 2p<br />

bound st<strong>at</strong>e. The second plot shows the vari<strong>at</strong>ion of the curv<strong>at</strong>ure, the energy<br />

m<strong>in</strong>imum, and the number of electrons <strong>in</strong> a 2p bound st<strong>at</strong>e correspond<strong>in</strong>g to the<br />

energy m<strong>in</strong>imum <strong>in</strong> the first plot.


FIGURE 5.27. The density and the sp<strong>in</strong> density <strong>in</strong>duced by a carbon impurity <strong>at</strong><br />

0.0005 background density.<br />

124


FIGURE 5.28. The density <strong>in</strong> the bound st<strong>at</strong>es of sp<strong>in</strong>-up and sp<strong>in</strong>-down electrons<br />

<strong>in</strong>duced by a carbon impurity <strong>at</strong> 0.0005 background density.<br />

125


FIGURE 5.29. The density and the sp<strong>in</strong> density <strong>in</strong> the conduction band <strong>in</strong>duced<br />

by a carbon impurity <strong>at</strong> 0.0005 background density.<br />

126


FIGURE 5.30. The density and the sp<strong>in</strong> density <strong>in</strong>duced by a carbon impurity <strong>at</strong><br />

0.00075 background density.<br />

127


FIGURE 5.31. The density <strong>in</strong> the bound st<strong>at</strong>es of sp<strong>in</strong>-up and sp<strong>in</strong>-down electrons<br />

<strong>in</strong>duced by a carbon impurity <strong>at</strong> 0.00075 background density.<br />

128


FIGURE 5.32. The density and the sp<strong>in</strong> density <strong>in</strong> the conduction band <strong>in</strong>duced<br />

by a carbon impurity <strong>at</strong> 0.00075 background density.<br />

129


FIGURE 5.33. The density and the sp<strong>in</strong> density <strong>in</strong>duced by a carbon impurity <strong>at</strong><br />

0.001 background density.<br />

130


FIGURE 5.34. The density <strong>in</strong> the bound st<strong>at</strong>es of sp<strong>in</strong>-up and sp<strong>in</strong>-down electrons<br />

<strong>in</strong>duced by a carbon impurity <strong>at</strong> 0.001 background density.<br />

131


FIGURE 5.35. The density and the sp<strong>in</strong> density <strong>in</strong> the conduction band <strong>in</strong>duced<br />

by a carbon impurity <strong>at</strong> 0.001 background density.<br />

132


FIGURE 5.36. The density and the sp<strong>in</strong> density <strong>in</strong>duced by a carbon impurity <strong>at</strong><br />

0.0011 background density.<br />

133


FIGURE 5.37. The density <strong>in</strong> the bound st<strong>at</strong>es of sp<strong>in</strong>-up and sp<strong>in</strong>-down electrons<br />

<strong>in</strong>duced by a carbon impurity <strong>at</strong> 0.0011 background density.<br />

134


FIGURE 5.38. The density and the sp<strong>in</strong> density <strong>in</strong> the conduction band <strong>in</strong>duced<br />

by a carbon impurity <strong>at</strong> 0.0011 background density.<br />

135


136<br />

Fig. 5.26 shows <strong>in</strong> this system th<strong>at</strong> the bound st<strong>at</strong>e energy as a function<br />

of the number of sp<strong>in</strong>-up electrons <strong>in</strong> a 2p bound st<strong>at</strong>e behaves unusually as a<br />

quadr<strong>at</strong>ic function. 20 Important parameters of quadr<strong>at</strong>ic functions are also shown<br />

<strong>in</strong> Fig. 5.26. The bound st<strong>at</strong>e energy decreases as the number of sp<strong>in</strong>-up electrons<br />

<strong>in</strong> a 2p bound st<strong>at</strong>e decreases from 3.0 electrons as expected. However, the bound<br />

st<strong>at</strong>e energy <strong>in</strong>creases as the number of sp<strong>in</strong>-up electrons <strong>in</strong> a 2p bound st<strong>at</strong>e<br />

decreases from the bound st<strong>at</strong>e energy m<strong>in</strong>imum due to the largely <strong>in</strong>duced sp<strong>in</strong>-<br />

down electrons near an impurity <strong>in</strong> the conduction band.<br />

This behavior is consistent with the potential.(Fig. 5.39) For <strong>in</strong>stance, the<br />

<strong>in</strong>duced sp<strong>in</strong> up and down electrons results <strong>in</strong> a behavior of the Coulomb potential<br />

for each band for 0.0005 background density as illustr<strong>at</strong>ed <strong>in</strong> Fig. 5.40. The total<br />

Coulomb potential near an impurity (r ≤ 1) <strong>in</strong> Fig. 5.41 decreases as the number of<br />

sp<strong>in</strong>-up electrons <strong>in</strong> a 2p bound st<strong>at</strong>e decreases and <strong>in</strong>creases as sp<strong>in</strong>-up electrons<br />

<strong>in</strong> a 2p bound st<strong>at</strong>e decreases from 2.5. One sees also th<strong>at</strong> the change <strong>in</strong> the total<br />

Coulomb potential (r ≤ 1) between 2.0 and 2.5 sp<strong>in</strong>-up electrons <strong>in</strong> a 2p bound<br />

st<strong>at</strong>e is not as large as th<strong>at</strong> between 2.5 and 3.0 sp<strong>in</strong>-up electrons <strong>in</strong> a 2p bound<br />

st<strong>at</strong>e. However, with the <strong>in</strong>crease of sp<strong>in</strong>-up exchange-correl<strong>at</strong>ion potential(Fig.<br />

5.42), one sees th<strong>at</strong> the total effective potential of sp<strong>in</strong>-up electrons <strong>at</strong> 2.0 electrons<br />

<strong>in</strong> a 2p bound st<strong>at</strong>e is now close to th<strong>at</strong> <strong>at</strong> 3.0 electrons. (See Fig. 5.39.) This<br />

potential behavior accounts for the vari<strong>at</strong>ion of the bound st<strong>at</strong>e energy <strong>in</strong> Fig.<br />

5.26 and the exchange-correl<strong>at</strong>ion <strong>in</strong>teraction <strong>in</strong> Fig. 5.42 causes the decrease of<br />

20 Note th<strong>at</strong> one can not use a simple first order of a Taylor series for this bound st<strong>at</strong>e<br />

energy as usual.


FIGURE 5.39. The effective potential of the sp<strong>in</strong>-up band vs the number of<br />

sp<strong>in</strong>-up electrons <strong>in</strong> a 2p bound st<strong>at</strong>e for 0.0005 background density.<br />

137


138<br />

FIGURE 5.40. The Coulomb potentials calcul<strong>at</strong>ed by <strong>in</strong>duced sp<strong>in</strong>-up and<br />

sp<strong>in</strong>-down electrons vs the number of sp<strong>in</strong>-up electrons <strong>in</strong> a 2p bound st<strong>at</strong>e for<br />

0.0005 background density.


139<br />

FIGURE 5.41. The Coulomb potentials vs the number of sp<strong>in</strong>-up electrons <strong>in</strong> a<br />

2p bound st<strong>at</strong>e for 0.0005 background density. The second plot shows the area<br />

which corresponds to the box <strong>in</strong> the first plot.


140<br />

FIGURE 5.42. The exchange-correl<strong>at</strong>ion potentials of <strong>in</strong>duced sp<strong>in</strong>-up and<br />

sp<strong>in</strong>-down electrons vs the number of sp<strong>in</strong>-up electrons <strong>in</strong> a 2p bound st<strong>at</strong>e for<br />

0.0005 background density.


the sp<strong>in</strong>-moment(the <strong>in</strong>crease of the anti-sp<strong>in</strong>-moment <strong>in</strong> the conduction band)<br />

s<strong>in</strong>ce sp<strong>in</strong>-up and down electrons experience the same Coulomb potential.<br />

141<br />

The phase shift <strong>in</strong> Fig. 5.43 shows a consistent behavior. Note <strong>in</strong> the overall<br />

sc<strong>at</strong>ter<strong>in</strong>g length and volume a decrease for sp<strong>in</strong>-up electrons and an <strong>in</strong>crease for<br />

sp<strong>in</strong>-down electrons. One knows th<strong>at</strong> the Friedel sum rule forces the phase shift<br />

<strong>at</strong> the Fermi energy to adjust to screen the excess charge of the core correctly.<br />

The way this system behaves due to the Friedel sum rule is th<strong>at</strong> the decrease <strong>in</strong><br />

the number of sp<strong>in</strong>-up electrons <strong>in</strong> a 2p bound st<strong>at</strong>e causes an <strong>in</strong>crease <strong>in</strong> the<br />

phase shift of the sp<strong>in</strong>-down electrons <strong>at</strong> the Fermi energy and the decrease <strong>in</strong><br />

the phase shift of the sp<strong>in</strong>-up electrons, which is consistent with the behavior of<br />

the sp<strong>in</strong>-density <strong>in</strong> the conduction band. With 3.0 electrons <strong>in</strong> a 2p bound st<strong>at</strong>e,<br />

the phase shift <strong>in</strong> Fig. 5.43 shows th<strong>at</strong> there is a sc<strong>at</strong>ter<strong>in</strong>g resonance <strong>in</strong> the high<br />

energy region (≥ ɛF ) of the phase shift of the sp<strong>in</strong>-down electrons. The resonance<br />

moves closer to the Fermi energy as the number of electrons <strong>in</strong> a 2p bound st<strong>at</strong>e<br />

decreases. Accord<strong>in</strong>gly, one sees large <strong>in</strong>creases <strong>at</strong> the Fermi energy <strong>in</strong> the density<br />

of <strong>in</strong>duced st<strong>at</strong>es of the sp<strong>in</strong>-down band, as illustr<strong>at</strong>ed <strong>in</strong> Fig. 5.44 and Fig. 5.45.<br />

These results can be expla<strong>in</strong>ed by self-consistency and the exchange-<br />

correl<strong>at</strong>ion <strong>in</strong>teraction. 21 If the <strong>in</strong>duced density <strong>in</strong> the bound st<strong>at</strong>es is reduced<br />

artificially, electrons <strong>in</strong> the conduction band simply can not be <strong>in</strong>duced more to<br />

ma<strong>in</strong>ta<strong>in</strong> the same total <strong>in</strong>duced density s<strong>in</strong>ce the same total <strong>in</strong>duced density yields<br />

the same Coulomb potentials and accord<strong>in</strong>gly the same <strong>in</strong>duced density <strong>in</strong> the con-<br />

21 These explan<strong>at</strong>ions need further <strong>in</strong>vestig<strong>at</strong>ion.


142<br />

FIGURE 5.43. Phase shifts of l = 0 and l = 1 for sp<strong>in</strong>-up and sp<strong>in</strong>-down electrons.


FIGURE 5.44. The density of <strong>in</strong>duced st<strong>at</strong>es for sp<strong>in</strong>-up and sp<strong>in</strong>-down electrons<br />

<strong>at</strong> 0.0005 background density.<br />

143


FIGURE 5.45. The density of <strong>in</strong>duced st<strong>at</strong>es for sp<strong>in</strong>-up and sp<strong>in</strong>-down electrons<br />

<strong>at</strong> 0.001 background density.<br />

144


145<br />

FIGURE 5.46. The exchange-correl<strong>at</strong>ion potential with different <strong>in</strong>duced densities<br />

<strong>in</strong> the sp<strong>in</strong>-up band. Background density is 0.0005. The <strong>in</strong>duced density <strong>in</strong> the<br />

sp<strong>in</strong>-down band is assumed to be 0.047.<br />

duction band. 22 The reduced density <strong>in</strong> the bound st<strong>at</strong>es, thus, tends to cause a<br />

more extended <strong>in</strong>duced density. Moreover, the potential due to the exchange-<br />

correl<strong>at</strong>ion <strong>in</strong>teraction 23 also can vary correspond<strong>in</strong>g to the change <strong>in</strong> the density.<br />

Fig. 5.46 shows one example of the exchange-correl<strong>at</strong>ion behavior <strong>in</strong> the case th<strong>at</strong><br />

the <strong>in</strong>duced density <strong>in</strong> the sp<strong>in</strong>-up band varies. The reduced density <strong>in</strong> the sp<strong>in</strong>-<br />

22 Once the density <strong>in</strong> the bound st<strong>at</strong>es is changed, the old bound st<strong>at</strong>es and sc<strong>at</strong>tered<br />

st<strong>at</strong>es which are orthogonal to each other do not s<strong>at</strong>isfy the self-consistency anymore.<br />

23 The exchange-correl<strong>at</strong>ion <strong>in</strong>teraction used <strong>in</strong> this work only depends on the local<br />

density.


146<br />

FIGURE 5.47. Partial wave decompositions of the Friedel sum rule. These plots<br />

show the change <strong>in</strong> the number of electrons <strong>in</strong>duced by a carbon impurity for each<br />

band. 0.0005 is used for a background density. Bound st<strong>at</strong>es are not <strong>in</strong>cluded.


up band gives rise to a more repulsive exchange-correl<strong>at</strong>ion potential and thus the<br />

sp<strong>in</strong>-down band has rel<strong>at</strong>ively more <strong>in</strong>duced electrons <strong>in</strong> the conduction band due<br />

to the Coulomb <strong>in</strong>teraction as the number of electrons <strong>in</strong> a 2p bound st<strong>at</strong>e of the<br />

sp<strong>in</strong>-up band decreases. Fig. 5.47 shows the response of the conduction bands <strong>in</strong><br />

terms of the phase shifts <strong>at</strong> the Fermi energy. One sees th<strong>at</strong>, as the the number of<br />

electrons <strong>in</strong> a 2p bound st<strong>at</strong>e of the sp<strong>in</strong>-up band decreases, the sp<strong>in</strong>-up conduction<br />

band loses more electrons <strong>in</strong> overall angular momentum st<strong>at</strong>es and the decrease of<br />

<strong>in</strong>duced electrons <strong>in</strong> the p st<strong>at</strong>e of the sp<strong>in</strong>-up band is compens<strong>at</strong>ed by ma<strong>in</strong>ly the<br />

same p st<strong>at</strong>e of the sp<strong>in</strong>-down band.<br />

147


148<br />

FIGURE 5.48. Bound st<strong>at</strong>e energies, the conduction band m<strong>in</strong>imums, the<br />

sp<strong>in</strong>-moment, and the immersion energies vs the magnetic field for a hydrogen<br />

impurity with 0.0025 background density.<br />

5.2.2. Non-zero external magnetic field<br />

A hydrogen impurity <strong>in</strong> a sp<strong>in</strong>-polarized system has been tested by apply<strong>in</strong>g<br />

an external magnetic field. Detailed <strong>in</strong>form<strong>at</strong>ion for the implement<strong>at</strong>ion can be<br />

found <strong>in</strong> Sec. 4.6. Results for three different background densities are obta<strong>in</strong>ed.<br />

Fig. 5.48 shows the <strong>in</strong>duced sp<strong>in</strong>-moment <strong>in</strong> a system, bound st<strong>at</strong>e energies, the<br />

conduction band m<strong>in</strong>imum, and the immersion energies for a background density<br />

of 0.0025. As the magnetic field <strong>in</strong>creases, the background density (conduction<br />

band width) of sp<strong>in</strong>-up (majority) electrons <strong>in</strong>creases and the background density<br />

of sp<strong>in</strong>-down (m<strong>in</strong>ority) electrons decreases. Therefore, ∆N <strong>in</strong> this case is the<br />

excess sp<strong>in</strong>-moment, the devi<strong>at</strong>ion from the moment due to the background by


149<br />

FIGURE 5.49. The exchange-correl<strong>at</strong>ion potential with different background densities<br />

of the sp<strong>in</strong>-up band. The total background density is 0.0025 and the <strong>in</strong>duced<br />

density is assumed to be 0.03 for each sp<strong>in</strong>-up and sp<strong>in</strong>-down band. This plot<br />

shows how the exchange-correl<strong>at</strong>ion potential for each band may vary with the<br />

<strong>in</strong>creas<strong>in</strong>g magnetic field.<br />

itself. There are two electrons <strong>in</strong> a 1s st<strong>at</strong>e without an external magnetic field as<br />

<strong>in</strong> a non-sp<strong>in</strong> polarized system. (See Sec. 5.1.) One sees th<strong>at</strong> bound st<strong>at</strong>e ener-<br />

gies behave as <strong>in</strong> non-sp<strong>in</strong> polarized systems <strong>in</strong> Fig. 5.48. Th<strong>at</strong> is, the bound st<strong>at</strong>e<br />

energy <strong>in</strong> the sp<strong>in</strong>-up band approaches the conduction band m<strong>in</strong>imum as the back-<br />

ground density of the sp<strong>in</strong>-up band <strong>in</strong>creases while the bound st<strong>at</strong>e energy <strong>in</strong> the<br />

sp<strong>in</strong>-down band becomes lower compared to the band m<strong>in</strong>imum. As the magnetic<br />

field <strong>in</strong> Fig. 5.48 approaches 0.09894, 24 where the sp<strong>in</strong>-down conduction band com-<br />

24 The unit for the magnetic field <strong>in</strong> this work is 3.3241346·10 5 T, which will be omitted<br />

<strong>in</strong> this paper for convenience.


pletely disappears, the calcul<strong>at</strong>ions are not performed correctly due to numerical<br />

problems (long wave length of Friedel oscill<strong>at</strong>ions) as mentioned <strong>in</strong> Ch. 4 and thus<br />

d<strong>at</strong>a are not available for a magnetic field between 0.08 and 0.10. One important<br />

phenomenon is the sp<strong>in</strong>-moment which decreases with the <strong>in</strong>creas<strong>in</strong>g magnetic field<br />

and then rapidly <strong>in</strong>creases after a magnetic field of 0.06, as illustr<strong>at</strong>ed <strong>in</strong> Fig. 5.48.<br />

Two important facts are, due to the difference between sp<strong>in</strong>-up and sp<strong>in</strong>-down<br />

electrons, the exchange-correl<strong>at</strong>ion <strong>in</strong>teraction and the Coulomb <strong>in</strong>teraction of the<br />

Friedel oscill<strong>at</strong>ions with two different Fermi wave vectors.<br />

150<br />

Fig. 5.49 shows th<strong>at</strong>, if the total background density rema<strong>in</strong>s the same, the<br />

exchange-correl<strong>at</strong>ion potential of the sp<strong>in</strong>-up band becomes more repulsive and<br />

the exchange-correl<strong>at</strong>ion potential of the sp<strong>in</strong>-down band rapidly becomes more<br />

<strong>at</strong>tractive due to the exchange-correl<strong>at</strong>ion <strong>in</strong>teraction as the background density<br />

of the sp<strong>in</strong>-down band approaches zero, which suggests th<strong>at</strong> there may be a rapid<br />

vari<strong>at</strong>ion <strong>in</strong> potentials of sp<strong>in</strong>-up and sp<strong>in</strong>-down electrons when the external mag-<br />

netic field <strong>in</strong>creases. The potential vari<strong>at</strong>ion <strong>in</strong> Fig. 5.49 is actually very large<br />

compared to the potential difference <strong>at</strong> the <strong>at</strong>omic 2p loc<strong>at</strong>ion (r ∼ 4) for a hy-<br />

drogen impurity between the high and low background densities (such as 0.04 and<br />

0.0025). As the external magnetic field <strong>in</strong>creases, this behavior yields not only a<br />

more repulsive potential <strong>in</strong> the vic<strong>in</strong>ity of an impurity but also a large peak <strong>in</strong> r<br />

times the potential near the 2p loc<strong>at</strong>ion for the background density of the sp<strong>in</strong>-<br />

up band, as illustr<strong>at</strong>ed <strong>in</strong> Fig. 5.50, which is the opposite behavior compared to<br />

the potentials of systems without a magnetic field <strong>in</strong> Fig. 5.51. 25 The number of<br />

25 The lower background density has a larger fluctu<strong>at</strong>ion <strong>at</strong> large distances <strong>in</strong> the r times<br />

potential for a system without the external magnetic field as <strong>in</strong> Fig. 5.51.


FIGURE 5.50. The effective potentials for the different magnetic field 0.01 and<br />

0.04. A hydrogen impurity with 0.0025 background density is used.<br />

151


152<br />

FIGURE 5.51. The effective potential for a hydrogen impurity. No external magnetic<br />

field is applied. The potential of the sp<strong>in</strong>-up band is identical to th<strong>at</strong> of the<br />

sp<strong>in</strong>-down band.<br />

<strong>in</strong>duced electrons, therefore, decreases <strong>in</strong> the sp<strong>in</strong>-up band due to the repulsive<br />

exchange-correl<strong>at</strong>ion <strong>in</strong>teraction and <strong>in</strong>creases <strong>in</strong> the sp<strong>in</strong>-down band due to the<br />

<strong>at</strong>tractive exchange-correl<strong>at</strong>ion <strong>in</strong>teraction and, as a result, the sp<strong>in</strong>-moment de-<br />

creases up to 0.05 magnetic field as shown <strong>in</strong> Fig. 5.48. The phase shifts <strong>at</strong> the<br />

Fermi energy show a consistent behavior <strong>in</strong> Fig. 5.52. Without any external mag-<br />

netic field, there is no difference between the sp<strong>in</strong>-up and sp<strong>in</strong>-down bands. This<br />

means th<strong>at</strong>, if the number of <strong>in</strong>duced electrons <strong>in</strong> the s-st<strong>at</strong>e of the conduction<br />

band is reduced by a higher background density, the higher angular momentum<br />

st<strong>at</strong>es must have a larger contribution to the number of <strong>in</strong>duced electrons, due to<br />

the Friedel sum rule as shown <strong>in</strong> Fig. 5.53. The phase shifts <strong>at</strong> the Fermi energy


153<br />

FIGURE 5.52. Partial wave decompositions of the Friedel sum rule. These plots<br />

show the change <strong>in</strong> the number of electrons <strong>in</strong>duced by a hydrogen impurity for<br />

each band. 0.0025 is used for a background density.


154<br />

FIGURE 5.53. Partial wave decompositions of the Friedel sum rule. no external<br />

magnetic field is applied. This plot for the sp<strong>in</strong>-up band is identical with th<strong>at</strong> for<br />

the sp<strong>in</strong>-down band.<br />

with a non-zero magnetic field, however, <strong>in</strong> Fig. 5.52, show a different behavior<br />

for the higher angular momentum st<strong>at</strong>es. One sees <strong>in</strong> Fig. 5.52 th<strong>at</strong> the vari<strong>at</strong>ion<br />

<strong>in</strong> the number of <strong>in</strong>duced electrons due to a different magnetic field for the high<br />

angular momentum st<strong>at</strong>es of the sp<strong>in</strong>-down band (≥ p) is small up to 0.05 mag-<br />

netic field compared to th<strong>at</strong> for a system without a magnetic field(Fig. 5.53). One<br />

sees also th<strong>at</strong> the phase shifts <strong>at</strong> the p-st<strong>at</strong>e <strong>in</strong> the sp<strong>in</strong>-up band become lower as<br />

the magnetic field <strong>in</strong>creases and thus lose more electrons <strong>in</strong> the p-st<strong>at</strong>e while the<br />

s-st<strong>at</strong>e lose electrons as well. This behavior is also opposite to a system without a<br />

magnetic field(Fig. 5.53). The strong exchange-correl<strong>at</strong>ion <strong>in</strong>teraction due to the<br />

magnetic field for this background density is believed to cause these phenomena.


155<br />

FIGURE 5.54. The <strong>in</strong>duced electronic densities <strong>in</strong> the conduction band for different<br />

magnetic fields. A hydrogen impurity with 0.0025 background density is<br />

used.


156<br />

FIGURE 5.55. The <strong>in</strong>duced electronic densities <strong>in</strong> the conduction band of sp<strong>in</strong>-up<br />

electrons for each angular momentum st<strong>at</strong>e. These plots correspond to the magnetic<br />

field 0.08 <strong>in</strong> Fig. 5.48 and Fig. 5.54. A hydrogen impurity with 0.0025 background<br />

density is used.


157<br />

FIGURE 5.56. The density of st<strong>at</strong>es of the sp<strong>in</strong>-up and sp<strong>in</strong>-down bands for<br />

different magnetic fields. The arrows <strong>in</strong> the first plot <strong>in</strong>dic<strong>at</strong>e the loc<strong>at</strong>ions of the<br />

sc<strong>at</strong>ter<strong>in</strong>g resonances. A hydrogen impurity with 0.0025 background density is<br />

used.


FIGURE 5.57. Comparison of kF of the sp<strong>in</strong>-down band with k values which<br />

correspond to the sc<strong>at</strong>ter<strong>in</strong>g resonances <strong>in</strong> Fig. 5.56.<br />

158<br />

There is another important <strong>in</strong>teraction which causes the rapid <strong>in</strong>crease <strong>in</strong><br />

the sp<strong>in</strong>-moment (Fig. 5.48) and the fluctu<strong>at</strong>ions <strong>in</strong> the high angular momentum<br />

st<strong>at</strong>es of the sp<strong>in</strong>-up band as <strong>in</strong> Fig. 5.52. S<strong>in</strong>ce there is a difference <strong>in</strong> Fermi<br />

wave vector between the sp<strong>in</strong>-up and sp<strong>in</strong>-down bands, one may easily expect th<strong>at</strong><br />

there is a be<strong>at</strong> p<strong>at</strong>tern <strong>in</strong> the <strong>in</strong>duced density. The sp<strong>in</strong>-down band has a larger<br />

range of Coulomb holes and peaks <strong>in</strong> the density than the sp<strong>in</strong>-up band due to<br />

the smaller Fermi wave vector as shown <strong>in</strong> Fig. 5.54. The sp<strong>in</strong>-up band tries to<br />

compens<strong>at</strong>e a large range of these Coulomb fluctu<strong>at</strong>ions <strong>in</strong> the <strong>in</strong>duced density<br />

of the sp<strong>in</strong>-down band and results <strong>in</strong> the <strong>in</strong>creased values of the phase shifts of<br />

high angular momentum st<strong>at</strong>es <strong>at</strong> the Fermi energies. On the other hand, the<br />

sp<strong>in</strong>-down band also tries to keep the same Coulomb holes as usual by remov<strong>in</strong>g


sp<strong>in</strong>-down electrons. 26 One sees th<strong>at</strong> the <strong>in</strong>creases for l = 3, 4, and 5 of the<br />

sp<strong>in</strong>-up band <strong>in</strong> Fig. 5.52 and consistently the <strong>in</strong>duced densities <strong>in</strong> the sp<strong>in</strong>-up<br />

band for the high angular momentum st<strong>at</strong>es, which are <strong>at</strong> loc<strong>at</strong>ions of the large<br />

Coulomb holes of the sp<strong>in</strong>-down band, are unusually large as illustr<strong>at</strong>ed <strong>in</strong> Fig.<br />

5.54 and Fig. 5.55. Also note th<strong>at</strong>, for large magnetic fields, there are significant<br />

decreases <strong>in</strong> the l = 1, and 2 phase shifts of the sp<strong>in</strong>-down band <strong>in</strong> Fig. 5.52.<br />

This <strong>in</strong>teraction dom<strong>in</strong><strong>at</strong>es the exchange-correl<strong>at</strong>ion <strong>in</strong>teraction as the sp<strong>in</strong>-down<br />

background density approaches zero and yields a large sp<strong>in</strong>-moment as <strong>in</strong> Fig.<br />

5.48. One can see this <strong>in</strong>teraction <strong>in</strong> the density of st<strong>at</strong>es as well. (See Fig. 5.56.)<br />

There is a sc<strong>at</strong>ter<strong>in</strong>g resonance <strong>in</strong> the density of st<strong>at</strong>es for the sp<strong>in</strong>-up band which<br />

becomes large as the background density of the sp<strong>in</strong>-down band approaches zero.<br />

One sees also <strong>in</strong> Fig. 5.57 th<strong>at</strong> the momentum value correspond<strong>in</strong>g to the sc<strong>at</strong>ter<strong>in</strong>g<br />

resonance closely follows the Fermi wave vector of the sp<strong>in</strong>-down band, which is<br />

rel<strong>at</strong>ed to the Coulomb compens<strong>at</strong>ion of the sp<strong>in</strong>-up band. After the sp<strong>in</strong>-down<br />

conduction band disappears, the sp<strong>in</strong>-moment is simply -1 s<strong>in</strong>ce only the sp<strong>in</strong>-<br />

down band has a bound st<strong>at</strong>e. The immersion energy varies <strong>in</strong> Fig. 5.48 opposite<br />

to the sp<strong>in</strong>-moment s<strong>in</strong>ce the <strong>in</strong>duced sp<strong>in</strong>-up electrons have a lower energy when<br />

the energy is shifted by the magnetic field.<br />

26 This Coulomb <strong>in</strong>teraction of the Friedel oscill<strong>at</strong>ions makes obta<strong>in</strong><strong>in</strong>g a numerical self-<br />

consistent solution much more difficult.<br />

159


160<br />

FIGURE 5.58. Bound st<strong>at</strong>e energies, the conduction band m<strong>in</strong>imums, the<br />

sp<strong>in</strong>-moment, and the immersion energies vs the magnetic field for a hydrogen<br />

impurity with two different background densities, 0.01 and 0.04.


161<br />

FIGURE 5.59. The exchange-correl<strong>at</strong>ion potential with different background densities<br />

of the sp<strong>in</strong>-up band. The total background density is 0.04 and the <strong>in</strong>duced<br />

density is assumed to be 0.03 for each sp<strong>in</strong>-up and sp<strong>in</strong>-down band. This plot<br />

shows how the exchange-correl<strong>at</strong>ion potential for each band may vary with the<br />

<strong>in</strong>creas<strong>in</strong>g magnetic field.<br />

The calcul<strong>at</strong>ions for a background density of 0.01 yield similar results as<br />

shown <strong>in</strong> Fig. 5.58. 27 The background density 0.04, however, gives different results<br />

as shown <strong>in</strong> Fig. 5.58. Note th<strong>at</strong>, without an external magnetic field, there are no<br />

electrons <strong>in</strong> the bound st<strong>at</strong>es for a background density of 0.04 and a 1s bound st<strong>at</strong>e<br />

exists <strong>in</strong> the sp<strong>in</strong>-down band for the magnetic field larger than 0.25. The ma<strong>in</strong><br />

27 In the results for 0.01 background density, there is a discont<strong>in</strong>uity <strong>in</strong> the sp<strong>in</strong>-moment<br />

near 0.19 magnetic field as shown <strong>in</strong> Fig. 5.58, which is believed to be a numerical<br />

problem due to the large sp<strong>at</strong>ial extension of the bound st<strong>at</strong>e <strong>in</strong> the sp<strong>in</strong>-up band.


162<br />

FIGURE 5.60. The <strong>in</strong>tegr<strong>at</strong>ed values of the density of st<strong>at</strong>es of the sp<strong>in</strong>-up band<br />

for different magnetic fields to f<strong>in</strong>d the total <strong>in</strong>duced number of electrons <strong>in</strong> the<br />

sp<strong>in</strong>-up conduction band. A hydrogen impurity with 0.04 background density is<br />

used. The band m<strong>in</strong>imums are shifted <strong>in</strong> such a way th<strong>at</strong> they have the same<br />

m<strong>in</strong>imum.<br />

difference is th<strong>at</strong> the sp<strong>in</strong>-moment slightly <strong>in</strong>creases as the magnetic field <strong>in</strong>creases<br />

from zero but shows the same behavior as results for the background densities of<br />

0.0025 and 0.01 after 0.035 magnetic field. This may be expla<strong>in</strong>ed by the exchange-<br />

correl<strong>at</strong>ion <strong>in</strong>teraction for this high background density and the difference <strong>in</strong> the<br />

band width(kF ) between sp<strong>in</strong>-up and sp<strong>in</strong>-down bands. S<strong>in</strong>ce this is such a high<br />

background density, the exchange-correl<strong>at</strong>ion potentials vary rel<strong>at</strong>ively slowly as<br />

the magnetic field <strong>in</strong>creases from zero compared to Fig. 5.49 (See Fig. 5.59.) and<br />

the change <strong>in</strong> the band width due to the magnetic field becomes more important.<br />

In this case, the number of <strong>in</strong>duced electrons <strong>in</strong>creases as the Fermi wave vector


163<br />

FIGURE 5.61. Partial wave decompositions of the Friedel sum rule. These plots<br />

show the change <strong>in</strong> the number of electrons <strong>in</strong>duced by a hydrogen impurity for<br />

each band. 0.04 is used for a background density.


FIGURE 5.62. The density <strong>in</strong>duced by a hydrogen impurity for each conduction<br />

band. The background density is 0.04 and the magnetic field is 0.3.<br />

of the sp<strong>in</strong>-up electrons <strong>in</strong>creases, s<strong>in</strong>ce there are more st<strong>at</strong>es available which can<br />

contribute to the <strong>in</strong>duced electrons. One can <strong>in</strong>tegr<strong>at</strong>e the density of st<strong>at</strong>es to<br />

f<strong>in</strong>d the number of <strong>in</strong>duced electrons, which is shown <strong>in</strong> Fig. 5.60. The <strong>in</strong>tegr<strong>at</strong>ed<br />

number of <strong>in</strong>duced electrons as a function of energy <strong>in</strong> the sp<strong>in</strong>-up band is slightly<br />

different between a magnetic field of 0.1 and 0.35, but the higher Fermi wave vector<br />

results <strong>in</strong> a ga<strong>in</strong> of more electrons as illustr<strong>at</strong>ed <strong>in</strong> Fig. 5.60. The phase shifts <strong>at</strong><br />

the Fermi wave vector <strong>in</strong> Fig. 5.61 for values of the magnetic field between 0.1<br />

and 0.35 show a significant change for high angular momentum st<strong>at</strong>es (such as<br />

l = 1 ∼ 4) compared to the change <strong>in</strong> the s-st<strong>at</strong>e as the magnetic field <strong>in</strong>creases.<br />

The <strong>in</strong>duced density of the sp<strong>in</strong>-up band is much larger than th<strong>at</strong> of the sp<strong>in</strong>-down<br />

164<br />

band for a magnetic field gre<strong>at</strong>er than 0.35 as illustr<strong>at</strong>ed <strong>in</strong> Fig. 5.62. However, as


the magnetic field <strong>in</strong>creases, the sp<strong>in</strong>-moment <strong>in</strong> Fig. 5.58 decreases for a magnetic<br />

field larger than 0.035 28 where the exchange-correl<strong>at</strong>ion potential of the sp<strong>in</strong>-down<br />

band may decrease very rapidly and thus the difference <strong>in</strong> the exchange-correl<strong>at</strong>ion<br />

potentials becomes very large as shown <strong>in</strong> Fig. 5.59. The same reason<strong>in</strong>g which is<br />

applied to analyze the results for a background density of 0.0025 can be applied to<br />

the results for a magnetic field larger than 0.035 for this high background density<br />

as well.<br />

165<br />

28 0.035 magnetic field corresponds to 0.032796 background density of the sp<strong>in</strong>-up band.


6. CONCLUSIONS<br />

Immersion energies are calcul<strong>at</strong>ed for H, Li, C, and excited systems of H and C<br />

us<strong>in</strong>g the Kohn-Sham equ<strong>at</strong>ions <strong>in</strong> the LDA scheme. The self-consistent solutions<br />

are verified and compared with the Friedel sum rule, Friedel oscill<strong>at</strong>ions, phase<br />

shifts, the sc<strong>at</strong>ter<strong>in</strong>g length, bound st<strong>at</strong>e energies, potentials, Janak’s theorem,<br />

and compressibilities.<br />

166<br />

In order to obta<strong>in</strong> the correct self-consistent solutions, numerical problems<br />

must be solved first. The numerical and theoretical errors <strong>in</strong> LDA due to the rapid<br />

density vari<strong>at</strong>ion <strong>in</strong> the vic<strong>in</strong>ity of an impurity are canceled out by subtract<strong>in</strong>g<br />

the reference energy of a free <strong>at</strong>om. The Friedel sum rule and Friedel oscill<strong>at</strong>ions<br />

play a crucial role <strong>in</strong> solv<strong>in</strong>g numerical problems dur<strong>in</strong>g the numerical iter<strong>at</strong>ion<br />

process. The long tail of the Friedel oscill<strong>at</strong>ions is strongly coupled to the <strong>in</strong>duced<br />

density near an impurity by the Coulomb <strong>in</strong>teraction. A mix<strong>in</strong>g r<strong>at</strong>io as a function<br />

of r is used to acceler<strong>at</strong>e the convergence by suppress<strong>in</strong>g this coupled <strong>in</strong>teraction<br />

<strong>in</strong> the mix<strong>in</strong>g step <strong>in</strong> the self-consistent calcul<strong>at</strong>ion. In the numerical calcul<strong>at</strong>ion,<br />

<strong>in</strong>f<strong>in</strong>ity must be replaced with a cut-off large number and numerical results must<br />

be corrected us<strong>in</strong>g the behavior of the Friedel oscill<strong>at</strong>ions.<br />

The background density <strong>at</strong> the immersion energy m<strong>in</strong>imum is 0.0024 for a<br />

non-sp<strong>in</strong> polarized system of H and 0.0033 for a non-sp<strong>in</strong> polarized system of<br />

C. The behavior of immersion energies as a function of the background density<br />

depends on the number of bound st<strong>at</strong>es as well as on the impurity. For a neutral<br />

impurity, which is forced to have the same number of electrons as the impurity<br />

charge, the immersion energies show a simple l<strong>in</strong>ear behavior and approach zero as<br />

the background density becomes zero. The immersion energies for excited systems<br />

are verified with Janak’s theorem.


The impurity is completely screened, which is consistent with the Friedel<br />

sum rule, regardless of the number of bound st<strong>at</strong>es or the background density. In<br />

the low energy sc<strong>at</strong>ter<strong>in</strong>g cases, one can predict the behavior of the phase shifts<br />

from the Friedel sum rule and the number of bound st<strong>at</strong>es. The phase shifts may,<br />

however, show a complic<strong>at</strong>ed behavior due to the narrow width of the conduction<br />

band <strong>at</strong> an extremely low background density. This behavior can be expressed <strong>in</strong><br />

terms of sc<strong>at</strong>ter<strong>in</strong>g lengths and compressibilities, which is verified with analytical<br />

theories.<br />

167<br />

At low background densities, the calcul<strong>at</strong>ions show th<strong>at</strong> the immersion en-<br />

ergy for the sp<strong>in</strong>-polarized system of a carbon impurity cont<strong>in</strong>ues to decrease as the<br />

background density decreases and does not have the immersion energy m<strong>in</strong>imum<br />

near the background density where the non-sp<strong>in</strong>-polarized system of a carbon im-<br />

purity has a m<strong>in</strong>imum. For these sp<strong>in</strong>-polarized systems, the exchange-correl<strong>at</strong>ion<br />

<strong>in</strong>teraction <strong>at</strong> low background densities becomes very important, caus<strong>in</strong>g a very<br />

strong tendency toward a sp<strong>in</strong>-polarized solution. It <strong>in</strong>duces a 2p bound st<strong>at</strong>e <strong>in</strong><br />

the sp<strong>in</strong>-up band only, and can yield a non-zero magnetiz<strong>at</strong>ion even without an<br />

external magnetic field. These phenomena lead to an unusual angular momentum<br />

st<strong>at</strong>e behavior <strong>in</strong> terms of the number of <strong>in</strong>duced electrons for excited systems of<br />

a carbon impurity <strong>at</strong> low background densities.<br />

With an external magnetic field, the strong Coulomb <strong>in</strong>teraction appears due<br />

to two different wave lengths of Friedel oscill<strong>at</strong>ions, for sp<strong>in</strong>-up and sp<strong>in</strong>-down<br />

bands. With this Coulomb <strong>in</strong>teraction, several <strong>in</strong>teractions such as the exchange-<br />

correl<strong>at</strong>ion <strong>in</strong>teraction and the change <strong>in</strong> the band width due to an external mag-<br />

netic field give rise to a complic<strong>at</strong>ed behavior of the <strong>in</strong>duced sp<strong>in</strong>-moment. For<br />

<strong>in</strong>stance, there is a rapid <strong>in</strong>crease of the sp<strong>in</strong>-moment as the band width of the<br />

sp<strong>in</strong>-down band goes to zero. This magnetic behavior of a sp<strong>in</strong>-polarized system


due to the exchange-correl<strong>at</strong>ion <strong>in</strong>teraction <strong>at</strong> low background densities is driven by<br />

the response of an electron gas <strong>in</strong> the conduction band and, as a result, has a long<br />

range effect compared to usual l<strong>at</strong>tice constants. These results, therefore, suggest<br />

th<strong>at</strong> the separ<strong>at</strong>ion distance between impurities is important, and the magnetic<br />

behavior of impurities may largely depend on the separ<strong>at</strong>ion distance.<br />

The model used <strong>in</strong> this work is spherically symmetric, the sp<strong>in</strong>-orbit <strong>in</strong>ter-<br />

action is ignored, and the self-<strong>in</strong>teraction correction is not implemented. A non-<br />

spherical model must be used for partially occupied angular momentum shells and,<br />

hence, may yield angular dependent results especially with an external magnetic<br />

field. The sp<strong>in</strong>-orbit <strong>in</strong>teraction also may lead to very different results s<strong>in</strong>ce non-<br />

zero angular momentum st<strong>at</strong>es can show the complic<strong>at</strong>ed behavior as expla<strong>in</strong>ed<br />

<strong>in</strong> Ch. 5. In the future, a non-spherical model should be studied with the sp<strong>in</strong>-<br />

orbit <strong>in</strong>teraction and the self-<strong>in</strong>teraction correction. Further, a non-local density<br />

scheme for the exchange-correl<strong>at</strong>ion <strong>in</strong>teraction should be used to <strong>in</strong>vestig<strong>at</strong>e the<br />

important dependance of our results on the exchange-correl<strong>at</strong>ion potential.<br />

168


BIBLIOGRAPHY<br />

[1] P. Hohenberg and W. Kohn, Phys. Rev. 136, B864 (1964)<br />

[2] W. Kohn and L.J. Sham, Phys. Rev. 140, A1133 (1965)<br />

[3] H.J.F. Jansen and A.J. Freeman, Phys. Rev. B 30, 561 (1984)<br />

[4] David J. S<strong>in</strong>gh, Planewaves, Pseudopotentials and the LAPW Method., KAP<br />

(1994)<br />

[5] M.J. Puska, R.M. Niem<strong>in</strong>en, M. Man<strong>in</strong>en, Phys. Rev. B 24, 3037 (1981)<br />

[6] E. Zarembar, L.M. Sander, H.B. Shore, and J.H. Shore, J. Phys. F7, 1763<br />

(1977)<br />

[7] J.D. Kress and A.E. DePristo, J. Chem. Phys. 87, 4700 (1987); 88, 2596<br />

(1988)<br />

[8] A. Sal<strong>in</strong>, A. Arnau, P.M. Echenique, and E. Zarembar, Phys. Rev. B 59, 2537<br />

(1999)<br />

[9] R. Díez Muiño and A. Sal<strong>in</strong>, Phys. Rev. B 60, 2074-83 (1966)<br />

[10] J. Friedel, Philos. Mag. 43, 153-89 (1952)<br />

[11] J. Friedel, Adv. Phys. 3, 446-507 (1953)<br />

[12] J. Friedel, Nuovo cimento 7, 287-311 (1958)<br />

[13] R.M. Dreizler and E.K.U. Gross, Density Functional Theory, Spr<strong>in</strong>ger-Verlag<br />

(1990)<br />

[14] S. Lundqvist and N.H. March, Theory of the Inhomogeneous <strong>Electron</strong> <strong>Gas</strong>,<br />

Plenum (1983)<br />

[15] C. Kittel, Quantum Theory of Solids, Wiley (1987)<br />

[16] John C. Inkson, Many-body theory of solids, Plenum (1984)<br />

[17] L. Hed<strong>in</strong> and B. I. Lundqvist, J. Phys. C:Solid St. Phys. 4, 2064-83 (1971)<br />

[18] Gerald D. Mahan, Many-Particle <strong>Physics</strong> 3rd ed., KA/PP, (2000)<br />

[19] U. von Barth and L. Hed<strong>in</strong>, J. Phys. C:Solid St. Phys. 5, 1629-42 (1972)<br />

[20] B. Y. Tong and L. J. Sham, Phys. Rev. 144, 1-4 (1966)<br />

[21] A. P. Albus, M.S. Thesis, <strong>Oregon</strong> St<strong>at</strong>e University (1999)<br />

169


[22] Henri J. F. Jansen, unpublished, <strong>Oregon</strong> St<strong>at</strong>e University<br />

[23] J. J. Sakurai, Modern Quantum Mechanics., Addison-Wesley (1994)<br />

[24] J. D. Jackson, Classical Electrodynamics., Wiley (1975)<br />

[25] Neil W. Ashcroft and N. David Merm<strong>in</strong>, Solid St<strong>at</strong>e <strong>Physics</strong>., Saunders College<br />

Publish<strong>in</strong>g (1976)<br />

[26] A. C. Hewson, The Kondo problem to heavy fermions., Cambridge (1992)<br />

[27] W. Kohn and C. Majumdar, Phys. Rev. 138, A1617-A1620 (1965)<br />

[28] Stephen <strong>Gas</strong>iorowicz, Quantum <strong>Physics</strong>., Wiley (1974)<br />

[29] Amit Goswami, Quantum Mechanics., Wm. C. Brown (1997)<br />

[30] R. H. Landau, M. J. Paes, Comput<strong>at</strong>ional <strong>Physics</strong>., Wiley-Interscience (1997)<br />

[31] R. H. Landau, Quantum Mechanics II., Wiley-Interscience (1996)<br />

[32] M.J. Puska and R.M. Niem<strong>in</strong>en, Phys. Rev. B 27, 6121 (1983)<br />

170<br />

[33] E. Babic, R. Krsnik, B. Leontic, M. Ocko, Z. Vucic, I. Zoric, and E. Girst,<br />

Solid St<strong>at</strong>e Commun. 10, 691 (1972); G. Bo<strong>at</strong>o, M. Bugo, and C. Rizzuto,<br />

Nuovo Cimento 45, 226 (1966); Y. Fukai, Phys Rev. 186, 697 (1967)<br />

[34] William H. Press, Saul A. Teukolsky, William T. Vetterl<strong>in</strong>g, Brian P. Flannery,<br />

Numerical Recipes <strong>in</strong> C, CAMBRIDGE UNIVERSITY PRESS (1997)<br />

[35] A. Bechler and R. H. Pr<strong>at</strong>t, Phys. Rev. A 28, 1190 (1983)<br />

[36] A. Bechler and R. H. Pr<strong>at</strong>t, J. Phys. A: M<strong>at</strong>h. Gen. 20, 133-141 (1987)


APPENDIX<br />

171


APPENDIX A. Atomic Rydberg units<br />

energy:<br />

Atomic Rydberg units are def<strong>in</strong>ed by = 2me = e 2 /2 = 1.<br />

The unit for length is the Bohr radius:<br />

a0 = 2<br />

mee 2 = 1 0.529 · 10−8 cm .<br />

172<br />

The unit for time is the r<strong>at</strong>io of an angular momentum and the Rydberg<br />

t0 = <br />

ER<br />

= 23<br />

mee 4 = 1 4.84 · 10−17 s .<br />

The unit for energy is the Rydberg energy:<br />

ER = e2<br />

2a0<br />

The unit for charge is:<br />

= mee 4<br />

The unit for charge density is:<br />

= 1 Rydberg 13.6 eV .<br />

22 q0 = 1 = e<br />

√ 2 1.1329105 · 10 −19 C .<br />

ρ0 = 1 = 7.6452571 · 10 11 C/m 3 .<br />

Other units for important physical quantities are as follows.<br />

The unit for electric field:<br />

The unit for current:<br />

E0 = 1 3.6360903 · 10 11 V/m .<br />

I0 = 1 2.3418037 mA .


The unit for conductivity:<br />

The unit for resistance:<br />

The unit for magnetic <strong>in</strong>duction:<br />

The unit for magnetiz<strong>at</strong>ion:<br />

σ0 = 1 2.2999241 · 10 6 S/m .<br />

R0 = 1 8.2164712 kΩ .<br />

B0 = 1 3.3241346 · 10 5 T .<br />

M0 = 1 4.4253673 · 10 7 A/m .<br />

173


APPENDIX B. Comparison with published d<strong>at</strong>a<br />

This section is dedic<strong>at</strong>ed to focus on the ma<strong>in</strong> difference between this thesis<br />

and Puska, Niem<strong>in</strong>en, and Man<strong>in</strong>en’s work( [5]).<br />

B.1. Numerical aspects<br />

174<br />

In Puska, Niem<strong>in</strong>en, and Man<strong>in</strong>en’s work, only 10 partial waves were used.<br />

The Pr<strong>at</strong>t improvement scheme(Broyden’s method) is used for the mix<strong>in</strong>g r<strong>at</strong>io.<br />

They used a first order approxim<strong>at</strong>ion of Taylor series to remove the dependence<br />

of the cut-off radius <strong>in</strong> the calcul<strong>at</strong>ion of the exchange-correl<strong>at</strong>ion energy.<br />

In this thesis, a cut-off value of 10 −8 for the phase shifts is used to determ<strong>in</strong>e<br />

the number of partial waves for each k. Hence, the number of partial waves is<br />

variable. Exponential-like functions are used for the mix<strong>in</strong>g r<strong>at</strong>io s<strong>in</strong>ce discont<strong>in</strong>u-<br />

ities <strong>in</strong> the potential <strong>in</strong> r-space are found <strong>in</strong> Broyden’s method. The <strong>in</strong>form<strong>at</strong>ion<br />

of the last two <strong>in</strong>put potentials is utilized <strong>in</strong> the mix<strong>in</strong>g scheme to consider the<br />

response of iter<strong>at</strong>ion process. The cut-off radius dependance of the calcul<strong>at</strong>ions is<br />

removed by consider<strong>in</strong>g the oscill<strong>at</strong>ion of <strong>in</strong>tegr<strong>at</strong>ed values, which is believed to be<br />

accur<strong>at</strong>e due to the regular behavior of the Friedel oscill<strong>at</strong>ions <strong>at</strong> large distances.<br />

The output potentials also must be corrected after each iter<strong>at</strong>ion of the numerical<br />

calcul<strong>at</strong>ions. This is performed <strong>in</strong> such a way th<strong>at</strong> the system s<strong>at</strong>isfies the Friedel<br />

sum rule and correct Friedel oscill<strong>at</strong>ions. It is also po<strong>in</strong>ted out th<strong>at</strong> the system<strong>at</strong>ic<br />

numerical error of core st<strong>at</strong>es of an impurity <strong>in</strong> an electronic gas can be canceled<br />

out by th<strong>at</strong> of the reference of a free <strong>at</strong>om energy. In order to obta<strong>in</strong> this cance-<br />

l<strong>at</strong>ion, one must use the same r-mesh for both calcul<strong>at</strong>ions. Detailed <strong>in</strong>form<strong>at</strong>ion<br />

can be found <strong>in</strong> Ch. 4.


B.2. Results<br />

175<br />

The immersion energies obta<strong>in</strong>ed <strong>in</strong> this thesis are slightly lower than those<br />

<strong>in</strong> Puska, Niem<strong>in</strong>en, and Man<strong>in</strong>en’s work. For <strong>in</strong>stance, the background density is<br />

about 0.025 <strong>at</strong> zero immersion energy for H while it is 0.02 <strong>in</strong> Puska, Niem<strong>in</strong>en, and<br />

Man<strong>in</strong>en’s work. However, the result for the background density <strong>at</strong> the immersion<br />

energy m<strong>in</strong>imum is very close to th<strong>at</strong> of Puska, Niem<strong>in</strong>en, and Man<strong>in</strong>en’s work.<br />

They showed 0.0026 for H and 0.0035 for C while, <strong>in</strong> this work, 0.0024 for H and<br />

0.0033 for C are obta<strong>in</strong>ed. While they focused on the partial-wave decompositions<br />

of the density of <strong>in</strong>duced st<strong>at</strong>es, the results for the partial-wave decompositions of<br />

the Friedel sum rule 1 are obta<strong>in</strong>ed and used to analyze the behavior of the poten-<br />

tials especially for sp<strong>in</strong>-polarized systems. The follow<strong>in</strong>g new results are obta<strong>in</strong>ed.<br />

Dielectric functions and compressibilities for H are calcul<strong>at</strong>ed and compared with<br />

the theory <strong>in</strong> this work. Immersion energies for excited systems are also calcul<strong>at</strong>ed<br />

and verified with the Janak’s theorem, and the results for sp<strong>in</strong>-polarized systems<br />

and sp<strong>in</strong>-coupl<strong>in</strong>g with an external magnetic field are also newly obta<strong>in</strong>ed <strong>in</strong> this<br />

work, which have not been performed <strong>in</strong> the older work.<br />

1 This k<strong>in</strong>d of discussion also can be found <strong>in</strong> [32].


APPENDIX C. Boundary conditions for non-spherical potentials <strong>at</strong><br />

small r<br />

where<br />

The general idea is taken from [8]. The equ<strong>at</strong>ions to be solved are<br />

d 2<br />

dr 2 + ɛn −<br />

V m<br />

<br />

ll ′ (r) =<br />

l(l + 1)<br />

r2 <br />

unlm(r) = <br />

l ′<br />

176<br />

V m<br />

ll ′ (r)unl ′ m(r) , (C1)<br />

dΩ Y ∗<br />

lm(Ωr)Veff(r, θ)Yl ′ m(Ωr) . (C2)<br />

The coupl<strong>in</strong>g terms must be exam<strong>in</strong>ed <strong>in</strong> order to f<strong>in</strong>d the boundary conditions of<br />

a set of l<strong>in</strong>early <strong>in</strong>dependent solutions of the coupled equ<strong>at</strong>ions <strong>at</strong> small r. The<br />

potential can be expanded due to its cyl<strong>in</strong>drical symmetry as<br />

V (r) = <br />

Vi(r)Pi(cos θ) , (C3)<br />

i<br />

where Pi(cos θ) is a Legendre polynomial. The behavior of the components Vi(r)<br />

of the potential <strong>at</strong> small r can be shown, which is<br />

lim<br />

r→0 Vi(r) ∼<br />

Us<strong>in</strong>g the expansion for V (r),<br />

V m<br />

<br />

ll ′ (r) =<br />

Note th<strong>at</strong><br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

r i if i > 0 ,<br />

− Z<br />

r<br />

if i = 0 .<br />

(C4)<br />

dΩ [Y ∗<br />

lm(Ωr)V0(r)P0(cos θ)Yl ′ m(Ωr) + Y ∗<br />

lm(Ωr)V1(r)P1(cos θ)Yl ′ m(Ωr)<br />

<br />

+Y ∗<br />

lm(Ωr)V2(r)P2(cos θ)Yl ′ m(Ωr) + · · · ] . (C5)<br />

dΩ Y ∗<br />

lm(Ωr)Pv(cos θ)Yl ′ m(Ωr) = βδ(l−v)+2i,l ′ , (C6)<br />

where i = 0, 1, 2, · · · until l ′ ≤ (l + v).


177<br />

Now, us<strong>in</strong>g the form of Vi(r) <strong>at</strong> small r <strong>in</strong> (C4), each term <strong>in</strong> (C5) contributes<br />

to V m<br />

ll ′ (r) <strong>at</strong> small r as<br />

<br />

m 1<br />

lim Vll ′ (r) = α0 δl,l ′ + α1r<br />

r→0 r<br />

+ α2r 2<br />

<br />

Y ∗<br />

lm(Ωr)P1(cos θ)Yl ′ m(Ωr)<br />

Y ∗<br />

lm(Ωr)P2(cos θ)Yl ′ m(Ωr) + · · · (C7)<br />

1<br />

= β0<br />

r δl,l ′ + β1δl±1,l ′r + β2δl±2,l ′r2 + · · · , (C8)<br />

where only the lowest order terms of r are kept for each coupl<strong>in</strong>g term l ′ and βi<br />

simply represents constants produced <strong>in</strong> <strong>in</strong>tegrals and coefficients <strong>in</strong> Vi(r). For<br />

spherically symmetric potential, the form of ulm <strong>at</strong> small r is<br />

lim<br />

r→0 ulm(r) = Ar l+1<br />

Us<strong>in</strong>g this form, the coupl<strong>in</strong>g terms become <strong>at</strong> small r<br />

<br />

lim V<br />

r→0<br />

m<br />

ll ′ (r)ul′ l ′ 1<br />

m(r) = β0<br />

l ′<br />

r rl+1 + β1r l+3 + β1r l+1<br />

One can see th<strong>at</strong> the coupl<strong>in</strong>g terms follow the form of<br />

(C9)<br />

+ β2r l+5 + β2r l+1 + β3r l+7 + β3r l+1 + · · · . (C10)<br />

r l+1 if j < l ,<br />

r 2j+l+1 if j ≥ l ,<br />

(C11)<br />

where j = 0, 1, 2, · · · , lmax. 2 S<strong>in</strong>ce there exist solutions follow<strong>in</strong>g these form of<br />

coupl<strong>in</strong>g terms, one can f<strong>in</strong>ally write power series form for l<strong>in</strong>early <strong>in</strong>dependent<br />

solutions of coupled equ<strong>at</strong>ions.<br />

2 lmax is the maximum l value used <strong>in</strong> the numerical calcul<strong>at</strong>ion to trunc<strong>at</strong>e the partial<br />

wave expansion <strong>in</strong> (3.3).


u j<br />

lm (r) = rj+|l−j|+1<br />

178<br />

∞<br />

b l ir i−1 . (C12)<br />

The lowest order terms r l+1 s<strong>at</strong>isfy autom<strong>at</strong>ically the boundary conditions <strong>at</strong> small<br />

r. For the higher order terms than r l+1 , there are lower order terms to be coupled<br />

<strong>in</strong> (C1).<br />

i=1


APPENDIX D. S<strong>in</strong>gular value decomposition(SVD)<br />

179<br />

The s<strong>in</strong>gular value decomposition(or SVD) is very popular, used for many<br />

different areas such as image compression but essentially a technique to deal with<br />

m<strong>at</strong>rices which are s<strong>in</strong>gular or very close to s<strong>in</strong>gular and thus for which Gaussian<br />

elim<strong>in</strong><strong>at</strong>ion or other decomposition methods may fail.<br />

Any real m×n m<strong>at</strong>rix A can be decomposed uniquely with a m×m orthogonal<br />

m<strong>at</strong>rix U and an n × n orthogonal m<strong>at</strong>rix V such th<strong>at</strong><br />

A = U · D · V T , (D1)<br />

where D = diag(σ1, σ2, · · · , σn) and follows order of σ1 ≥ σ2 ≥ · · · ≥ σn. The<br />

scalars σ1, σ2, · · · , σn are called the s<strong>in</strong>gular values of A and columns of U and V<br />

are eigenvectors of AA T and A T A, respectively.<br />

Let’s consider a system of l<strong>in</strong>ear equ<strong>at</strong>ions.<br />

A · x = b . (D2)<br />

The nullspace is def<strong>in</strong>ed as a subspace of x such th<strong>at</strong> A · x = 0 if A is a s<strong>in</strong>gular.<br />

The range is def<strong>in</strong>ed as a subspace of b which can be obta<strong>in</strong>ed by mapp<strong>in</strong>g x with<br />

A. The dimension of the range is actually the rank of A. The dimension of the<br />

range is also can be obta<strong>in</strong>ed by subtract<strong>in</strong>g the dimension of the nullspace from<br />

m. If A is nons<strong>in</strong>gular, its range is all of the vector space b. If A is s<strong>in</strong>gular, the<br />

rank r is less than m and σr+1 = σr+2 = · · · = σn = 0. Suppose th<strong>at</strong> A is s<strong>in</strong>gular.<br />

If b = 0, any column of V which corresponds to zero σj becomes a solution of<br />

Ax = 0. A solution is <strong>in</strong> this case x = a1vr + a2vr+1 + · · · + an−rvn where vj is a<br />

j column of V and a 2 1 + a 2 2 + · · · + a 2 n−r = 1. If the vector b is not zero and lies <strong>in</strong><br />

the range of A, solutions can be found by<br />

x = V · [diag(1/σj)] · U T · b (D3)


with a condition of the smallest length |x| 2 which removes any redundant solutions<br />

s<strong>in</strong>ce any column of V with a zero σj(solutions correspond to the nullspace) can<br />

be added to x <strong>in</strong> any l<strong>in</strong>ear comb<strong>in</strong><strong>at</strong>ion. One important trick <strong>in</strong> (D3) is to equ<strong>at</strong>e<br />

1/σj to zero if σj = 0. The readers should refer to [34] for more details.<br />

• An example for A · x = 0 (calcul<strong>at</strong>ed <strong>in</strong> M<strong>at</strong>hem<strong>at</strong>ica):<br />

The m<strong>at</strong>rix A is given by<br />

⎛ ⎞<br />

1 3 6 10<br />

⎜ ⎟<br />

⎜ ⎟<br />

⎜ 2 4 5 3 ⎟<br />

A = ⎜ ⎟<br />

⎜ ⎟<br />

⎜ 2 6 12 20 ⎟<br />

⎝ ⎠<br />

4 8 10 6<br />

180<br />

. (D4)<br />

S<strong>in</strong>ce the last two rows of A are just twice the first two rows, the rank of A is 2.<br />

The m<strong>at</strong>rix A can be decomposed us<strong>in</strong>g SVD as mentioned before. U, D, and V<br />

are<br />

⎛<br />

⎞<br />

⎜<br />

−0.390888 0.217271 −0.729017 0.518202<br />

⎟<br />

⎜<br />

⎟<br />

⎜ −0.217271 −0.390888 0.518202 0.729017 ⎟<br />

U = ⎜<br />

⎟<br />

⎜ −0.781776 0.434542 0.364509 −0.259101<br />

⎟<br />

⎝<br />

⎠<br />

−0.434542 −0.781776 −0.259101 −0.364509<br />

⎛<br />

⎞<br />

⎜<br />

30.5886 0 0 0<br />

⎟<br />

⎜<br />

⎟<br />

⎜ 0 8.0212 0 0 ⎟<br />

D = ⎜<br />

⎟<br />

⎜ 0 0 0 0<br />

⎟<br />

⎝<br />

⎠<br />

0 0 0 0<br />

⎛<br />

⎞<br />

⎜<br />

−0.134925 −0.351883 0.695905 −0.611301<br />

⎟<br />

⎜<br />

⎟<br />

⎜ −0.333743 −0.568331 0.265892 0.703503 ⎟<br />

V = ⎜<br />

⎟<br />

⎜ −0.560942 −0.405685 −0.627282 −0.356765<br />

⎟<br />

⎝<br />

⎠<br />

−0.74549 0.623376 0.227011 0.064138<br />

, (D5)<br />

, and (D6)<br />

, (D7)


S<strong>in</strong>ce the rank of A is 2, σ3 = σ4 = 0 which correspond to the nullspace. The<br />

181<br />

calcul<strong>at</strong>ion can be confirmed as<br />

U · D · V T ⎛ ⎞<br />

⎜<br />

1 3 6 10<br />

⎟<br />

⎜ ⎟<br />

⎜ 2 4 5 3 ⎟<br />

= ⎜ ⎟ = A . (D8)<br />

⎜ ⎟<br />

⎜ 2 6 12 20 ⎟<br />

⎝ ⎠<br />

4 8 10 6<br />

Last two columns of V give solutions of A · x = 0, which can be confirmed as<br />

⎛ ⎞ ⎛<br />

⎜<br />

0.695905<br />

⎟ ⎜<br />

−1. × 10<br />

⎜ ⎟ ⎜<br />

⎜ 0.265892 ⎟ ⎜<br />

⎟ ⎜<br />

A · ⎜ ⎟ = ⎜<br />

⎜ −0.627282<br />

⎟ ⎜<br />

⎟ ⎜<br />

⎝ ⎠ ⎝<br />

0.227011<br />

−6<br />

1. × 10−6 −2. × 10−6 2. × 10−6 ⎞<br />

⎟<br />

⎠<br />

. (D9)<br />

⎛ ⎞ ⎛<br />

⎜<br />

−0.611301<br />

⎟ ⎜<br />

−2. × 10<br />

⎜ ⎟ ⎜<br />

⎜ 0.703503 ⎟ ⎜<br />

⎟ ⎜<br />

A · ⎜ ⎟ = ⎜<br />

⎜ ⎟ ⎜<br />

⎜ −0.356765 ⎟ ⎜<br />

⎝ ⎠ ⎝<br />

0.064138<br />

−6<br />

−1. × 10−6 −4. × 10−6 −2. × 10−6 ⎞<br />

⎟<br />

⎠<br />

. (D10)<br />

The right hand sides <strong>in</strong> the above equ<strong>at</strong>ions show very small values which are<br />

actually zero with<strong>in</strong> the numerical tolerance. Note th<strong>at</strong> for any column vj of V ,<br />

v T j · vj = 1.


APPENDIX E. Behavior of phase shifts for E → ∞<br />

FIGURE 6.1. Behavior of phase shifts for k → ∞.<br />

182<br />

The high energy behavior of phase shifts is tested numerically with a hydrogen<br />

impurity and a background density of 0.0025. A discussion of phase shifts <strong>in</strong><br />

Coulomb-like potentials can be found <strong>in</strong> several places. [31] [35] [36]<br />

For a potential of the type [36]<br />

V (r) = − ag(λr)<br />

r<br />

, (E1)<br />

approxim<strong>at</strong>e expressions for high energy phase shifts are, up to a first order,<br />

δl(k) δcl + a<br />

[ln(2k/λ)] , (E2)<br />

k


where δcl is the po<strong>in</strong>t-Coulomb phase shift, which is, as k is <strong>in</strong>creas<strong>in</strong>g, vanish<strong>in</strong>g<br />

much faster than the second term <strong>in</strong> (E2). The results <strong>in</strong> Fig. 6.1 are obta<strong>in</strong>ed<br />

with the assumption, λ = 1/Rm, which is correct for g(λr) = 1 − r/Rm. 3<br />

183<br />

One can see <strong>in</strong> Fig. 6.1 th<strong>at</strong> numerically obta<strong>in</strong>ed results well agree with the<br />

analytic form.<br />

3 The potential is zero for r ≥ Rm.


APPENDIX F. Sc<strong>at</strong>ter<strong>in</strong>g length and bound st<strong>at</strong>e energy<br />

FIGURE 6.2. Sc<strong>at</strong>ter<strong>in</strong>g length versus 1/ |EB|, where EB is the bound st<strong>at</strong>e<br />

energy rel<strong>at</strong>ive to the conduction band m<strong>in</strong>imum.<br />

184<br />

It is shown <strong>in</strong> 3.6 th<strong>at</strong> the sc<strong>at</strong>ter<strong>in</strong>g length a can be rel<strong>at</strong>ed to the bound<br />

st<strong>at</strong>e energy as <strong>in</strong><br />

a = r0 + 1<br />

kB<br />

. (F1)<br />

If the effective potential range r0 is sufficiently small, as <strong>in</strong> the screened cases, the<br />

sc<strong>at</strong>ter<strong>in</strong>g length can be estim<strong>at</strong>ed from the bound st<strong>at</strong>e energies us<strong>in</strong>g (F1).<br />

with (F1).<br />

Fig. 6.2 shows th<strong>at</strong> the numerical results calcul<strong>at</strong>ed <strong>in</strong> this work are consistent

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!