Local kinetic-energy density of the Airy gas

Similar documents
= N 2 = 3π2 n = k 3 F. The kinetic energy of the uniform system is given by: 4πk 2 dk h2 k 2 2m. (2π) 3 0

Testing the kinetic energy functional: Kinetic energy density as a density functional

IDEAL AND NON-IDEAL GASES

The Application of Density Functional Theory in Materials Science

Space charge build-up in XLPE-cable with temperature gradient

= atm. 760 mm Hg. = atm. d. 767 torr = 767 mm Hg. = 1.01 atm

Accuracy assessment of the scalar network analyzer using sliding termination techniques

Section 3: Crystal Binding

Q-Chem: Quantum Chemistry Software for Large Systems. Peter M.W. Gill. Q-Chem, Inc. Four Triangle Drive Export, PA 15632, USA. and

Accuracy of the coherent potential approximation for a onedimensional array with a Gaussian distribution of fluctuations in the on-site potential

Mechanics 1: Conservation of Energy and Momentum

Uniform electron gas from the Colle-Salvetti functional: Missing long-range correlations

Pore Radius and Permeability Prediction from Sonic Velocity

Jitter Measurements in Serial Data Signals

CLASSICAL CONCEPT REVIEW 8

5.61 Physical Chemistry 25 Helium Atom page 1 HELIUM ATOM

CHAPTER 9 ATOMIC STRUCTURE AND THE PERIODIC LAW

Chapter Test B. Chapter: Measurements and Calculations

Results of wake simulations at the Horns Rev I and Lillgrund wind farms using the modified Park model

Spatially separated excitons in 2D and 1D

Aalborg Universitet. CLIMA proceedings of the 12th REHVA World Congress Heiselberg, Per Kvols. Publication date: 2016

Free Electron Fermi Gas (Kittel Ch. 6)

On the first Townsend coefficient at high electric field

Differential Relations for Fluid Flow. Acceleration field of a fluid. The differential equation of mass conservation

Chemistry 13: States of Matter

Prelab Exercises: Hooke's Law and the Behavior of Springs

NMR and IR spectra & vibrational analysis

MATRIX ALGEBRA AND SYSTEMS OF EQUATIONS. + + x 2. x n. a 11 a 12 a 1n b 1 a 21 a 22 a 2n b 2 a 31 a 32 a 3n b 3. a m1 a m2 a mn b m

THE IDEAL GAS LAW AND KINETIC THEORY

EWEA CREYAP benchmark exercises: summary for offshore wind farm cases

Application of the work function to study the percentage composition of aluminum alloys

Gases and Kinetic-Molecular Theory: Chapter 12. Chapter Outline. Chapter Outline

State Newton's second law of motion for a particle, defining carefully each term used.

Measurement of Charge-to-Mass (e/m) Ratio for the Electron

AP CHEMISTRY 2009 SCORING GUIDELINES

Definition and Evaluation of Mean Beam Lengths for Applications in Multidimensional Radiative Heat Transfer: A Mathematically Self-Consistent Approach

Trends of the Periodic Table Diary

REVIEW QUESTIONS Chapter 8

3. Derive the partition function for the ideal monoatomic gas. Use Boltzmann statistics, and a quantum mechanical model for the gas.

A pure covalent bond is an equal sharing of shared electron pair(s) in a bond. A polar covalent bond is an unequal sharing.

Using light scattering method to find The surface tension of water

Prediction of airfoil performance at high Reynolds numbers

Wednesday 16 January 2013 Afternoon

The properties of an ideal Fermi gas are strongly determined by the Pauli principle. We shall consider the limit: µ >> k B T βµ >> 1,

HEAT UNIT 1.1 KINETIC THEORY OF GASES Introduction Postulates of Kinetic Theory of Gases

MATRIX ALGEBRA AND SYSTEMS OF EQUATIONS

Quantifying the Influence of Social Characteristics on Accident and Injuries Risk: A Comparative Study Between Motorcyclists and Car Drivers

Electrostatic Fields: Coulomb s Law & the Electric Field Intensity

Vector Spaces; the Space R n

Exam 4 Practice Problems false false

Study the following diagrams of the States of Matter. Label the names of the Changes of State between the different states.

- particle with kinetic energy E strikes a barrier with height U 0 > E and width L. - classically the particle cannot overcome the barrier

Ferromagnetic resonance imaging of Co films using magnetic resonance force microscopy

CHAPTER 26 ELECTROSTATIC ENERGY AND CAPACITORS

Orbits of the Lennard-Jones Potential

Copyright 2011 Casa Software Ltd. Centre of Mass

2. Room temperature: C. Kelvin. 2. Room temperature:

Duration and Bond Price Volatility: Some Further Results

Name: Section Registered In:

States of Matter CHAPTER 10 REVIEW SECTION 1. Name Date Class. Answer the following questions in the space provided.

Electrochemical Kinetics ( Ref. :Bard and Faulkner, Oldham and Myland, Liebhafsky and Cairns) R f = k f * C A (2) R b = k b * C B (3)

Name Class Date. In the space provided, write the letter of the term or phrase that best completes each statement or best answers each question.

Physics 9e/Cutnell. correlated to the. College Board AP Physics 1 Course Objectives

8.2 Elastic Strain Energy

Temperature Measure of KE At the same temperature, heavier molecules have less speed Absolute Zero -273 o C 0 K

Numerical analysis of Bose Einstein condensation in a three-dimensional harmonic oscillator potential

Numerical Solution of Differential Equations

Aalborg Universitet. Publication date: Document Version Også kaldet Forlagets PDF. Link to publication from Aalborg University

PRACTICAL IMPLICATIONS OF LOCATION-BASED SCHEDULING

Vacuum Technology. Kinetic Theory of Gas. Dr. Philip D. Rack

The long-range WindScanner system how to synchronously intersect multiple laser beams

Reflection and Refraction

The Viscosity of Fluids

CHEMISTRY STANDARDS BASED RUBRIC ATOMIC STRUCTURE AND BONDING

arxiv:cond-mat/ v1 [cond-mat.soft] 25 Aug 2003

Chapter 18 Temperature, Heat, and the First Law of Thermodynamics. Problems: 8, 11, 13, 17, 21, 27, 29, 37, 39, 41, 47, 51, 57

Why? Intermolecular Forces. Intermolecular Forces. Chapter 12 IM Forces and Liquids. Covalent Bonding Forces for Comparison of Magnitude

5 Numerical Differentiation

Lecture 3 Fluid Dynamics and Balance Equa6ons for Reac6ng Flows

MATH10212 Linear Algebra. Systems of Linear Equations. Definition. An n-dimensional vector is a row or a column of n numbers (or letters): a 1.

VALIDATION, MODELING, AND SCALE-UP OF CHEMICAL LOOPING COMBUSTION WITH OXYGEN UNCOUPLING

h 2 m e (e 2 /4πɛ 0 ).

7. DYNAMIC LIGHT SCATTERING 7.1 First order temporal autocorrelation function.

Lecture 3: Optical Properties of Bulk and Nano. 5 nm

THE KINETIC THEORY OF GASES

modeling Network Traffic

arxiv:nucl-th/ v2 16 Oct 2006

GCSE COMBINED SCIENCE: TRILOGY

M.Sc., Ph.D., Postdoc(USA) Professor Theoretical Physics et.in

Untitled Document. 1. Which of the following best describes an atom? 4. Which statement best describes the density of an atom s nucleus?

Online Changing States of Matter Lab Solids What is a Solid? 1. How are solids different then a gas or a liquid?

Kinetic Theory of Gases. Chapter 33. Kinetic Theory of Gases

Active learners in sustainable electronics and it

Forensic Science Standards and Benchmarks

Candidate Number. General Certificate of Education Advanced Level Examination June 2010

6 J - vector electric current density (A/m2 )

atm = 760 torr = 760 mm Hg = kpa = psi. = atm. = atm. = 107 kpa 760 torr 1 atm 760 mm Hg = 790.

How To Understand Enzyme Kinetics

Chem 1A Exam 2 Review Problems

Review D: Potential Energy and the Conservation of Mechanical Energy

Transcription:

Downloaded from orbit.dtu.dk on: Sep 18, 216 Local kinetic-energy density of the Airy gas Vitos, Levente; Johansson, B.; Kollár, J.; Skriver, Hans Lomholt Published in: Physical Review A (Atomic, Molecular and Optical Physics) DOI: 1.113/PhysRevA.61.52511 Publication date: 2 Document Version Publisher's PDF, also known as Version of record Link to publication Citation (APA): Vitos, L., Johansson, B., Kollár, J., & Skriver, H. L. (2). Local kinetic-energy density of the Airy gas. Physical Review A (Atomic, Molecular and Optical Physics), 61(5), 52511. DOI: 1.113/PhysRevA.61.52511 General rights Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. Users may download and print one copy of any publication from the public portal for the purpose of private study or research. You may not further distribute the material or use it for any profit-making activity or commercial gain You may freely distribute the URL identifying the publication in the public portal? If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Local kinetic-energy density of the Airy gas L. Vitos and B. Johansson Condensed Matter Theory Group, Physics Department, Uppsala University, S-75121 Uppsala, Sweden J. Kollár Research Institute for Solid State Physics, H-1525 Budapest, P.O. Box 49, Hungary H. L. Skriver Center for Atomic-scale Materials Physics and Department of Physics, Technical University of Denmark, DK-28 Lyngby, Denmark Received 3 November 1999; published 13 April 2 The Airy gas model is used to derive an expression for the local kinetic energy in the linear potential approximation. The expression contains an explicit Laplacian term 2 5 ( /2m) (r) that, according to jellium surface calculations, must be a universal feature of any accurate local description. Applied to the noble gases the expression reduces the errors by a factor of 5 over previous results obtained by the linear potential approximation. PACS numbers: 31.15.Ew, 71.15.Mb PHYSICAL REVIEW A, VOLUME 61, 52511 During the last two decades density-functional theory 1 has played a decisive role in the calculation of physical properties of atoms, molecules, and solids. In this context it is important to note that all available ground-state energy density functionals lead to deficient results on account of an inappropriate representation of the kinetic-energy contribution. This deficiency may, of course, be overcome by applying the approach of Kohn and Sham 2 as it is in most density-functional calculations. The appoach, however, is based on an implicit functional that requires a complete selfconsistent solution to the Kohn-Sham equations. As a result, the determination of the kinetic-energy contribution becomes the time-limiting step in density-functional calculations. An explicit kinetic-energy functional is therefore of crucial importance in future applications of density functional theory. Furthermore, the knowledge of a kinetic energy functional for the noninteracting particles has become one of the primary aims within density-functional theory as formulated within the metageneralized gradient approximation to the exchange-correlation density functional 3. Starting from the homogeneous electron gas including, e.g., gradient corrections, there have been many attempts to develop explicit kinetic-energy functionals 4. In general these functionals have been developed and tested for use in atomic calculations, and they have proven to be inadequate in the treatment of, for instance, surfaces of solids 5. Of particular relevance for the present purpose is the explicit kinetic-energy functional derived in the pioneering work by Baltin 6 on the basis of the Wigner-Kirkwood approach 4 and the linear potential approximation. It has the form t lin s rt TF s rf nr1/2, 1 nr/3 where t s TF is the Thomas-Fermi kinetic-energy density, and the function f of the scaled gradient is tabulated in Ref. 6. The functional was subsequently applied in calculations of the kinetic energy of the noble gases, and it was found to overestimate the Hartree-Fock results by as much as a factor of 2 7. As a result of this failure further developments of the approach appear to have been abandoned. Recently, Kohn and Mattson 8 put forward the concept of the edge electron gas as an appropriate starting point for the treatment of systems with edge regions. The simplest realization of the edge electron gas is the Airy gas model, and in the present paper we use this model to derive an explicit expression for the kinetic-energy density. Since the Airy gas model is based on the linear potential approximation, the resulting expression is equal to that obtained by Baltin 6 except for the fact that we isolate an explicit Laplacian term of the form 2 2m 2 nr, where 5, 2 which we exclude when the functional is inverted to obtain an explicit expression for the kinetic-energy density in terms of the scaled gradient. That is, we write the kinetic-energy density in the form of Eq. 1, but with a different f, plus Laplacian 2. Since the contribution from the Laplacian to the total kinetic energy of a confined system must vanish the present functional yields kinetic energies which differ considerably from those obtained by a functional in which also the Laplacian is expanded in terms of the scaled gradient. As a result, the error in the kinetic energy of the rare gases is reduced in the present approach by approximately a factor of 5 relative to previous calculations by the linear potential approximation 7. The starting point for the derivation of the Airy gas kinetic-energy functional is the potential v eff z Fz for zl for Lz, 3 15-2947/2/615/525114/$15. 61 52511-1 2 The American Physical Society

VITOS, JOHANSSON, KOLLAR, AND SKRIVER PHYSICAL REVIEW A 61 52511 which is linear in z, independent of x and y, and has a hard wall at L far from the electronic edge at z. The slope of the effective potential FdV eff /dz leads to a characteristic length scale l 2mF 2 1/3, 4 and the electron and kinetic-energy densities are then given by nzl 3 n, t s z 2 2m l5 t, 5 6 FIG. 1. The scaling function P(s) of Eq. 14 obtained for the Airy gas and the second-order gradient expansions. where z/l, t s ( /2m) occ () 2, and are the Kohn- Sham orbitals. Further, n 1 Ai 2 d, 7 t s 1 2 2 d d 2d Ai 1 4 Ai 2 d. The properties of the Airy function and its derivatives may be used to transform the kinetic energy density function 8 into t s 3 5 n 5 n, which upon application of the scalings of the ith derivative 8 9 n (i) zl i3 n (i), i,1,2,..., 1 characteristic of the Airy gas model, leads to the local Airy gas LAG expression for the kinetic-energy density of the real electron gas: t s LAG z 3 5 nz 2 2m z l 3 5 2m nz. 11 According to Eq. 3 and 4, the term in the square bracket on the right-hand side of Eq. 11 is equal to v eff (z), and the kinetic-energy density then reads t s LAG z 3 5 nzv effz 5 2m nz. 12 Alternatively, one may insert the Thomas-Fermi kineticenergy density t s TF z 3 5 2m 32 /3 n 5/3 z into the first term of Eq. 11 to obtain 13 t s LAG zt s TF zp sz 5 where the scaled gradient 2m nz, 14 nz snz 23 1/3 nz n 4/3 23 1/3 n s, 4/3 15 on account of scaling relations 1, is conserved when going from the real electron gas to the Airy gas. Finally, we have used the properties of the Airy function to derive an explicit expression for z/l 3 in terms of the density and its derivative. We thereby find the expression t s z 3 5 2m 5 nz 3nznz2nznz nz 4nznz 3nznz 2m nz. 16 for the kinetic-energy density of the Airy gas which avoids the numerical inversion of the scaling function. However, this explicit expression is only useful in cases where the real potential is close to the linear form assumed in the Airy gas model. Note that, for an Airy gas, the three expressions 12, 14, and 16 are, of course, equivalent. However, in the application to real systems Eq. 12 is the most accurate and Eq. 16 the least accurate of the three forms. The scaling function P(s) which appears in Eq. 14 may be calculated in the Airy gas model from P 3 /3 n/3 17 and s() defined in Eq. 15. In Fig. 1 we have plotted P(),s() and this function provides the connection between the real system and the Airy gas through the local scaled gradient ssn(z) which is conserved in the transformation to the Airy gas model. The inversion procedure is the same as that used by Baltin 6, except that here it is applied only to the first term t s TF (z)p s(z) in the kinetic- 52511-2

LOCAL KINETIC-ENERGY DENSITY OF THE AIRY GAS PHYSICAL REVIEW A 61 52511 FIG. 2. The integrated local errors Eq. 18, of the kineticenergy densities for a self-consistent jellium surface (r s 3) as functions of the prefactor of the Laplacian term. In the figure GEA (2) refers to the second order gradient expansion functional t s a t TF ( /2m) 1 9 t W, where t W is the von Weizsäcker functional, LAGa refers to t s a 3 5 nv eff cf. Eq. 12, and LAGb refers to t s a (z)t s lin, (z), where t s lin, (z) reduces to Eq. 1 for. and to t s TF (z)p(s) of Eq. 14 for.4. In the inset the integrated local errors of LAGb for physically interesting density range r s 2 6 are shown. energy density expression while the Laplacian term is kept unchanged. Below we shall justify this approach. A comparison of the kinetic energy functionals available in the literature shows that in most cases a Laplacian term of the form of Eq. 2 with varying prefactors is included; see, e.g., Ref. 4. Here we do not refer to 1 2 ( /2m) n(r), which is the trivial difference between the two natural definitions of the local kinetic-energy density. In want of a theoretical principle by which a unique value of the prefactor may be found, we have used the jellium model of a metal surface 9 for which accurate numerical solutions may be found to determine. That is, we have calculated the local difference between the exact Kohn-Sham kinetic-energy density t s KS (z)( /2m) occ () 2 for a jellium surface of intermediate density r s 3., and that obtained by a number of kinetic-energy functionals including Laplacian terms with a range of prefactors, i.e., 1 s t s a z 2 2m nzt s KS zdz, 18 where t s a (z) is the Thomas-Fermi part of the actual kinetic-energy density, e.g., t s TF (z)p s(z), and s the total Kohn-Sham surface kinetic energy. The results of the jellium calculations shown in Fig. 2 point to.4 as a unique value, where the asymptotic FIG. 3. The kinetic-energy densities for a self-consistent jellium surface (r s 3) as a function of the distance from the surface in units of Bohr radius, a ) obtained from the Airy gas expression Eq. 12 compared to the exact Kohn-sham results. It is seen that the airy gas expression Eq. 12 is almost exact for x3 and x. In the inset are shown errors in percent for the zeroth Thomas-Fermi, the second-order gradient expansion, and the low gradient limit of Eq. 14. linear behavior of all three functionals included in the figure extrapolate to zero error. It is further seen that at.4 the Airy gas functional 12 provides the best overall description of the kinetic-energy density while Baltin s functional, i.e.,, leads to a considerable error in the local description. Since.4 is also the value found in the real-space derivation of the Airy gas functional we suggest that a Laplacian with the prefactor 2 5 must be a universal feature of any local kinetic energy functional. Note that there is no a priori reason why the kinetic-energy density calculated from the Kohn-Sham orbitals for a jellium surface should correspond exactly to the prefactor 2 5 for the Laplacian term. The accuracy of the Airy gas functional 12 is demonstrated in Fig. 3, where we show the kinetic-energy density TABLE I. The difference in percent in the atomic kinetic energies evaluated from the self-consistent LDA 11 Kohn-Sham densities relative to the exact Kohn-Sham results. GEA (),GEA (2), LAG, and LLP label results obtained using the Thomas-Fermi approximation, the second-order gradient expansion 4, the Local Airy gas approximation, 1 and the locally linear potential approximation 6,7. Atom GEA () GEA (2) LAG LLP Ne 8.6.8 2.3 68.3 Ar 6.8.4 1.2 53.8 Kr 5.1.2.1 38.9 Xe 3.3 1.6 1.4 31.9 52511-3

VITOS, JOHANSSON, KOLLAR, AND SKRIVER PHYSICAL REVIEW A 61 52511 of the jellium surface model calculated in four different approaches. The only significant deviation between the Airy gas functional 12, and the exact Kohn-Sham result is seen to occur just below the jellium edge where the effective potential has a large positive curvature which cannot be captured by the linear potential approximation. In contrast, the Thomas-Fermi and second-order gradient functionals have significant errors even far from the jellium edge. It is clear that an accurate local description also leads to an accurate total kinetic energy. In fact, the Airy gas functional and the second-order gradient expansion functional yield total kinetic energies for the jellium surface which, however, are of the same order of magnitude because the oscillations in the local gradient expansion lead to a cancellations of errors. Since the contribution from a Laplacian term on account of Green s theorem vanishes for a confined system, it follows that only the Thomas-Fermi part of the kinetic-energy density 14 is important in calculations of the total kinetic energy. Consequently, one should not expand the Laplacian of the density in terms of the scaled gradient because such an expansion will lead to a nonvanishing contribution to the total kinetic energy from the expanded Laplacian term. That this is in fact the case is clearly shown by the results of the atomic calculations for the noble gases presented in Table I. Here we find that the Airy gas functional 1 has the accuracy of the second-order gradient expansion, as in the case of the jellium surface calculations, in contrast to the previous calculations by the linear potential approximation 6,7 which show significant errors. We conclude, that the Airy gas model and the equivalent linear potential approximation may in fact form an appropriate and well-defined starting point for density-functional theories of the kinetic energy density of the inhomogeneous electron gas. We gratefully acknowledge interesting and fruitful discussions with Professor Walter Kohn. Thanks are also due to Doctor Karsten Jacobsen and Doctor Andrei Ruban for their valuable comments. L.V. and B.J. are grateful to the Swedish Natural Science Research Council and the Swedish Foundation for Strategic Research for financial support. Part of this work was supported by the research Project No. OTKA 2339 of the Hungarian Scientific Research Fund. The Center for Atomic-scale Materials Physics is sponsored by the Danish National Research Foundation. 1 P. Hohenberg and W. Kohn, Phys. Rev. B 136, 864 1964. 2 W. Kohn and L.J. Sham, Phys. Rev. A 14, 1133 1965. 3 J. P. Perdew, S. Kurth, A. Zupan, and P. Blaha, Phys. Rev. Lett. 82, 2544 1999. 4 R. M. Dreizler and E. K. U. Gross, in Density Functional Theory Springer-Verlag, Berlin, 199, and references therein. 5 L. Vitos, H. L. Skriver, and J. Kollár, Phys. Rev. B 57, 12611 1998. 6 R. Baltin, Z. Naturforsch. Teil A 27, 1176 1972. 7 S. K. Ghosh and L. C. Balbás, J. Chem. Phys. 83, 5778 1985. 8 W. Kohn and A. E. Mattsson, Phys. Rev. Lett. 81, 3487 1998. 9 N. D. Lang and W. Kohn, Phys. Rev. B 1, 4555 197. 1 In these calculations we used the Padé approximation P(s) (1a 1 s 2 a 3 s 6 )/(1a 2 s 2 a 3 s 4 ) for the scaling function. For a 1.8944, a 2.6511, and a 3.431 the relative error between the exact result and the parametrized form for s1 was less than.5%. 11 J. Perdew and Y. Wang, Phys. Rev. B 45, 132441992. 52511-4