a” ‘ r‘ t 3% " p" e... . Mud Hr: ’ '. L m! J' ,9;- .I' “.1. 1'”, {b.5135 MICHIGAN STA‘I'E Ur RSITY LIBRARIES minim!” ”i“Minimum 31 93 00904 6271 II This is to certify that the dissertation entitled Stucture and Dynamics of Low-Dimensional Systems presented by Weiqing Zhong has been accepted towards fulfillment of the requirements for Ph. D . degree in Physics W/ [ Major professor e aqn (/31 (A? (993 Professor David Tomanek Dat MSU 13 an Affirmative AcrinnVEquul Opportunity Institution 0‘ 12771 LIBRARY Michigan State University PLACE lN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or betore date due. DATE DUE DATE DUE DATE DUE __ Ii I:JL___i E] . ll II I i — L hi i MSU Is An Aflirmative Action/Equal Opportunity Institution emana-DJ STRUCTURE AND DYNAMICS OF LOW-DIMENSIONAL SYSTEMS By WEIQING ZHON G A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the Degree of DOCTOR OF PHILOSOPHY Department of Physics and Astronomy 1993 ABSTRACT STRUCTURE AND DYNAMICS OF LOW-DIMENSION AL SYSTEMS by Weiqing Zhong I studied structural and dynamical properties of low dimensional systems using numerical approaches at different levels of sophistication. I focused my interest on the physical phenomena associated with the interaction of atoms with surfaces. I studied two topics which are intimately connected with this interaction. One of them is a procedure to quantitatively interpret atomic force microscopy (AFM) images of solid surfaces. The other topic is the effect of adsorbate on structural, electronic and dynamical properties of the substrate. When studying theory for the AFM, I considered a model system consisting of a Pd metal tip probing the graphite surface. Using ab initio density functional the— ory (DFT), I calculated the interaction between this Pd AFM tip and graphite and _. mapped it onto a parameterized energy functional. The calculation of AFM tip tra- jectories at different loads revealed that atomic resolution is only achievable for loads exceeding 5 x 10'9N. Larger loads, on the other hand, are likely to lead to surface destruction. Calculated variation of potential energy along a trajectory has been used subsequently to determine atomic scale friction. Assuming no wear and a complete dissipation of the energy gained by the Pd tip atoms, I found a very small friction coefficient p z 10" for loads near 10"8 N and an increase in p with increasing load in agreement with experiments. I selected H/Pd as the model system to study the effect of adsorbate on the structural and dynamical properties of the substrate. I based the description of the H- Pd system on density functional calculations, which were subsequently mapped onto a parametrized many—body alloy Hamiltonian. Using this Hamiltonian, I calculated the equilibrium structure of the clean and hydrogen covered Pd (001) and Pd (110) surfaces, as well as the corresponding surface phonon spectra. The most pronounced effect of hydrogen is a strong softening of the Rayleigh wave on Pd(001), which is indirectly related to “hydrogen embrittlement” observed in the bulk. I addressed this latter problem using molecular dynamics. I studied the equilibrium structure, elastic properties, and in particular the mechanical breakdown of bulk Pd under tensile stress, as a function of temperature and hydrogen concentration. My results indicate that the microscopic origin of “hydrogen embrittlement” is an increased ductility and plasticity in regions saturated by hydrogen, in agreement with the postulated Hydrogen Enhanced Local Plasticity mechanism. ACKNOWLEDGMENTS I would like to express deep gratitude to my advisor Prof. David Tomanek for his guidance, understanding, and encouragement during my five years of graduate study. I thank him for invaluable training in all aspects of research, including research methods, research manners, organization of ,work, and presentation of results. I will benefit from this training for the rest of my life. His inspiration and insight highly motivated my work and made this research enjoyable. I am grateful to Professor S. D. Mahanti, Professor P. Danielewicz, Professor R. Tobin, Professor W. Repko, Professor G. F. Bertsch, and Professor W. Pratt for their service on my Ph.D. guidance committee. I would express my special thanks to Professor S. D. Mahanti for his careful guidance and advice. I thank Professor J. S. Kovacs for valuable advice and J. King and S. Holland for constant help. I would like to thank Professor A. Bulgac, Professor G.F. Bertsch, Professor T. Ka- plan, Professor P. Duxbury, and Professor M. Thorpe for valuable suggestions and stimulating discussions. My thanks also go to my dear friends Dr. Y. Cai, Q. Yang, Dr. Y. S. Li, Dr. G. Overney, Dr. Z. Sun, N. J. Ju, Dr. Y. Wang, J. M. Song, B. Zhang, B. Lian, Dr. R. Y. Fan, N. Mousseau, J. D. Chen, X. H. Yu, and Q. F. Zhu for their friendship and help. I am grateful to my parents and brothers for their love, encouragement, and support during all my life. My wife Hongjun Yan deserves my deepest gratitude. I would also like to acknowledge financial support from Michigan State University, the Office of Naval Research, and the Center for Fundamental Materials Research. iii PUBLICATIONS W. Zhong and D. Tomanek, First-Principles Theory of Atomic-Scale Friction, Phys. Rev. Lett. 64, 3054 (1990). W. Zhong, G. Overney, and D. Tomanek, Theory of Atomic Force Microscopy on Elastic Surfaces, The Structure of Surfaces III, edited by S.Y. Tong, M.A. Van Hove, X. Xide and K. Takayanagi, Springer-Verlag, Berlin (1991), p243. W. Zhong, G. Overney, and D. Tomanek, Limits of Resolution in Atomic Force Microscopy Images of Graphite, Europhys. Lett. 15, 49 (1991). G. Overney, W. Zhong, and D. Tomanek, Theory of Elastic Tip—Surface Inter- actions in Atomic Force Microscopy, J. Vac. Sci. Technol. B9, 479 ( 1991). D. Tomanek, W. Zhong, and H. Thomas, Calculation of an Atomically Mod- ulated Friction Force in Atomic Force Microscopy, Europhys. Lett. 15, 887 (1991). D. Tomanek and W. Zhong, Palladium-Graphite Interaction Potentials Based on First-Principles Calculations, Phys. Rev. B43, 12623 (1991). W. Zhong, Y. S. Li, and D. Tomanek, E'fl'ect of Adsorbates on Surface Phonon Modes: H on Pd(001) and Pd(110), Phys. Rev. B44, 13053 (1991). G. Overney, D. Tomanek, W. Zhong, Z. Sun, H. Miyazaki, S. D. Mahanti, and H.-J. Giintherodt, Theory for the Atomic Force Microscopy of Layered Elastic Surfaces, J. Phys. : Cond. Mat. 4, 4233 (1992). W. Zhong, Y. Cai, and D. Tomanek, Mechanical Stability of Pd—H Systems: A Molecular Dynamics Study, Phys. Rev. B 46 8099 (1992). iv 10. 11. 12. 13. 14. 15. W. Zhong, G. Overney, and D. Tomanek, Structural Properties of Fe Crystals, Phys. Rev. B 47 95 (1993). G. Overney, W. Zhong, and D. Tomanek, Structural Rigidity and Low Frequency Vibrational Modes of Long Carbon Tubules, Z. Phys. D 27 93 (1993). W. Zhong, D. Tomanek, and George F. Bertsch Total Energy Calculation for Extremely Large Clusters: The Recursive Approach, Solid State Comm. 86 607 (1993). W. Zhong, Y. Cai, and D. Tomanek, Computer Simulation of Hydrogen Em- brittlement in Metals, Nature 362 435 (1993). G. Overney, W. Zhong, and D. Tomanek, Structural Optimization of Stage—I Alkali-Metal-Graphite Intercalation Compounds Using Density Functional The- ory, in preparation. D. Tomanek, W. Zhong and E. Krastev Stability of Multi-Shell Fullerenes, sub- mitted for publication. Contents LIST OF TABLES iv LIST OF FIGURES v 1 Introduction 1 1.1 Theory for the Atomic Force Microscopy ................ 2 1.2 Structure and dynamics of H—Pd systems ................ 7 1.2.1 Effect of hydrogen on surface phonon modes of Pd ....... 7 1.2.2 Mechanical stability of Pd-H systems .............. 9 1.3 Structure of the Thesis .......................... 11 2 Theory 13 2.1 Ab initio Density Functional Formalism ................. 13 2.1.1 Density Functional Theory .................... 13 2.1.2 Local Density Approximation .................. 15 2.1.3 Computational details ...................... 17 2.2 Many-Body Alloy Hamiltonian ...................... 20 2.3 Lattice dynamics ............................. 23 vi 2.4 Molecular dynamics . . 3 Theory for the Atomic Force Microscopy 3.1 Calculation of the Pd-graphite interaction ............... 3.2 Limits of resolution in AFM images of gra—phite ............ 3.3 First-principles theory of atomic-scale friction ............. 3.4 Ideal friction machines 3.5 Conclusions ...... OOOOOOOOOOOOOOOOOOOOOOOOOO 4 Structure and Dynamics for H Loaded Pd 4.1 Application of the Many-Body Alloy Hamiltonian to the Pd—H system 4.1.1 Construction of the Many-Body Alloy Hamiltonian ...... 4.1.2 Structure and stability of clean and hydrogen covered Pd(001) and Pd(110) surfaces ....................... 4.2 Phonon structure in the bulk and at the surface of Pd—H systems i 4.3 Mechanical stability of Pd-H systems .................. 4.3.1 Thermal expansion and the melting transition in Pd ...... 4.3.2 Mechanical stability of H-free Pd under tensile stress ..... 4.3.3 Mechanical stability of H loaded Pd under tensile stress . . . . 4.3.4 Discussion . . . 4.4 Summary and conclusions ........................ LIST OF REFERENCES vii 24 29 31 38 48 55. 63 67 68 69 72 80 90 90 96 101 107 111 114 List of Tables 4.1 Interaction parameters used in the Many-Body Alloy Hamiltonian for the Pd-H system .............................. 69 4.2 Relaxations at clean and hydrogen covered Pd surfaces ......... 75 viii List of Figures 1.1 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9 Schematic picture of the Atomic Force Microscope (AFM) ...... 4 Adsorption energy of Pd on graphite as a function of the adsorption height ..... ‘ .............................. 32 Dependence of the Pd—graphite interaction energy on the charge den- sity at the adsorption site ........................ 35 Radial plot of the charge density of a carbon atom ........... 37 Calcualted adsorption energy of Pd on graphite as a function of the adsorption height ............................. 39 Force between a Pd AF M tip and a graphite surface .......... 42 Atomic-scale corrugation of the monatomic Pd AFM tip scanning graphite surface ................................... 43 Total charge density of an AF M tip of Pd interacting with the elastic graphite surface ................... ' ........... 46 Potential energy and force acting on an AF M tip of Pd along a trajec- tory on the graphite surface ....................... 49 Microscopic friction coefficient as a function of load .......... 52 3.10 Macroscopic friction force as a function of load applied to a large object 54 ix 3.11 Two models for the Friction Force Microscope ............. 3.12 Potential energy, friction force, and averaged friction force in the “max- imum friction microscope” ........................ 3.13 Microscopic friction mechanism in the “realistic friction microscope” . 3.14 Average friction force between a Pd tip and graphite, as a function of the load on the tip and the force constant of the cantilever ...... 4.1 Cohesive energy in bulk Pd and PdH as a function of lattice constant, based on the MBA Hamiltonian ..................... 4.2 Surface energy changes for clean and H-covered Pd (001) and Pd (110) surfaces as a function of first interlayer spacing ............. 4.3 Adsorption energy of H on the Pd (001) surface as-a function of the adsorption height ............................. 4.4 Adsorption energy of H on the Pd (110) surface as a function of the adsorption height ............................. 4.5 Phonon dispersion relations for bulk Pd and PdH ........... 57 59 60 64 70 74 77 78 81 4.6 Effect of adsorbed hydrogen on the surface phonon modes of the Pd(001) 84 4.7 Effect of adsorbed hydrogen on the surface phonon modes of the Pd(110) 85 4.8 Equilibrium Pd atomic volume in bulk Pd as a function of temperature, and same in bulk PdH as a function of hydrogen concentration . . . . 4.9 Atomic trajectories and the pair correlation function in bulk Pd at different temperatures .......................... 4.10 Distribution of atomic binding energies in bulk Pd ........... 91 95 4.11 Dependence of the deformation of Pd and PdH, on tensile stress . . . 97 4.12 Dependence of the bulk moduli of Pd and PdH; on temperature and hydrogen concentration a: ......................... 100 4.13 Distribution of binding energies of Pd atoms in bulk Pdes at T = 300 K at zero and critical tensile stress ................. 104 4.14 Distribution of atomic volumes in bulk PdHogs at T = 300 K. . . . . 105 xi Chapter 1 Introduction The major task of theoretical physics is to understand the laws which govern na- ture, use them to explain observed phenomena, and predict the properties of new physical systems. The subject of interest for theoretical condensed matter physics are electronic, magnetic, structural, and dynamical properties of solids and liquids. The most popular technique to address these properties from first principles is the Density functional theory [Lun 83] in the Local Density Approximation (LDA), which has been shown to describe ground state properties of solids with a high accuracy. Since this theory is free of adjustable parameters, I will use it as a basis for all the calculations in this Thesis. The large computational requirement associated with ab initio methods such as the LDA limits the range of its applicability to relatively small systems with a high symmetry. In order to describe the dynamical behavior of large systems with low symmetry, I calculate the interatomic forces using a parametrized Many-Body Alloy Hamiltonian [Zho 91a], which I developed as an efficient interpolation scheme between ab initio results. In spite of its simplicity, this Hamiltonian describes the interatomic interactions in simple metals and late transition metals with sufficient accuracy. The total energy calculations in this Thesis are performed using the combination of the LDA and MBA formalisms. The dynamical behavior is then calculated using the in- teratomic forces in the equations of motion. For a canonical ensemble, this procedure yields lattice dynamics at zero and finite temperatures, and the elastic response to externally applied loads. In this thesis, I have applied these calculation tools to study structural and dy- namical properties of low dimensional systems. I focused my interest on the physical phenomena associated with the interaction of atoms with surfaces. One important application of this interaction is the quantitative interpretation of atomic force mi- croscopy (AF M) images of solid surfaces. For the model system consisting of a Pd metal tip and graphite. I calculated the AFM images, the limits of nondestructive imaging, and the origin of atomic-scale friction. The second important topic is the effect of adsorption on the properties of the substrate. I considered Pd as a model sys- tem, and studied the effect of adsorbed hydrogen on the surface structure and phonon spectra. I found these results to be of significance when interpreting bulk phenom- ena associated with adsorbed hydrogen, such as the “hydrogen embrittlement. In the following Section 1.1 and 1.2, I will introduce these two topics, emphasizing the open questions and the general importance of these calculations. The structure Of my Thesis will be addressed in Section 1.3. 1.1 Theory for the Atomic Force Microscopy Following its invention by Binnig et al in 1986, the Atomic Force Microscope (AF M) [Bin 86, Bin 87] has been rapidly evolving into a powerful tool to examine the mor- phology and local rigidity of conducting and insulating surfaces alike [Tom 89]. The AFM uses an “atomically” sharp tip to scan the sample surface at a sample-to-tip separation of a few angstroms. Unlike the more established Scanning Tunneling Mi- croscope (STM) [Bin 82], which is sensitive to the electronic density of states near the Fermi energy [Ter 83], the AF M probes the force field Fm between an “atom- ically sharp” tip and the substrate (Figure 1.1). This force Fm is detected by the deflection of a soft cantilever which supports the tip. During a horizontal scan of the surface, the AF M measures and records the equilibrium vertical tip position, while Fe,“ is been kept constant by regulating the deflection of the cantilever. The reso- lution of atomic-scale features at the surface is possible if the equilibrium tip height 2 is detectably different at inequivalent (e.g. on-top, hollow) surface sites. This is the case if the corresponding corrugation £322,005 A, which is the sensitivity of the AFM under ideal conditiOns. Since the AFM provides real-space image, same as the STM, the ambiguity and complexity of interpreting diffraction patterns is avoided. The power of the AFM lies in its ability to detect local structure rather than periodic patterns. These can be point defects, steps, grain boundaries, and similar on conducting or non-conducting surfaces. The AF M has been used suCCessfully to image a variety of systems with atomic resolution. Very good results have been achieved in ionic system such as NaCl [Mey 90], and layered materials such as graphite [Bin 87]. While the experiments have primarily focused on achieving atomic resolution, two questions have remained unanswered so far. The first is, under which conditions atomic-scale features can be observed by a microscopic tip, and how these features relate to the surface topography. The second open question, which will be addressed in this Thesis, relates to the conditions, under which nondestructive imaging can be achieved. The major challenge for the theory is to determine whether it is possible to achieve atomic resolution using the AF M on a specific material, under well controlled and idealized conditions. The potential for atomic resolution is limited in two ways. E. Force 00 Q Will) 0000 00000 First, if the load applied to the tip, Fe", is very small, the tip-substrate separation is large, and the corrugation A2 is too small to be observed. For “dull” tips with multiple apex atoms, the corrugation is further decreased due to the tip/substrate incommensurability. At large values of Fe“, where the tip is capable of probing the corrugation of the charge density, the tip-surface distance is very small. Under these conditions, a destruction of the surface or the tip is likely. The objective of my calculations is to determine if there is an optimum force, which yields detectable corrugations on the atomic scale, but which is still non-destructive to the surface. The AFM has been used to investigate the friction on the atomic scale [Mat 87]. This experiment investigated the friction force experienced by the tip during a scan with no wear across a perfect surface. In general, friction forces are observed in nature whenever two bodies in contact are in relative motion [Lan 60]. The force F f is related to the applied load Fm between the two bodies by Ff = I‘chr (11) The friction coefficient a has been found to range typically between 10‘2 for very smooth interfaces and 1 for rough interfaces. Friction forces have dissipative character: they couple the macroscopic mechanical degrees of freedom to microscopic degrees of freedom. In this way, macroscopic mechanical energy is dissipated. At rough interfaces, dislocation motion and plastic deformations are mainly responsible for the dissipation of macroscopic energy. This fact makes the friction process with wear one of the most complex and least understood process in nature. Friction without wear, on the other hand, occurs between perfect and weakly interacting surfaces and is much easier to understand. Since in this case, no atomic rearrangement occurs, it is phonons and electronic excitations which serve as the only source for energy dissipation. Atomic-scale friction, which can be quantitatively studied using the AFM operating in a nondestructive manner, falls in this second category. [Sla 93] The idealized conditions for this atomic-scale friction open the problem to an inde- pendent fist-principles theoretical study [Zho 90]. In absence of plastic deformations, friction originates in an atomic-scale corrugation of the interaction potential which is probed by the tip when scanning the surface under an external force Fm. The two contributions to this potential are the variation of tip-substrate bond strength and the work against Fm if the tip-substrate distance varies along the trajectory. The maximum friction force can be estimated using the variations of the total potential energy of the tip-substrate system during the scan. A microscopic description of friction without wear must address the fact that friction is a non-conservative process. In other words, the friction force depends on the direction of motion between the two bodies in contact. A closed-loop integral over such a force yields a nonzero value which corresponds to the dissipated energy. In other words, the friction force cannot be obtained from a gradient of a potential. A quantitative study of friction requires the investigation of the microscopic mechanisms for energy dissipation and of the effectiveness with which macroscopic degrees of freedom are coupled to microscopic degrees of freedom. In order to study these process quantitatively, a Friction Force Microscope (FFM) has been constructed and used to measure the atomic-scale friction [Ove 92]. A very precise FFM can determine the horizontal and vertical force on the tip simultane- ously Since different portions of otherwise flat surface may have different friction coefficients, the FFM may provide the most detailed information and enhanced con- trast as compared to the AFM image of such a surface. Hence, friction imaging may prove to be a very useful application of the FFM [Ove 92]. 7 1.2 Structure and dynamics of H—Pd systems The interaction of hydrogen with transition metals is a fundamentally interesting topic with wide ranging technological applications [Ale 78]. The dissociative adsorption of molecular hydrogen at a metal surface and the subsequent surface and bulk diffusion of atomic hydrogen is a prototypical surface process. This is a well-suited model system for the study of the hydrogen-metal bond at different stages of the reaction. At least of equal importance is the microscopic understanding of the effect of hydrogen on the structural and dynamical properties of the substrate or host metal. This effect, which has received less attention in the literature, manifests itself on the surface as H-induced surface relaxation, surface reconstruction, as well as a change of the surface phonon spectra. Hydrogen atoms are known to diffuse into many metals easily. In the bulk, hydrogen can change the equilibrium structure and elastic properties of the host metal. Technologically, hydrogen atoms in the bulk of transition metals can also reduce the mechanical stability of the system significantly, which is known as hydrogen embrittlement. The particular interest in Pd is motivated to a large degree by the ability of this metal to form hydrides and thereby to act as a medium for hydrogen storage. In this Thesis, I will concentrate on studying the effect of adsorbed hydrogen on the surface phonon modes of Pd, and the microscopic origin of hydrogen embrittlement in Pd. I will show that both phenomena are intimately related. 1.2.1 Effect of hydrogen on surface phonon modes of Pd Extensive studies of interaction between adsorbed H and the Pd surface include the characterization of the H-metal bond, hydrogen induced surface relaxation and re- construction, and the effect of hydrogen on the the electronic structure of the metal substrate [Tom 86b, Sun 89, Tom 91a, Cat 83, He 88, Beh 80, Rie 83, Rie 84, Tar 86, N yb 83, Bes 87]. So far, most research effort has been focussed on the adsorption ge- ometry and on the electronic and structural properties of the adsorption system. Surface phonons have received far less attention in the recent literature, in spite of the wealth of information they contain about the nature of bonding at surfaces. This is caused mainly by the difficulty to measure and calculate reliable surface phonon dispersion curves throughout the whole surface Brillouin zone. Only recently, He time-of-fiight spectroscopy [Toe 87, Lab 87, Doa 83, Har 85] and electron-energy-loss spectroscopy(EELS)[Leh 83, Roc 84, Wut 86, Iba 87, Yos 88] have been used to measure the dispersion curves of surface phonons on a variety of metal substrates. The quantitative interpretation of these data is lacking in most cases, since predictive ab initio calculations (such as “frozen-phonon” calculations) are computationally very involved. Only in selected cases, Local Density Approxima- tion (LDA) [Koh 65] calculations have been performed for the high-symmetry modes [Ho 86]. The majority of published phonon calculations use lattice dynamics based on sim— ple two- and three-body potentials [Joh 72, Bor 84, Bor 85, Hal 88]. These types of calculations use bulk and surface interatomic force constants and distances as inde- pendent parameters which are chosen to fit experimental results [Lah 87]. In general, these calculations show a good agreement with the observed data. The predictive power is limited by the generally large number of force constant parameters which depend on the model, the system and the surface studied. While these calculations can provide a rough guidance in the interpretation of experimental results, direct comparisons between different models are of limited use. More recently, the Embedded—Atom -Method (EAM) [Daw 84, Foi 85, Foi 87, Daw 89, Foi 89] has been used to calculate phonon dispersion relations on surfaces such as Cu(100), Cu(lll) and Ag(111) [Nel 77, Nel 89, Luo 88]. The major differ- ence with respect to the calculations quoted above is that all parameters have been obtained by fitting the measured bulk properties. The EAM has proven to be quite successful in the prediction of surface phonon spectra and the corresponding changes of interatomic force constants and distances at surfaces. The major weakness of these calculations is the limitation to single-component systems, since charge transfers be- tween different sites are assumed to be zero. Of less importance is the fact that the success of the EAM technique depends on the type and quality experimental data and that the fit is based mainly on such observed bulk properties. To avoid these weak points, I developed a model Many-Body Alloy (MBA) Hamil- tonian for the Pd-H system, based on ab initio calculations of bulk materials (both single component systems and alloys). This Hamiltonian can be easily applied to many alloy systems, and the parametrization can be uniquely determined from a set of ab initio calculations. This mapping does not leave any adjustable parameters. The MBA technique is a very useful tool to determine the total energy and forces acting on individual atoms at the surface and in the bulk. In particular, using lattice dynamics, the surface phonon modes of clean and hydrogen adsorbed Pd surface can be calculated. This allows for a detailed analysis of the effect of hydrogen on the surface phonon spectra of Pd. 1.2.2 Mechanical stability of Pd-H systems Fundamental understanding of the mechanical stability of transition metals under var- ious conditions is of great technological importance. Very little is known about the atomic-level response of bulk metals to applied tensile stress, at varying temperature 10 and hydrogen concentration. The study of these effects is expected to answer some key questions related to macroscopic materials properties, especially the so-called “hydro- gen embrittlement”. This hydrogen-induced reduction of the mechanical strength of metals causes great concerns when considering hydrogen storage in metals, stability of fusion reactors and underwater structures, and space technology. In spite of a significant effort to resolve this important problem, there is still sub- stantial controversy regarding the microscopic origin of “hydrogen embrittlement” in a given system [Bir 79]. One of the oldest and most commonly referred to mechanisms of hydrogen embrittlement is the “decohesion mechanism” which associates this effect with a decreased metal bond strength in the presence of hydrogen [Ste 60, Foi 86].. The “hydrogen related phase change mechanism” has been suggested as the origin of “hydrogen embrittlement” in systems where a brittle hydride phase is stabilized by the presence of hydrogen and the crack tip stress field [Wes 69]. The “Hydrogen Enhanced Local Plasticity” (HELP) mechanism postulates a hydrogen-induced local plasticity enhancement at crack tips which facilitates fracture formation [Bea 72]. The most straight-forward way to address the above questions is to use a molec- ular dynamics (MD) technique. Such a calculation describes the evolution of the system with time, based on a direct numerical solution of the equations of motion for individual atoms. This procedure goes beyond lattice dynamics [Mar 63] which is limited to small atomic displacements and can not address problems such as the melting transition or fracture. The appropriate MD technique for these questions will describe the evolution of the system at constant nonzero temperatures using N osé dy- namics. It can also model the response of a system to an externally applied tensile stress [All 90, Car 90]. Such a simulation provides microscopic information about the dynamics of the system, including structural changes during phase transitions (such 11 as melting or crack formation due to tensile stress under different conditions). 1.3 Structure of the Thesis In this Thesis, I will address primarily two subjects related to interaction between adsorbed atoms and surfaces. They are the Atomic Force Microscopy and atomic- scale friction at graphite surface, and the structure and dynamics of Pd-H systems. In Chapter 2, I will review the computational tools I used. I will start with the Density Functional Theory, in particular its implementation in the Local Density Approximation (LDA) technique. I will briefly discuss the computational details, such as the basis and use of ab initio pseudopotentials. Then, I will derive the Many- Body Alloy Hamiltonian for metal and alloy systems. At the end, I will present the formalism for Nosé and Rahman—Parrinello Molecular Dynamics simulations, which allows for the description of the dynamics of a canonical ensemble exposed to an external anisotropic stress field. In Chapter 3, I will develop the theory of Atomic Force Microscopy. I will first present the ab initio results for the Pd-graphite interaction. Next, I will present a parametrization of the Pd-graphite interaction potential, which is inspired by Embed- ded Atom Method and which is not restricted to high symmetry sites. Combined with continuum elasticity theory, these results will be used to determine whether atomic resolution can be achieved nondestructively in AF M experiments on the graphite sur- face. Then, I calculate the atomic scale friction of a sharp Pd tip on graphite and determine the friction coefficient. I propose two idealized friction machines to explain the possible microscopic mechanisms of energy dissipation in the friction process. In Chapter 4, I calculate the equilibrium structure and dynamics of Pd-H systems. This chapter is divided into two parts. In the first part, I construct and test the many- 12 body alloy Hamiltonian for the Pd-H system. I calculate the equilibrium structure and surface phonon spectra of clean and hydrogen covered Pd surfaces, in order to determine the effect of hydrogen on the equilibrium structure and dynamics of Pd. In the second part, I use molecular dynamics to study the dynamical properties of Pd-H systems at finite-temperature. I will study the melting transition of pure Pd in the bulk bulk and the mechanical stability of bulk Pd at different temperatures and hydrogen concentrations. I will show that the mechanisms of hydrogen embrittlement can be understood by carefully analyzing the MD simulation results. Chapter 2 Theory 2.1 Ab initio Density Functional Formalism The basic theory governing the electronic and structural properties of solids is quan- tum mechanics. In systems with many particles (such as ~ 1023 in solids), the exact solution of quantum mechanical equation is essentially impossible to obtain, due to the complex many-body interactions which couple many degrees of freedom. Even though the ground state energy can be expressed in terms of one- and two-particle Green’s functions, the computation of these quantities involves a set of differential equations which couple all n-particle Green’s functions. The basic difficulty to de- scribe even ground-state properties of solids is significantly reduced in the Density Functional Theory. 2.1.1 Density Functional Theory In this Thesis, I will derive the Density Functional Theory (DFT) formalism very briefly, and refer the reader to other excellent reviews for more details [Lun 83, Cal 84, Jon 89, Mah 90, Dre 90]. The DFT is based on two theorems proposed by Hohenberg and Kohn in 1964 [Hoh 64], which state that: 13 14 l. The electron density n(1"') in the ground state is a functional of the potential no 2. The potential V0") is a unique functional (to within a constant) of the electron density n(r'). This is equivalent to saying that the exact ground state properties of a system can be calculated using a variational approach involving only the electron density, rather than an antisymmetric wave function. I follow the derivation by Levy [Lev 79]. The Hamiltonian describing the motion of N electrons in an external potential me is H=T+%+Km on where T and Vcc are the kinetic and electron-electron interaction operators, respec- tively. For all densities n(r') which can be obtained from an antisymmetric wave function 112(r'i, r3, . . . , rTv), Levy defined the functional Fin] = minWIT + Veelib) = ( 3.;an + Veell‘pihin), (2-2) where the minimum is taken over all 1b that give the density n, and tbmgn minimize F[n]. Let us denote Egg, $35, and nas(r‘) to be ground state energy, wave function, and electron density of the system under external field Vm. Then, Bin] 2 [dv’Vmb’inb’HFin] (2.3) ( 3“an + We ‘1" I/extlfpr’hin) > 5'05, (2-4) . 15 which follows from the variation principle applied to the ground state. This relation applies in particular to the ground state wave function swhich gives WIGSIT + Vcc + thl'l’cs) S (1033.?ng + Vise + Kali/1333:)- (25) "GS The charge density corresponding to $35 and lbw-n are the same, so the interaction with the external potential can be subtracted on both sides of the inequality, WGSIT + Welt/’05)- W) 33373” + Kali/’i'lflrfl- ‘ (2-6) Combined with Eq. (2.4), we find (thoslT + Kelwcsl= < 333.? IT + Welt/’33:) = Fines} (27) So, the ground state energy Ens = fdmzdflnosbf) + F[nos] (2.8) is a functional of ground state charge density. This equation, combined with Eq. (2.4) completes the proof of the basic theorems. These theorems provide a general method for calculating ground state properties. However, theses theorems do not suggest any particular form of the functional F [n] To solve this problem, Kohn and Sham introduced the Local Density Approximation (LDA) in 1965 [Koh 65]. 2.1.2 Local Density Approximation In their famous paper [Koh 65], Kohn and Sham decomposed the energy functional as E[n] = To[n] + F + E“, (2.9) 16 where (7'1") (7‘2) [(7‘1 - r2]. 2W. 1")n( 1")+//dr'idr'§n (2.10) Here, To is the kinetic energy of a system of non-interacting quasi-electrons with the density n(F). F is the Coulomb energy of the system of quasi-electrons. E“ is the exchange-correlation energy, which gives the contributions to the total energy except for the kinetic and Coulomb terms. Variational theory, applied on this energy functional, gives T__o__[n]6 Tn +I/ezt‘l'VH'f'ch—I-‘zo (211) Here, V” is Hartree potential derived from the second term of F, VIC = 6Exc[n]/6n, and u is the Lagrange multiplier which constrains the total number of particles to be constant. On the other hand, the variational equation for a system of non-interacting quasielectrons reads 6To[n] + V 6n - p = 0. (2.12) Inspection of Eqs. (2.11) and (2.12) shows that they are the same if V: chg+VH+Vw The solution of Eq. (2.12) can be obtained using a Schr6dinger equation. For a system of quasi-electrons, this leads to a set of Kohn-Sham equations R2 l-gng-V’ + v... + vim + molt-(o = 6.1M") (2.13) and "(a = z [WU-"llz- (2.14) Note that e,- and 1b, are not necessarily the eigenvalues and eigenfunctions of the system; only the electron density n(1'~') is the correct physical quantity. 17 Kohn-Sham equations involve a universal functional, the exchange-correlation en- ergy E,c[n], and its functional derivative V1.60"), which are only known exactly for a homogeneous electron gas. As a manageable simplification which has been introduced in the Local Density Approximation (LDA), the functional Erc[n] is replaced by a local function Exc(n). Hence, E“ = fdf'negc(n(f")), (2.15) where age is the exchange-correlation energy density for the homogeneous electron gas. Many parametrizations of file have been proposed, such as that of Wigner[Wig 38], Kohn and Sham[Koh 65], Hedin and Lundqvist[Hed 71], Ceperly and Alder[Cep 80] etc.. In our calculation, we usually use Hedin-Lundqvist parametrization, which have been proven to give good results. The LDA is a accurate in systems with a slowly varying charge density. It is also accurate in systems with a large charge density, where the kinetic and Hartree terms dominate E“, and in systems with small charge densities, where gradient of n(i~') are small. This formalism has been successfully applied to many systems, including atoms, molecules, clusters, surfaces and bulk of solids. 2.1.3 Computational details In the real calculations based on the DFT, the initial guess of the charge density is obtained using a superposition of atomic charge densities. Using this charge density as input, V“ can be obtained using the LDA. This provides all of the information needed to set up the Kohn-Sham equations, which are usually solved as matrix equations in a given basis. The solution of the Kohn-Sham equations gives e,- and 11),. Then the Fermi energy can be determined and the charge density obtained as output. In the 18 next iteration, this charge density is used as input, and the procedure is repeated until self-consistency is achieved. In the self-consistent field method, the input and output potential is required to be the same. One practical consideration in a realistic calculation is the selection of a basis. The plane wave basis is the simplest and easiest to implement, since all basis functions are orthogonal to each other. Plane waves give a complete set of basis functions; a finite basis is typically limited by the cutoff energy EC. This basis is ideal for nearly free electron systems like simple metals, where the effective potential does not vary too much. The number of basis functions needed is proportional to Bf”, where d is the dimensionality of the system. The computational load increases as Eff/2, since solution of Kohn-Sham equations involves matrix diagonalization. This computational effort makes a plane wave basis impractical for materials with localized electrons, i.e. transition metals and semiconductors. In my Linear Combination of Atomic Orbitals (LCAO) [Laf 71, Cal 72, Che 84] calculation, I use a local Gaussian basis in order to improve efficiency [Har 82]. The form of the basis functions is ezp(—a,-r2) for 3 states f.- = (2:,y, z)ea:p(—a,-r2) for p states (:ry, yz, 22:, a:2 — y2, r2 — 322)e.rp(—a,-r2) for d states This basis provides an efficient representation of the atomic wave functions in systems with localized electrons. More than one radial Gaussians are typically used for each orbital to allow for variational freedom. Completeness of the basis is tested by increasing the number and location of Gaussian functions and monitoring the convergence of the calculated total energy. In my calculations, usually three to four Gaussian decays for each atomic orbital (s, p,, py, ...) are found to give a sufficiently complete basis set. The decay constants a for a given atomic type are chosen to 19 minimize the total energy of the bulk system. Since the LCAO basis contains non- orthogonal basis functions, the solution of the Kohn-Sham equations is more involved. Beside using Gaussian functions for basis functions, it is very useful to fit the screened atomic potentials and charge density to a set of Gaussians as well. This procedure allows for an efficient evaluation of crystal LDA-Bloch functions up to very high Fourier components corresponding to a large value of the energy cutoff EC. Most electronic and structural properties of solids are determined by the behavior of valence electrons, while the core electrons are essentially inert. Valence electrons feels a very weak potential outside the core due to the screening by the core electrons. Replacing the true ionic potential by a pseudopotential, which models the nuclear and the nonlocal core potential, simplifies the calculation considerably. I use the ab initio pseudopotential form proposed by Hamann, Schliiter, and Chiang [Ham 79, Bac 82], which has following desirable properties: 1. Pseudopotentials are continuous non-divergency functions with a continuous first derivative everywhere. 2. Real and pseudo eigenvalues for valence states agree for a chosen “prototype” atomic configuration. 3. Pseudo wavefunctions are nodeless. 4. Real and pseudo wavefunctions agree beyond a chosen “core radius” re. 5. The integral from 0 to r of the real and pseudo charge densities agree for r > rc for each valence state (norm conservation). 6. The logarithmic derivatives of the real and pseudo wave functions and their first energy derivatives agree for r > rc. 20 Properties 5 and 6 are important to ensure optimum t‘ransferability of the pseudopo- tential between different chemical environments. Norm conservation enables the use of the pseudo charge density instead of the real charge density in UP T calculations. Property 6 ensures the correct scattering properties of the potential. No adjustable parameters occur in the construction of the Hamann-Schliiter-Chiang pseudopoten- tials. The LCAO Gaussian basis and ab initio pseudopotentials are very powerful which make precise LDA calculations for transition metals or semiconductors feasible. Nev- ertheless, the computational load increases rapidly with increasing complexity of the system (reduced symmetry, increased size of the unit cell). For very large systems, parametrized calculation schemes are more appropriate. A very efficient scheme to calculate the total energy of a system is the Many-Body Alloy Hamiltonian to be discussed below. 2.2 Many-Body Alloy Hamiltonian The Many-Body Alloy (MBA) Hamiltonian is an extension of a total energy scheme, which has been successfully used previously to study the electronic and structural properties of small clusters, surfaces of metals and dilute metal alloys [Tom 83, Tom 86a, Tom 85b, Tom 85a, Spa 84]. Its derivation has been published recently by W. Zhong, Y.S. Li, and D. Tomanek [Zho 91a]. As discussed earlier [Tom 83, Tom 86a, Tom 85b, Tom 85a], the total cohesive energy of a solid can be decomposed into individual atomic binding energies Ecoh(i), 8.8 E... 2.5A). This interaction can be well parametrized in terms of the charge density of the graphite host. 3.2 Limits of resolution in AFM images of gra- phite As pointed out in Chapter 1, the possibility of atomic resolution in the AFM is limited in two ways. Too small forces lead to insufficient corrugations, and too large forces may destroy the surface. So far, these limitations of the atomic resolution have not been addressed in the literature. Present theoretical information is limited to calculations of the interaction between an infinite “periodic” carbon or aluminum tip and a rigid surface [Cir 90, Bat 88], and the interaction between a single AFM tip and an elastic surface represented by a semi-infinite continuum [Tom 89] or by a model system of finite thickness [Abr 89, Lan 90]. 39 End (9") Figure 3.4: Adsorption energy E“ of Pd on graphite as a function of the height 2 of Pd atoms above the graphite surface. F irst-principles results are given by 0 and e for the hollow and the on-top site, respectively. These data are compared to the present results, obtained using Eqs. (3.2), (3.3) and (3.4) and given by the solid and the dashed line for the hollow and the on-top site, respectively. Upper inset: Schematic top view of the adsorption geometry. Lower inset: An enlarged section of the graph near equilibrium adsorption. (From Ref. [Tom 91b]. ©American Physical Society) 40 In this Section, I discuss for the first time the theoretical limits of atomic resolution in AFM. I will use the ab initio results of the interaction between a “sharp” monatomic Pd tip and the surface of graphite discussed in Section 3.1, to predict corrugations for a varying load Fm applied on the AFM tip. These calculations also predict tip-induced elastic substrate deformations which limit the range of applicable loads Fert- Using the adsorption energies End given by an ab initio calculation, the force on a “sharp” monatomic AFM tip is given by _ aEad(z) fez-t — '— 62 ~ ' (3'5) Here, microscopic forces per atom are denoted by f and macroscopic forces applied to large objects by F. I also investigated the effect of the long-range Van der Waals forces on the tip-substrate interactions which are not described correctly by the LDA especially at large tip-substrate distances. The Van der Waals force between an extended conical tip and a flat surface is estimated using the expression Fv4w(z) = A3 x tanza/(6z), where a is half the opening angle of the cone [And]. In this expression, A” is the Hamaker constant and z is the distance between the cone tip and the surface. In this calculation, I consider a = 30° and A3 = 3 x 10‘19 J, which is a typical value for metallic systems. For tip-substrate distances 2 > 3 A, the Van der Waals forces are very small, typically Fww < 10-10 N. At smaller distances, these forces can be neglected when compared to the closed-shell and internuclear repulsion which are described correctly within LDA. Since each of these regions is dominated only by one type of interaction, I determine the total tip-substrate force Fm as a superposition of the force described by LDA and the Van der Waals force. In Fig. 3.5(a), I present results for the total 41 force F = fez, between a conical AFM tip with 1 Pd apex atom and a graphite substrate, together with a schematic top view of the geometry. The LDA calculations yield nearly the same weak Pd-graphite interaction poten- tial (Edd < 0.1 eV) in the “H” and “T” sites beyond the equilibrium Pd-graphite bond length 2,, z 3 A and consequently the same weak interaction force. This is plau- sible since End is closely related to the total charge density )0, shown in Fig. 3.5(b), which has only a very small site-dependence in the attractive region of the potential at z > 2,, due to Smoluchowski smoothing [Smo 41]. Consequently, the corrugation A2 in this region is too small to be detected by the AF M. In the weakly repulsive region of the potential, for 2 A < z < 3 A, the on-top site is slightly favored with respect to the hollow site due to a weak chemisorption bond with substrate C2125 orbitals. For atom bond length 2 < 2 A, the strongly repulsive Pd-graphite interac- tion is mainly determined by the closed-shell repulsion which energetically favors the hollow site. Hence, in the repulsive region of the potential, the Pd tip comes closest to the substrate near the “T” site for small loads. For large loads, the tip is closest to the surface near the “H” site. Fig. 3.6(a) shows the expected AFM corrugation Az during an my scan of the graphite surface, for feet = 10‘8 N. The equilibrium bond length 2(fm) at other than the calculated high-symmetry sites has been determined using a separate model calculation, described in Section 3.1 and Ref. [Tom 91b], which relates End to the total charge density of the graphite host at the Pd adsorption site. This energy functional is assumed to be universal and reproduces End in high-symmetry sites accurately. In this calculation, I assumed a perfectly rigid substrate and a monatomic Pd tip. A top view of this tip at the graphite hollow site is shown schematically in the inset of Fig. 3.6(b). 30- iriio‘9 N) N O Figure 3.5: (a) Calculated force F between a Pd AFM tip with one apex atom and the surface of hexagonal graphite, as a function of the Pd-graphite distance 2. The solid and dashed lines correspond to the sixfold hollow (H) and the on-top (T) sites, respectively. An enlarged section of the graph near the equilibrium is shown in the inset. A second inset shows the adsorption geometry in top view. A possible trajectory of an AFM tip along x is shown by arrows. (b) Valence charge density of the Pd/ graphite system. The results of the LDA calculation are for the on-top adsorption site near the equilibrium adsorption distance 2”, and are shown in the :2 plane perpendicular to the surface. The ratio of two consecutive charge density contours p(n + l)/p(n) is 1.2. (From Ref. [Zho 91b]. ©Les Editions Physique) 43 III I III]. ‘ n“. 0.094,,” \ . “3.0.21, ”‘ . I, I, r/ l I ’r ‘ 0 9% ',,’I/,' o ’I I” a 0, III r"! I I . . . I I! I,’I,’;I,,, ‘. o ”obi/n. -n\| s..: I ” “n so I II: -\\\“| . 5’07" us‘.°.’,”I/r \| O I .e ‘ M35 ’1" . .' '3 W \WW\ T 1T 11'] . T. T. H T T H 0.00 2.50 5.00 7.50 10.00 12.50 x(K) Figure 3.6: (a) Surface corrugation Az experienced by a monatomic Pd AFM tip scanning the my surface plane of rigid graphite under the applied AFM load (per atom) fm = 10‘8 N. (b) A2 (with respect to the “H” site) along the surface .1: direction for a “sharp” l-atom tip, for fa. = 10'” N (dotted line), 5 x 10'9 N (dashed line), 10"8 N (solid line), and 2 x 10'8 N (dash-dotted line). The inset shows the geometry of the tip-graphite system in top view. The AF M tip is shown above the hollow site, and the shaded area represents the Pd atom. (From Ref. [Zho 91b]. ©Les Editions Physique) 44 Fig. 3.6(b) shows the AF M corrugation A2(.r) for different loads fat. The tip trajectory along the surface zit-direction, shown by arrows in the inset of Fig. 3.6(b), contains the “T” and “H” sites and yields the largest corrugation. As discussed above, the favored surface site changes with changing load. For the sake of simple comparison, I set A2(hollow site)=0 in Fig. 3.6(b) and obtain a sign change of A2 near for: = 2.5 x 10‘9 N. These calculations show that atomic resolution in the constant force mode in the AFM, corresponding to A22,0.05 A, requires loads fcx¢i5 x 10‘9 N. Since the corrugation A2 along a trajectory connecting adjacent “T” sites is very small (see Fig. 3.6(b) ), the observation of individual carbon atoms is unlikely, which has been confirmed by the experiment [Mey 88]. For an n-atom tip, which is commensurate with the substrate, the average load per AFM tip atom is fat = Fat/n and the equilibrium tip position 2 can be estimated from Eq. (3.5). It should be noted that under certain conditions, such a “dull” multi- atom tip can still produce atomic corrugation for fat similar to a monatomic tip. This is the case for an ideally aligned tip with a close-packed (111) surface, since the unit cells of Pd and graphite are nearly identical in this case. The range of applicable loads Fm is limited by the condition that substrate dis- tortions near the AF M tip should be small and remain in the elastic region. Since full- scale LDA calculations of local AFM-induced distortions of a semi-infinite graphite surface are practically not feasible, I adopt the following approach. First, I use con- tinuum elasticity theory [Lee 89], with elastic constants obtained from ab initio calcu- lations [Tom 89], to determine the relaxation of carbon atoms at the graphite surface in response to the AF M load applied through an AF M tip. [This continuum approach is applicable in the linear response regime and has been successfully used previously to calculate local rigidity, local distortions and the healing length of graphite near 45 an AFM tip and near intercalant impurities [Tom 89, Lee 89]. In a second step, the atomic structure in the total charge density of the Pd/ graphite system is regained from a superposition of atomic charge densities given by LDA-atom calculations. The semi-infinite system of graphite layers is characterized by the interlayer spac- ing d, the in-plane C-C bond length dc-c, the flexural rigidity D, the transverse rigid- ity K (proportional to C44) and c-axis compressibility G (pr0portional to C33)[Tom 89, Lee 89]. The LDA calculations for undistorted graphite yield d = 3.35 A and dc—C = 1.42 A, in excellent agreement with experiment [Zab 89]. In the contin- uum calculation, I further use D = 7589 K, K = 932 KA'2 and G = 789 KA'“ which have been obtained from calculated graphite vibration modes [Tom 89] and the experiment [Zab 89]. The total charge density of the graphite surface, distorted by a Pd AF M tip, is shown in Fig. 3.7. A comparison of charge density contours with results of the self-consistent calculation in Fig. 3.5(b) proves a posteriori the applicability of the linear superposition of atomic charge densities. Fig. 3.7(a) shows that the substrate distortion in response to a monatomic AFM tip at a load Fm = fez: = 10’9 N is moderate. According to Fig. 3.6(b), the corresponding corrugations during an AF M scan with this load are A2 z 0.03 A and thus below the limit of detection. A larger load Fm = [at = 5.0 x 10"9 N leads, according to Fig. 3.6(b), to marginally detectable corrugations A2 a: 0.06 A, but distorts graphite much more, as shown in Fig. 3.7(b). ‘ For larger applied forces Fm = f“. > 5.0 x 10’9 N, which would lead to sizeable corrugations, the local distortions of graphite exceed the elastic limit. A rough es- timate of these distortions, based on continuum elasticity theory, indicates that for f mZS x 10‘9 N, the distance between graphite layers near the AFM tip approaches L 46 >>X TX ’ i091 " @X ‘ Figure 3.7: Total charge density p of the monatomic Pd AFM tip interacting with the elastic surface of graphite near the hollow site. Contours of constant p are shown in the :2 plane perpendicular to the surface, for (a) Fm = fest = l x 10'9 N and for (b) Fm = f,“ = 5 x 10'9 N. The ratio of two consecutive charge density contours p(n + l)/p(n) is 1.4. The location of the applied load acting on the Pd atom is indicated by a triangle. (From Ref. [Zho 91b]. ©Les Editions Physique) 47 the value of intra—layer C-C distances. This new diamOnd-like bonding geometry leads to a rehybridization of carbon orbitals and will necessarily result in an irreversible substrate deformation. An independent estimate of the critical AFM force for this plastic deformation can be obtained from a first-principles calculation [Fah 86] of the graphite-diamond transition as a function of external pressure. These results, corre- sponding to an “infinitely extended tip”, indicate a critical force per surface atom of fext = 10’9 N for this transition. Due to the large flexural rigidity of graphite, this force increases by half an order of magnitude for a one-atom tip, in agreement with the above result. A realistic AFM tip is more complex than the model tip discussed above and could consist of a micro-tip of one or few atoms on top of a larger tip. A substantial por- tion of this larger tip could, through the “cushion” of a possible contamination layer, distribute the applied load more evenly across a large substrate area, reduce the large curvature near the tip (see Fig. 3.7(b) ) and increase the minimum interlayer sepa- ration. This effect would increase the upper limit of applicable loads fm compatible with elastic substrate deformations and would lead to atomic resolution [Mey 88]. Summarizing these results, I used ab initio calculations to determine corrugations observable in Atomic Force Microscopy of graphite. I found that in the constant- force mode, atomic resolution is marginally possible for AFM loads (per atom) close to 5 x 10’9 N. For smaller loads, the corrugation A2 is too small to be observed. For loads which are too large, graphite deformations exceed the elastic limit and subsequently result in the destruction of the substrate. 48 3.3 First-principles theory of atomic-scale fric- tion In this Section, I present a predictive first-principles theory of atomic-scale friction. I use ab initio results of the interaction energy between a perfect layer of atoms and a graphite layer, as a function of their relative distance . I use this information to determine the friction coefficient u between these systems in contact. My results include quantitative predictions of u as a function of the external force, a qualitative explanation of the increase of u with increasing surface roughness, and a qualitative explanation of the decrease of u with increasing relative velocity between the objects in contact in case of sliding friction. For a microscopic understanding of the friction process, I first consider the motion of an atomic layer along the surface. This layer is in registry with the substrate and experiences an external force per atom fez, normal to the surface. At each site, the equilibrium adsorption length 2 is given by the condition fest = ‘%E0d(z)- (3.6) I consider a straight trajectory along the :1: direction in the surface connecting nearest neighbor hollow sites which are separated by A3: (see the inset in Fig. 3.8(a)). The position-dependent part of the potential energy V of the system has two main com- ponents. These are variations of the adsorption bond energy and the work against the external force fm applied to the adsorbate, due to variations of the adsorption length. Hence, V(Iifert) = Ead(za Z(¢afezt)) + fectz(xafext) _ %(fezt)1 (37) 49 10"N) ° .° . . O r-b ' \ ‘0 I, “ i, ‘t I ‘ , -0o1 -‘ I, | g, “ (‘1 u ‘s a “ ’I \ f ' \l H B .’ H B \‘ -0.3 1 . . - 0-0 2 5 5.0 . x(A) Figure 3.8: (a) Potential energy V(x) of the Pd-graphite system as a function of the position of the Pd layer along the surface x-direction, for external forces fm = 3 x 10’9 N (dotted line), 6 x 10'9 N (dashed line) and 9 x 10'9 N (solid line). The inset shows the adsorption geometry and trajectory of the Pd layer in side view. (b) Atomic-scale structure of the force along the surface f, (dashed line) and the friction force f, = |f,| (solid line) for fm = 9 x 10‘9 N. (From Ref. [Zho 90]. ©American Physical Society) 50 where I arbitrarily set the potential energy to zero at the hollow site by defining I/O(fe:rt) = Ead(xHaz(fext)) + fext 2(IHff61‘t)‘ (3.8) In Fig. 3.8(a) I show V(:c) for different external forces. I obtain the periodic potential V(:r) at other than the calculated high-symmetry sites from a Fourier expansion over the reciprocal lattice. I keep the lowest components of this expansion and determine the expansion coefficients from the calculated values for the “T” and “H” sites. How- ever, the calculated frictional coefficient depends only on extrema of the potential V(:c), which is assumed to be at “T” and “H” sites. From my calculation, In I find that the mechanical component dominates and is only partly compensated by adsorption energy differences. As a result of variations of V along 2:, there is a position-dependent force f3 along the a: direction, shown in Fig. 3.8(b), which is given by f1:(xa fest) = %V(xafext)- (3.9) The maximum value of f, describes the static friction governing the onset of stick- slip motion. The sliding friction on the other hand must be obtained as a weighted average over this force from the energy dissipated in friction along 2:. I first con- sider a conservative part of this process, corresponding to potential energy increase AVmu(fm) = Vmu(f,¢¢) - Vm;n(fm) along Ax, which yields a positive value of ft. The non-conservative part corresponds to a decrease of V(:c) along a: and a negative value of f,. Friction losses AE, along Ax must not exceed the maximum increase in the potential energy, hence AE, _<_ AVnm. (3.10) 51 For very slow tracking velocities, any gain in potential energy during the dissipative part of the friction process is efficiently transferred into surface phonons and electron- hole pairs. Then, I can consider both sides of Eq. (3.10) to be equal, which corresponds to f, = [ft], shown in Fig. 3.8(b). This is the first quantitative prediction of atomic- scale structure in the friction force which has been observed recently [Mat 87]. The energy dissipated in friction along Ax can also be related to the average friction force < f f > along the trajectory, as AEf =< f;>A:1:. (3.11) Using Eq. (3.10) for AEf, I obtain 1. = Ax awmr _ win Applying the definition of the friction coefficient 14 = f f / fat, I find _ _ AVma, - 7:..— - ea.- ‘3'”) In Fig. 3.9 I show u as a function of fat. I find a general increase of u with increasing external force. The minimum in u(fm) near fez; = 5 x 10‘9 N is caused by the switching of the minima in V(a:) from H to B, shown in Fig. 3.8(a). Near fez. = 5 x 10'9 N, the lowest order Fourier components of the potential V vanish. According to the interpolation scheme discussed above, this leads to u = 0 (shown in Fig. 3.9). In reality, a small nonzero value of u is expected due to higher-order Fourier components of V (which are usually much smaller than those considered). From former results presented in Section 3.2 and and Ref. [Tom 89], I conclude that if the external force (per atom) exceeds 10‘8 N, the graphite surface is very strongly deformed [Mam 86] and likely to be ruptured [Ski 71]. Since no plastic deformations have been observed 52 0.06 :5. 4.) d 0.05 0 00-4 a 0.04 H 8 o 0.03 8 3;; 0.02 0 "-0 a 0.01 0.00 ‘ L l L 0 5 10 15 20 25 30 fu,(10'°N) Figure 3.9: Microscopic friction coefficient )1 between a Pd AF M tip and graphite as a function of load per tip atom fm. (From Ref. [Zho 90]. ©American Physical Society) 53 in the AF M studies [Mat 87]. the applied forces were probably in the region fort < 10"8 N. For these values of fat, the calculated friction coefficient of the order p z 10‘2 agrees with the experiment [Ski 71, Mat 87]. In order to obtain a meaningful comparison with observable friction forces, I have to make further assumptions about the macroscopic interface and the elastic response of the substrate to external forces. In the simplest case, I consider an atomically flat interface, where N atoms are in contact with the substrate, and neglect elastic deformations. Then, the external force per atom fat is related to the total external force F m by 1 feet = NFext- (3.14) In Fig. 3.10 I use the calculated u(fm) to plot the total friction force F f for such a perfectly flat interface consisting of 1500 Pd atoms. Since )1 increases with increasing value of fat, the F f versus Fm relationship is nonlinear, which has also been observed in the AFM experiment [Mat 87]. In the case of large external forces and an elastic substrate such as graphite, elastic theory predicts [Lan 86] the substrate deformations to be proportional to Fez/,3. In case of a spherical tip [Mat 87], the tip-substrate interface area and the corresponding number of atoms in contact is proportional to F3153. Then, the force per atom fat is proportional to Fez/,3. Hence for increasing external forces, variations of the effective force per atom and of p are strongly reduced due to the increasing interface area. This is illustrated by a dashed line in Fig. 3.10, where I used N = 1500 atoms for Fm = 10'6 N. These results are in good agreement with the AFM results for a large nonspecific tungsten tip with a radius R = 1500 A - 3000 A on graphite [Mat 87], but show a slightly larger increase of the friction force than observed for the range of 20 N 0 ’5 ? 15 L C i. fl v A 10 f to CI. [ , V metros -. ' spear-sis; 5 l was m l -' cousrmr . is: m. . .~ ] O 5 10 15 20 rux104N) Figure 3.10: (a) Calculated macroscopic friction force F f as a function of the external force Fm for a large object. The solid line describes a “flat” object, the surface of which consists of 1500 atoms in contact with a rigid substrate. The dashed line de- scribes friction of a large spherical tip and also considers the effect of elastic substrate deformations on the effective contact area. The dotted line corresponds to a constant friction coefficient )1 = 0.012. (From Ref. [Zho 90]. ©American Physical Society) (b) Experimental measured average friction force as a funCtion of load on the tip as it slides across the surface. (From Ref. [Mat 87]. ©American Physical Society) 55 external forces investigated. This theory can be used also to explain the dependence of p on the roughness of the interface and on the relative velocities of the two objects in contact. At a rough surface, the number of atoms N in contact with the substrate is smaller than at a flat surface which leads to an increase of fat and hence of p. Also, with increasing relative velocity, the coupling between macroscopic and internal microscopic degrees of freedom (phonons, electron-hole pairs) gets less efficient. Then, AE; < AVm“ in Eq. (3.10) which causes a decrease of < f f > and u. Summarizing the results in this Section, I determined the atomic-scale friction associated with a layer of Pd atoms moving across a graphite substrate from ab initio total energy calculations. I evaluated the friction energy caused by variations of the chemical bond strength and work against an external normal load. The calculated value of the friction coefficient is very small, in the order p z 10'2 for loads near 10‘8 N. This small value can be explained by the weak dependence of Pd-graphite interaction on the adsorption site. I also found )1 to increase with load in agreement with recent Atomic Force Microscopy experiments. 3.4 Ideal friction machines In this last Section, I determine the upper limit of the average friction force < F f >, originating from the potential energy barriers AV(Fm) during the “surface diffusion” of Pd on graphite. There, I assumed that the energy gained by the tip atom is completely dissipated into microscopic degrees of freedom. In the following, I will address the mechanism which leads to the excitation of microscopic degrees of freedom and hence to energy dissipation. I will make use of the first-principles Pd-graphite interaction potential. Based on these data, I will determine atomic-scale modulations fi'l 56 of the friction force F, along the horizontal trajectory of the Pd tip on graphite in the quasistatic limit of relative velocity v —> 0. I will show that F f depends sensitively not only on the corrugation and shape of the Pd-graphite interaction potential, but also on the specific construction parameters of the F FM. The latter point is of utmost importance if friction forces obtained with different microscopes are to be compared to one another. Two models of a Friction Force Microscope are shown in Fig. 3.11. In both models, the suspension of the tip moves quasistatically along the surface x-direction, with its position 2);; as the externally controlled parameter. The tip is assumed to be stiff with respect to excursions in the surface y-direction. I restrict the discussion to the case of tip-induced friction and assume a rigid substrate, which applies for friction measure- ments on graphite [Mat 87, Zho 90]. In the “maximum friction microscope” [Zho 90], the full amount of energy needed to cross the potential energy barrier AV along As is dissipated into heat. This process and the corresponding friction force can be ob- served in an imperfect Atomic Force Microscope which is shown in Fig. 3.11(a). A vertical spring connects the tip and the external microscope suspension M. The hor- izontal positions of the tip and the suspension are rigidly coupled, :c, = 2M = :c. For 0 < :1: < Ax/ 2, a constant load Fm on the tip is adjusted by moving the suspension up or down. For Ant/2 < a: < Ax, however, the tip gets stuck at the maximum value of 2,. At Am, the energy AV stored in the spring is abruptly and completely released into internal degrees of freedom which appear as heat. The calculated potential energy V(:c) during this process is shown in Fig. 3.12(a). I considered a monatomic Pd tip sliding along a trajectory connecting adjacent hollow and bridge sites on graphite, for a load Fm = 10'8 N. In order to predict F f, I have used the results of my previous ab initio calculation for Pd on graphite presented in 57 Figure 3.11: Two models for the Friction Force Microscope (FFM). In both models, the external suspension M is guided along the horizontal surface a: direction at a constant velocity v = dzM/dt —v 0. The load Fm on the “sharp” tip (indicated by V) is kept constant along the trajectory 2,(:1:;) (shown by arrows). (a) A “maximum friction microscope”, where the tip is free to move up, but gets stuck at the maximum 2, between Az/2 and Ar. (b) A “realistic friction microscope”, where the position of the tip 2:, and the suspension 2M may differ. In this case, the friction force F J is related to the elongation x: - 12M of the horizontal spring from its equilibrium value. (From Ref. [Tom 91c]. ©Les Editions Physique) 58 Section 3.1 and Ref. [Zho 90], which have been conveniently, parametrized [Tom 91b]. The force on the tip in the negative :r-direction, as defined in Fig. 3.11(a), is given by _ 5V(xM)/82M if 0 < 2M < Ax/2 Elf“) " { 0 if Ax/2 < 2M < A2: (3'15) and shown in Fig. 3.12(b). The nonzero value of the average friction force < F f >, indicated by the dash-dotted line in Fig. 3.12(b), is a clear consequence of the mech- anism which allows the tip to get stuck. F f(xM) is a non-conservative force since it depends strongly on the scan direction. In absence of the “sticking” mechanism, F f is given by the gradient of the potential everywhere, as indicated by the dashed line in Fig. 3.12(b). It is independent of the scan direction and hence conserva- tive. In this case, F ; inhibits sliding for 0 < 2354 < Az/2 and promotes sliding for Ax/2 < 2M < Ax, so that the friction force averages to zero. A more realistic construction of Friction Force Microscope is shown in Fig. 3.11(b). In this microscope, the AFM-like tip-spring assembly is elastically coupled to the suspension in the horizontal direction, so that 2:, may differ from 2M. For a given 1:5, the tip experiences a potential V(:r¢,2¢) = V,-m(.r¢,2¢) + Fm 25 consisting of the tip-surface interaction V5,“ and the work against Fm. The tip trajectory 2¢,m,n(a:t) during the surface scan is given by the minimum of V(:r,, 2,) with respect to 2,. For this trajectory, V(:c¢) = V(:1:5, 2¢,m,-,,) represents an effective tip-substrate potential. This potential V(:c¢) depends strongly on Fm and is corrugated with the period- icity of the substrate due to variations of the chemical bond strength and of 25min. It is reproduced in Fig. 3.13(a) in the case of a monatomic Pd tip on graphite and Fm = 10’8 N. The corrugation of the potential V(;1:t) will elongate or compress the horizontal spring from its equilibrium length which corresponds to 2:, = 3M. The V(X)(eV) 1.6 t(b) I r 3 , 1.0 ”- -( A z 05 b o-e-e -e-e-e-e-e ..... g '1' o 0.0 , V4 ‘ a V \ I, [gr-0.5 - x‘ .’ . “1.0 ' ‘\‘-"I’ .1 -1.5 l n 1 0 142$ 2 A 3 1(3) Figure 3.12: Potential energy of the tip V(:r) (a), the friction force F,(:1:) and the average friction force < F I > (b) in the “maximum friction” microscope with a monatomic Pd tip on graphite. The arrows indicate the tip trajectory corresponding to a relaxed vertical position 2. for a constant load on the tip Fm = 10"8 N. (From Ref. [Tom 91c]. ©Les Editions Physique) 60 0.8 (a) F... 10'8 N 0.6 P 0.4 " d V(xt )(eV) /// // it}? zAxs MA) Figure 3.13: Microscopic friction mechanism in the “realistic friction microscope”. The calculations are for a monatomic Pd tip on graphite and Fm = 10"8 N. Results for a soft spring, giving nonzero friction, are compared to a zero-friction microscope with a hard spring. (a) Potential energy of the tip V(z.). (b) A graphical solution of Eq. (3.19) yielding the equilibrium tip position at the intersection of the derivative of the potential 6V(2g)/63¢ (dashed line) and the force due to the horizontal spring Ff. Solid lines, for different values of 1:", correspond to a soft spring with c = 10.0 N/ m, and dashed lines correspond to a hard spring with c = 40.0 N / m. (c) The calculated equilibrium tip position z¢(zM). (d) The friction force F, as a function of the FFM position 254 and the average friction force < F f > for the soft spring (dash-dotted line). (From Ref. [Tom 91c]. ©Les Editions Physique) 61 friction force is given by PAW) = -c(a=: - m). (3.16) where c is the horizontal spring constant. The total potential energy VM of the system consists of V(:c¢) and the energy stored in the horizontal spring, 1 2 V,O¢(:r¢, 3M) = V(a:t) + §c(a:t — 2M) . (3.17) For a given horizontal position ml of the FFM suspension, the equilibrium position of the tip at, is obtained by minimizing Vtot with respect to mt. This gives aviot _ aV(I¢) 6:5: -— 7;:— + C(51); — xM) = 0 (3.18) or, with Eq. (3.16), _ _ (ii/(1%) Ff — CED; + cxM — 61¢ . (3.19) A graphical solution of Eq. (3.19) is shown in Fig. 3.13(b) and the resulting relation x¢(x M) is shown in Fig. 3.13(c). If the force constant c exceeds the critical value cc,“ = —[02V(:r¢)/8z?]m,-,,, which for Pd on graphite and Fm = 10‘8 N is cc,“ = 23.2 N /m, I obtain a single solution x, for all 354. This situation is indicated by the dotted line in Figs. 3.13(b) and 3.13(c) for a hard spring with c = 40.0 N /m. The friction force F, is given by Eq. (3.16) and shown by the dotted line in Fig. 3.13(d). F; is independent of the scan direction and hence conservative, resulting in < F f >= 0. Hence no friction should occur in the AFM, which is the limiting case of an F FM for c—voo. A more interesting case arises if the horizontal spring is soft, c < cm}. This sit- uation is shown by the solid line in Figs. 3.13(b) and 3.13(c) for c = 10.0 N/m. In 62 this case, the solution :r.(1:M) of Eq. (3.19) displays a sequence of instabilities. These instabilities lead to a stick-slip motion of the tip with increasing mM, similar to “pluck- ing a string”. The hysteresis in the x,(:cM) relation, shown in Fig. 3.13(c), results in a dependence of the force F I on the scan direction. The friction force F f(:r M) in this case is shown by the solid line in Fig. 3.13(d). It is a non-conservative/dissipative force and averages to a non-zero value < F f >= 3.03 x 10'10 N, given by the dash- dotted line. The energy released from the elongated spring into heat is represented by the shaded area in Fig. 3.13(d). The present theory predicts occurrence of friction only for very soft springs or a strongly corrugated potential V(a:¢). The latter fact can be verified experimentally since the corrugations AV(1:¢) increase strongly with increasing applied load [Zho 90]. Consequently, for a given c, the friction force is zero unless a minimum load Fm is exceeded. On the other hand, for a given Fm, no friction can occur if c exceeds a critical value can-(Fest). A similar situation occurs during sliding between large commensurate flat sur- faces of A on B. In that case, c is given by the elastic constants of A at the inter- face [Toml 29], hence can not be changed independently. Since c is rather large in many materials, zero friction should be observed for moderate applied loads in the absence of wear and plastic deformations. For a multiatom “tip” which is commen- surate with the substrate, the tip-substrate potential is proportional to the number of tip atoms at the interface, n, and so is the critical value cc,“ for nonzero friction. In this case, the effective FFM spring depends both on the external spring and the elastic response of the tip material. The inverse value of cc,“ is given by the sum of the inverse values of the corresponding spring constants. For a large tip which is incommensurate with the substrate, no friction should occur [McC 89]. 63 The average friction force < Ff > as a function of the load Fm and the force constant c is shown as a contour plot in Fig. 3.14 for a monatomic or a larger com- mensurate Pd tip on graphite. Clearly, the applicable load range is limited by the underlying assumption of contact without wear. This figure illustrates that not only the friction force F f, but also the friction coefficient )1 =< F f > /Fm, depend strongly both on the interaction potential between the two materials in contact and on the intrinsic force constant c of the Friction Force Microscope. This clearly makes the friction force dependent on the construction parameters of the FFM. There is also one advantage in this fact: c can be chosen in such a way that nonzero friction occurs even at small loads Fest- Summarizing the results in this Section, I calculated the atomic-scale modulation of the friction force and the corresponding stick-slip motion at the interface during the relative motion between Pd and graphite. I proposed two idealized versions of a Friction Force Microscope. I showed that the friction force depends not only on the Pd-graphite interaction potential, but even more critically on the construction parameters of such a microscope. 3.5 Conclusions In this Chapter, I have studied theory of atomic force microscopy, using Pd AFM tip and graphite substrate as the model system. I have calculated interaction between Pd and graphite using ab initio density functional theory within local density approximation, with Pd atom placed above hollow and on-top sites of graphite. The precise calculations show that Pd—graphite interaction is characterized by strong close shell repulsion at small distances, which strongly favors hollow site, and very weak chemical bond at larger distance with on- 64 c/n (Nm—l) F.,./n (10‘5N) Figure 3.14: Contour plot of the average friction force < F f > between a Pd tip and graphite, as a function of the load Fm and the force constant c. All forces and the force constant are normalized by the number of tip atoms 71 in contact with the substrate. (From Ref. [Tom 91c]. ©Les Editions Physique) 65 top site slightly favored. The total energy results at high symmetry points can be well mapped to a parameterized form, as a function of graphite host charge density. The optimum combination of ab initio and parameterized results provides a precise and complete picture of Pd/ Graphite interaction. Using these results, I have studied the limit of resolution of Pd AFM tip on the graphite surface. The calculated trajectories of Pd probing graphite surface under different loads reveals that, the minimum load on Pd is about 5 x 10’9N to produce observable corrugation. Using continuum elasticity theory, I have estimated that the load exceeding 10’3N leads to too much deformation and is thus destructive to the surface. So the atomic resolution of graphite from Pd AFM tip is only marginally achievable. The atomic scale friction between an Pd tip and graphite surface is obtained from the variation of Pd tip energy along the scanning trajectory. There are two contributions to this energy, one is from the work against the load when corrugation is none zero, another comes from adsorption energies difference at different sites. Assuming all the kinetic energy gained by the AFM tip dissipates into heat, I obtain the first quantitative prediction of friction force using ab initio techniques. The calculated friction coefficient is in the order of 10'2 for loads near 10’8 N, and it increases with increasing loads. The calculation is in good agreement with recent experiment. I further investigate the possible mechanisms of energy dissipation during the fric- tion process. I proposed two models for friction force microscope (FFM). In the more realistic version, I was able to calculate the atomic scale modulation of the friction force, and thus explain the stick-slip motion of the tip. I found the averaged friction force depends not only on the tip-surface interaction potential, but also strongly on 66 the construction parameters of the FFM. Chapter 4 Structure and Dynamics for H Loaded Pd As I already mentioned in Chapter 1, the structure and dynamics of H loaded Pd is of great importance to both basic science and technology. While the ultimate goal is to perform the corresponding study completely from first principles, the computational involvement is presently prohibitive for such an endeavor. Therefore, I base the study of the Pd—H system on the Many-Body Alloy (MBA) Hamiltonian, which was derived in Chapter 2, with parameters reproducing the results of an LDA calculation for this system [Tom 86b, Sun 89, Tom 91a]. In Section 4.1, I use the Hamiltonian to calculate the structural properties of bulk Pd and PdH, as well as clean and hydrogen covered Pd surfaces. The results of the surface studies, which have no additional adjustable parameters, are found to be in good agreement with corresponding LDA calculations and experimental data. In Section 4.2, I use lattice dynamics to calculate the dispersion relations for bulk Pd and PdH, and compare the results to experimental data. The effect of H adsorbates on the phonon spectra of Pd (001) and Pd (110) surfaces is investigated by performing the corresponding lattice dynamics calculations. In Section 4.3, dynamical properties of Pd-H systems at finite temperature are 67 !. 68 investigated using molecular dynamics. I present results for the equilibrium struc- tural and elastic properties of Pd at different temperatures and hydrogen concentra- tions, and show quantitative results for the mechanical failure and crack formation in hydrogen-free and hydrogen-loaded Pd due to a uniaxially applied load. I will show that a careful analysis of these results can provide valuable insight into the mechanism of “hydrogen embrittlement”. This Chapter contains material which has appeared in the following three publi- cations: e W. Zhong, Y. S. Li, and D. Tomanek, Effect of Adsorbates on Surface Phonon Modes: H on Pd(001) and Pd(110), Phys. Rev. B44, 13053 (1991). e W. Zhong, Y. Cai, and D. Tomanek, Mechanical Stability of Pd-H Systems: A Molecular Dynamics Study, Phys. Rev. B 46 8099 (1992). e W. Zhong, Y. Cai, and D. Tomanek, Computer Simulation of Hydrogen Em- brittlement in Metals, Nature 362 435 (1993). 4.1 Application of the Many-Body Alloy Hamil- tonian to the Pd—H system In this Section, I determine the parameters of the MBA Hamiltonian for the Pd—H system, based on previously published ab initio results[Tom 86b, Sun 89, Tom 91a] for the equilibrium structure and cohesive energy of bulk Pd and PdH. It should be noted that this procedure does not introduce any free parameters. In the following, I will apply this Hamiltonian next to determine structural properties and vibrational spectra of clean and hydrogen covered Pd surfaces. 69 Table 4.1: Interaction parameters used in the Many-Body Alloy Hamiltonian for the Pd-H system. Interaction qag pag rag,o(A) 653(6V) {05(eV) Pd-Pd 3.40 14.8 2.758 0.08376 1.2630 H—H 3.22 5.28 2.300 0.1601 0.9093 H-Pd 2.20 5.50 1.769 0.6794 2.5831 4.1.1 Construction of the Many-Body Alloy Hamiltonian In the MBA Hamiltonian, each of the H—H, H—Pd and Pd—Pd interactions is char- acterized by a set of five parameters: {0,63,24,12 and r0 (four of these parameters are independent). The Pd—Pd interaction is obtained from the previous ab initio calculation[Tom 91a] of the cohesive energy Eco}, as a function of the lattice constant a for bulk Pd. I consider nearest neighbor interactions only and obtain a simplified expression for the bulk cohesive energy, 7‘?de _1)]}1/2 r0,Pde rPde _1)] r0,Pde E...(Pd bulk) = —{ 2.... 53.....epr—2qp.,p.( + 21>qu ‘53.?de eXPl-P( (4-1) Here, Zbuu‘ = 12 for the fcc structure and mpdpd = a\/2/2 is the nearest neighbor distance. The calculated cohesive energy Eco}, of bulk Pd as a function of the lattice constant a is given by the solid line in Fig. 4.1 and compared to corresponding LDA results of Ref. [Tom 91a]. The parameters used in Eq. (4.1) are given in Table 4.1. The parameters for H—H interaction can be determined in a similar way as those for the Pd—Pd interaction, by mapping the MBA Hamiltonian to the ab initio results for molecular hydrogen. The corresponding parameters are given in Table 4.1. To determine the parameters for the Pd—H interaction, I apply the MBA Hamil- tonian to bulk PdH. Considering the nearest-neighbor Pd—H, H—H and Pd-Pd in- teractions, I formally decompose the cohesive energy of the PdH crystal with N aCl 70 0.5 '- 0.4 '- 0.3 AEcoh (eV) 0.2 0.1 0.0 3.7 3.8 3.9 4.0 4.1 4.2 4.3 4.4 4.5 a (A) Figure 4.1: Cohesive energy changes AEcoh = E60,, - Ecohp in bulk Pd and PdH as a function of the lattice constant a. Values obtained using the MBA Hamiltonian for Pd (solid line) and PdH (dashed line) are compared to LDA results of Ref. [Tom 91a] for the corresponding systems, given by e and 0. (From Ref. [Zho 91a]. ©American Physical Society) 71 structure as Ecoh = Ecoh(H) "l" Ecoh(Pd)- (4.2) The binding energies of H and Pd atoms in this structure are given by Ecoh(H) = -{ ZH(Pd)§. The force constants describing the restoring forces acting on a Pd atom displaced along a high-symmetry direction are reduced from c = 10.6 eV/A2 in bulk Pd to only 6.8 eV/A2 in PdH. The effect of the the presence of hydrogen on the bulk modulus is comparably small. I obtain B = 2.03 x 1012 dyn/cm2 for bulk Pd which compares well with the experimental value[Kit 86] B = 1.81 x 1012 dyn/cm2 and the LDA value of 2.15 x 1012 dyn/cm“. The corresponding value in. PdH is only 3% smaller, B = 1.98 x 1012 dyn/cmz, which is in fair agreement with the LDA result 1.95 x 1012 dyn/cm2 of Ref. [Tom 91a]. Since my main interest is the effect of hydrogen on the Pd modes, I do not include the hydrogen-derived optical modes in Fig. 4.5. These modes have a very high fre- quency and are well separated from the acoustic Pd-derived modes due to the mass disparity of these atoms. While the hydrogen atoms can be basically thought of as Einstein oscillators, the weak coupling between hydrogen atoms broadens the opti- cal states to a z 3 T Hz broad band. Since the H—H nearest neighbor distance in PdH 113,1"! = 2.06 A is much larger than the H-H interaction range (is 1 A), I have neglected the direct H-H interaction in this calculation, but have accounted for the dominant indirect interaction mediated by Pd atoms. A closer analysis of the opti- cal bands reveals that two out of the three bands show no dispersion and represent Einstein modes of hydrogen atoms with no direct coupling. The above mentioned dispersion of the third branch reflects the degree of Pd mediated indirect interaction between H atoms, which is almost equally strong for the first and second neighbors. These effects in the optical band can also be seen in the measured phonon dispersion relations of the related systems PdDo,33[Row 74] and PdTo_7[Row 86]. 83 The calculated phonon spectra for the Pd(100) and Pd(110) surfaces are shown in Figs. 4.6 and 4.7, respectively. These surfaces are represented by relaxed 25- layer Pd slabs. Phonon spectra of the clean (001) and (110) Pd surfaces, shown in (a), are compared to results for hydrogen covered surfaces, shown in (b). In order to distinguish surface from bulk states, I are showing phonon bands of bulk Pd in Figs. 4.6(c) and 4.7(c). For the sake of simple comparison, the bulk bands are projected onto the same two-dimensional Brillouin zones in (c) as used in (a) and (b). Note that this procedure yields continuous bulk bands at each iii-point in the Brillouin zone. The apparent discreteness in Figs. 4.6(c) and 4.7(c) results from a finite Iii-point sampling in a direction perpendicular to the Brillouin zone. For both surfaces, my calculations indicate the presence of surface modes which appear either in the bulk band gaps or are split off from the bulk band edges. Based on the comparison of phonon spectra for Pd(001) in Fig. 4.6(a) and the projected bulk spectra in Fig. 4.6(c), the presence of the surface introduces a soft Rayleigh mode 51 which is substantially softer than any bulk mode in the Brillouin zone. At the M point, this mode corresponds to vibrations of surface atoms per- pendicular to the surface. The origin of the mode softening is a decreased interlayer interaction at the surface, which is only partly compensated by the surface contrac- tion. The analysis of my results indicates that the zone-edge frequency of the 51 mode increases from 3.72 THz for the unrelaxed surface to 4.11 THz for the relaxed surface. A second surface mode S.., which is barely split from the bulk band at M, is a transversal mode corresponding to in-plane vibrations. In addition to these soft modes, the surface introduces a phonon mode 56 with u z 6 THz in theigap of the bulk spectrum near X. This mode corresponds to in-plane vibrations of topmost layer atoms, coupled to out-of-plane vibrations of second layer atoms. 84 6.00 ~‘~\§ A ‘ \ 5%) N 4 00 " 3’31}: E ‘ 2‘. l: 3’ 7&3: / ll 2.00 . ’ ., J( '8 I '1 n . >0 ~11 v (T112) fl} ’ 7//// W / 0']. j ,’ ///;-,. ,., ,/tl v (1111) Figure 4.6: (a) Calculated phonon dispersion relations for a clean 25-layer Pd slab with a (001) surface. (b) Corresponding results for a H-covered Pd slab (mono- layer coverage, hollow site). (c) Bulk phonon dispersion relations of Fig. 4.5(a), projected onto the two-dimensional surface Brillouin zone used in (a) and (b). (From Ref. [Zho 91a]. ©American Physical Society) 85 /”*\ A: \ 1“ \ )I A h v Ll ’f 1)) will; \‘i (in. I I 1.! § ’ I fl \ \ l l .\ ... “L \\\ $\ 1 l l /f. . I") I // “(fitfll \ \1 . .O‘A‘lu ll" / ”I I ( i A v('l'Hz) 2.00 '1 “Hill I \ J/ Y a \Q . \\ ‘\ 141112) Figure 4.7: (a) Calculated phonon dispersion relations for a clean 25—layer Pd slab with a (110) surface. The experimental data of Ref. [Lah 87] are given by e. (b) Cor- responding results for a H-covered Pd slab (monolayer coverage, long-bridge site). (c) Bulk phonon dispersion relations of Fig. 4.5(a), projected onto the two—dimensional surface Brillouin zone used in (a) and (b). (From Ref. [Zho 91a]. ©American Physical Society) 86 The phonon spectrum of a hydrogen-covered Pd (001) surface has been shown in Fig. 4.6(b). I have assumed a monolayer coverage corresponding to occupying all hollow sites by H atoms. The Pd surface has again been represented by a 25- layer slab. Similar to the bulk Pd-H system, I am mainly interested in the effect of hydrogen on the Pd surface modes, and do not show the H-derived high-frequency optical modes which are well separated from the Pd modes. The most striking change in comparison to the H-free surface shown in (a) is a massive softening of the Rayleigh mode. At M, the frequency of the 51 mode decreases from 4.11 THz to 2.77 THz due to hydrogen adsorption. On the other hand, hydrogen does harden other modes, such as the bulk band gap mode at u z 6 THz which has been discussed above. A detailed analysis of the eigenstates shows that surface phonon modes with a vibration amplitude perpendicular to the surface experience a large amount of softening. Other surface modes with amplitudes restricted to the surface layer, such as the Se mode, experience a hardening due to the restricted movement in presence of hydrogen atoms. The eigenvector analysis also indicates that hydrogen adsorption also increases the confinement of surface modes to the few topmost layers and shortens the penetration depth into the bulk. As done above for Pd (001), in Fig. 4.7, I compare the phonon spectra of clean and hydrogen-covered Pd(110) surfaces to bulk phonon spectra. For the hydrogen covered surface, I have assumed a monolayer coverage corresponding to the occupation of all long-bridge sites (the equilibrium adsorption site) and represented the Pd surface by a 25-layer slab. A comparison between the calculated phonon spectra for clean Pd(110) and in- elastic He scattering data of Ref. [Lah 87] is shown in Fig. 4.7(a). The softest surface mode is the Rayleigh mode SI. My calculation reproduces the general features of this 87 mode quite well, but the calculated frequencies are z 10% higher than the observed data. While I am aware that the parametrized MBA Hamiltonian may not describe the detailed changes of metal bonding at surfaces to a very high accuracy, I can not exclude the possibility of a slight hydrogen contamination of the Pd sample used in Ref. [Lab 87], which is very hard to detect and would also soften the surface Rayleigh mode. Since Pd(110) shows the largest surface contraction, the softest surface phonon modes are quite close to the lowest lying bulk bands throughout the surface Brillouin zone. The strongest softening can be observed at the I7 point, where I obtain three surface modes 51, S; and E well below the bulk band. My analysis of the eigenstates indicates that these modes correspond to in—plane (along the surface a: and y direc- tions) and out-of-plane (along the z direction) vibrations of .the topmost layer. The lowest mode is an in-plane mode with an amplitude along the y direction. Similar to the (001) surface, the surface contraction generally hardens the surface phonon modes. This calculation shows that at the I7 point, because of the relaxation, the lowest surface modes with topmost layer amplitudes along the a: , y, and z direc- tions are shifted from 2.36 THz, 2.29 THz and 2.77 THz to 2.71 THz, 2.23 THz and 2.94 THz, respectively. From the comparison between Figs. 4.7(a) and (c), one can see that the presence of the (110) surface introduces several other vibration modes beyond the Rayleigh mode. The bulk phonon spectrum, shown in Fig. 4.7(c), contains gaps near X, at u z 5.5 T112, and near 17, at V a: 4.5 THz. On Pd( 110), three surface states appear in both gaps. The 5.4 THz gap mode at X corresponds to topmost layer atoms vibrating along the surface a: direction, and the 4.7 THz gap mode at 1" corresponds to an analogous vibration along the surface y direction. The surface also introduces 88 high frequency states slightly above the top of the bulk bands, well visible at the X and 17 points. These modes have considerable amplitudes deep into the bulk, and correspond to alternating in-plane and out-of-plane vibrations on the individual layers. Calculated surface phonon dispersion relations for the H covered Pd(110) surface are shown in Fig. 4.7(b). As for bulk PdH and H/Pd(001), I do not show the H- derived high-frequency optical modes for H/Pd(110) in Fig. 4.7(c). The general trends discussed above for the effect of hydrogen on the surface modes of the Pd(001) surface hold also for the (110) surface. Specifically, One can observe a softening of surface modes involving an out-of-plane motion of topmost layer atoms, and a hardening of modes involving an in-plane vibration of surface atoms along the a: direction at X and the y direction at 17. Other vibration modes experience very little change. Hydrogen-induced bonding changes at the Pd(110) surface can be understood by investigating the three lowest phonon modes SI, S; and E at the I7 point. The 51 mode, which involves in-plane vibrations of the topmost layer along the surface y direction and out-of-plane vibrations of the second layer, has been pushed up in frequency from 2.23 THz for the clean surface to 2.71 THz for the H covered surface. The main reason for this hardening is a restricted freedom of motion of topmost layer Pd atoms due to hydrogen atoms adsorbed in the long-bridge site. The S; mode, corresponding to a vibration of topmost layer Pd atoms along the close-packed surface :1: direction, has been softened by the presence of hydrogen from 2.71 THz for the clean surface to 2.46 THz for the H covered surface. Finally, the E' mode, corresponding to out-of-plane vibrations of the topmost layer, experiences the strongest softening from 2.94 THz for the clean surface to 2.56 THz for the H covered surface. As compared to the (001) surface, phonon modes at the Pd(110) surface are less 89 affected by hydrogen adsorption. This is especially truefor the softening of the Rayleigh mode. This last conclusion may not hold in case of an adsorption in the quasi-threefold site discussed in the previous Section. In that case, I would expect an increasing Rayleigh mode softening with increasing hydrogen coverage, which has been observed at the F point on the related Ni(110) surface in high—resolution EELS experiments[Leh 87]. From this calculation, one can see that hydrogen adsorption has a profound effect on both the equilibrium structure and dynamical properties of Pd surfaces. These two effects are closely related. The presence of hydrogen adsorbates reverses the topmost layer contraction, and the large anharmonicity of the inter-layer interactions causes the effective inter-layer force constant to decrease with increasing inter-layer distance. This results in a softening of the Rayleigh surface phonon mode with an out-of~plane amplitude in the topmost layer. Since the hydrogen-induced expansion is larger on the more densely packed (001) surface than on the (110) surface, the Rayleigh mode softening is also stronger on the Pd(001) surface. At both Pd(001) and Pd( 110) surfaces, the equilibrium adsorption height of hy- drogen is close to zero, i.e. hydrogen atoms are buried inside the topmost Pd layer, where they affect the effective Pd-Pd force constants most. There is very little sur- face stress associated with this adsorption site, due to the small size of the H atoms. Surface phonon modes with in-plane vibration amplitudes are nearly independent of surface relaxations, but are affected by the presence of hydrogen. This latter effect results from the hydrogen-induced change of Pd—Pd force constants, and also the restricted freedom of motion due to the extra adsorbed atoms which hardens some of these modes. 90 4.3 Mechanical stability of Pd-H systems 4.3.1 Thermal expansion and the melting transition in Pd The first nontrivial application of the present formalism is to determine the equilib- rium volume of Pd as a function of temperature, and to study the melting transition. In the corresponding molecular dynamics simulation, I consider an originally cubic volume (or simulation box) containing 500 Pd atoms on an fcc lattice. Periodic boundary conditions are used to eliminate surface effects, and the external pressure is set to zero. I start the simulation by first equilibrating the system for 10,000 time steps (corresponding to 2 x 10'113) in a VTN canonical ensemble which is in con- tact with a heat bath at T = 300 K. This simulation is followed by another 20,000 time steps in the T tN canonical ensemble at zero external pressure. The equilibrium volume V per atom is related to the volume of the simulation box, and is obtained from a running average during the simulation. The statistical error based on the last 10,000 time steps is found to be negligible. After V has been determined for the given temperature, the heat bath is gradually warmed up to another temperature, and the equilibrium volume per atom is obtained from a statistical average over 10,000 time steps at the new temperature. My results, presented in Fig. 4.8(a), indicate a thermal expansion of the lattice in the whole temperature range studied. In the temperature range between T = 2000 - 2050 K, the equilibrium volume per Pd atom shows a discontinuous increase, indicative of a first-order phase transition. This phase transition corresponds to the melting point, as can be verified by inspecting the trajectories of individual Pd atoms at temperatures below and above the critical temperature, (shown in Figs. 4.9(b) and 4.9(c). Near the melting point, my calculations show a narrow hysteresis describing 91 N .... N O H CO I \ L H (D U .... K) I : ,] Volume (A3) H O f/ 1 iT\1(theory) TM(eth ) H (I 14 1 L l l l 0 500 1000 1500 2000 2500 3000 T (K) 18.0 17.0 Volume (Aa ) ... 9‘ O L 14.0 ‘ L 0.0 0.2 0.4 0.6 X 0.8 1.0 Figure 4.8: (a) Equilibrium volume V of a Pd atom in bulk Pd as a function of temperature T. The dashed line is a guide to the eye connecting the calculated data points. Observed linear expansion of Ag[Pea 58], the neighboring element of Pd, is shown by the dotted line. (b) Equilibrium volume V per Pd atom in bulk PdH,, as a function of the hydrogen concentration 2:. The dotted line corresponds to the observed[Pei 78] expansion coefficient of Pd as a function of 0:. Results of the present molecular dynamics simulations are given by s in (a) and (b). (From Ref. [Zho 92]. ©American Physical Society) 0015345070 ' ' = 3001( r (A) °oiééiééiu 92 r (A) 0 T = 205le' 012345878 r (A) Figure 4.9: Atomic trajectories in bulk Pd at the temperatures (a) T = 300 K and (b) T = 1800 K below the melting temperature TM, and (c) at T = 2050 K above TM. Corresponding pair correlation functions g(r) at T = 300 K, T = 1800 K and T = 2100 K are shown in (d), (e), and (f), respectively. (From Ref. [Zho 92]. ©American Physical Society) 93 an overheated solid or an undercooled liquid, depending on whether the system is being heated or cooled. The small difference of S 10% between the calculated melting temperature and the experimental value[Kit 86] of TM = 1827 K is impressively small in view of the fact that the interaction potentials are based on T = 0 static properties of Pd. A small positive difference between the calculated and the observed melting point is also expected due to the finite size of the unit cell and absence of large defects or a surface. In the temperature range T < 1500 K, the thermal expansion of the lattice is nearly linear, corresponding to a thermal (linear) expansion coefficient of a; = 1.7 x 10"5 K". While I could not find any corresponding experimental results for Pd, this value lies very close to a; = 1.89 x 10‘5 K‘1 which has been observed in Ag[Pea 58], a neighbor of Pd in the periodic system. At higher temperatures, the volume-temperature relation shows strong deviations from linearity. This relation- ship bears information about the Pd-Pd interaction potentials at large interatomic separations. Above the melting point, the expansion coefficient of the system experi- ences an abrupt increase above the solid phase value, correlated with a sharp increase of the interparticle distance at TM. In order to illustrate the effect of temperature on the dynamical behavior of the system, I showed the trajectories of individual Pd atoms [projected onto the (100) plane] in Figs. 4.9(a)-(c) at different temperatures. At T = 300 K, all atoms appear to be pinned to their equilibrium site and show negligible fluctuations about this po- sition, reflected in a very small Debye-Waller factor. At T = 1800 K, the fluctuations of Pd atoms about their equilibrium sites are already quite appreciable. The size of these fluctuations is a considerable fraction of the lattice constant, which, according to Lindemann’s criterion, is indicative of the proximity to the melting point. A statis- 94 tically negligible fraction of atoms is also seen to diffuse faraway from their site, but the crystalline order still exists. Above the melting point, all atoms diffuse relatively freely throughout the sample, as shown in Fig. 4.9(c), and the long range order is destroyed. A more quantitative measure of the crystalline order is the pair correlation func- tion g(r) which is shown in Figs. 4.9(d)—(f) for the above temperature values. Clearly, the validity of conclusions related to the long-range order in the crystal is limited by the finite size of the simulation box. At T = 300 K, g(r) consists of a set of sharp peaks which will result in a sharp diffraction pattern. At T = 1800 K, g(r) still shows a substantial amount of structure even for large interatomic distances r, indicative of long-range order, but the peaks are smeared out and begin to overlap. Above the melting transition, the most characteristic feature in the pair correlation function is a peak at the nearest-neighbor distance. For larger interatomic distances, g(r) contains almost no information about the atomic structure and approaches the constant value g(r) = 1 very rapidly. This would justify a treatment of the liquid as a continuous medium beyond the nearest-neighbor distance. As I will discuss later on in connection with fracture, interesting details about the cohesion of the system under different conditions can be learned from the distribution of atomic binding energies. In Figs. 4.10(a) and (b), I compare the binding energy distribution for bulk Pd atoms at T = 300 K and T = 2100 K, just above the melting point. In absence of applied tensile stress, my simulations indicate that binding energies in the system can be characterized by a rather featureless single- peaked distribution function indicating that most atoms have an indistinguishable environment. The small width of this peak at T = 300 K, shown in Fig. 4.10(a), reflects an almost perfect crystalline order at this temperature. Above the melting 95 T ? 25 :0 ‘5' 20(3) ‘80 f-3 'E 33. € 15- .50 E ‘3 10 ' -40 ‘5. 3: :- ‘3 5' -2o .: s ] E '3', 0 1 0 1 :. 20 25 3.0 3.5 40 31/ Eeoh(ev) 1 3 10 ,- 100 _., 1b) - .3 81' ‘1‘ .80 E d ' . ”Q 2 I E 6- «60 E- .3. : ' é : 1 ‘r '1' 1‘0 5 3: 'o 1 5' = .‘ :_-_ :5 2 . I. 7'20 E— 3, 0, #0 J: ‘5. 2.0 . . 4.0 3 T 0 :1 (C) 100 g 8) .30 E- ‘3 a: :9. s- ‘00 3'- :53 2 ‘3 4* «10 "P. .. s- : w 3 2» +20 9,: 0 or a '3 '5. 2.0 2.5 3.0 3.5 4.0 3 Each (0") Figure 4.10: Distribution of atomic binding energies Eco). in bulk Pd at (a) T = 300 K and (b) T = 2100 K at zero pressure. (c) Corresponding data for Pd at T = 300 K at the point of critical uniaxial tensile stress pc for fracture, discussed in Section 4.3.2 The solid lines give the probability distribution, the dashed lines show the integrated probability. (From Ref. [Zho 92]. ©American Physical Society) 96 point, this peak is strongly broadened and shifted towards smaller binding energies, as shown in Fig. 4.10(b). The former eflect comes from the large variety of binding sites in the liquid, the latter one reflects the loss of cohesion mainly due to a uniform lattice expansion. 4.3.2 Mechanical stability of H-free Pd under tensile stress One of the most challenging problems related to the mechanical stability of metals is to obtain a quantitative understanding of the fracture process due to uniaxially applied tensile stress. I address this problem by studying the deformation of a metal block (the simulation box) under a uniaxial load, as shown schematically in Fig. 4.11(a). I expect the length 2 of the metal block first to increase monotonically with increasing load. Once a critical value of the tensile stress pc is reached, the material can no longer support the load and breaks into parts. In the following, I describe the molecular dynamics simulation of the elastic and plastic deformations in the Pd metal block which is exposed to increasing uniaxial tensile stress, as a function of temperature and - in the following subsection - as a function of hydrogen concentration. In the present MD simulation, I consider a canonical (TtN) ensemble of 500 Pd atoms in a simulation box which has cubic shape at zero applied stress, but which can vary its shape freely in the case of anisotropic pressure. For a given temperature of the heat bath, I determine the shape changes [especially the elongation of the cell Az, see Fig. 4.11(a)] in response to uniaxial tensile stress 'which is described by the stress matrix (4.6) GOO ”606 0 2= 0 0 Results for the elongation A2 of the MD unit cell as a function of p are shown in 97 .0 A A A A A 0.0 2.5 5.0 7.5 10.0 12.5150 p(GPa) ' - 5". r3200 (0)1151 - rte-01‘ a'osrrz =0. 25 10.0’ t v ‘ 0. A . L A . 00.0 2.8 8.0 7.8 10.0 12.8 18.0 9(GPn) Figure 4.11: (a) Schematic picture of Pd deformations under tensile stress, showing the length 2 and deformation A2 of the unit cell. (b) Deformation A2 of bulk Pd as a function of the external tensile stress p, for different temperatures. (c) Deformation A2 of hydrogen loaded bulk Pd at T = 300 K, as a function of the external tensile stress p, for different hydrogen concentrations. The molecular dynamics results are given by the data points. The lines are guides to the eye. (From Ref. [Zho 92]. ©American Physical Society) 98 Fig. 4.11(b) for temperatures between 77 K and 1800 K. In this simulation, I increased the uniaxial stress p in finite steps of initially Ap = 2.0 GPa from zero to a load just below the point of fracture. From there on, I decreased the steps Ap to 0.5 — 1 GPa in order to increase the accuracy of the calculated critical tensile stress pc. At each value of p, the system has been allowed typically 5,000 time steps (or 10-113) to equilibrate. I observed that the equilibration took longer near the point of fracture and extended the simulation accordingly. The equilibrium shape of the MD unit cell has been obtained by averaging over the last 5,000 time steps. The first important result of this simulation is the order of magnitude for the criti- cal tensile stress pc to initiate fracture, typically a few GPa. As shown in Fig. 4.11(b), pc decreases with increasing temperature, from 11 GPa at T = 77 K to 1 GPa at T = 1800 K. On the other hand, I find an increase in the plasticity, given by Bz/ap, with increasing temperature at constant load. I conclude that the Young’s modulus Y, defined by Y = 6p/(92, decreases as the temperature rises. Both effects indicate that the material becomes softer and easily deformable with increasing temperature, which agrees with the everyday experience. Microscopically, this softening corresponds to the increased probability of activated atomic diffusion leading to a new equilibrium geometry (plastic deformations to a “thin wire” and fracture under excessive uniaxial load). In the elastic region p << pc, the Az-p relationship is nearly linear for all tem- peratures. This Hooke’s law type behavior is expected based on the MBA interaction potential which is dominated by harmonic terms close to the equilibrium. It is interesting to note that the above MD simulations, performed under uniaxial stress, contain the information about the bulk modulus B which describes the elastic response to isotropic pressure. For a cubic crystal, the elastic response to a very small 99 applied uniaxial stress p is p = 01161 + 201262 (4.7) along the direction of the load. Here, 61,62 are the strain components along the load direction and perpendicular to the load direction, respectively, and C11, C12 are elastic stiffness constants. No stress occurs perpendicular to the load direction [see Fig. 4.11(a)], so 0 = 0126] + 01162 + 01262 . (4.8) From Eqs. (4.7) and (4.8) one can derive V(p) - W? = 0) p = (011+ 2012) V(p = 0) (4.9) Using B = (1/3)(C11 + 2C12) for a cubic crystal leads to -1 B = 1V (a!) . (4.10) 3 6p F0 As shown in Fig. 4.12(a), the present results for the absolute value of B and the temperature dependence of B are in good agreement with the experimental data of Ref. [Lan 79]. At low temperatures, the bulk modulus is found to be essentially independent of temperature. With rising temperature, however, the present results indicate a strong decrease of B. I found it instructive to inspect the distribution of binding energies for a signature of atomic fracture at very large tensile loads. In Fig. 4.10(c), I show the distribution at the point of critical tensile stress pc, for a developed fracture. As compared to the stress-free situation shown in Figs. 4.10(a) and (b), the distribution of binding energies shows several distinctive peaks. The lowest binding energies correspond to 100 200 p. a (a) ...: ----------------- . 160 ' ~ ”;~ ~: 140 ~ xx] 7 A 120 " “\ 7 (U ‘s l 0- 100 ~ ‘~ 4 8 \ I m 80 " “x ‘i 60 " “9, '1 40 r i‘: 20 r o L L 1 0 500 1000 1500 2000 T00 200 (b) 190 180 170 180 180 140 130 120 110 ' loo 1 1m L 1m 1 l L l l 0.0 0.2 0.4 0.8 0.8 1.0 Figure 4.12: (a) Temperature dependence of the bulk modulus B of Pd. The dashed line is a guide to the eye connecting the calculated points, given by e. The dotted line shows the experimental data of Ref. [Lan 79]. (b) Dependence of the bulk modulus of PdH, on the hydrogen concentration 2:. (From Ref. [Zho 92]. ©American Physical Society) 101 sites at the surface of the crack. The highest binding energies, same as in the stress- free sample shown in Fig. 4.10(a), correspond to bulk sites in the intact fragments. 4.3.3 Mechanical stability of H loaded Pd under tensile stress Hydrogen is well known to dissociate at transition metal surfaces and to penetrate easily into the bulk metal, releasing the heat of hydride formation in many systems such as Pd[Ale 78]. This process makes such metals an ideal medium for hydrogen storage. On the other hand, the presence of hydrogen is known to have an adverse effect on mechanical properties of metals, specifically facilitating the formation of cracks under tensile stress[Bir 79]. In order to obtain a microscopic understanding of the processes associated with hydrogen-assisted crack formation, I simulated the response of hydrogen loaded Pd to uniaxial tensile stress in a molecular dynamics calculation. I have chosen a cubic simulation box containing 500 Pd atoms as the initial MD unit cell, and occupied the 500 octahedral interstitial sites in the lattice at random with hydrogen atoms. I have considered three different H concentrations in Pde, namely :1: = 0.1, :1: = 0.25, and :c = 0.6, and performed all simulations at room temperature, T = 300 K. At each H concentration, I first let the system equilibrate over a period of more than 30,000 time steps (corresponding to 1.5 x 10‘113). First, I study the volume changes due to hydrogen at zero pressure. The free vol- ume V(z) of Pd atoms in PdH,” shown in Fig. 4.8(b), has been determined for each H concentration a: by averaging over 10,000 time steps after reaching the equilibrium. I found the free volume to increase almost linearly with increasing hydrogen concen- tration x, which agrees with the observed relation[Pei 78] V(x) = V(0)(1 + 0.192;). 102 In order to understand the effect of dissolved hydrogen on the mechanical stability of Pd, I studied the response of the system to uniaxial tensile stress using molecular dynamics. As in the hydrogen-free samples, the system was first allowed to equilibrate under fixed external tensile stress p and constant temperature of the heat bath T = 300 K. Then, the stress-induced elongation A2 of the simulation box was determined by averaging over 10,000 time steps (corresponding to 5 x 10’125). The results, presented in Fig. 4.11(c), show an almost linear relationship between p and A2. I found the slope of the Az(p) curves to be almost independent of a: at low values of p, indicating that changes of the hydrogen concentration have little effect on the Young’s modulus Y. These results also indicate that the critical tensile stress pc for the onset of fracture, corresponding to the “end points” of the Az(p) curves, decreases with increasing hydrogen concentration. I found that hydrogen can reduce the critical tensile stress for fracture pc substantially when compared to the hydrogen- free system, but that the order of magnitude of pc in the different systems is the same. This hydrogen-induced reduction of the mechanical strength is sometimes called “hydrogen embrittlement”[Bir 79]. As I will discuss later on, the microscopic results indicate that this is a misnomer; I find hydrogen to enhance the ductility and plasticity of the metal matrix locally, thereby weakening the structure as a whole. One can also use the MD results for uniaxial tensile stress to estimate the bulk modulus of the system, following the procedure outlined in the previous subsection. The results, presented in Fig. 4.12(b), indicate that the bulk modulus B decreases strongly in the presence of hydrogen. Another quantity of interest is the Poisson’s ratio )1 which is the ratio of the unit cell deformations along the direction of the applied load, A2, and perpendicular to it, Am. In a cubic system, )1 relates the 103 Young’s and the bulk modulus as B = Y/ (3 — 6p), which defines p as L 63' 11 = (4.11) NIH My above results for Y(2:) and B(:r) indicate that the Poisson’s ratio decreases strongly with increasing hydrogen concentration in the metal. As in the hydrogen—free system, I investigated the distribution of binding energies for a signature of atomic fracture at different tensile loads. My results for the PdHo,25 system are summarized in Fig. 4.13(a) for zero tensile stress and in Fig. 4.13(b) for critical tensile stress. The structure of the binding energy distribution in the stress- free case reflects the distribution of inequivalent Pd atoms with 0—6 hydrogen nearest neighbors. The relatively featureless distribution of binding energies at the point of critical tensile stress p = pc, displayed in Fig. 4.13(b), is in sharp contrast to the hydrogen-free case shown in Fig. 4.10(c), and is more reminiscent of the results for molten Pd, shown in Fig. 4.10(b). The absence of distinct fracture-related features in the binding energy distribution in Fig. 4.13(b) indicates that all Pd atoms reside in a relatively homogeneous atomic environment. This environment is close to the molten system; it has an amorphous structure and can easily be plastically deformed. I conclude that increased hydrogen concentration has a similar effect on the mechanical properties of Pd as a temperature increase, namely enhanced ductility and plasticity. An independent microscopic signature of fracture on the atomic scale can be of- ten found in the distribution of Wigner-Seitz volumes associated with the individual atoms in the crystal. Statistical presence of large atomic volumes indicates the oc- currence of fracture with no additional assumptions about the number, position or morphology of one or more simultaneous cracks. This information is displayed in Fig. 4.14 for Pd and H atoms in PdHogs at p = pc. The volumes of Pd atoms, given 104 '> 25] :7 l (8.) :° ‘5‘ 20] z- : [ a-s. : '3 73 15: 5;: :2 l f: “'3 10 . .3 3» 5- :: l. a: '3 5 E _g :2 a 0 ‘< a. 2.0 2.5 3.0 E:eoh (8V) T E 10 (b) I, 100 5 8 . 5 80 § ‘3 : s g 6- 5 ‘60 g» 72‘ 5 °‘ 1 4 L ‘40 E 3: E— : 2 _ ‘ 0a 3 , 20 ___c-_: s :- g o - e . 0 7‘ ca. 2.0 2.5 3.0 3.5 4.0 Ecoh (8V) Figure 4.13: Distribution of binding energies of Pd atoms in bulk PdHo_25 at T = 300 K (a) at zero pressure and (b) at the point of critical uniaxial tensile stress pc [compare with Fig. 4.11(c)]. The solid lines give the probability distribution, the dashed lines show the integrated probability. (From Ref. [Zho 92]. ©American Physical Society) 105 .0 on 0 fl 1 .0 10 CI 0.15 0.10 0.05 0.00 Volume distribution (A’s) Figure 4.14: Distribution of atomic volumes in bulk PdHo,25 at T = 300 K. The solid line shows the distribution of atomic volumes associated with Pd atoms, obtained using the Wigner-Seitz cell construction. The dashed line is obtained by first deter- mining, which Pd sites are associated with each of the H atoms in the lattice, and displaying the distribution of these Pd volumes. The dotted line gives the ratio of the values given by the dashed and the solid lines, divided by the volume and multiplied by 3. (From Ref. [Zho 92]. ©American Physical Society) 106 by the solid line, show a distribution which is characterized by a sharp peak near 15 A3, corresponding to atoms in bulk-like environment, and a wide structureless tail towards larger volumes, associated with atoms near the crack surface. This finding again confirms the previous conclusion that Pd atoms have no preferential binding arrangement in the hydrogen loaded sample, which shows a plastic behavior under critical tensile stress. The preferential hydrogen sites in this structure are determined in the following way. I enlarged the original Wigner-Seitz volumes of Pd atoms by a factor of 1.5 in each direction and for each hydrogen atom in the crystal, I generated a list of Pd sites which contain this atom] in their enlarged unit cell. This definition allows for more than one Pd site to be associated with a given hydrogen atom. Next, I combined the lists of Pd sites associated with each H atom and plotted the distribution of the corresponding Pd Wigner-Seitz volumes. The results are given by the dashed line in Fig. 4.14. A comparison with the solid line for the Pd atoms shows no strong preference of hydrogen atoms for specific sites in the metal structure. The dotted line in Fig. 4.14 represents the probability that a Pd atom, characterized by its atomic volume, is likely to have one or more hydrogen atoms as its closest neighbors. In case that there would be no preferential sites for hydrogen atoms, this curve should be flat. These results indicate a strong preference of hydrogen atoms for sites near highly coordinated Pd atoms in the bulk. From the position of the peak in the dotted curve, which lies at a slightly larger volume than that of bulk Pd atoms, one can infer that hydrogen atoms are more likely to occupy subsurface sites or sites close to the crack tip than sites in the bulk of Pd. 107 4.3.4 Discussion The above molecular dynamics calculations, based on the N osé and Rahman-Parrinello formalism, suggest a useful and consistent picture of the atomic-scale processes which occur in PdH, at different temperatures and hydrogen concentrations as, specifically in response to large uniaxial tensile stress. Even though the interaction potentials are based on static ab initio calculations at T = 0, the finite temperature results of these simulations are in good agreement with experimental data. This increases my confidence that the MBA Hamiltonian describes the interactions in the Pd—H system correctly to a large degree. Since this Hamiltonian has a solid theoretical background, it can provide microscopic insight into the nature of interatomic interaction under dif- ferent conditions. These simulations show that the introduction of hydrogen into bulk Pd at room temperature has a similar effect on the structural properties and elastic behavior as a temperature increase in the hydrogen-free metal. Increased hydrogen loading and increased temperature both increase the free volume linearly, decrease the bulk modulus, and reduce the critical tensile stress for fracture. It is plausible to some degree that the presence of hydrogen simulates a temperature increase, since the light H atoms have a much faster dynamics that Pd atoms and can easily excite Pd vibrations in elastic collisions. This effect is enhanced by the fact that hydrogen atoms reduce the bonding strength between Pd atoms and soften the vibrational modes of the Pd lattice[Zho 91a]. I found that hydrogen-induced changes of structural properties and elastic response become more pronounced with increasing hydrogen concentration, as shown in Figs. 4.8(b), 4.12(b) and 4.11(c). At the point of fracture, I found that the presence of hydrogen enhances the plasticity of the system significantly, causing hydrogen-assisted “melting” even at T = 300 K, which is associated with a 108 substantial diffusion of Pd atoms. I found it useful to‘compare the present results for hydrogen loaded Pd under critical tensile stress to hydrogen-free Pd at the melting point in an animated video movie, and found Pd diffusion at the crack surface of the hydrogen loaded system to be comparable with the atomic diffusion in melting Pd metal. As mentioned above, a fundamental difference between the effect of hydrogen loading and temperature increase lies in the fact that hydrogen modifies the interac- tion between Pd atoms. This difference is most obvious in the response to uniaxial tensile stress, shown in Figs. 4.11(b) and 4.11(c). In the hydrogen-free system, one can observe both the bulk and the Young’s modulus to decrease with increasing tem- perature, while the Poisson’s ratio p is nearly constant, consistent with Eq. (4.11). On the other hand, increasing the hydrogen concentration in Pd at a constant tem- perature T = 300 K still causes the bulk modulus to decrease, but has little effect on the Young’s modulus. In this case, the Poisson’s ratio decreases with increasing hy- drogen loading, indicating an increasing resistance against shape deformations. This effect could also assist in the initial formation of cracks. Based on the results shown in Fig. 4.11(c), I find that the critical value of tensile stress pc drops at an increasing rate with increasing hydrogen concentration, indicating a continuous ductility increase in the system with increasing number of H atoms in the vicinity of the Pd metal bonds. The microscopic origin of the ductility increase of Pd in presence of hydrogen is the rehybridization of Pd orbitals in the hydrogen loaded metal. As was found in a previous ab initio calculation of atomic binding in bulk Pd and PdH[Sun 89], the hydrogen-induced binding changes of Pd stem mainly from a filling of antibonding states in the Pd4d band, accompanied by a depletion of the partly filled Pd5s band and a small charge transfer towards hydrogen. These effects are considered, albeit ...“, 109 in a very approximate way, in the many-body alloy Hamiltonian. Unlike in models based on pairwise interactions, this Hamiltonian does address the different electronic hopping processes to neighboring atoms from the point of view of electronic band formation in the attractive part of the total energy. Hence the binding energy of a Pd atom is not simply proportional to the number of nearest neighbors, but depends in a more complex way on the hybridization with the neighboring atoms and the band filling, which appears to be essential for the understanding of bonding changes in this system. The calculated values for the critical tensile stress are about one order of mag- nitude too high when compared to experimental data for single- and polycrystalline samples. There are two reasons for this overestimate of pc. First, it is impossible to study the dynamics of the fracture process in a molecular dynamics calculation for realistic time scales and large systems. The time spans presently accessible by such simulations fall at least ten orders of magnitude short of a realistic time for the formation of a crack, which is typically seconds. Second, in a realistic system, the fracture process is assisted and proceeds by dislocation motion in the crystal. The presently used unit cell containing 500 Pd atoms, which is large for MD standards, is clearly too small to show the spontaneous formation and motion of dislocations. I also tried to address the effect of dislocations in a somewhat artificial way, namely removing a single atom from the unit cell, thus creating a defect site and a possible seed for a dislocation or a crack. The corresponding MD simulation did not show a reduction of pc when compared to the initially “perfect” systems. This is not surprising since the typical atomic fluctuations near the point of fracture are large and comparable to the size of a single atomic vacancy. In hydrogen loaded Pd, the local decohesion and structural relaxations are medi- 110 ated by hydrogen which has a very large diffusion constant. I studied the interplay between the time scales for structural changes and the diffusion of hydrogen by con— sidering isotope substitution. I found that replacing H by D atoms in PdH,c has no effect on pc, probably due to the large difference between the time scales for hydrogen motion and structural relaxation, and possibly also the incoherent motion of hydrogen atoms in the metal which averages out local changes of elastic properties. The above microscopic results for the elastic response of hydrogen-loaded Pd to uniaxial tensile stress strongly support one of the previously postulated mecha- nisms for “hydrogen embrittlement”, namely the Hydrogen Enhanced Local Plasticity (HELP) mechanism[Bir ]. This mechanism, which has been used to interpret exper— imental data, postulates that hydrogen concentrates preferentially near the tip of a starting crack. Hydrogen subsequently locally softens the metal matrix in the vicinity of the crack tip, which leads to an increase in the velocity of dislocation motion. This process can lead to a softening over microns over very short time scales. The system will become microscopically ductile, but will appear as brittle on macroscopic length scales. While the present calculations have been performed for a specific system, namely hydrogen loaded Pd, I expect that the effect of hydrogen on the stability of other fcc metals will be qualitatively the same. The situation in bcc metals (such as Fe) may be somewhat different, since hydrogen is observed to stiffen rather than soften these structures[Bir ]. These questions are presently being addressed in corresponding MD simulations[Zho 93b]. 111 4.4 Summary and conclusions In conclusion, I applied the Many-Body Alloy (MBA) Hamiltonian to Pd—H systems. All parameters have been obtained from ab initio Density Functional calculations, with no adjustable parameters for surface properties. I tested this Hamiltonian first and calculated the equilibrium structure and binding energy of bulk Pd and PdH, as well as H-free and H covered Pd(001) and Pd(110) surfaces. I found the calcu- lated results to be in good agreement with experimental data and results of ab initio calculations where available. Next, I studied the effect of hydrogen on the vibration spectra of bulk Pd and the Pd(001) and Pd(110) surfaces, by constructing the dynamical matrix based on the MBA Hamiltonian. I found that in the bulk systems, hydrogen softens the Pd vibration modes, as seen in the comparison of bulk Pd and PdH phonon spectra. The results for the clean and H covered (001) and (110) surfaces of Pd are not as clear-cut. I found the most pronounced eflect of hydrogen coverage to be the softening of the surface Rayleigh mode with out-of-plane vibration amplitudes on the topmost layer. Other surface modes, such as in-plane vibrations of the topmost layer, are affected to a lesser degree or occur at higher vibration frequencies in the presence of hydrogen. Finally, I have used the N osé and Rahman-Parrinello molecular dynamics formal- ism to study the equilibrium structure and elastic properties of bulk Pd as a function of temperature and hydrogen concentration. I have used this formalism first to predict the elastic constants, thermal expansion and melting temperature of hydrogen-free and hydrogen loaded bulk Pd. I found my results to be in good agreement with available experimental data. Introducing uniaxial tensile stress as an independent variable into this formalism 112 has enabled me also to study the elastic deformations as a function of the applied load at different temperatures and hydrogen concentrations. At small applied loads, I found that the bulk and the Young’s moduli decrease with increasing temperature in hydrogen-free bulk Pd. Increased hydrogen concentration at constant temperature has a very similar effect as the temperature increase in the hydrogen-free metal: The system gets softer, which is reflected in a decreased bulk modulus. While hydrogen softens the Pd-Pd bonds, its presence does not affect the Young’s modulus. Conse- quently, the Poisson’s ratio decreases with increasing hydrogen loading, indicating an increasing resistance towards shape deformations. This behavior might assist in the formation of cracks. At large values of the uniaxial tensile stress, one can observe the onset of crack for- mation. I find that the critical tensile stress for fracture decreases both with increasing temperature and increasing hydrogen concentration. Near the point of fracture, how- ever, the elastic response of hydrogen-free and hydrogen-loaded Pd is vastly different. Following the fracture, Pd atoms can be found in well-defined “crystalline” sites in the fragments or at the crack surface for the Pd system which is free of hydrogen. In hydrogen-loaded Pd, the metal structure can be called amorphous at the point of fracture. One can find a broad distribution of Pd sites in this system which can easily be deformed plastically. I conclude that the hydrogen-induced reduction of the mechanical stability of Pd (and likely also other metals) originates from an increased ductility and plasticity in parts of the sample with a large hydrogen concentration, such as regions near grain boundaries and dislocations. These conclusions agree with one of the previously postulated mechanisms for “hydrogen embrittlement”, namely the Hydrogen Enhanced Local Plasticity (HELP) mechanism[Bir ]. More detailed studies will be necessary to confirm this behavior also in bcc metals (such as iron). A 113 comparison with corresponding experimental data, stemming from atomic resolution studies of single crystals would be highly desirable. Bibliography [Abr 89] F.F. Abraham and LP. Batra, Surf. Sci. 209, L125 (1989). [Alb 87] T.R. Albrecht and CF. Quate, J. Appl. Phys. 62, 2599 (1987). [Ale 78] G. Alefeld and J. Valkl, Eds., Hydrogen in Metals I and II, Vols. 28 and 29 of Topics in Applied Physics (Springer—Verlag, Berlin, 1978). [A1190] M.P. Allen and DJ. Tildesley, Computer Simulation of Liquids (Oxford Press, New York, 1990). [And ] M. Anders and C. Heiden (submitted for publication). [And 80] H.C. Andersen, J. Chem. Phys. 72, 2384 (1980). [Arn 66] RD. Arnell, J.W. Midgley, and D.G. Teer, Proc. Inst. Mech. Eng. 179, 115 (1966). [Bac 82] GB. Bachelet, D.R. Hamann, and M. Shliiter, Phys. Rev. B 26, 4199 (1982). [Bar 85] C.J.Barnes, M.Q. Ding, M. Lindroos, R.D. Diehl, and DA. King, Surf. Sci. 162, 59 (1985). [Bas 89] R. Bastasz, T.E. Felter, and WP. Ellis, Phys. Rev. Lett. 63, 558 (1989). [Bat 88] LP. Batra and S. Ciraci, J. Vac. Sci. Technol. A 6, 313 (1988). 114 115 [Bea 72] CD. Beachem, Metall. Trans. 3, 437 (1972). [Beh 80] R.J. Behm, K. Christmann and G. Ertl, Surf. Sci. 99, 320 (1980). [Bes 87] F. Besenbacher, I. Stensgaard and K. Mortensen, Surf. Sci. 191, 288 (1987). [Bin 82] G. Binnig, H. Rohrer, Ch. Gerber and E. Weibel, Phys. Rev. Lett. 49, 57 (1982). [Bin 86] G. Binnig, C.F. Quate and Ch. Gerber, Phys. Rev. Lett. 56, 930 (1986), and Appl. Phys. Lett. 40, 178 (1982). [Bin 87] G. Binnig, C. Gerber, E. Stoll, T. R. Albrecht, C. F. Quate, Europhys. lett. 3, 1281 (1987). [Bir 73] H. K. Birnbaum, M. Grossbeck, and S. Gahr, in: Hydrogen in Metals, M. Bernstein and A. Thompson, eds., A.S.M. Metals Park, Ohio, 1973, p. 303. [Bir 79] H.K. Birnbaum, in: Environmentally Sensitive Fracture of Engineering Ma- terials, Z. A. Foroulis, eds., T.M.S. New York, 1979, p. 326. [Bir 84] H.K. Birnbaum, J. Less Common Metals 103, 31 (1984). [Bir ] Howard K. Birnbaum, private communication. [Bob 90] K.P. Bohnen, Th. Rodach, and KM. H0, in The Structure of Surfaces [11, edited by M.A. Van Hove, K. Takanayagi, and X. Xie (Springer-Verlag, Hei- delberg, 1990), and references cited therein. [Bor 84] V. Bortolani, G. Santoro, U. Harten, and JP. Tonnies, Surf. Sci. 148, 82 (1984). 116 [Bor 85] V. Bortolani, A. Franchini, F. Nizzoli, and G. Santoro, Surf. Sci. 152/153, 811 (1985). [Bre 89] D.W. Brenner, Phys. Rev. Lett. 63, 1022 (1989). [Cal 72] Langlinais and J. Callaway, Phys. Rev. B 5 124 (1972). [Cal 84] J. Callaway and NH. March, Solid State Phys. 38, 135 (1984). [Car 90] A.E. Carlsson, in Solid State Physics, Vol. 43, Academic Press (New York, 1990), p. 1. [Cat 83] M.G. Cattania, K. Christmann, V. Penka and G. Ertl, Gazz. Chim. Ital. 113,433 (1983). [Cep 80] D.M. Ceperley and B.J. Alder, Phys. Rev. Lett. 45 566 (1980). [Cha 73] DJ. Chadi and M.L. Cohen, Phys. Rev. B 8, 5747 (1973). [Cha 86a] C.T. Chan, D. Vanderbilt, and S. G. Louie, Phys. Rev. B 33, 2455 (1986). [Cha 86b] C.T. Chan, D. Vanderbilt, S.G. Louie and JR. Chelikowsky, Phys. Rev. B33, 7941 (1986). [Che 84] J. R. Chelikowsky and S. G. Louie, Phys. Rev. B 29, 3470 (1984). [Cir 90] S. Ciraci, A. Baratoff and LP. Batra, Phys. Rev. B 41, 2763 (1990). [Dav 83] H.L. Davis, J.R. Noonan, Surf. Sci. 126, 245 (1983). [Daw 84] MS. Daw and M.I. Baskes, Phys. Rev. B 29, 6443 (1984). [Daw 89] MS. Daw, Phys. Rev. B39, 7441 (1989). [Doa 83] RB. Doak, U. Harten and JP. ’Toennis, Phys. Rev. Lett. 51, 578 (1983). 117 [Dre 90] RM. Dreizler and E.K.U. Gross, Density Functional Theory (Springer- Verlag, 1990). [Fah 86] S. Fahy, S.G. Louie and M.L. Cohen, Phys. Rev. B 34, 1191 (1986). [Foi 85] SM. Foiles, Phys. Rev. B32, 3409 (1985). [Foi 86] SM. Foiles, M.I. Baskes, and MS. Daw, Phys. Rev. B 33, 7983 (1986). [Foi 87] S.M. Foiles, Surf. Sci. 191, L779 (1987). [Foi 89] SM. Foiles and JB. Adams, Phys. Rev. B40, 5909 (1989). [Gea ] C.W. Gear, private communication. [Con 89] S. Gould, K. Burke, P.K. Hansma, Phys. Rev. B 40 5363 (1989). [Gru 89] M. Grunze and H.J. Kreuzer, Eds., Adhesion and Friction, Springer Series in Surface Sciences, Vol. 17 (Springer-Verlag, Berlin 1989). [Hal 88] BM. Hall, D.L. Mills, M.H. Mohamed, and LL. Kesmodel, Phys. Rev. B38, 5856 (1988). [Ham 79] DR. Hamann, M. Schliiter and C.Chiang, Phys. Rev. Lett. 43, 1494 (1979). [Har 82] B. Harmon, W. Weber, and D. R. Hamann. Phys. Rev. B 25, 1109 (1982). [Har 85] U. Harten, J .P. Toennis, Ch. W611 and G. Zhang, Phys. Rev. Lett. 55, 2308 (1985). [He 88] J.—W. He, D.A. Harrington, K. Griffiths, and RR. Norton, Surf. Sci. 198, 413 (1988). [Hed 71] L. Hedin and B.J. Lundqvist, J. Phys. C4, 2064 (1971). 118 [Ho 86] K.-M. Ho and K.P. Bohnen, Phys. Rev. Lett. 56, 934 (1986). [Hoh 64] P. Hohenberg and W. Kohn, Phys. Rev. 136, B864 (1964). [Her 92] C. Horie and H. Miyazaki (submitted for publication). [Iba 87] H. Ibach, J. Vaccum Sci. Technol. A5 419 (1987). [Joh 72] RA. Johnson, Phys. Rev. B6, 2094 (1972). [Jon 89] RD. Jones and O. Gunnarsson, Rev. Mod. Phys. 61, 689 (1989). [Kit 86] Ch. Kittel, Introduction to Solid State Physics, 6th edition (John Wiley, New York, 1986). [Koh 65] W. Kohn and L.J. Sham, Phys. Rev. 140 A1133 (1965). [Kus 90] D. Kusnezov, A. Bulgac and W. Bauer, Ann. Phys. 204, 155 (1990). [Laf 71] Lafon, R.C. Chaney, and C.C. Lin, in Computational Methods in Band The- ory, edt. P.M. Marcus, D.F. Janak, and AR. Williams (Plenum, New York, 1971), p. 284. [Lah 87] A.M. Lahee, J.P. Toennies and Ch. W611, Surf. Sci. 191, 529 (1987). [Lan 60] L.D. Landau and EM. Lifshitz, Course of Theoretical Physics, Volume 1: Mechanics (Pergamon, Oxford 1960), p. 122. [Lan 86] L.D. Landau and EM. Lifshitz, Course of Theoretical Physics, Volume 7: Theory of Elasticity, (Pergamon, Oxford 1986), p. 53. [Lan 79] Landolt—Bérnstein (new series), edited by K.-—H. Hellwege (Springer—Verlag, New York, 1979), Vol. III/11, pp. 10, 114. 119 [Lan 89a] U. Landman, W.D. Luedtke, A. Nitzan, Surf. Sci. 210, L177 (1989). [Lan 89b] Uzi Landman, W.D. Luedtke, and M.W. Ribarsky, J. Vac. Sci. Technol. A 7, 2829 (1989). [Lan 90] U. Landman, W.D. Luedtke, N .A. Burnham, and R.J. Colton, Science 248, 454 (1990). [Lee 89] S. Lee, H. Miyazaki, S.D. Mahanti and SA. Solin, Phys. Rev. Lett. 62, 3066 (1989). ' [Leh 83] S. Lehwald, J.M. Szeftel, H. Ibach, T.S. Rahman and D.L. Mills, Phys. Rev. Lett. 50, 518 (1983). [Leh 87] S. Lehwald, B. Voigtliinder, and H. Ibach, Phys. Rev. B 36, 2446 (1987). [Lev 79] M. Levy, Proc. Natl. Acad. Sci. (USA) 76, 6062 (1979). [Lun 83] S. Lundqvist and N .H. March, Theory of the Inhomogeneous Electron Gas (Plenum, 1983). '[Luo 88] N. Luo, W. Xu, and S.C. Shen, Solid State Commun. 67, 837 (1988). [Lyn 86] S.P. Lynch, J. Mat. Sci. 21, 692 (1986). [Mah 90] G.D. Mahan and KR. Subbaswamy, Local Density Theory of Polarizability (Plenum, 1990). [Mam 86] H.J. Mamin, E. Ganz, D.W. Abraham, R.E. Thomson, and J. Clarke, Phys. Rev. B 34, 9015 (1986). [Mar 63] A.A. Maradudin, E.W. Montroll, and G.H. Weiss, Theory of Lattice Dy- namics in the Harmonic Approximation (Academic Press, New York, 1963). 120 [Mat 87] C. Mathew Mate, Gary M. McClelland, Ragnar Erlandsson and Shirley Chiang, Phys. Rev. Lett. 59, 1942 (1987). [McC 89] G.M. McClelland in Adhension and friction, edited by M. Grunze and H. j. Kreuzer, Springer Series in Surface Science, Vol. 17 (Springer-Verlag, Berlin) 1989. [Mey 88] E. Meyer, H. Heinzelmann, P. Griitter, Th. Jung, Th. Weiskopf, H.-R. Hidber, R. Lapka, H. Rudin, and H.-J. Giintherodt, J. Microsc. 152, 269 (1988). [Mey 90] G. Meyer and N. M. Amer, Appl. Phys. Lett. 56, 2100 (1990). [Mil 71] A.P. Miller and B.N. Brockhouse, Can. J. Phys. 49,704 (1971). [Nel 77] J.S. Nelson, E.C. Sowa, M.S. Daw, Phys. Rev. Lett. 61, 1977 (1988). [Nel 89] J.S. Nelson, M.S. Daw, and EC. Sowa, Phys. Rev. B40, 1465 (1989). [Nos 84] s. Nosé, J. Chem. Phys. 81, 511(1984); Mod. Phys. 52,255 (1984). [Nyb 83] C. Nyberg and C.C. Tengstiil, Phys. Rev. Lett. 50, 1680 (1983). [Ori 77] RA. Oriani and RH. Josephic, Acta Metal]. 25, 979 (1977). [Ove 92] RM. Overney, E. Meyer, J. Frommer, D. Brodbeck, R. Luthi, L. Howald, H.-J. Guntherodt, M. Fujihira, H. Takano, and Y. Gotoh, Nature 359, 133 (1992) [Par 80] M. Parrinello and A. Rahman, Phys. Rev. Lett. 45, 1196 (1980). [Par 81] M. Parrinello and A. Rahman, J. Appl. Phys. 52, 7182 (1981). [Par 82] M. Parrinello and A. Rahman, J. Chem. Phys. 76, 2662 (1982). 121 [Pea 58] W.B. Pearson, A Handbook of Lattice Spacings and Structures of Metals and Alloys, Pergamon, New York, 1958. [Pei 78] H. Peisl in Hydrogen in Metals I, Vol. 28 of Topics in Applied Physics, edited by G. Alefeld and J. Vélkl (Springer-Verlag, Berlin, 1978), p. 69. [Pet 87] J.B. Pethica, W.C. Oliver, Physica Scripta T 19, 61 (1987). [Pra 89] LR. Pratt and J. Eckert, Phys. Rev. B 39, 13170 (1989). [Ray 84] J.R. Ray and A. Rahman, J. Chem. Phys. 80, 4423 (1984). [Ray 88] J.R. Ray, Comp. Phys. Rep. 8, 109 (1988). [Rie 83] K.H. Rieder, M. Baumberger and W. Stocker, Phys. Rev. Lett. 51, 1799 (1983). [Rie 84] K.H. Rieder and W. Stocker, Surf. Sci. 148, 139 (1984). [Roc 84] S. M. Rocca, S. Lehwald, H. Ibach and T.R. Rahman, Surf. Sci. 138, L123 (1984). [Row 74] J.M. Rowe, J.J. Rush, H.C. Smith, M. Mostoller, and HE. Flotow, Phys. Rev. Lett. 33, 1297 (1974). [Row 86] J.M. Rowe, J.J. Rush, J.E. Schirber and J.M. Mintz, Phys. Rev. Lett. 57, 2955 (1986). [Ski 71] J. Skinner, N. Gane, and D. Tabor, Nat. Phys. Sci. 232, 195 (1971). [Sko 87] M. Skottke, R.J. Behm, G. Ertl, V. Penka, and W. Moritz, J. Chem. Phys. 87,6191 (1987). 122 [Sla 93] M. Slameron, R. Overeny, E. Meyer, M.O. Robbins, P.A. Thompson, C.C. Crest, J.A. Harrison, C.T. White, R.J. Colton, D.W. Brenner, J. Belak, D.B. Boercker, I.F. Stowers, U. Landman and W.D. Luedtke, MRS Bulletin, May 1993, p. 15. [Smo 41] R. Smoluchowski, Phys. Rev. 60, 661 (1941). [Spa 84] D. Spanjaard and M.C. Desjonqueres, Phys. Rev. B30, 4822 (1984). [Spr 62] J. Spreadborough, Wear 5, 18 (1962). [Ste 60] EA. Steigerwald, F.W. Schaller, and AR. Troiano, Trans. Metal]. Soc.. A.I.M.E. 218, 832 (1960). [Sto 89] P. Stoltze, J .K. N ¢rskov and Uzi Landman, Surf. Sci. Lett. 220, L693 (1989). [Sun 89] Z. Sun and D. Tomanek, Phys. Rev. Lett. 63, 59 (1989). [Tab 84] T. Tabata and H.K. Birnbaum, Scripta Metal]. 18, 231 ( 1984). [Tar 86] B. Tardy, J.C. Bertolini, CR. Acad. Sci., Ser. 2, 302 813 (1986). [Ter 83] J. Tersoff and DR. Hamann, Phys. Rev. Lett. 50, 1998 (1983), and Phys. Rev. B31, 805 (1985). [Toe 87] J.P. Toennies, J. Vacuum Sci. Technol. A5, 440 (1987); ibid. A2, 1055 (1984). [Tom 83] D. Tomének, S. Mukherjee and K.H. Bennemann, Phys. Rev. B28, 665 (1983); ibid. 829, 1076 (1984) (E). [Tom 85a] D. Tomanek, A.A. Aligia and C.A. Balseiro, Phys. Rev. B32, 5051 (1985). [Tom 85b] D. Tomanek and K.H. Bennemann, Surf. Sci. 163, 503 (1985). 123 [Tom 86a] D. Tomanek, Phys. Lett. 113 A, 445 (1986). [Tom 86b] D. Tomének, S.C. Louie, and C.T. Chan, Phys. Rev. Lett. 57, 2594 (1986). [Tom 86c] D. Tomének and S.C. Louie, Phys. Rev. Lett. 57, 2594 (1986). [Tom 89] D. Tomének, G. Overney, H. Miyazaki, S.D. Mahanti and H.J. Giintherodt, Phys. Rev. Lett. 63, 876 (1989) and ibid. 63, 1896(E) (1989). [Tom 91a] D. Tomének, Z. Sun, and S.C. Louie, Phys. Rev. B43, 4699 (1991). [Tom 91b] D. Tomanek and W. Zhong, Physical Review B (in press). [Tom 91c] D. Tomanek, W. Zhong, and H. Thomas, Europhys. Lett. 15, 887 (1991). [Tom] 29] C.A. Tomlinson, Phil. Mag. S. 7, Vol. 7, 905 (1929). [Wes 69] D.G. Westlake, Trans. Am. Soc. Metals 62, 1000 (1969). [Wig 38] ER Wigner, Trans. Faraday Soc. 34 678 (1938). [Wut 86] M. Wuttig, R. Franchy and H. Ibach, Solid State Commun. 57, 445 (1986). [Yan 91] Liquiu Yang, Talat S. Rahman and Murray S. Daw, Phys. Rev. B44, 13725 (1991). [Yos 88] J. Yoshinobu, M. Onchi and M. Nishijima, Phys. Rev. B 38, 1520 (1988). [Zab 89] H. Zabel in Graphite Intercalation Compounds, Topics in Current Physics, edited by H. Zabel and S.A. Solin (Springer-Verlag, New York, 1989). [Zho 90] W. Zhong and D. Tomanek, Phys. Rev. Lett. 64, 3054 (1990). [Zho 91a] W. Zhong, Y.S. Li and D. Tomanek, Phys. Rev. B 44, 13053 (1991). [Zho 91b] W. Zhong, G. Overney, and D. Tomanek, Europhys. Lett. 15, 49 (1991). 124 [Zho 92] W. Zhong, Y. Cai, and D. Tomanek, Phys. Rev. B 46, 8099 (1992). [Zho 93a] W. Zhong, G. Overney, and D. Tomének, Phys. Rev. B 47 (1993). [Zho 93b] W. Zhong and D. Tomének, in preparation. "‘9199559“