vao a . I! .1... 1% F thl‘h—ilcvl .5Hh- rdl.fil . . 9... «mm. 262%. ‘.'l a .0 13;. t .u .sdl 2 7 A 53.9.”. .3271 :5.‘ My}... 3.7.3. 1.5:... 93.3.1..- urns; » \1! I . {25.15 . .. 2... .v .733 . ‘ ; 16:25an7. 2 «Sin | » . . 35.32.. :- 1 'm. P): of an MCSCF calculation molecular electrostatic potential Multi-Reference Configuration Interaction number of electrons number of iterations normalization constant projection operator distance from center of mass to the nucleus farthest from the center of mass momentum weight of determinant k in a many determinant wave function QMC if m SCF SMO STO VQMC VSEPR quantum Monte Carlo charge distance between two electrons distance between an electron and a nucleus distance between two nuclei self consistent field spectroscopic molecular orbital Slater type orbital CPU time expectation value of kinetic energy expectation value of potential energy variational quantum Monte Carlo valence shell electron pair repulsion variance charge on nucleus n INTRODUCTION This thesis is arranged as a series of studies each of which is a separate chapter. Each of these can be taken as an independent individual study. However, as a collection these studies present representative samples of four different aspects of quantum chemistry. Chapter I presents a study of the electronic structure of the mono hydroxide and oxide cations of Ti, V and Cr. This is a fairly traditional quantum chemistry study although a difficult case. Open shell, coordinatively unsaturated transition metal compounds present significant technical difficulties in applying quantum chemistry techniques. The automated algorithms found in popular software packages such as Spartan and Gaussian do not correctly handle these systems. Simpler levels of theory, such as I-IF theory, do not give even qualitatively reasonable results for these compornrds. Thus more complex calculations must be employed. Chapter II focuses on the nature of electron lone pairs, and employs scientific visualization techniques. These techniques are a more recent form of data analysis utilizing object oriented graphics software. This study 1 2 provides a good overview of the capability of visualization techniques as well as attempting to answer a question of scientific and pedagogical importance. Chapter III is an evaluation of the potential for variational quantum Monte Carlo (VQMC) techniques for replacing high order Moller-Plesset perturbation theory and other many determinant techniques for high accuracy descriptions of transition metal systems. The theory behind VQMC and explicitly correlated wave functions as well as issues related to the efficient computer implementation of these calculations is discussed. Several new algorithms for the eflicient integration of heavy element wave frmctions are proposed. Chapter IV is a discussion of the existing theorems describing the behavior of fire total electron density. The most general and powerful form of any mathematical treatment is its theorems and their proofs. This provides an excellent counter point to the visualization of electron density for specific molecules illustrated in chapter 11. These theorems represent the most rigorous description of quantum chemistry. This is the opposite extreme from simplified models such as VSEPR and the artist’s renditions often found in text books. Chapter I TRANSITION METAL OXIDE AND HYDROXIDE CATIONS A. Introduction The frmdamental understanding of chemistry is built on the systematic characterization of categories of chemical compounds. Ab initio techniques provide information about compormds which is extremely difficult to obtain experimentally. This ab initio study examines the electronic structure of a number of compounds which are not well described by traditional models such as Lewis dot structures and VSEPR theory [1]. This characterization is important not only for understanding these gas phase molecules, but for its eventual extrapolation to coordination chemistry, materials science and bioinorganics. A number of early transition metal oxide and hydroxide cations are characterized at the Multi-Configurational Self Consistent Field (MCSCF) and Multi-Reference Configuration Interaction (called MRCI or specifically MCSCF+1+2 in this study) levels of theory [2]. The qualitative description is primarily obtained from the MCSCF calculations. More accurate quantitative results are obtained by doing a configuration interaction calculation from this reference space (MCSCF+1+2). Prior to this study, Tilson and Harrison examined the ScO+ and ScOH+ ions at the same levels of theory that are used here [3]. They found that the ScO" ion has a 12+ grormd state. It is characterized as having a triple bond formally consisting of two electron pair bonds and one dative bond. The ScOIF ion has a 2A grormd state in which the Sc - 0 bond is formally a triple bond consisting of one electron pair bond and two dative bonds. The CrO+ ion has been examined in a number of prior theoretical studies [4-6]. New CrO+ calculations were done for this study in order to have all results at a consistent level of theory. Theoretical studies of the neutral oxide molecules have been published [7,8], but the studies in references 3—6 are the only previous cation calculations. The bond energies of these compormds have been measured experimentally using mass spectrometric techniques [9-11]. However, there has been no experimental geometry determination. Other than these experiments and the theoretical results mentioned in the previous paragraph, no previous studies have been made of these molecules. The linear molecules are all arranged with the atoms on the Z axis, while the bent molecules are in the XZ plane. For the linear molecules, the d orbitals are labeled according to the symmetry of the orbital (e.g., (1,,z would be labeled xx). A number of electronic configurations were computed for each molecule in order to determine the lowest energy electronic structure. Section G describes the results of tests of a number of wave functions. Section H gives a detailed description of the basis set and wave frmction used to obtain the final results. B. Qualitative description of bonding Considering how the hydroxide radical might bond to a metal leads to three possibilities. First is the bent structure prevalent in organic chemistry which is pictorially represented as A second possrbility is the formation of 1: bonds. Forming a dative 1: bond with the lone pair electrons on oxygen gives rise to the linear double bonded structm'e The third possibility is sp hybridization of the oxygen atom to give the maximum possrhle bonding interaction, a triple bond with one strong shared bond and two weaker dative bonds pictorially represented as The natural orbitals of the MC SCF wave function are used to obtain a description of the bonding within the hydroxide cations. Note that there is not a rigorous way to assign which bonds are dative and which are electron pair bonds. For the linear molecules, the 1t bonds must be equivalent by symmetry and are kept so in the calculation. The titanium atom has a ground state electron configuration of 4sz3d2, while the lowest energy state of Ti” is 4s'3d2. The Ti’ ion has an excited state with a 3:13 configuration 33 mH higher in energy [12]. When titanium is bonded to other atoms, a detailed analysis seldom leads to the assignment of a [significant amormt of 45 character in either the bonds or lone pair electrons of the grormd state [13,14]. The TiOH+ ground state was formd to be a doubly degenerate 3A state with a linear geometry. The pictorial representation and Lewis structure for thisstateare The oxygen atom exhibits sp hybridization as discussed above. Considering the formation of the Ti-O bond fiom Ti+ and an OH radical would lead to the characterization of the bond as a triple bond with one electron pair bond and two dative bonds with both electron pairs coming from the oxygen. We make the bond assignment based on Ti” since the dissociation to Ti+ is much lower in energy than dissociation to Ti or Ti”. This triple bond characterization gives the maximum possible interaction between the Ti and O atoms assuming energetically accessible orbital occupations. A very low lying 32‘ state with the 8. and 8. orbitals singly occupied was also found. A doubly degenerate 2A ground state is predicted for the TiO+ molecule. This state is represented as A low lying 2): state with the do orbital singly occupied was also found. Without the presence of the hydrogen atom, there is no energetic advantage to hybridizing the O atom as evidenced by very little mixing of the 0 2s orbital with it’s p orbitals. Based on having three valence electrons around a Ti” ion and four p electrons around an O atom, the bond is characterized as a triple bond with two electron pair bonds and one dative bond from the oxygen. This bonding again gives the maximum accessrble interaction between the atoms. In forming a VOH+ structure analogous to the TiOH+ structure, the additional electron could be put in the empty d orbital, paired with one of the unpaired electrons or put in the 4s orbital. The lowest energy state formd was a 42’ state obtained by filling the last available d orbital. The pictorial representation and Lewis structure for the ground state are 10 8,. + 00V :ZO—H 5+ c The ScOI-F molecule studied previously follows the same metal - oxygen bonding structure as seen here in TiOH+ and VOH+ [3]. A VO” compound with a metal oxygen bonding structure similar to that of TiO+ will have two unpaired electrons, similar to TiOH“. However, the computed grormd state is a 32' state, represented as Forming a CrOH+ compound with bonding similar to that of VOH+ requires either occupying the 4s orbital on Cr to obtain a pentet state or pairing electrons on the metal center to get a lower spin triplet state. The lowest energy structure of the latter type is a 32‘ state with the do orbital ll doubly occupied. This is higher in energy than the separated Cr+ (6S) ion and OH radical (see section G). Two stable states of CrOH+ were found which do not follow the same bonding pattern that was observed in the previous hydroxide structures. Note that the ground state electronic configuration of the Cr atom is 4s‘3d5 rather than 4s23d4 indicating an additional stabilization due to having a half filled 3d shell. Likewise, the ground sate of Cr+ has a 3d5 configruation. The preservation of the halffilled 3d shell around the Cr atom leads to a linear 5’1'] structure with only one 1: bond and a bent 5A; structure with a single 0' bond. These two states were found to be very close in energy with the 5A. structure having a slightly lower energy. The pictorial representation and Lewis structure for the bent grormd state are cl. é coco 12 dyz. + o,, . 7 Cr ---- O (1,2,,2 o dxz - 137°\'\ H The CrO+ structure also deviates from the trend of the previous oxides in order to maintain a half filled 3d shell around the Cr atom. Two structures were seen, a 42‘. structure and a 411 structure slightly lower in energy, represented as 5- o 75x OCT: 6—“... 5+. This molecule has been studied a number of times previously [4-6]. Our results agree with the most accurate of the previous studies. 13 C. Thermochemist_ry of dissociation If the oxides could be formed from the corresponding hydroxides by the tmcoupling of the electrons in the OH bond of the ground state hydroxide, the result would be a 4A state for TiO+, a 52 state for VO+ and a 6IT state for GO“. Each of these is a higher spin state than is predicted for the ground state electronic structure of these molecules. Energy diagrams presented in F igrues 1.1 and 1.2 depict the hydroxide at the bottom and the separated atoms at the top. Two routes are presented to reach the separated products either by first dissociating the metal - oxygen bond or first dissociating the oxygen - hydrogen bond. There are two levels associated with the MO+ + H system. The higher energy structure is the high spin state of the oxide as discussed in the previous paragraph. The lower energy structure is the ground state electronic configrnation. Both states reflect optimized bond lengths. Both TiOIF and VOH+ show the metal - oxygen multiple bonding structure which was first observed in ScOI-F [3]. The metal oxygen bond in the grormd state of TiOI-I+ and VOIF is weaker than in the oxide as in the Sc analogue. This trend is consistent with the characterization of the oxides I4 +IO_._. > IE madn— I + <~ +O_._./ # I + > :6 MN: I + Na +O>/ IO + +> I .7 Wm +0) IE 3.3— I+O++> sense .32 noon mirage: +30> - 3 occur.— AE 35cm .38. l and—c7 r and—2- r end—2.. rad—2- rem—S- r 3.32. Iced—2. r 3.22. 16 having two strong shared bonds and one weaker dative bond while the hydroxides have one shared and two dative bonds. Due to the differences in bonding structures between the Cr compormds and those of Ti and V, the Cr energetics are not expected to follow exactly the same pattern. However, a similar energy level diagram can be obtained for the chromium compounds as shown in Figure 1.3 . The (TI state of CrO+ maintains the half filled shell environment around Cr, but does not give the maximum possible interaction between the two atoms as formd in the ground state. D. Entitative results The calculated bond lengths and dissociation energies are shown in tables 1.1 and 1.2 . Past studies have shown that the MCSCF predictions can be expected to follow the correct trends in bond energy and give good geometric results [3,14,15]. The MCSCF+1+2 results are expected to give more accurate bond energies. 17 I .7 EV +90 sense .32 Rocco $1,582 E06 - 3 2:5 ch >326 .83. r 2..»— _ T r «6.3 _ T r 8.3 _ T r 2.2:. I GO»— _ T r 3.2 _ T 18 Table 1.1 - Metal - Oxygen bond lengths and dissociation energies molecule Scorr+ rion” VOH CrOH ScO TiO vo CrO r(M-0> , MCSCF 3.5093 3.428 3.372 3.403 3.0953 3.044 3.011 3.101 “M02. MRCI 3.5053 3.378 3.335 3.386 3.1203 3.076 3.048 3.124 162.43 140.8 118.5 45.0 214.13 144.0 95.8 45.3 De(M'O) MRCI" 172.33 155.4 142.4 63.3 232.93 172.7 145.3 78.65 Values are for the predicted grotmd state configurations ScO+ 'I‘iO+ VO+ CrO+ * MCSCF calculations have the 7: bonds correlated 1 2+ 273 3 z- 4n ScOH" TiOH+ CrOI-I+ lengths in bohr, energies in mH, references superscripted De(M-O) MCSCF' D0(M'O) Experiment" 190°, 140‘0 1779, 18010 1659, 171 ‘0 119“), 116“ 2639, 254“ 2559, 257'1 2209, 209‘1 136“ " MRCI calculations are single and double excitations from the MCSCF "‘ experimental results from mass spectrometry 19 Table 1.2 - Oxygen - Hydrogen bond lengths and dissociation energies lengths in bohr, energies in mH, references superscripted «0m , Dre-r0, Dram .. molecule MCSCF MC SCF Experiment Soon+ 1.8483 91.63 4810 TiOH+ 1.798 114.3 85‘0 VOH+ 1.798 140.1 11310 CrOH” 1.793 116.6 14410 OH 1.800 117.4 1629 Values are for the predicted ground state configruations ScOH+ 2A TiOH+ 3A VOH+ 42' @011+ 5A. OH 211 * MCSCF calculations have the metal - oxygen n bonds correlated, giving an SCF description of the OH bond " experimental results from mass spectrometry 20 Both the MCSCF and MCSCF+1 +2 calculations reflect a shorter metal - oxygen bond length in the grormd state of the oxides than in the corresponding hydroxide. This shorter bond is accompanied by an increase in the bond dissociation energy. Crystal structures for compounds containing titanium - oxygen bonds yield bond lengths in the range 3.06 - 4.17 bohrs [16,17]. Since the oxide falls at the low end of this range it is reasonable to assign a triple bond to it. The hydroxide falls in the middle of the experimental range consistent with it’s weaker titanirnn - oxygen bond. Experimental vanadium - oxygen bond lengths fall in the range 2.93 - 4.29 bohrs [16,17]. Having the oxide bond length at the low end of the range and hydroxide bond length in the middle of the range supports having an electronic structure similar to that of the titanium analogues. Chromium - oxygen bond lengths fall in the range 2.86 - 4.03 bohrs [16,17]. The shorter oxide bond length is consistent with the double bond assignment in this molecule. Likewise the hydroxide bond length is consistent with the assignment of a single 0 bond. 21 E. Wave function analysis A Mulliken population analysis and contour plots were produced from the natural orbitals of the MCSCF calculations. The Mulliken population analysis scheme is expected to be reasonable for these calculations since the basis set did not contain diffuse functions. The net charges show a pattern of a half electron polarization of the bonds (Table 1.3). This half electron polarization has been seen in studies of many other transition metal - main group diatomic and triatomic molecules [3,14,15]. Table 1.3 - MCSCF Net Atomic Charges Molecule |_S_t_a_t_§ Meg Qtyggn Hydrogen ScOH+ 2A +1.41 0.77 +0.36 ref3 riOH+ 3A +1.52 -0.82 +0.30 VOI-F 4:- +1.49 -079 +0.29 CrOH+ 5Al +1.45 -0.76 +0.31 Sco+ 12* +1.28 -0.28 ref3 TiO” 2A +1.50 050 v0" 32' +1.44 -0.44 Cro+ 4n +1.43 .043 22 For the diatomics, the a bond is primarily composed of the 2pZ orbital on oxygen. A contour plot of the TiO+ a bond natural orbital is shown in Figure 1.4 . Ti has is on the left in all contour plots. The other oxides give a similar picture. The 1: bonds for the oxides are primarily composed of the p1t orbital on oxygen bonding to the d, orbital on the metal (Figure 1.5). All of the bonding orbitals are polarized towards the oxygen with the a bond more polarized than the 7: bond as shown by the population analysis in Tables 1.4 and1.5. Table 1.4 - Metal valence orbital populations (MCSCF) State 48 4p, 4p" 4p], 3do 3dr:x 3d7ry 3d8. 3d5+ Scori+ (2A)ref3 0.03 0.10 0.03 0.03 0.10 0.12 0.12 1.0 TiOH+ (3A) 0.13 0.01 0.04 0.04 0.96 0.15 0.15 1.0 VOH+ (42') 0.08 0.02 0.04 0.04 1.01 0.16 0.16 1.0 1.0 CrOH+ (5A1) 0.10 0.05 0.04 0.05 0.31 0.95 1.04 1.0 1.0 Sco+ ('z*)ref3 0.02 0.19 0.05 0.05 0.55 0.43 0.43 1 "HO" (2A) 0.01 0.00 0.03 0.03 0.34 0.54 0.54 1.0 v0+ (32') 0.01 0.00 0.03 0.03 0.30 0.60 0.60 1.0 1.0 Crot (411) 0.02 0.00 0.02 0.05 0.65 0.78 1.05 1.0 1.0 23 Figure 1.4 - TiO+ MCSCF sig bond I. I 1. II III- a cii--|009f I .m/ or I /. II. I I! I / t\ \ \ \ \ I a l' s\§ 0‘: u\ \\ \ .l 1‘ 0‘ .\ \t \ 5.00 50 6. -3.50 2 24 Figure 1.5 - TiO+ MCSCF pi bond .". .0' ----‘ ' ' 5.00 -5.00 6.50 -3.50 Z 25 Table 1.5 - Oxygen and hydrogen valence orbital populations (MCSCF) State 28 2pc, 2px 2py Hydrogen ScOH’ (2A) 1.74 1.37 1.83 1.83 0.64 ref3 TiOH+(3A) 1.81 1.41 1.80 1.80 0.70 VOH' (42') 1.81 1.39 1.79 1.79 0.71 CrOH+ (5A1) 1.77 1.43 1.66 1.90 0.69 Sco+ (1):“) 1.80 1.44 1.52 1.52 ref3 r10+ (2A) 1.97 1.68 1.42 1.42 vo+ (32-) 1.98 1.71 1.36 1.36 Cro+ (Ti) 1.99 1.34 1.19 1.89 The two linear hydroxides, TiOIF and VOH+ have 1: bonds with electron distributions similar to those of the oxides. The 0 bonds in these two compormds show a dative M-0 0 bond and a covalent OH 0 bond such as the ones for TiOH+ in Figures 1.6 and 1.7 . In these compormds, the 7: bonds are more polarized than the 0 bonds. The node in the a bond of Figure 1.6 can reasonably be considered a polarization near the nucleus. 26 Figure 1.6 - TiOH+ MCSCF sigma bond 6.00 l l L ~6.00 6.00 zaxis 27 Figure 1.7 - TiOH+ MCSCF sigma OH bond anlS -6.00 -6.00 6.00 zaxrs 28 Figure 1.8 - CrO+ MCSCF pi bond .°‘~ 6.50 -3.50 IS Z 29 Figure 1.9 - CrO+ MCSCF pi* bond ’0' [I I O 0.0. t/olr . . ’ . "-.-.--/ \ 6.50 5.00 -3.50 -5.00 30 The CrO+ 7t system in the y direction contains 3 electrons, two in the bonding orbital, which is polarized even further towards the oxygen (Figure 1.8) and one in the 7r“ which is polarized towards the Cr atom (Figure 1.9). The bent CrOI-F 5A] state has three natural orbitals in the valence a system, which describe a o(Cr-O), a 0(O-H) and a lone pair on oxygen. (Figures 1.10, 1.11 and 1.12 respectively.) All three figures are in the plane of the molecule with oxygen at the origin, Ti at z=-3.403 bohr and H at z=1.311 bohr, y=l.223 bohr. Figures 1.4 through 1.12 all depict the positive contours with the solid lines and negative contoms with dashed hues for contour values +0.002, +0.005, +0.01, +0.02, +0.04, +0.08, +0.16, +0.32, +0.64, +1.28. F. Low lm’ states Low lying excited state wave frmctions were obtained through the use of symmetry constraints and appropriate initial guesses. The energies of low energy excited states are summarized in Table 1.6 . The two TiOH+ states are only different by the occupation of the impaired electrons. We were tmable to get the MCSCF calculation to converge to the 32 state. This suggests that the 32 state is not the predicted ground state at the MCSCF level of theory. 31 Figure 1.10 - CrOH+ MCSCF sigma bond /. 6.00 -6.00 6.00 -6.00 32 Figure 1.11 - CrOH+ MCSCF sigma OH bond 6.00 -6.00 -6.00 33 Figure 1.12 - CrOH+ MCSCF O LP 6.00 6.00 34 The MCSCF+1+2 energies for CrOH+ are too close in energy to be certain whether the ground state is the bent 5A] state or the linear 52 state. The chromium oxide cation also has a very low energy excited state [4-6]. The CrO+ ground state has been a matter of debate in the literature [4- 6]. These results agree with the most accruate methods applied thus far, but are not significantly better than the previous works. 35 Table 1.6 - Total electronic energies (in hartrees) molecule state energy level titheorv norr: :A -923.74605 MCSCF non 3A -923.79362 MCSCF+1+2 non+ 2‘ -923.75673 MCSCF+1+2 VOIF 42- 4018.20456 MCSCF VOH“ 4): -1018.26626 MCSCF+1+2 CrOH: :Al -1118.58523 MCSCF CrOH+ 5Al -1118.67616 MCSCF+1+2 CrOH+ 511 411857381 MCSCF CrOH 11 -1118.67476 MCSCF+1+2 no“ 2A -92313249 MCSCF no“ 2A -923.17622 MCSCF+1+2 no“ ‘11 92302322 MCSCF r10“ “A -923.14174 MCSCF+1+2 v0“ 32: -1017.56520 MCSCF vo“ 3): -1017.62547 MCSCF+1+2 v0“ 52 -1017.51691 MCSCF vo“ 5: -1017.60985 MCSCF+1+2 00: :11 411796931 MCSCF Cro+ 4r1 -1118.04068 MCSCF+1+2 00+ 42 -1117.94804 MCSCF CrO 2 -1118.02701 MCSCF+1+2 cm“ 611 4117.93228 MCSCF etc“ 611 -1118.06084 MCSCF+1+2 36 G. Wave function tests All of these calculations have been done with an extension of Dunning’s reoptimization of the Wachter basis set [19]. The precise description of this set is given in the next section. This set was chosen because it has proven to be reliable in many previous studies [3,14,15]. This also has the advantage of keeping the results of this study comparable to these previous studies. The more difficult question was how to include correlation in the wave frmction. It has been well documented that correlation is crucial to obtaining even a qualitatively correct description of similar chemical systems [3,4,6,14,15,18]. This of course immediately opens the question of how to construct a good wave frnrction. Excellent results could be obtained with heavily correlated methods such as firll CI’s or complete active space self consistent field (CASSCF) calculations. Unfortrmately these methods are too computationally intensive to be done on the crurent generation of work station class computers. In HF calculations, the difficulty is in computing the integrals which scale as N4 where N is the number of basis frmctions and calculations with the same basis set are being compared. In many-determinant calculations the number of 37 integrals has not changed but becomes a relatively minor problem compared to the computation of the configmation list (list of individual determinants) which can scale as poorly as N8. In the course of this study a number of types of wave frmctions were used. Some of these are useful only as intermediate steps and thus omitted from the previous discussion. For the sake of completeness all of these levels of theory are discussed here. 1. SCF calculations SCF or HF calculations (the terms are synonymous in this day) are known to be ofien qualitatively incorrect to the point of predicting the wrong ground state and ordering of states for metal - main group systems. However, SCF calculations are still a useful first step in the process of arriving at the final results. SCF wave frmctions are often used as an initial guess for more complex calculations. Also, although the ordering of nearby states may be incorrect, states which show a very large difference in energy at the SCF level can be expected to maintain a 1arge spacing with more accurate wave functions. Table 1.7 lists the energies of various electronic states at the SCF level of theory. Bonding orbitals are designated by Greek letters (1! or o) and 38 orbitals centered only on the metal are designated by English letters (8 or (1). These initial calculations did not include a geometry optimization although an effort was made to use as reasonable a geometry as possible. All of the metal hydroxides listed in Table 1.7 were calculated as linear molecules except for the 5A1 CrOH+ calculation, which had a 105 degree bond angle. The TiOH+ calculations use a Ti-O bond length of 3. 12 bohr and an O-H bond length of 1.78 bohr. The VOHI+ calculations use a V-0 bond length of 3.43 bohr and an OH length of 1.80 bohr. The CrOI-I+ calculations use an O-H length of 1.79 bohr and a Cr-O bond length of 3.37 bohr, except for the 5A1 which used a 3.369 bohr length and the 5IT which used a 3.323 bohr bond length. All of the hydroxide calculations showed Ti-O and O-H 0 bonds (included in the core) except for the 5 IT CrOH+ calculation which had one 1! bond with no a bond. The TiO+ calculations were done with a 2.93 bohr bond length. The VO+ calculations used a 3.372 bohr bond length. The CrO+ calculations were done with a 3.0 bohr bond length. 39 Table 1.7 - Total SCF electronic energies (in hartrees) molecule TiOH+ non" TiOH* TiOI-F VOIF VOH” VOH* Crorl“ (bent) CrOH+ CrOH” CrOH+ CrOI-F TiO+ TiO+ TiO+ V0+ VO+ CrO+ Cr0+ CrO+ m 3A- 32- '2 '2 m 2 (core) 7:,‘ 2 x (core) 7: 2 (core) 7:, (core) ”2 I 2 (care) It, 2 (care) It, 2 (care) It, 2 (care) It, 2 (care) It, 2 (core) 7r,r !(core) 71': 2 (core) 71',r 2 (care) It, 2 (care) It, 2 (care) It, 2 (care) It, 2 (care) It: 2 (core) 7:, 2 (care) It, 2 (care) 7:, y ‘9' z a; d;,_y. dj.) 7:; d3) 71; df=) 7:2 d' d',) n2 d', , d‘, 41,) y x—y : .212. . 1,) y x-y It: 43' 61:, ally) (core) 02 d; d:,_y, d; (11),) d; 0112-..: “C: ally) a; d}, d;,_y. egg) 2 I l I I 7:, 4s dirty, d .2 d”) 7!: d3, 43' (1;) ”i 11;) 7’: d1.) 1148‘) 7:2 d',_y, lily) y x 7:2 d‘, , d:,> y x-y 2 2 :5 dirty) 7:2 d'2 , d'2 d; y x—y z 7:2 d‘ d', 4,) I”: 2 2 l a, dx,_y, or”) energy -923.67796 -923.67319 -923 .633 l 6 -923.57630 -1018.l4967 -1018.03075 -1018.02580 -1118.55670 -lll8.55520 -lll8.44184 -lll8.41322 -1118.26561 -923.04632 -922.97239 -922.96448 -lOl7.39671 -1017.36484 -1017.34811 -1017.79055 -lOl7.73586 ~1017.72798 2. MCSCF calculations The titanium and vanadium MCSCF calculations had both metal - oxygen 7: bonding orbitals correlated. Correlating the more polarizable 7r bonds while keeping an SCF description of the 0' bond was found to be optimal for the scandiurn analogues by Tilson and Harrison [3]. Test calculations with the sigma bond correlated as well, verified the validity of this assmnption for these molecules. The MCSCF wave frmctions are descrrbed in more detail in section H. 3. GVB calculations Generalized valence bond calculations (GVB) use only two determinants per pair of correlated electrons. These calculations were explored as a reference space for a MRCI calculation in order to make the larger calculations easier to do. Test calculations for the chromium compounds showed a loss of 85 ml-I compared to the MCSCF calculations. Since it was possible to obtain suficient disk space to use the MCSCF reference space, the GVB calculations were not pursued due to their questionable accrnacy. 41 4. MCSCF+1+2 calculations Based on previous work, multi-reference configuration interaction calculations are expected to be the most quantitatively acclnate description of the geometry and energetics of these systems. A valence CI was done from the MC SCF reference space. For the titanium and vanadium compounds, the 1t bonds and 1mpaired electrons were included in the active space. For the chromium compormds the metal - oxygen sigma bond was also included in the active space. This wave frmction gave the optimal acclnacy with a reasonable amount of computational resources. These calculations are discussed further in section H. S. SCF+1+2 calculations Inorganic chemists often think of metal main group bonds as being entirely ionic coordinative bonds. If this were indeed an adequate description, configuration interaction calculations from an SCF reference space would be expected to perform well. This is due to the fact that the SCF wave function usually describes ionic systems well. For the grormd state of TiOH+, the SCF energy is 68.1 mH higher than the MCSCF energy. An SCF+1+2 calculation with an optimized geometry gave a total energy of -923.78739 H, only 6.2 mH higher than the MRCI. 42 This result indicates that this system is very sensitive to electron correlation and that SCF+1+2 calculations give a reasonably good description of the molecule. The MCSCF+1+2 calculations are used for the rest of this study due to their increased accruacy, ability to dissociate to the adiabatic products and quality of population analysis from the MCSCF orbitals. H. Final wave function dggription The metal atom basis contraction was (14sllp6d) -> (5s4p3d). The basis set used was Dunning’s [l9] reoptimization of the Wachter [20] basis set with the 6th d exponent recommended by Hay [21]. The oxygen basis was contracted as (11s7p1d) —-) (433p1d). The Duijneveldt [22] 1157p was used with an additional diffuse d fimction with an exponent of 0.85 . The hydrogen contraction was (481p) —) (Zslp) using Huzinaga’s [23] 45 basis with Drmnings [24] 2s contraction coefficients. An additional p exponent of 1.0 was added as suggested by Harrison [25]. Calculations at the MCSCF level correlated some of the bonding orbitals against their antlbonding orbitals. Configuration interaction calculations were performed using the MCSCF active space as the reference 43 space. This MC SCF+1 +2 method has been shown to give fairly accurate bond lengths and bond energies with a minimmn amount of computation. In the titanium compound MCSCF calculations, pairs of orbitals were correlated as follows. no" 2A’ ] (core)(7rr,7r32)2 ("ya”b2 51—) 3 TiOH+ A' I (care) (unit; )2 (zy,7r:,)2 61. 01) The 0'1 refers the do (dzz) orbital. By keeping the 0' bonds in the core, we are using an SCF description of these bonds even at the CI level. The description of the vo+ 32' and von+ ‘2‘ differ from their titanium analogs only in the presence of the additional unpaired electron in the active space. The difi‘erence in bonding in the chromium compounds required a consistent correlation scheme to be constructed as 4 CrO+ II (core) (030*)2 (Irx,7r;)2p§ 61.61 d;> 5 CrOH+ A |(core) (030")2 61 6: d; 61;.) l The TiOH+ MCSCF wave function contained 25 configuration state functions (CSF). The no" MCSCF contained 17 CSF’s. The VOH” and 44 vo+ MCSCF’s contained 34 csr’s and 25 CSF’s respectively. The CrOH+ and CrO+ MCSCF wave fimctions contained 7 and 34 CSF ’s respectively. The non" MCSCF+1+2 wave function contained 36,686 configuration state functions. The no+ MCSCF+1+2 contained 10,748 CSF’s. The VOH+ and vo“ Cl’s contained 64,734 CSF’s and 27,927 CSF’s respectively. The CrOH+ and CrO+ MCSCF+1+2 wave functions contained 61,574 and 59,034 CSF’s respectively. All calculations were carried out with the Columbus software suite [26]. 1. Conclusions A number of conclusions about these molecules can be drawn fiom this study. The early transition metal mono oxide cations exihibit nearly the shortest bonds known for oxygen with these metals. The oxide bonds are very strong bonds which can reasonably be assigned to be triple bonds. The mono hydroxide cations of Sc, Ti and V have linear grormd states. The metal - oxygen bonds in these molecules are short compared to the range of experimentally determined bond lengths. These bonds are characterized as being triple bonds consisting of one electron pair bond and two dative bonds 45 in order to give the maximum possible interaction between the metal and the OH ligand. The most stable electronic states of chromium compormds are those states which preserve the half filled 3d shell environment around the Cr atom. This leads to electronic and geometric structures considerably different from those of the corresponding vanadium compounds. Contour plots and population analysis indicate very polar bonds characterized typically by a 0.5 electron polarization. This characterization makes these systems nearly perfectly centered between completely covalent and completely ionic behavior. LIST OF REFERENCES 10. ll. 12. 13. LIST OF REFERENCES D. D. Ebbing “General Chemistry” Houghton Mifllin (1987) J. F. Harrison, D. C. Yormg submitted to J. Phys. Chem. J. L. Tilson, J. F. HarrisonJ. Phys. Chem. 9_5, 5097 (1991) J. F. Harrison J. Phys. Chem. 20, 3313 (1986) P. G. Jasien, W. J. Stevens Chem. Phys. Lett. 151, 72 (1988) Y. Takahara, K. Yamaguchi, T. Fueno Chem. Phys. Lett. 1_5_8_, 95 (1989) C. W. Bauschlicher, Jr., P. Maitre Theor. Chim. Acta. .9_0, 189 (1995) R. Bergstrbm, S. Limell, L. Eriksson Int. J. Quantum Chem. 52, 427 (1996) D. E. Clemmer, N. Aristov, P. B. Armentrout J. Phys. Chem. 21, 544 (1993) T. F. Magnera, D. E. David, J. Michl J. Am. Chem. Soc. 111, 4100 (1989) H. Kang, J. L. Beaucltamp J. Am. Chem. Soc. 198, 5663 & 7502 (1986) C. E. Moore “Atomic Energy Levels ” National Bureau of Standards (1949) F. A Cotton ‘O‘Chemical Applications of Group Theory ” John Wiley & Sons (1990) 14. 15. l6. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 47 J. F. Harrison, K. L. Kunze “Organometallic Ion Chemistry” Kulwer, 89 (1996) D. B. Lawson, J. F. Harrison J. Phys. Chem. 199, 6081 (1996) F. A. Cotton, G. Wilkinson “Advanced Inorganic Chemistry ” John Wiley & Sons (1988) A. G. Orpen, L. Brammer, F. H. Allen, 0. Kennard, D. G. Watson, R. Taylor J. Chem. Soc. Dalton Trans. S1 (1989) B. H. Botch, T. H. Dunning, J. F. Harrison J. Chem. Phys. 15, 3466 (1981) T. H. Dlmning private communication. A. J. H. Wachter J. Chem Phys. _5_2, 1033 (1970) P. J. HayJ. Chem. Phys. 66, 4377 (1977) F. B. Duijneveldt IBM “Technical Research Report No. Rl-945 ” IBM Research Laboratory, San Jose, CA (1971) S. Huzinaga “Approximate Atomic Functions 11.” Research Report; Division of Theoretical Chemistry, Department of Chemistry, Univeristy of Alberta (1971) T. H. Drmning J. Chem. Phys. 53, 2823 (1970) M. Rivera, J. F. Harrison, A. Alvarado-Swaisgood J. Phys. Chem. 24, 6969 (1990) R. Shepard, I. Shavitt, R. M. Pitzer, D. C. Comeau, M. Pepper, H. Lischka, P. G. Szalay, R. Ahlrichs, F. B. Brown, J. G. Zhao. Intemat. J. Quantum Chem. S2; 149 (1988) Chapter II LONE PAIR ELECTRONS ON THE WATER MOLECULE A. Introduction Lone pairs in water and related systems are frequently described as being regions of enhanced electron density [1]. Indeed, the familiar "rabbit ears" representation, also called equivalent orbitals (E0), epitomizes this idea and allows the systemization and rationalization of many chemical properties [2]. E0 SMO The lone pair electrons on the water molecule can be modeled as spectroscopic molecular orbitals (SMO's) or equivalent orbitals (EO's). 48 49 Alternatively, the lone pairs can be viewed as two different orbitals called spectroscopic molecular orbitals (SMO). This is the view given by the symmetry adapted orbitals in most computational chemistry programs. This model has in the past been cited for the existence of two peaks in the photoelectron spectrrnn, but it is now recognized that both models predict two peaks if examined closely [3]. In spite of the enormous appeal and success of these representations, little is known about the physical nature of the lone pairs. Most discussions invoke the molecular orbital model and focus on the nature of individual orbitals [4,5]. Since these orbitals are not unique, these discussions are somewhat subjective [6]. In this chapter, the physical manifestation of the lone pairs in an isolated, gas phase 1-120 molecule is explored in terms of several aspects of the wave function. First the limitations of orbital based models are discussed. Then two invariants of the 1-120 wave frmction with well understood physical interpretations are examined: the electron density, p(?) and the molecular electrostatic potential (MEP) derived from this density. Finally, two properties obtained by mathematical manipulations of the wave frmction are discussed: density difl'erence plots and the Laplacian of the electron density. 50 The electron distrrbution to the "rear" of the oxygen (in the "lone pair" region) is examined to see if there is any observable structure to support the idea of the lone pairs as regions of enhanced electron density. B. Visualization Technig us This project utilizes a nmnber of visualization techniques. These techniques are computer graphics extensions of traditional graphs. All of these images are views of actual data, not artist’s renditions as are sometimes encountered. A typical graph of f(x) versus x can be viewed as having one coordinate dimension, x (where x is being thought of as a physical distance), and one data value, f(x), at each point in space (corresponding to each value of x). This traditional plot of Ex) would be considered a one dimensional graph in the language surrormding visualization techniques. In contrast a plot of land elevation would be thought of as having two coordinate dimensions (latitude and longitude) and one data value (the elevation) at each point and is thus called a two dimensional data set. The traditional way to plot two dimensional data sets is as a contour plot as shown in Figure 2.1 and 2.2 . A more intuitively quantitative plot, called a mesh plot, is obtained by using the data value as a displacement in the third 51 Figure 2.1 - Contour plot of the electron density in the plane of the HZO molecule. The innermost contour corresponds to 0.142 e/A3 and subsequent contorn's decrease 1miforrnly by 0.00570 e/A3. 52 Figure 2.2 - Contour plot of the electron density in the plane perpendicular to the molecular plane and bisecting the HOH angle with the hydrogen atoms to the left. The innermost contour corresponds to 0.142 e/A3 and subsequent contours decrease uniformly by 0.005 70 e/A3. 53 Figure 2.3 - Mesh plot of electron density in the plane of the water molecule, oriented with the hydrogens towards the viewer. The peak on the oxygen atom has been truncated in order to make the hydrogens visible. 1“] l‘~ [\IV l\/\/ xI‘K‘I 5.1" 54 dimension. Figure 2.3 is a mesh plot of the electron density for the water molecule in the plane of the molecule. A third visualization technique for two dimensional data is to assign colors to data values in order to give a colorization plot as shown in Figures 2.4 - 2.6, which represent high densities as red and low densities as blue. In the case of electron density, there is a density associated with every point in three dimensional space. The three dimensional analogue to a contour of electron density would be a che, called an isosru'face, as shown in Figrne 2.7 . C. Orbital Renrggngtions Wave frmctions for many electron systems are often constructed from one electron frmctions, called orbitals. For example, the Hartree-Fock wave filnction is a single determinant of one electron orbitals. Orbital based descriptions have proven to be useful for the derivation of many simplified models such as VSEPR theory [2] and the Woodward-Homnan e rules [7]. However, orbitals are not unique. Most ab initio and semiempirical programs use symmetry adapted orbitals. These are the linear combinations of basis frmctions necessary to give orbitals which transform according to the irreducible representations of 55 Figure 2.4 - Colorization of total electron density in the plane of the water molecule Water total electron density o ‘ I.- .. .' t 2 . ' ‘ I ’_ ...., ‘ e ' 56 Figure 2.5 - Colorization of the total electron density at right angles to the plane of the water molecule Water total electron density 3 57 Figure 2.6 - Colorization of the total electron density both in the plane of and perpendicular to the water molecule Water total electron density \ 58 mvfiood .ommmumwm. um...“ .. .1 .u . s84 .. ... . t... . 1. —. ...- .1... . . a .15»... . . . . . .3... WA... 1&— v . . .. . C 3.3+. . .c. . N c .. .. A. . . . . . mgod a. . . n...a ... |.. V r a .. ,. c.. . r. .7. 3.0.0 AncchammémcobuoE mam—Bu 5.5020 :33 .833 vvvod 3 2.9m 59 the molecular point group. These orbitals reduce the cost of computation by giving a block diagonal Hamiltonian matrix with one block for each irreducrble representation [6]. Another set of useful orbitals are localized orbitals. Localized orbitals may be constructed by finding the set of orbitals with the minimum Coulombic repulsion between different orbitals. This results in having one orbital for each pair of bonding, lone pair and core electrons. This intuitive description is often invoked when discussing molecular structure [2]. There are two lone pair orbitals in the water molecule 1mder either of these systems. The symmetry adapted orbital system gives an A. and a B1 orbital (SMO’s). The localized orbital system gives an identical orbital for each of the lone pairs (EO’s). Two equivalent orbitals can be obtained merely by taking positive and negative linear combinations of the symmetry adapted orbitals. Past discussions of lone pair electrons have described the lone pairs in terms of these orbital descriptions [4,5]. Since orbitals are truly only convenient ways of partitioning a wave fimction into subfirnctions, arguments over which orbitals are correct are tantamormt to arguing over whether 4 is 2+2 or 3+1. D. To lElectron Densi The total electron density is an observable obtained from the wave flmction [8]. The electron density is the number of electron charges per unit volume (e/A3 in this study). It is experimentally observed via X-ray crystallography, scanning microscopy techniques and electron scattering [9- 12]. To date, easily computed electron densities are more accurate than any experimentally obtainable densities. Many texts have traditionally described the lone pair electrons as regions of high electron density. The electron density arising from a single orbital may indeed match this description. However, the experimentally observable total electron density is c0mputed as the sum of all of the orbital densities rather than a single orbital density. Figures 2.1 and 2.2 show two views of the electron density in the 1120 molecule. The innermost contour corresponds to 0.142 e/A3 and subsequent contorus decrease uniformly by 0.005 70 e/A3. Clearly, the electron density is rather featureless, especially in the lone pair region. Figure 2.7 shows a sequence of three dimensional isosurfaces of the density beginning at 0.148 e/A3 and decreasing to 0.00148 e/A3. We see that contour levels 0.148 and 0.0739 e/A3 represent density essentially localized 61 on oxygen with the hydrogens beginning to contribute at 0.0592 e/A3. By the time we get to the contour level 0.00148 e/A3 we have a rather structureless electron distribution. The size of this last che is, at most, 3.23 A across. Although the density is featureless, it is not spherical in the lone pair region. Figure 2.8 shows the 0.0327 e/A3 isosurface with an inscribed sphere centered about the O atom. The inscribed sphere is colored green and has been increased in size 1mtil it just breaks through the isosurface. If the isosru'face was equidistant from the oxygen, the inscribed sphere would break the srn'face at all points simultaneously. Consequently the white areas are firrther from the nucleus than the green areas. If we increase the radius of the inscribed green sphere very slightly we see the green sphere completely outside of the isosurface. The electron distribution is very nearly spherical and we See no pronomrced variation in the electron density that reminds us of the lone pairs. Note that the ragged line where the sphere meets the density isosurface is due to the visualization software approximating a sphere by a high order polygon. However, the slight asymmetry in the electron distribution, as shown in Figure 2.8, is indeed consistent with the notion of lone pairs oriented approximately tetrahedrally with respect to the OH bonds. 62 Figure 2.8 - The 0.0327 electron per cubic Angstrom isosurface with an inscribed sphere (green) centered at oxygen. The radius of the inscribed sphere is just large enough to break through the isosurface. 63 There is now a large body of computational, experimental and axiomatic evidence to contradict the idea of lone pair electrons being regions of high electron density. See chapter IV for fruther discussion of the total electron density. E. Molgular Electrostatic Potential The molecular electrostatic potential (MEP), 41, is an observable, intimately related to the electron density p(?) [13]. The MEP at any point in space, R , arormd the HZO molecule is given by MFMV w“ Zr R - - _ ¢()= I]r-1‘é|+ H le_R| where, p(?) is the electron number density at r, RJ- is the vector to and Zj is the charge on the jth nucleus. A point particle with an infinitesirnally small charge, q, will have a potential energy q ¢(R) at the point R. There are regions arormd the water molecule where AR) is positive and regions where it is negative. Positive point charges will be attracted to the regions of negative potential and negative charges to the regions of positive potential. AR) was evaluated using the same electron density function as is displayed in Figures 2.1 - 2.8 . Figures 2.9 - 2.13 show contours and Figure 2.9 - Contours of the electrostatic potential in the plane of the HZO molecule. 65 Figure 2.10 - Contours of the electrostatic potential in the plane perpendicular to the plane of the 1-le molecule. 66 Figure 2.11 - Colorization of the electrostatic potential in the plane of the water molecule. Water electrostatic potential . a... s. ;,r i. - 1:3 “xiv“ 67 Figure 2.12 - Colorization of the electrostatic potential in the plane perpendicular to the plane of the water molecule. Water electrostatic potential 68 Figure 2.13 - Colorization of the electrostatic potential in the plane of and in the plane perpendicular to the plane of the water molecule. Water electrostatic potential _w,.:¢l¢nf "Vi-a3 . . 4’."- A 1‘2! “the" A i 4,... r ~ ~. 1 1": ‘ ‘ ~'A' 9. ‘3‘” 7 , jam“? .1! 69 colorizations of d) in the molecular plane and in the plane bisecting the HOH angle. In Figures 2.11-2.13, the black region around the nuclei is a region of positive electrostatic potential, which becomes red as large positive values are encountered near each nucleus. The colored areas are regions of negative electrostatic potential with blue being near zero and the largest magnitude negative regions represented by a black patch inside of a red region. There are several interesting observations to be made. First, we see that the negative contours are clustered to the rear of the water molecule away from the positive hydrogens. Secondly, there is considerable structure intheMEPandwhilethereareminimainbothplanesthe globalmim'maare in the plane bisecting the HOH angle. These two regions, symmetrically located above and below the plane containing the nuclei, are 1.26 A from the oxygen nucleus and subtend an angle of 98° with the oxygen. Figure 2.14 shows the three dimensional nature of these regions of low electrostatic potential. The contour values are given in energy units (eV), which corresponds to the energy a tmit positive charge would have at these points. The -1.90 eV contour encloses a volume which extends continuously from one side of 1120 to the other. At -1.97 eV we see the regions beginning to separate and at ~2.04 eV the minima in the MEP are symmetrically 70 Figure 2.14 - Three dimensional nature of regions of low electrostatic potential. Water electrostatic potential —1.973 eV —2.041 eV 71 arranged above and below the molecular plane. The absolute minimum in ¢(R) is approximately -2.27 eV at which value the enclosed volumes are zero. Figure 2.15 shows a positive value of the MEP in blue (+272 eV) and a negative region in yellow (-2.041 eV). These minima define directions in space not too dissimilar to those associated with the approximately tetrahedrally located lone pairs often associated with water. This orientation is also consistent with known hydrogen bonding in gas phase clusters and liquid water as well as the crystal structure of ice. F. Electron Densm' Diflerence A density difference plot is a plot of the electron density in the molecule minus the electron densities of the isolated atoms. Thus positive regions represent a gain in electron density as might be expected for chemical bonding regions and lone pair regions. Likewise negative regions represent a net loss of electrons. In order to obtain a density difference plot, a judgement must be made as to how the individual atoms are oriented. For example, the two unpaired electrons on an oxygen atom could be pointed towards the hydrogens of the water molecule, or the atomic electron densities could be spherically 72 Figure 2.15 - The molecular electrostatic potential (MEP) in water. The blue surface corresponds to a MEP of +27 .2 eV while the yellow surface corresponds to a MEP of -2.04l eV. 73 averaged. The choice of orientation will afi‘ect the behavior seen in the density diflerence making this method somewhat subjective. Analysis of the density difference indicates an enhanced density relative to the separated atoms in the bonding regions and a single lone pair like region in the plane of the molecule as shown in Figure 2.16 . Although this is consistent with the SMO description of lone pair electrons, it is a mathematically derived quantity rather than a physical property of a single isolated molecule. G. La cian of E n The Laplacian of the electron density, \72 p , shows the behavior of the electron density relative to an exponential decay. A perfect exponential decay, fix) = exp( -x ), will have all of it’s derivatives continuous and monotonic giving no peaks in plots of derivatives of any order. If the electron density is a nearly exponential decay with an additional shoulder, that shoulder will appear as a peak in a plot of the second derivative. Thus there will be a negative minimum in the Laplacian of the electron density where there is the maximrmr positive deviation from exponential decay. Bader and coworkers have utilized this property to interpret electron lone pairs as maxima in plots of -V2p [14]. Note that this is a maximmn relative to an 74 Figure 2.16 - The 0.27 e/A3 isosurface of the density difference for water shows regions of increased electron density relative to the separated atoms in the plane of the molecule behind the oxygen atom (top lobe) and in the bonding regions (bottom two lobes). 75 exponential decay, but is still a region of monotonically decaying electron density. Figm'e 2.17 shows that the lone pair regions in plots of the Laplacian of electron density have a shape that is similar to the regions of negative electrostatic potential. Note however, that these regions are 0.33 A from the oxygen nucleus, significantly closer than the 1.26 A distance to the minimum in the electrostatic potential (Figures 2.14 & 2.15). Like the density difl'erence plot, the disadvantage of this technique is that it is mathematically derived rather than a direct physical interpretation. H. Computational Details The electron density and the electrostatic potential have been calculated from high quality ab initio wave fimctions at the self consistent field (SCF) and configuration interaction (CI) levels [6]. The oxygen atom was described by a Duijenveldt lls 7p basis set [15] contracted to 65 5p along with three tmcontracted d primitives [16] with exponents of 2.25, 0.75 and 0.25. The 5 basis had the 7th, 8th, 10th and 11th primitives uncontracted. The p basis had the 2nd, 3rd, 6th and 7th primitives rmcontracted. The hydrogen atom is described by a Polarized Valence Double Zeta (PVDZ) basis set [17] with the last two primitives tmcontracted and a single p fimction with an exponent ofl.0 . The basis sets used are listed in Tables 2.1 and 2.2. 76 Figure 2.17 - The -3.973 isosurface of V2 p 77 Table 2.1 - Oxygen atom basis set a contraction p contraction W Eur-sisal firms-14 Ema 31195.6 ’ 0.000210 114.863 0.002266 4669.38 0.001628 26.8767 0.017192 1062.62 0.008450 8.32077 0.07534] 301.426 0.034191 2.97237 0.212775 98.5153 0.110311 0.423600 0.398523 35.4609 0.269493 0.150740 0.184703 13.6179 0.423545 5.38618 0.283038 p contraction 1.53873 0.027484 Exponent Qgircicnt 26.8767 1.000 a contraction mm 2mm 9 contraction 31 195.6 0.000048 £529.19!!! goefi_c_ient 4669.38 0.000369 8.32077 1.000 1062.62 0.001936 301.426 0.007860 p contraction 98.5153 0.026719 13m ELM-1912!! 35.4609 0.070298 0.423600 1.000 13.6179 0.146011 5.38618 0.147898 p contraction 1.53873 -0.238701 m m 0.150740 1.000 a contraction m We! 13.6179 1.000 a contraction d contraction Man! W W 5.38618 1.000 2.25 1.0 s contraction d contraction Liane-rm Meet W germ 0.605500 1.000 0.75 1.0 a contraction d contraction cm mm mm 0.220540 1.000 0.25 1.0 78 Table 2.2 - Hydrogen atom basis set s contraction Elam we 13.01 0.019685 1.962 0.137977 0.4446 0.478148 0.1220 0.501240 8 contraction Exmnent Coeficient 0.4446 1.000 s contraction Exmnent oeficient 0.1220 1.000 p contraction W Coeficient 1.000 1.000 The SCF total energy of -76.06160H and the CI total energy of -76.34446H compare favorably with previous calculations [16] which predict, at best, -76.06429 H for the SCF total energy and 46.35328 H for the CI total energy. The optimized geometries used were a bond length of 0.955 A and a bond angle of 105.1° for the CI wave fimction and a length of 0.942 A and angle of 106.2° for the SCF wave fimction. The experimental geometry is a bond length of 0.9587 A and an angle of 103.9° [18]. The reported results were obtained with the CI wave fimction but we find essentially the 79 same results using the SCF function. The calculations were performed using the Colrnnbus [19] and Gaussian 90 [20] programs and the subsequent visualizations used the Advanced Visualization System (AVS) [21]. 1. Conclusions Although useful, orbital descriptions are inconclusive due to being non tmique. The electron density, which has been cited as the physical manifestation of lone pair electrons, does not have a maximum in the region to the rear of the water molecule. Density difference plots and the Laplacian of electron density show structm'e consistent with lone pair electrons, but do not have a well defined physical interpretations in themselves. The molecular electrostatic potential bears both a structure and physical interpretation consistent with the definition of lone pair electrons. Subsequent studies have formd the MEP to be consistent with the lone pair model in a variety of other molecules [22]. These results suggest that it might be more physically accmate to consider the lone pairs in H20 as a derived property of p(?) e.g., the MEP, KR) , rather then as the electron density itself. While the concept of a lone pair is extraordinarily valuable in interpreting, predicting and systematizing many aspects of chemistry, its characterization, as regions of space with 80 significantly enhanced electron density is a considerable oversimplification. Indeed, as we have shown, lone pairs (at least in H20) correspond to regions of minimum electrostatic potential rather than regions of high electron density. LIST OF REFERENCES 10. ll. 12. 13. LIST OF REFERENCES D. Eisenberg, W. Kauzmann “The Structure and Properties of Water” Oxford University Press (1969) See for example any text that discusses the VSEPR theory of molecular geometry. A representative one being, P. W. Atkins, J. A. Beran “General Chemistry” Scientific American Books (1992) J. Sirnons J. Chem. Educ. 62, 522 (1992) R. B. MartinJ. Chem. Educ. 66, 668 (1988) M. J. Laing J. Chem. Educ. 64;, 124 (1987) I. N. Levine “Quantum Chemistry” Prentice Hall, (1991) J. March “Advanced Organic Chemist ” John Wiley & Sons (1992) V. H. Smith in “Electron Distributions and the Chemical Bond” (Eds: P. Coppens and M. B. Hall), Plentnn, 3 (1981) S. Shrbata, F. Hirota, N. Kakuta, T. Muramatsu Int. J. Quant. Chem. 16, 281 (1980) D. A. Skoog “Principles of Instrumental Analysis” Satmders College Publishing (1985) C. M. Lieber Chem. Eng. News M 28 (1994) S. N. Magonov, M.-H. Whangbo Adv. Mat. 6, 355 (1994) P. Politzer, J. S. Mm'ray in “Transactions of the American Crystallographic Association” (Ed: R. Blessing) 26, 23 (1990) 81 14. 15. 16. 17. 18. 19. 20. 21. 22. 82 R. F. W. Bader “Atoms In Molecules; A Quantum Theory” Oxford (1994) F. B. van Duijneveldt “IBM Research Report RI 945” (1971) R. R. Lucchese, M. P. Conrad, H. F. Schaefer, 111 J. Chem. Phys. 68, 5292 (1978) T. H. Dtmning, Jr., P. J. Hay in “Methods in Molecular Electronic Structure Theory” (Ed: H. F. Schaefer, III) Plenlnn, (197 7) R. L. Cook, F. C. De Lucia, P. Helminger J. Mol. Spectrosc. _56, 62 (1974) R. Shepard, I. Shavitt, R. M. Pitzer, D. C. Comeau, M. Pepper, H. Lischka, P. G. Szalay, R. Ahlrichs, F. B. Brown, J. G. Zhao. Internat. J. Quantum Chem. _S_2_2_, 149 (1988) J. J. Frisch, M. Head-Gordon, G. W. Trucks, J. B. Foresman, H. B. Schlegel, K. Raghavaohari, M. Robb, J. S. Binkley, C. Gonzales, D. J. Defrees, D. J. Fox, R. A. Whiteside, R. Seeger, C. F. Melius, J. Baker, R. L. Martin, L. R. Kahn, J. J. P. Stewart, S. Topiol, J. A. Pople Gaussian 90, Revision J, Gaussian, Pittsburgh, (1990) A VS Advanced Visualization Systems, Waltham, (1992) L. Appleby “The Study of Lone Pair Electrons Using Electron Density and Molecular Electrostatic Potential” (thesis) Michigan State University (1996) Chapter III VARIATIONAL QUANTUM MONTE CARLO METHODS A. Introduction The wave mechanics formulation of quantum mechanics was originated in 1926 by Erwin Schrodinger [1]. This theory gives a mathematical description of the behavior of electrons in a molecule. Quantum chemical computations typically accormt for the kinetic energy of the electrons, the Coulombic attraction of electrons to the nuclei, the repulsion between the electrons and the repulsion between the nuclei. The equation describing this system is called the Schrodinger equation and is written, within the Bom- Oppenheimer approximation, as [T‘P=E‘-P (1) elect _V2+ elect nucl_elect1 H=Z— "+2 2,1 +2 Z—"— (2) i===W ‘3) and the integration is over all space and spin coordinates. The expectation value of the energy will always be greater than or equal to the exact energy as long as the boundary conditions are satisfied. Thus the calculation with the lower energy is closer to the exact solution. Typically the wave fimction parameters are optimized by determining the values that give the lowest energy. A number of fimctional forms for approximate wave fimctions have been proposed. One common technique is to use one electron fimctions which have no explicit dependence upon the electron - electron distance. A wave function formed fiom a single determinant of one electron fimctions takes into account the average value of the electron - electron repulsion only. This is the Self Consistent Field (SCF) method [3,7]. Hartree-Fock calculations can now be done routinely for small organic molecules. The difference between the HF energy and the exact energy is on the same order of magnitude as the energy of a chemical bond. HF calculations can give reliable results where there is a systematic canceling of errors (i.e. difl'erence in energy between two conformers). However, even a 87 very good HF calculation is not always accurate enough to give the desired information. A HF calculation can be used as the starting point for a more complex calculation such as a Moller-Plesset perturbation theory method and other multiple determinant methods such as a configuration interaction calculation [3]. Unfortrmately these more complex calculations require far more powerful computational resources than do HF calculations and may still not be as accurate as desired. Another fimctional form for the wave fimction was proposed by Hylleraas [8] for two electron atoms. He constructed a wave fimction from one electron ftmctions multiplied by a correlation ftmction. The correlation fimction is a fimction of the electron - electron distance. In this scheme the electron - electron repulsion is being accormted for more accurately than in HF theory. In the early 1960’s both styles of variational wave fimction were being examined by many workers [9]. It was shown that the Hylleraas style calculations were more accurate and converged faster than the HF method and its extensions [10]. Unfortunately the Hylleraas wave fimctions resulted in integrals too complex to evaluate analytically for any systems other than atoms and homonuclear diatomic molecules [1 1]. Due to the generalizabflity of the one electron calculations, explicitly correlated wave functions were all but completely abandoned. The reason for a renewed interest in explicitly correlated calculations is the development of a numerical Monte Carlo technique for evaluating the necessary integrals. The Monte Carlo method was first proposed by Metropolis, Rosenbluth, Rosenbluth, Teller & Teller (casually referred to asM(RT)2 due to the initials of the authors) in 1953 [12]. The early application of the Monte Carlo technique to molecular quantum mechanics calculations was pursued by Conroy in the 1960’s [13- 15]. These early calculations could be done on only the smallest systems due to the ineficiency of the computer algorithms and capability of the computers available at that time. Quantum Monte Carlo calculations really became feasrble for molecular systems with the creation of a greatly improved computer algorithm by Ceperley et. al. in 1977 [16]. All numerical integrals replace analytic integration with a smnmation of finite elements. If the locations of these finite elements are chosen on a lmiform grid in each dimension, the number of points to be sampled would be 89 the mnnber of points in one dimension raised to a power, which is the amber of dimensions. In the case of quantlnn chemical calculations, the number of dimensions is three times the number of electrons. This makes for a prohibitively large nmnbcr of points. In a Monte Carlo integration, the locations of the finite elements to sum are chosen randomly to give an approximate integral. The error incrnred in this statistical integration can be computed as a standard deviation between successive “identical” integrations. As the number of points chosen becomes very large, the answer converges to the exact integral (standard deviation goes to zero). B. Rand6m numbers 1. Importance Random numbers are treated mathematically by the fields of statistics and probability. These fields give mathematical descriptions of how random systems will behave based on mathematical descriptions of what “random nlnnbers” are. Results of quantum Monte Carlo calculations will not be reliable lmless there is a method available to obtain random mnnbers which behave in the ways which statistics assumes them to behave. The best way to describe what “good” random nmnbers should be like is to describe behavior that should and should not be exhibited. The behavior that should be seen is, as follows: A sequence of random numbers should give every possible number in its range of values and on average each possrble nmnber should appear the same number of times. A mnnber should appear in a random sequence, on average, as often as there are possible values retln'ned (e. g. the mnnber 7 in a sequence of random nlnnbers in the range 0 - 32000 should appear with the frequency of about once in every 32000 entries). The space of random ntnnbers should be spanned evenly (e.g. in a sequence of random numbers in the range 0 - 32000, a number in the range 0 - 4000 should on average appear every eighth iteration). The behavior that should NOT be seen is, as follows: Random numbers should not repeat a set sequence. Random numbers should not show an oscillation where every other number is high or low. Random mnnbers should not give long strings of high or low values. Serial subsequences of numbers should not repeat. 91 0 Random numbers should not give the same sequence of numbers every time they are used. The last point is usually handled by starting from some initial nmnber, called a seed, which is taken from the system clock. This ensures that no two sequences start at the same number because it is never the exact same date and time twice. The generation and testing of random nlnnbers will be discussed in the following sections. 2. Generation For the generation of random numbers, the seminal text is Knuth’s work [17] which is old but still a classic work of computer science and computational mathematics. This is a series of texts titled “The Art of Computer Programming”. Random number generation and testing are addressed in chapter 3 of volume 2, titled “Seminumerical Algorithms”. One of the earliest computer systems obtained random mnnbers by using an analog to digital converter to sample the noise on the incoming power lines. While this is a rmique and inventive solution, it is not really the best. Electric utilities use power generating equipment with a slight fi'equency droop, which means that the frequency of AC electricity is slightly lower during peak use times than dming light usage times. Thus even the 92 noise may be skewed to deliver statistically higher or lower values depending upon the time of day, day of week, outdoor temperature, etc. Most random number generators are pseudorandom number generators. These are algorithms which generate numbers by nonlinear mathematical equations in which the results of the previous iteration are used as the inputs for the next iteration. These are not perfectly random because the same sequence results a second time if the same initial starting point is used. However, as long as a mechanism is available to prevent starting fi'om the same point, these generators can generate sequences of numbers which obey all of the conditions which may be applied to defining random numbers. The most common type of random mnnber generators are called “linear congruential generators”. These typically generate mnnbers between 0 and m-l by using the results of one iteration, 1,, to calculate the next result, I,» 1, according to the formula 1,-+1=(a1,°+c)modm (4) where a and c are constants. These random number generators can give extremely poor or extremely good results depending upon the choice of values for the constants [17,18]. One fact that is always true is that a linear 93 congruential generator must eventually repeat its sequence of random numbers. For some applications, no additional acclnacy can be gained once the random number sequence begins to repeat. An example of this is an integration in which the random number is used as a location at which to evaluate the flmction. Once the sequence begins to repeat, the space has been spanned as thoroughly as possible with the algorithm. In other applications, the random number represents a relative displacement fi'om the current location of a particle. In this formulation, the results are not being repeated unless the random numbers begin their repeat cycle when the particles are exactly at their original locations, as is the case for variational quantum Monte Carlo calculations. A second popular random mnnber generation scheme is based on a Fibinacci sequence. The most well known Fibinacci generator is the ran3 generator given in “Nlnnerical Recipes” [19]. It uses a Fibinacci series which is lagged by 55 terms. This generator returns floating point values in the range 0 to 1 exclusive, (0,1) . This gives excellent results for many types of calculations without repeating its sequence. The disadvantage of using a Fibinacci generator is that the results can be extremely poor if the process being simulated can be mathematically mapped onto a F ibinacci sequence [20]. 3. Testing There are two classes of tests for random number generators. The first is a mathematical analysis of the equations used to generate the mnnbers. This is discussed by Knuth [17] but will not be pursued in this discussion. The second class of tests is an analysis of the sequence of numbers produced by the random number generator. All random number analysis methods are based upon testing for confidence intervals. For example, the results of a set of identical trials should be within one standard deviation of the analytic answer in 68% of the trials, within two standard deviations in 95% of the trials and within three standard deviations in 99.73% of the trials. An alternative confidence interval test utilizes the chi squared intervals. The standard deviation and chi squared tests yield mathematically equivalent information. The choice of a confidence interval test is usually made based on mathematical expediency. The standard deviation method was chosen for this work because it is the most commonly used in chemical applications. 95 The random number generators tested were the ran3 generator and ones built into the GNU C++ and Silicon Graphics C++ compilers. Traditionally, the random number generators provided with compilers were very poor linear congruential generators. These were created in the anticipation of needing random elements in video games and such but never intended to give statistically valid results. All three random number generators gave excellent results. The ran3 generator was finally chosen for the sake of future portability of the software. Therandomnmnberteststhaterdstareeachaway ofdirectly testing foroneofthe properties ofrandomnmnbersetsthatarelistedinthe section on the importance of random numbers. A comprehensive listing of well rmderstood tests is presented by Knuth [17]. Some of these test are merely alternative ways of verifying the same property. When a well known random number generator, such as ran3, is used there have already been documented tests of desired properties. It is however still the case that researchers concerned with statistical results should have an understanding of random number testing and verify the tests appropriate to their application. Eleven of the tests presented by Knuth were examined. Those that were not redlmdant were performed in depth. Additional spot checks of redlmdant properties were done. The single most important test is to verify the results for a known case with the algorithm and code that will be used for experimental cases. This is illustrated by the ran3 generator which gives excellent results for many applications, but gives very poor results when applied to Ising lattice calculations [20]. The reliability of the ran3 generator was verified extensively for use with variational quantum Monte Carlo calculations using both Gaussian type (GTO) basis sets and Slater type (STO) basis sets. This was done by nmning calculations which can be done without the use of random number generators to verify the results. 97 C. Theory Quantum Monte Carlo (QMC) methods are broken into three categories; variational quantum Monte Carlo (V QMC), difi'usion quantlml Monte Carlo (DMC) and Green’s fimction quantlmr Monte Carlo (GFMC). Only the VQMC method uses an analytic wave function and obeys the variation principle [21]. DMC calculations use a numerically defined wave fimction rather than integrating an analytic flmction. This wave function is approximate because exact bolmdary conditions, in the form of nodal surfaces, are not known. The energy expression is expressed as a diffusion equation in imaginary time. The Green’s frmction Monte Carlo technique uses an approximate Green’s flmction integration as well as a numerical wave fimction. For the pmposes of this study, there are strong advantages to using approximate analytic wave functions rather than numerical wave frmctions. Specifically, it is possrhle to examine the distribution of electrons within the wave fimction by examining the orbital coefficients. This gives a means to analyze bonding, hybridization and population analysis. Also the variational principle gives a measrn‘e of the suitability of a wave flmction for describing a molecular system. The rest of this chapter will focus on the variational quantum Monte Carlo technique only. Expectation values are computed as an integral over the wave flmction (equation 3). An alternative form of this integral is to use the probability distribution (the square of the wave flmction) instead of using the wave fimction to perform the integration. Thus the expectation value of energy, (E), would be computed as follows = Ife1 rand then keep the new positions, r = rm, x = x new' 5.e. Evaluate and add in the local energy. 19‘? e-e+—‘p— (19) 6. Divide by the number of iterations. E. = {7 (20) 114 A variation on this is the use of a sampling frequency. This means that step 5 .e is only performed on every mth iteration and the result of step 6 is multiplied by m. This is an attempt to avoid over sampling the region along the path of the particles Monte Carlo walk relative to other configurations. This is of the most concern when very small step sizes are used as in DMC calculations. Importance sampling techniques replace steps 5.3 c & d with equations utilizing a knowledge of the fimctional form of the probability distribution. These techniques will be discussed later in this section. All of the significant CPU time in the calculation is in the loop in step 5. Any improvement in the speed of the steps in this loop will be an improvement in the over all speed of the calculation. Any additional work that can be done out side ofthis loop will be more than made up for by improvements in the CPU time. As mentioned previously, the algorithm on the previous page is not practical due to a N! time complexity. In 1977 Ceperley et. a1. created a greatly improved algorithm in terms of the inverse Slater matrix [16]. . The Slater matrix, D, is the matrix of occupied orbitals evaluated at electron positions as follows 115 Dij = (Pj(xi) (21) The inverse Slater matrix is the matrix of cofactors of the Slater matrix divided by the determinant of the Slater matrix, which is the wave flmction. .1 _ Cii n _ 1F (22) Once the inverse Slater matrix is known, the acceptance ratio and evaluations of terms in the local energy flmction become a sum down a column of the inverse Slater matrix according to the following formulas '73,) an on V??? = Evil/50.19;! (r) (24) $552, {$3142 Bit) (25) 31%,): Evian-(0920) (26) Further discussion of equations 23, 24 and 26 is in appendix B] of reference [21]. Equation 25 was not found in the literature. This is implemented by moving one electron at a time, so only one column of the Slater matrix must be recomputed on any iteration if the move is accepted. no The most time consuming step of this algorithm is taking the inverse of the Slater matrix when the move is accepted. This step scales as N3, however each of these Operations are very fast. In tests of up to 21 electrons, the N2 behavior of computing the local energy and gradients was still the dominant factor in determining the calculation time (see the results and discussion section). Table 3 .2 lists the time complexity of selected steps in the VQMC calculation for the algorithm described here Table 3.2 - Time complexity of VQMC steps opfltion time complexity Matrix inversion N3 particle - particle distances N‘2 Energy expression MN2 Random number calls 3 Parameter gradients P N is the mnnber of electrons M is the number of occupied spatial orbitals P is the number of parameters 117 These time complexities are seen for a constant number of iterations in the Monte Carlo procedlue. The CPU time varies linearly with the number of Monte Carlo iterations, which is the single biggest factor in determining the total computation time. In DMC and GFMC calculations, many virtual particles are used, far more than the number of electrons. For these calculations a more complex procedure for updating the inverse Slater matrix increases the efficiency of the calculation as described in appendix B2 of reference [21]. For VQMC calculations with a very large number of electrons, this method would be advantageous. In general the variation of parameters is done by determining the derivative of energy with respect to the parameter by [16] g:(r(x)el(r))-(r(x))(ez(r)) (27) I’(x)=2%ln(‘l’(x)) (28) This is usually used in a gradient based optimization method such as a steepest descent method. Improved performance can be obtained by saving the gradient from the previous iteration and using the two gradients to 113 compute a cruvatrue for use in a quasi-Newton formulation. Non gradient based methods, such as a simplex optimization can be used for a very small number of parameters but are not as efficient as the gradient based methods [19]. An adaptation of Pulay's direct inversion of the iterative subspace (DIIS) method has not yet been implemented for QMC calculations [30]. AnimportantfactorinthespeedofaQMCprogramisthe preprocessing of needed mnnbers. The primitive and contraction coefficient normalizations can be done before starting the iterative Monte Carlo procedlne. Likewise, intermediate steps in the computation of gradients and the local energy expression are saved so that only the column corresponding to the electron that was moved need be recomputed. These types of algorithmic changes can increase the computation speed by orders of magnitude. Specific examples will be given in the next section. Another valuable algorithmic tool in creating eficient programs is the use of difi'erent sampling techniques. In the simplest Metropolis sampling technique the new location for an electron is chosen randomly within a small cube around the present location of that electron. The acceptance / rejection test ensrues that this gives a proper sampling of the probability distribution. This random choice of the electron movement does not take into account any 119 knowledge of where there will tend to be a high probability of finding the electrons. There are a number of techniques, called importance sampling techniques, which do take into account the nature of the probability distribution when choosing new electron locations. Some of these introduce additional approximations into the calculations while others are still giving a numerical integration which must converge the exact integral. Fokker-Plank importance sampling is an approximate technique which does not utilize a rejection step [21]. This method is an adaptation of a statistical mechanics technique. The new location for the electron being moved is chosen according to the formula xm = x + step @312: g + step_ size gasdev (29) where gasdev is a random number in a Gaussian distribution with a mean of zero and a variance of one. Tests of this method, presented in the next section, showed it to be too approximate to be useful for chemical computations on even the smallest molecules. Importance sampling with rejection, also called Metropolis importance sampling, is an importance sampling technique which does not introduce any additional approximation into the calculation [21]. This technique uses the 120 Fokker-Plank equation to move the electrons and adds a rejection step. The rejection step serves to maintain an exact sampling of the probability distribution for finite step sizes. When modeling transition metals, an additional problem arises which is negligible for light elements. There is a significant difl'erence in the behavior of core versus valence electrons as evidenced by the large difference in the electron density between these regions. This creates dimculties in choosing a reasonable step size for the calculation. A step size which is very small would be reasonable in the inner core region where the electrons are densely packed. However, a small step size would take an extremely long time to sample the entire valence region. Likewise, a step size large enough to efficiently sample the valence would often overshoot the high probability region in the core thus requiring more iterations to get a reasonable sampling of the core region. A number of variable step size techniques have been proposed for application to heavy elements. Variable step size teclmiques often introduce an additional error into the calculation referred to as a step size bias. This is a statistical error at the point where an electron is seeing a change in the step size, which is due to the 121 electron making a move for which there is zero probability of making the reverse move as illustrated in Figure 3.2 . The split 1 technique was introduced by Belohorec, Rothstein and Vrbik in a paper examining CuH vrbrational properties [33]. This technique uses four diflerent step sizes for difierent electrons as follows Is 2s 2p 0.016 1 3s 3p 0.04 r 3d 0.12 1.’ 4s 1 The assignment of electrons to orbitals was made based on the ordering of the electrons from the nucleus. Thus the ten electrons closest to the nucleus would be moved with a 0.016 I step size, etc. The above technique incorporates a significant statistical error at the point where two electrons are interchanged between shells and no error other wise. We have proposed a method dubbed "variable t" utilizing the opposite philosophy of a very small error at almost every iteration. This method chooses the step size based on the distance fi'om the nucleus according to step_size = r( 1.0 - 0.99 exp ( -L r) ) (30) 122 Figure 3.2 - Step size bias A particle moving (the arrow) from a region with a large step size (the large square) to a region with a small step size (the small square) has zero probability of returning to it’s original position. This introduces a statistical error, called a step size bias. 123 L=+ (31) Where L is chosen to give the single STO that reproduces (r). In practice the accumulation of very small errors using this technique was larger than the sum of less frequent larger errors given by the split 1: method. Based on this information and experience with transition metal calculations, we proposed a modification of the split 1: method called "split 1' b". This method varies the step size by one order of magnitude rather than the two orders of magnitude used in the split 1: method. It also classifies ' the electrons into three groups rather than four according to ls 25 2p 0.1 1 3s 3p 0.3 r 3d 45 I As shown in the next section, this proved to be the most efficient method tested for all electron VQMC calculations on transition metals. If a core - valence separation is to be utilized to treat heavy atoms, an obvious step would be to eliminate the core electrons completely by the use of core potentials. 124 Core potentials use a Gaussian expansion to describe the effective potential of the nucleus and core region on the valence electrons. Orthogonality between the core and valence is accormted for by the use of projection operators. Hay and Wadt's [34] averaged relativistic efi‘ective core potentials (AREP) were chosen based on previous experience with HF and correlated calculations. For Sc - Zn their core potentials are nonrelativistic potentials fitted to numerical Hartree-Fock wave flmctions. For Y - Hg they report relativistic potentials fitted to Cowan-Grifin calculations. This distinction is noted because calculationsusingtheHayandWadtsetareofien seeninthe literature which refer to the core potentials generically as relativistic potentials even when first row transition metals are being studied. These are called averaged potentials because they use only an orbital angular momentlnn projection [3 = |l)(l| rather than the l and m projectors P = |lm)(lm| [3 5]. One of the most promising ways to improve the performance of computer applications is to use a parallel computer or a collection of work stations to break the calculation into pieces and let difl'erent processors do part of the computation in parallel. QMC techniques are probably some of 125 the most readily parallelized calculations known. These techniques can be parallelized in a very course grained way in order to spread the calculation over separate machines without being hindered by the speed of communication over a network link. The most cornse grained way to parallelize a QMC algorithm is at the level at which the calculation is repeated several times in order to get average values and their standard deviations. Each of these repeat calculations can be done on a separate processor. Once each processor has done a complete calculation, the results are used to determine the actual values (average) and accuracy of those values (standard deviation). This scheme is useful when a relatively small number of very powerful processors is available. In order to utilize a very large number of processors, the iterations in the Monte Carlo procedure can be broken up and passed to different processors. The intermediate summations can then be totaled to get the calculation result. 126 E. Results and Discussion The reason for lmdertaking a study of the VQMC method is to evaluate it's potential for high acctnacy calculations on small molecules containing transition metals. These compounds are extremely sensitive to the effects of electron correlation and must be described by a reasonably well correlated method in order to get even qualitatively correct results [36]. These highly correlated calculations can become extremely computationally intensive (Table 3 .2). The previous work showing the high acclnacy obtainable by QMC calculations and the more favorable time complexity indicate that these methods may provide an efficient way of doing very accurate calculations. In order to 1mderstand the transition metal results it is important to 1mde the behavior of the underlying VQMC method. The hydrogen atom will be used to test the VQMC method by testing the effect of changing various aspects of the simulation. Since the hydrogen atom wave flmction used can be integrated analytically, this also serves as the single most important test of the suitability of the ran3 random number generator for VQMC applications. The helium atom is examined as the simplest system where a correlation flmction must be used. Finally a mnnber of So atom 127 calculations are presented in order to test the application to heavy atoms. The extension to molecular systems will not be examined in this work. 1. Hydrogen atom tests Note that VQMC gives the correct exact energy for the H atom when A the exact wave flmction is used. The local energy expression is £31— (Equation 7 ). When the exact wave flmction is used, the numerator becomes -0.5 times the wave flmction giving a local energy of -0.5 H, the exact total energy for the H atom. If many iterations are done, this constant is averaged with itself which of course gives an average of ~0.5 H. The same result would be obtained for any molecule if the exact solution were known. The previous paragraph seems trivial at first, but it illustrates several points which should be mentioned for comparison to other computational methods. First, QMC methods are ab initio methods meaning that no experimental data are used in the formulation of the method. Second, the method will admit an exact solution if the variational frmction spans the space of the exact fimction. (Density Flmctional theory has not yet yielded an exact solution to the hydrogen atom.) The following is an approximate solution to the hydrogen atom consisting of a single Gaussian type orbital. 128 ‘1! = N(ot) exp( -or r2 ) (32) N(ot) = ( 2 or / 1r )3“ (33) The analytic integration gives approximately (E) = 1.5 or - 1.59576912 or"2 (34) The local energy for this wave fimction is A —’fPE=3ot-2o2r2-1/r (35) which integrated by VQMC should yield the same result as Equation (34). Choosing the value of or to be 1.0 should give an expectation value of (E) = 4.095769122 H. The following were obtained by VQMC 100 tests of 10“ iterations, EN = 0.09 :t 0.03 H 100 tests of 106 iterations, EN = 0.096 a: 0.003 H 10 tests of 109 iterations, EN = 0.09575 :l: 0.00003 H The number of tests is the number of times that the calculation was repeated. The energy reported is the average energy from these repeat calculations and the standard deviation in their values. 129 For a single GTO H atom, a near optimal value of or would be 0.3 giving (E) = 0424039 H. Doing a VQMC calculation of 10 tests of 106 iterations gave E N = -0.4239 1 0.0006 H. These H atom calculations give a check on several things. First, they are a check of the correct implementation of VQMC in the computer program written for this study, called "rwh". Second, they show that the VQMC result is converging to the analytic result as the number of iterations is increased. Third, the two tests with 10‘5 iterations illustrate the fact that the standard deviation for a given size calculation should decrease as the wave fimction is improved. In fact, calculations have been done where the variation of parameters was performed by minimizing the deviation rather than minimizing the energy [21]. Fomth this is a check of the fact that the ran3 random number generator is appropriate for VQMC applications. All of the above H atom calculations were done choosing the initial position for the electron within a cube of side length one bohr centered on the nucleus. Also, the new location for the electron was chosen randomly within a box of side length one bohr centered on the present location of the electron. It would be prudent to test these settings to determine what values will give the most accurate calculation in the form of the lowest standard deviation. 130 Table 3 .3 gives the results of running 10 tests of 106 iterations for each of several difl'erent initial box sizes. The settings in all of these calculations are a = 0.3 and box size for moving the electron of one bohr. Table 3.3 - Effect of changing the initial box size Lox_s_ize (B) EA (H) Std. dev. (E) 0.01 -0.425 0.001 0.1 -0.424 0.001 1.0 -0.424 0.001 2.0 —0.424 0.002 5.0 -O.423 0.001 10.0 -0.424 0.001 100.0 -0.50 0.05 1000.0 -80 50 Several conclusions may be drawn from Table 3 .3 . First, the calculation is not extremely sensitive to the small changes in the value of this setting. Second, in this case a value close to the distance between the nucleus and the peak in the radial distribution frmction comes out about in the middle 131 of the range of reasonable values. Finally, a very lmreasonable value might have a very adverse efl'ect on the calculation as with the last calculation in the table. The efiect of changing the size of the box within which new electron positions are chosen was tested. This was done by keeping the same settings as the above sequence along with an initial box size of one bohr. The results of these tests are in Table 3.4 . 132 Table 3.4 - Effect of changing the box size hox s_i_ze (B) EN (H) Std. dev. (H) 0.001 -1.3 0.8 0.01 .042 0.08 0.1 .042 0.01 1.0 .0424 0.001 2.0 -0.4238 0.0009 5.0 .0425 0.001 10.0 0425 0.005 100.0 -0.7 0.2 1000.0 -1.1 0.3 Although the calculation will converge with nonoptimal values, it is worthwhile to use reasonable values. A comparison of the number of iterations to the calculation accuracy is given in Table 3.5 These tests were each done with 10 calculations, a = 0.3, box size and initial box size of 1.0 B. 133 Table 3.5 - Dependence upon number of iterations iterations EN (H) Std. dev. (H) 103 -035 (12 104 -044 0.03 105 -043 0.02 106 -0423 0.003 107 -0.424 0.001 108 -o.4239 0.0003 109 .04240 0.0001 10” -0.42406 0.00006 It is also important to have some indication that a calculation is giving a reasonable result. One test of a reliable calculation is a virial theorem test. The virial theorem states that the expectation value of the potential energy divided by the expectation value of the kinetic energy should equal negan've two. (_Vl _ (71 2 (36) 134 The virial theorem check is used in conventional HF calculations to verify that the wave frmction is a reasonable description of the system. This usage still holds for VQMC calculations along with a second usage. The virial theorem test in a VQMC calculation serves as a check that a sufl'lcient nmnber of iterations have been done to give an adequate statistical sampling of the probability distribution. The most extreme example of the need for this test is presented by the hydrogen atom in which the exact wave frmction is used as discussed at the beginning of this section. Only a single Monte Cario iteration is necessary to give the exact energy of -0.5 H. However, all of the other properties of the atom are still expectation values over the probability distribution. To illustrate this two tests of 10 iterations were done. The average virial theorem value was -1.4774780. Doing 10 tests of 104 iterations gave a virial theorem value of -1.939564 and ten tests of 107 iterations gave -1 .999373 . 2. Helium atom tests a. Literature survey A nmnber of Hylleraas style calculations have been done for the He atom. These are calculations that use an explicitly correlated wave frmction, similar to VQMC calculations. The distinction between the two is that the 135 Hylleraas calculations used analytic integration of the expectation value integrals while VQMC calculations use a numerical Monte Carlo integration. Hylleraas calculations have the advantage of being much faster than VQMC calculations. The disadvantage is that the flmctional form of the wave frmction is limited to functions which can be integrated analytically. This is why Hylleraas calculations are not practical or necessarily even possible for molecules. The simplest Hylleraas calculation uses a wave fimction of the form ‘I’=exp(-§r,)exp(-§r2)(l+br,2) (37) The optimal values of C = 1.849 and b = 0.364 give an energy of 2.892 H [8]. Kinoshita used a correlation function which was a 39 term polynomial expansion in r1, r; and n; [29]. He computed an energy of -2.9037225 H. Pekeris used a 1078 term expansion to compute an energy of -2.903724375 H for the 1S gromld state of the He atom. He also computed an energy of -1 .47486588483 H for the 3S first electronic excited state of the He atom [3 7]. The best available He ground state energy seems to be from a 308 term expansion used by Thakkar and Koga [38]. They computed an energy of -2.9037243770341144 H. 136 A number of other studies have been done which focused on obtaining similar accuracy with a smaller expansion form for the correlation function [39-45]. These studies obtained excellent accuracy with 200 - 300 term polynomial expansions. These Hylleraas calculations were all done with correlation functions, which were polynomial expansions due to the constraint of having to be able to do the integrals analytically. The choice of correlation flmction for VQMC calculations is usually based on making the local energy expression easy to evaluate as well as being an accurate representation of the wave frmction. Wave flmctions which are most convenient for VQMC calculations are those of the form given in Equation (10). Furthermore it is most convenient to evaluate local energy expressions if the correlation frmction, f(rij), has the form fro.) = exp( g0,» (38) The simplest case of this form is illustrated by Tsien and Pack [46] who used a wave frmction of the form ‘P=exp(-§r,)exp(-§r2)exp(ar12) (39) They formd the optimal values of the constants to be I; = 1.8581 and or = 0.2547 to give an energy of -2.8896l8 H. 137 Pritchard used a three term exponential correlation ftmction along with three STO orbitals that yielded an energy of -2.903148 H [47]. Lester et. al. tried using a 632p GTO expansion for the orbitals with three s type GTO's for a correlation frmction to get an energy of -2.90039 H [48]. 138 b. Correlation function studies Correlation flmctions are not as well understood as more well studied methods such as HF theory. The known boundary conditions have been stated in the “Theory” section. There are still aspects of these calculations open to debate and others not yet explored. Calculations that do not obey all of the bormdary conditions given in the section on theory were performed. This has been done in order to gain an tmderstanding of how each condition afl'ects the calculation results. The calculations that will be presented utilize wave flmctions consisting of determinants of one electron flmctions times a product of correlation fimctions as in Equation (10). This immediately opens the question of what one electron flmctions should be used. There are three obvious choices. 1. Hydrogenic frmctions. 2. Harlree-Fock frmctions. 3. thctions optimized for use in the VQMC calculations. The use of hydrogenic frmctions would seem to be indicated by the flmctional form of the Hamiltonian. The Hamiltonian has kinetic energy and nucleus - electron attraction terms which depend upon one electron only. The 139 repulsion terms depend upon two electrons. Having a wave function that also had one and two electron terms that corresponded to these would seem logical. The one electron flmctions which are eigenfunctions of the approximate Hamiltonian formed by omitting electron - electron repulsion terms are the hydrogenic atom flmctions. The second way to think of VQMC is as a second step correction to the Hartree-Fock wave frmction as CI and CC calculations are done. This would give an efficient mechanism for optimizing the one electron functions in a HF calculation where the integrals can be done more eficiently. Under this scheme, the electron correlation would be put in twice, first in net efi‘ect only by the HF wave flmction then explicitly by the correlation function. This might lead to dificulties due to the correlation flmction having to make up for the short comings of the HF wave flmction as well as putting in the electron correlation. The third option of simply optimizing the one electron fimction for use in the VQMC wave flmction would admit either of the previous possibilities if they were the optimal frmctions. However, this is also the route which would always require the most computation time. Since excessive CPU time is the 140 major draw back of the VQMC method, it would be best to avoid this option if a more emcient technique can be found. The three choices of one electron frmctions mentioned here are still a matter of debate in the literature. There are not yet any rigorous proofs of the correctness of any of these options. From a pragmatic standpoint any of these choices can yield accruate energies if the correlation ftmction is sufficiently - complex to compensate for any inaccmacies in the one electron frmction. This is a difiicult question to resolve. A marginal correlation function may yield a better energy than a good correlation frmction if the shortcoming of the marginal fimctionisofasortthatcanbecompensated forbyashifiinthe one electron fimction and vice versa. The rest of this section will be concerned with the shape of the correlation flmction, firm). As such it is most convenient to use terminology relative to the correlation function. The equations will continue to be written in the same notation that has been used thus far. However, the terminology will reflect a one dimensional system with the origin on one of the electrons of the helium atom and lengths being distances to the other electron. Thus f(O) will be referred to as the origin and Koo) will be considered to be infinity. 141 For the sake of the following discussion, the position of the electrons relative to the nucleus is irrelevant. The next issue is the value of the correlation flmction at the origin, f(O). At first glance it might be expected that the value should be zero at this point in order to give a zero probability of the electrons being at the exact same point in space (early studies using correlation flmctions assumed, incorrectly, that this was true [49]). However, there is a theorem stating that the correlation flmction should have a finite value at this point [26]. There is a theorem stating that the correlation flmction should have a finite positive slope at the origin thus giving it a cusp [27]. Functions will be examined which have a zero slope analogous to using GTO functions to approximate a STO. Once a normalization scheme has been defined, the exact value of the slope at the origin can be determined. Some workers have done calculations which enforce this slope [50]. The studies presented by us will allow the correlation frmctions to optimize variationally in order to give the best overall description of the wave frmction. This is the more frequently used technique. At large distances the correlation function should go to a constant finite value. This fact was originally presented by Hartree and Ingman [26]. Three 142 later works argued that the correlation function goes to infinity [27,51,52]. It is now generally recognized that the original work was correct. The shortcomings inherent in the incorrect proofs will be discussed. Finally, the correlation flmction should be monotonic (continuously) increasing fi'om f(0) to fit»). This is usually a tacit assumption in the choice of correlation frmction. It is possible to obtain a reasonable energy for the He atom using a correlation frmction that disobeys this condition, as shown by Roothaan and Weiss [28] and Equation (50). The following paragraphs summarize the energies which can be compared to the results presented subsequently for the 1S grormd state of the helium atom. The very crude approximation obtained by completely omitting electron - electron repulsion yields the following equations which are solvable exactly. 2 2 fi=-_V_I__Y_L__2___7; (4o) 2 2 r1 r2 ‘i’=exp(-2rl)exp(-2r2) (41) E=-4H (42) 143 Using the above wave flmction with the 1/r12 electron - electron repulsion term in the Hamiltonian, the energy (E) = -2.7495 H is obtained. Adding more ftmctions with optimized coefiicients and exponents will eventually result in reaching the HF limit of (E) = -2.86168 H. For this discussion, Thakkar and Koga's value of (E) = -2.9037243770341144 H will be considered to be an exact solution. The most commonly used test of performance is the variation principle which states that the most accurate energy will be the lowest energy obtained. In the case of VQMC calculations, both a variational upper bound and a variational lower bound to the energy can be computed (Equation 15). As the wave fimction is improved, both of these should converge towards the exact energy. There is the possibility of a situation in which the upper bound is closer for one calculation whereas the lower bound is closer for another calculation. In this case the calculation with the closer lower bormd probably has a better overall description of the wave frmction while the other has obtained a better energy with large positive and negative deviations in various regions of space. 144 A program, called rwhe, was written for performing VQMC calculations on two electron atoms. The program uses a wave function of the form given in Equation (10). In order to compare correlation functions with a range of properties the one electron frmctions will be held fixed. As a simplest case the following wave function form will be used ‘P=exp(-2r,)exp(-2r2)f(rlz) (43) The worst case example would be the correlation ftmction f( r12 ) = exp( 1/r12 ) (44) which is plotted in Figure 3.3 . This gives a wave frmction which is not normalizable due to the correlation frmction going to infinity. Attempting a VQMC calculation with this frmction resulted in a failed calculation due to a floating point overflow error. A function that is finite, but monotonic (continuously) decreasing (Figm'e 3.4) rather than increasing is fl r12 ) = eXP( - r12 ) (45) Doing a VQMC calculation of 10 tests of 106 iterations gave EN = -0.97 :t 0.03 H. The high energy and large deviation are both the result exp( l/r12 ) 145 1 Figure 3.3 - Plot of the flmction firm) r «N i Nu i fin .4. N r a.— rr a; r b.— r e.— i n.— .r v.— r n.— i N.— .r m— i _ r m6 i ad r rd i 0.9 r md 1. ed r n6 - Nd r fie 1.00E+01 -- 1.00E+02 T OBOE-Kl) 9.0013401 1* 2.00E+01 3r 8.00E+01 .. 7.00E+01 «r 6.00E+01 ~» 5.0013+01 + 4.00901 .. 3.0013+01 «r h A C.’ exp(-r12) 146 Figure 3.4 - Plot of the ftmction f(rlz) \ 0.9 4 0.8 + 0.7 .. 0.6 ~~ 0.5 '1’ 0.4 it 0.3 it 02 4 0.1 .. 2‘13 r m - n 1r a." r QN e QN r. a." rr EN 1. EN r o." \N r 9N T n.” r1 1 m.” . e." U . e." r 0N . MN .. «a 1m e «a i —.N e r _.N r N = fir N r 04 \m r m.— r . r r . .. M.” K ,M.” . n . 4 0 ~ .0 ma 0 ~ 4 2 r m t 2 if V; M r v.— : @— a1 MA . e . r N ~ .m rr N — r. _.— Cm hr —.— r — rr _ . 3 M . .3 T as D“. . ad a he 5 r ho rfi 0.0 3 r ooc . 3 e .. 3 L. v.6 W 11 v.9 r M.° o .. M.° r N.O F r N.O 1T —.° . r —.o o . at it 4 4. a a 4 o 0 8 7 6 5 4 3 2 J 0 0 0 0 0 o 0 0 0 £23 147 of this being a very poor correlation frmction to the point where it is detrimental to the calculation rather than an improvement. The following flmction (Figure 3.5) t( r12 ) = exp( -l/r12 ) (46) is monotonic increasing but has both an incorrect value at the origin, f(0) = 0, and an incorrect slope g! = O. This gave an energy of r=0 EN = -2.365 :I: 0.003 H. The opposite test would be the filnction (Figure 3.6) K ’12 ) = exp( r12 ) (47) This fimction has a finite value and slope at the origin but goes to infinity at long distances. This is not a violation of the normalizability boundary condition because the total wave fimction still goes to zero at large electron separations. This frmction resulted in the energy E N = -2.435 i 0.006 H. Tsien and Pack [45] used the similar frmction f( r12 ) = exp( 0.2547 r12 ) (48) to obtain an energy of EN = -2.870 :t 0.001 H. Note that this is just slightly below the HF limit. Next we will look at a frmction similar in behavior to Equation (46). 148 exp( r12 ) Figure 3.6 - Plot of the frmction f(rlz) .. ad tr EN i ON 1.. W.N a. Na 1. _N r. a.— rr a.— m. E nr 0.— rv n.— n. E i m.— .a n; i ad 1T rd 1. 0d 3. m6 Lt V.° 11 M.° i nd 4. _d 25 ~~ 20 «>- 15 + 3.. 10 ~~ —05 2 1 r Figure 3.7 - Plot of the frmction f(rn) = exp(-02 . m." r a." .. 5.a r a." r m." . 1N T MN it N." r _.N r N a m.— r a.— r b.— r o.— rm.— . 9.— rr m.— .r N._ r m— r dd r ad 1r. 5.a 4 9o 4 ad 8.. 149 f(r 12) = exp(“ ’1?) (49) This will have the correct behavior at long distance and incorrectly go to zero at the origin if a and B are both less than zero (Figure 3.7). This frmction does, however, have more freedom to find an optimal function shape compared Equation (46). The optimal values or = -0.2 and [3 = -0.5 were obtained. This gave an energy of -2.795 1 0.009 H. Using the correlation flmction from the previous paragraph, holding the values of 01 and [3 fixed and allowing the orbital exponent to optimize to 1.75 gave an energy of -2.860 :1: 0.003 H which is still slightly higher than the HF limit. A flmction with even more freedom to give the correct shape near the origin would be (Figure 3.8) f(r,,) = exp(—0.079;; + 0.3 64499r3’”) (50) This gave an energy of EN = ~2.901 i 0.003 H. This gives 90% of the correlation energy beyond the HF limit. This is an excellent energy for a two term expansion. Since the behavior of this frmction far from the origin is incorrect, it is not adaptable to molecular calculations. It does, however, 0.75 12 + 03 64499!‘ 1.5 12 r exp(-0.075 150 .4 Figure 3.8 - Plot of the ftmction f(rn) 1.6 ~- l.4 .. 12 at 0.8 4r 0.6 4» 0.4 ~~ 02 T A be Cr! z- a - a rr QN rr 0N 4r ad 1r ad rr «MN 4r 6." .r 9N \ll/P. 4r QM .. on n r l. n. r t v." 5 .010. 1. an it MN 0 1H 1.. M.N i . f K it . .- u __ t n L.. Q— \lal. r. d.— 11 w.— rel it a.— Lr 0.— m 11 °.~ 1r W.— T. m at W.— .. 2 a 2 11 N.— m 4.. N.— r. —.— Pm hr ”.— lT — Lil .. as W. a. as 11 ”.c P hr ”.° rfi 5.a - hf 5.a . 9 r . 1.. 06 «m a we 5. m6 e .r md rr v.6 r tr V6 in Md .m r Md .4 3 F .T 3 i _d r —d o n .1 . a a a a o 0 4 2 1 8 6. 4 2 0 .I. l 0 0 0 0 3a 151 serve as an excellent model for the correct shape of the correlation frmction near the origin. A correlation flmction often used in molecular computations is the frmction (Figm'e 3.9) «or «rm—2:.) 91> This frmction is called a Jastro frmction, Jastro-Bijl fimction or Pade frmction. The origin of this frmction is somewhat ambiguous. This is indeed the flmctional form obtained from a first order Pade approximant. This flmction does not appear in the papers by Jastro [49] or Bijl [53] which are used as references to its origin. The earliest paper formd using this frmction is by Moskowitz and Kalos [54]. Regardless of its origin, this is a simple filnction that obeys all of the known bormdary conditions on the correlation frmction. When used for light atoms with HF orbitals, this function consistently gives 50% of the correlation energy. For heavy atoms, problems arise as discussed in the following section. Typical values of the parameters are a = 0.5 and B = 1.8 which give the energy EN = -2.7981 $00009 H. The results are relatively insensitive to scaling both parameters as shown by the fact that the values or = 4.3322 and B = 4.6717 give the energy EN = -2.8184 :t 0.0009 H. 152 Many correlation filnctions were tried before the correct behavior was known [28,39,46]. To date, only a handful of fimctions have been tested which obey all of the known bormdary conditions [21,55]. As such there is not yet an established systematic way of choosing correlation frmctions of successively increasing accuracy. 153 The behavior of the correlation frmction at long distances, f(oo), has been debated in the literature. Hartree and Ingman [26] correctly predicted that the correlation frmction should go asymptotically to a constant value, f(oo) = c. They based this conclusion on a combination of physical insight and mathematical argument. The physical insight used was the fact that the correlation frmction accounted for the coulombic repulsion between the electrons; thus, a constant value would give no interaction at long distances. The mathematical argmnent employed was that a H; wave function would separate into two H atom wave flmctions at long distances if the constant value of the correlation frmction at long distances was a non-zero finite value. The exact value that the wave frmctions goes to would be determined by the choice of a normalization scheme. Their realization that their method was not truly rigorous left the way open for other groups to attempt a more complete analysis. Baber and Hassé [51] obtained the incorrect result lim,u_,,f(r,,) = exp(]: r,,) (52) thus giving f(oo) = oo since they estimated k to be positive for He. They did not explain how they derived this expression, but did show how they set up the difi'erential equations. It seems likely that their means for determining 154 which terms in the differential equations were negligible at long distances was the source of the incorrect result. Green et. al. [52] followed a similar route in defining a rather complex series fimction for the large separation limit of firm) which also went to infinity. Pluvinage [27] used perturbation theory to obtain the incorrect result lim,u_,,,f(r,2) oc exp(rn) (53) thus giving Koo) = oo. He had ignored the cross terms of the form 01' f'in order to make the equations separable for his zero‘11 order approximation. He was quick to blame his relatively poor energy of (E) = -2.878 H on a poor zeroth order approximation. He did not deem it necessary to present any secondary analysis to verify that Equation (53) was qualitatively correct. The above incorrect results had as their primary verification the fact that wave flmctions with the specified behavior gave reasonable numerical results for individual atoms. However, at long distances in an atom the total wave frmction decays exponentially towards zero. Thus the large separation behavior of the correlation frmction has very little effect on the total wave flmction and thus the numerical results. In a molecule, however, the correlation fimction has a significant weight in the wave frmction at large distances where it is describing the 155 correlation between electrons centered on two different nuclei. It has been formd that for molecular calculations, correlation ftmctions obeying the original conditions set forth by Hartree and Ingman [26] must be used in order to obtain reasonable results [21]. There is much work yet to be done pertaining to the study of correlation frmctions. There are not yet llbraries of well tested fimctions analogous to the binaries of basis frmctions available for HF calculations. There are also still unanswered questions, as we have pointed out, regarding the best form of the wave fimction and exact shape of the correlation frmction. Although correlation flmctions are an important part of VQMC, this study will move on to exploring the dificulties inherent in applying VQMC techniques to transition metal atoms. This is being done for pragmatic reasons. It would be rmwise to spend an enormous amount of time ironing out difliculties with the He atom just to find some insurmormtable difiiculty preventing the study of other systems. c. Other te_st_s Even though a computer program gets correct results it may not be doing so as eficiently as possible. Since VQMC calculations can use enormous amounts of CPU time, it is worthwhile to attempt to optimize the 156 program as much as possible to use a minimal amount of CPU time. This is done by utilizing functions that are equivalent in result but faster. For example, computers perform multiplication operations faster than divisions so it is faster to multiply by 0.5 than to divide by 2. A good set of principles for designing programs for fast computation speed are 1. Use the algorithm with the best time complexity. 2. Don’t compute the same number twice. 3. Don’t put anything inside of a loop that does not have to be there. 4. Use the fastest operation possible. The relative speed of operations is as follows fimctions inside of loops slowest procedure calls trigonometric fimctions conditional statements division multiplication addition or subtraction assignment fastest 157 The VMCatom program was written first to get the correct answer by directly transforming equations into C++ som'ce code. It was then rewritten to optimize performance. As successive improvements were made a time test was done consisting of a He atom calculation with 10 tests of 10000 iterations using Clementi’s SS STO basis [56] and a Jastro fimction. Table 3 .6 lists the CPU time necessary to do this calculation as successive improvements to the algorithms were made. 153 Table 3.6 - CPU time test for optimized algorithms CPU time (minzsec) algorithm changed 1.28 no compiler optimization 1:28 -01 compiler Optimization 1:11 ~02 compiler optimization 0:57 primitive normalization constants hard coded 0:57 avoid repeat calculation of determinants 0:57 mathematical form of Jastro frmction 0:55 ~03 compiler optimization 0:43 avoid repeat calculation on rejected steps 0:38 recompute only one row of Slater matrix 0:18 normalize all primitives in advance 0:16 save intermediate steps of local energy calc. 0:13 avoid repeat orbital calculations on accepted steps 0:08 implement inverse Slater matrix method 159 Computer science classes no longer put a heavy emphasis on optimization of algorithms. This is because computers are now fast enough to make compiler optimization suficient for many applications. Table 3 .6 illustrates the fact that this is not yet the case for quantum chemistry software. Fokker-Plank importance sampling introduces an additional approximation to the calculation. In order to test the magnitude of this error, the He atom test calculation that was used for time testing was performed with F okker-Plank importance sampling. The energy for this test should be -2.88 :1: 0.02 H. Fokker-Plank importance sampling with an optimal step size of 0.2 bohr gave an energy -2.628 :1: 0.008 H. While the standard deviation did decrease somewhat, the loss in accuracy was too great to be acceptable. No further use of Fokker-Plank importance sampling was made. 160 3. Scandium atom tests Quantum mechanics calculations on transition metals are particularly difficult. Transition metal calculations are very sensitive to electron correlation. A good description of electron correlation can be obtained by the use of CI calculations, which are extremely time consuming as shown in Table 3.1 . In order to address many chemical questions it is necessary to include spin coupling and mass defect terms in the Hamiltonian as well. These relativistic efi‘ects are usually negligible for first row elements but become significant for heavy elements. The choice to explore QMC methods was made based on the favorable time complexity (Table 3.1). A better time complexity will ensure that a method is more efiicient for suficiently large systems. For small systems this may not be true since the overhead in the calculation may be a larger contributing factor than the over all time complexity. A second advantage of QMC calculations is that additional terms can very readily be added to the Hamiltonian. For example, the program modifications necessary to perform ECP calculations were done in two days time. Although relativistic terms were not included in this study, the ease of adding them is an advantage which is expected to be utilized in the future. 161 The original goal of being able to do accurate calculations on compounds containing transition metals, has translated into the operational goals of 1. Evaluate the accuracy of VQMC methods. 2. Have the ability to use small test jobs to predict the quality of final results. 3. Evaluate the efficiency of VQMC methods. The first of these goals has already be well documented in the literature [21,32]. The second is accomplished with the use of Equation (12). This leaves the third yet to be addressed by comparing the computational work, CW, obtained for various methods (Equafion 13). Although many QMC studies have been done on light elements, the application to heavy elements presents additional difiiculties [21]. To date there have been four published QMC calculations on transition metals [33,57- 59]. Christiansen published a paper in which be computed excitation energies for Sc and Y atoms [57]. These were done as 11 electron calculations using core potentials created by Hurley et. al. [60] and J 31 ohn et. al. [61]. VQMC calculations were done using the J astro correlation frmction (Equation 51). 162 Belohorec, Rothstein and Vrbik published a calculation for the CuH molecule [33]. The primary focus of their work was the vibrational behavior. They introduce the split r method, described in the section on computer implementation, which they used with a damped core calculation. Damped core calculations use VQMC to describe the core region and diffusion Monte Carlo to describe the valence. They also utilized data blocking which is similar in result to the sampling fiequency tested in this study although somewhat more sophisticated. They did not publish a total energy or valence energy for their calculation which is highly Imusual for any study regardless of what properties are being studied. Mitas did a large study of [Ar] core ECP methods [58]. He used the same ECP set for HP, VQMC and DMC calculations on Sc, Ti, Mn, Fe, Cu and Cuz. He used numerical orbitals which were made piecewise smooth with cube spline fimctions. The correlation frmction used was an expansion of products of squares of Jastro type fimctions. Mitas computed ionization energies, electron affinities and the energy of excitation to low lying excited states. The QMC methods yielded reasonable agreement to experiment. His QMC calculations were consistently a significant improvement over HF calculations. 163 In a second study [59] Mitas used [Ne] core potentials for describing the Fe atom with a range of computational techniques. The methods examined were HF, VQMC, DMC, CCSD(T), QCISD(T) and CISD (These last three are variations on the multiple determinant CI method). The energies computed and reliability of results were the same as in work in the previous paragraph. The QMC methods yielded more accurate excitation energies than the large multiple determinant calculations. The following sections present a mnnber of algorithms applied to the 2D grormd state of the Sc atom. Each technique will be analyzed for both speed and accuracy. Two basis sets for the Sc atom are used. The all electron calculations use Clementi’s lls 6p 5d STO basis set ((E)HF = -759.73555 H) [56]. The core potential calculations use Hay and Wadt’s eflective core potentials (non- relativistic for Sc) [34]. The [Ne] core potentials representing the inner 10 electrons and the [Ar] core potentials representing the inner 18 electrons have been tested. The quantity E N — var{el,,} is given because it is a variational lower bormd on the total energy available fi'om VQMC calculations. 164 The acceptance ratio listed is the ratio of the nmnber of iterations on which the step was accepted to the total number of iterations. An acceptance ratio very close to one indicates that a larger step size may be more efficient. An acceptance ratio below 0.5 indicates that a smaller step size may be more efficient. A number of small calculations are presented in sections a - k. Each of these calculations is done with ten tests of one million iterations in order to compare CPU times. More accurate results are presented in section 1. Sections a - e present HF calculations (no correlation frmction) with VQMC integration. These provided a comparison of the speed and accuracy of various algorithms. HF calculations are done so that the accmacy can be determined by comparing the results to the known correct value for Clementi’s basis set. Sections f and g present correlated all electron calculations. This gives an illustration of the effect of correlating different shells of electrons. Sections h - k present core potential calculations with and without correlation frmctions. The results of sections a - k are summarized in Table 3.9 in the “Conclusions” section. 165 a. Simple Metromlis sampling This is the Metropolis algorithm with the new electron location chosen within a cube of a set size, referred to as a step size. The following is a summary of the performance and computational details for this calculation EN=4w7aosH EN —var{e1,,} = -12424 H CW = 184 Hmin (9) = - step size = 0.5 B (optimal value) acceptance ratio = 0.58 CPU time = 4 hr 33 min (SGI with MIPS 4400) This is the simplest algorithm to implement. Fmthermore it introduces no additional approximations into the calculation. It may not, however, be the most efficient computation. p. Split 1 technique This technique was original introduced for use in damped core calculations. It bears testing since this calculation is the only all electron QMC calculation published thus far. The parameter r was varied to find the 166 value that gave the lowest standard deviation and thus the most efficient calculation. EN —var{e1,,} = -14756 H CW = 2934 Hmin 0%) = -1.86 r = 2.5 (optimal value) acceptance ratio = 0.90 CPU time = 5 hr 26 min (SGI with MIPS 4400) The energy is two standard deviations from the correct value -759 H. The virial theorem check is likewise a bit far from the correct value of -2. Since the computational work is greater than that for simple Metropolis sampling, this method does not represent an improvement for VQMC calculations. 0. Vm'able t technique This technique uses a step size which is a fimction of the distance from the nucleus as given by Equations (30) and (31). It was created in an attempt to circumvent some of the step size bias due to a large discontinuity in step 167 size right at the interface between the core and valence region. The price for avoiding a large discontinuity in step size was that there was a small step size bias at practically every iteration rather than only at the steps where core and valence electrons interchange places. EN: -738. a 3 H EN -var{e1,,} = -52534 H CW = 2835 Hmin 0%) = -134 r = 0.5 acceptance ratio = 0.82 CPUtime = 5 hr 15 min (SGI with MIPS 4400) The energy of -7 38 H compared to the correct result of -759 H indicates that this approach introduced far more error than was eliminated. The virial theorem check of -1.34 gives a second confirmation of this fact. 91. Split 1 b technique This technique was proposed based on an analysis of the two previous techniques. EN: -758.5 a 0.6 H 168 EN —var{el,,} = -l3875 H CW=104Hmin 0% = -192 r = 2.0 (optimal value) acceptance ratio = 0.67 CPU time = 4 hr 48 min (SGI with MIPS 4400) Based on the computational work measm'e of . efficiency, this is an improvement over simple Metropolis sampling. This same method was also used to test the efl‘ect of introducing a th sample frequency in which the local energy is only summed on every m iteration. The results of sample frequency tests are given in Table 3 .7 169 Table 3.7 - Sample fiequency tests Sample fimuencv Er_1e_rg1_(H) 1 -758.5 i 0.6 2 -758 :1: l 5 -759 a: 2 10 -759 :1: 1 21 -759 :h 1 42 -759 :l: 1 Less fi'equent sampling did not decrease the standard deviation. Any improvements to the accuracy of calculation were not visible in the case of these small test calculations. e. Split 1, immrtance sampling with rejection Light element calculations have shown importance sampling techniques to be a more eficient way of performing numerical integrations. Importance sampling was used with the split r method because simple Metropolis sampling and split 1 b were not numerically stable (virial theorem and energies off by an order of magnitude). EN: -684. a 5 H 170 EN —var{e1,,} = .4951 H CW = 13050 Hmin 0% = -123 r = 0.5 (optimal value) acceptance ratio = 0.89 CPUtime= 8hr 42min (SGIwithMIPS4400) For Sc, this technique does not show the same increase in speed that has been observed for light elements. The energy and virial theorem checks are also very- poor. Tests with lighter atoms verified that this method was implemented correctly. 171 f. Split 1 b, correlating 3s 3p 3d 4s Here a Jastro correlation frmction is applied to all pairs of the outer ll electrons. This increases the accuracy of the calculation by including correlation. The Jastro frmction parameters, or and B, were varied to give the lowest total energy. EN: “760. i 2 H EN —var{e1,,} = -14297 H CW = 2156 Hmin 0%) = -197 a= 2.5 (optimal) fl= 3-0 (Optimal) r = 2.0 acceptance ratio = 0.69 CPU time= 8hr59min (SGI with MIPS 4400) The additional CPU time required for computing the correlated energy expression is evident by the comparison to the 4 hour CPU time in section d. 172 g. Split 1 b, correlating 28 2p 3s 3p 3d 48 A J astro correlation frmction was applied to all pairs of the outer 19 electrons. This gave the following results. EN: ’758.7 It 0.9 H EN -var{e1,,} = -13750 H CW = 787 Hmin 0%) = -1.92 a = 0.04 (optimal) )5 = 1-84 (Optimal) r = 2.0 acceptance ratio = 0.67 CPU time= 16hr 11 min (SGIwithMIPS4400) This not only required more CPU time, but gave a higher energy than the calculation in section f. This loss of accmacy can most logically be explained as being due to the inability of a single set of J astro parameters to represent the correlation in both valence and inner core regions. 173 h. lflelcore ECP VQMC with simple Metropolis sampling was used to do a HF calculation describing 11 electrons explicitly. This was done for the sake of timing comparison. It is interesting to note that ECP frmctions can be added into the VQMC energy expression extremely easily. In fact, the increase in speed due to the use of an ECP basis is greater for VQMC calculations than for convention HF GTO calculations. EN = -45.94 i 0.05 H EN —var{el,,} = -96 H CW = 0.27 H min step size = 1.0 B (optimal) acceptance ratio = 0.67 CPU time = 1 hr 47 min (SGI with MIPS 4400) The virial theorem test is not reported for core potential calculations because the virial theorem does not hold for these calculations, QMC or otherwise. 174 i. INeI core ECP with Jastro flmction A Jastro function was applied to all pairs of the 11 outer electrons to obtain EN = -45.96 a 0.06 H EN —Var{eln} = '96 H CW = 0.89 Hmin a = 0.01 (optimal) B = 0.5 (optimal) step size = 1.0 B acceptanceratio= 0.68 CPUtime= 4hr 6min (SGIwithMIPS4400) j. |Ar| core ECP A HF calculation integrating the outer 3 electrons explicitly gave EN = -153724 :1: 0.02 H EN -var{e1,,} = -67 H CW = 0.0084 Hmin step size = 2.5 (optimal) acceptance ratio = 0.65 175 CPU time = 21 min (SGI with MIPS 4400) k. |Ar| core ECP with Jastro ftmction A Jastro frmction applied to all pairs of the 3 outer electrons gave EN = -1.55 :l: 0.06 H EN -var{e1,,} = —64 H CW = 0.10 Hmin or = 2.5 (optimal) B = 3.0 (optimal) step size = 2.5 acceptance ratio = 0.66 CPU time = 28 min (SGI with MIPS 4400) 176 l. Accurate VQMC Mpg Due to the necessary CPU time for performing all electron calculations, even with the most efiicient methods tested, core potential calculations were used to test the accuracy of the energies obtained by VQMC. Table 3 .8 lists the valence energies and excitation energies obtained. The [Ar] core ECP calculations used 10 tests of 107 iterations. The [Ne] core ECP calculations used 10 tests of 108 iterations for the ground state and 22 tests of 108 iterations for the 4F state. The rest of the calculation details were thesameasinsectionsh-k. 177 Table 3.8 - Energies using ECP [Ar] core with J astro valence energies State Energyflfl 4s2 3d1 21) -1545 :1: 0.004 3d3 4F -1399 :1: 0.008 [Ne] core with Jastro valence energies Stag Energyfl) 4s2 3d1 21) 45.936 2 0.004 3d3 4F 45.786 :1: 0.008 4s2 3d1 -) 3d3 excitation energy (eV) HE V MC Exp. [Ar] core 4.42 3.97 4.19 [Ne] core 4.47 4.08 4.19 178 Table 3.9 - Computational work and energy CPU time‘ Mm Milan) E_N(Ii) Gauguin) [Ar] core potential 0.0085 -1.54 i 0.02 0:21 [Ar] core potential with correlation 0.10 -l .55 :1: 0.06 0:28 [Ne] core potential 0.27 -45.94 :t 0.05 1:47 [Ne] core potential with correlation 0.89 -45.96 :1: 0.06 4:06 Split tau b 104 -758.5 d: 0.6 4:48 Simple Metropolis sampling 184 -759.7 d: 0.8 4:33 Split tan b, correlating 19 electrons 787 -758.7 :1: 0.9 16:11 Split tan b, correlating ll electrons 2156 -760. :t 2 8:59 Variable tau 2835 -738. :t 3 5:15 Split tau 2934 -753. i 3 5:26 Split tau, importance sampling 13050 -684. :1: 5 8:42 with rejection "' SGI Onyx with MIPS 4400 processor 179 F. Conclusions The computational work for each of the methods discussed is summarized in Table 3.9 . As expected, the core potential calculations were the fastest. Although very fast, the [Ar] core potentials are not accurate enough for many chemical applications. The “split 1 b” method, introduced here, is the most efficient all- electron VQMC algorithm tested. Simple Metropolis sampling does sm'prisingly well for all electron transition metal calculations. The use of importance sampling, which has seen much use for light elements, was the least efficient method. Accurate ECP calculations are presented in Table 3 .8 . As expected, the correlated [Ne] core frmctions came the closest to the experimental results. Fmther improvement of the accuracy of these calculations can be achieved by the use of more complex correlation frmctions and optimization of orbital coefi'rcients. The [Ne] core potentials may be an optimal balance of speed and accuracy for many applications. The VQMC method has given us the ability to correlate 3s, 3p, 3d and 4s electrons for transition metals. The current generation of work station computers that are readily afi‘ordable by academic institutions typically have 180 several hundred Mb of RAM and several Gb of disk space. For such machines, VQMC represents the only available way to do such calculations with this level of accruacy. The most accurate [Ne] core ECP calculations required in excess of a week worth of CPU time on the fastest SGI work stations available. As such these calculations are just becoming practical for small parallel processing machines. The major downfall of all of the QMC methods is the requirement for large CPU times. However, this detriment is being overcome by increases in CPU performance and availability of parallel computing facilities. Indeed, VQMC calculations have the potential to be the next work horse of the computational chemistry industry. At the risk of being wrong, the following listing is an attempt to predict what advances to the VQMC method will be the most crucial in developing methods with the accuracy and eficiency already in demand by the chemical industry. The list is numbered in order of most important developments first. 1. Development of accurate correlation frmctions. 2. Convenient user interfaces. 3. Parallelization. 4. More eficient sampling schemes. LIST OF REFERENCES 10. 11. 12. 13. LIST OF REFERENCES E. Schrodinger Annln. Phys. :72, 361 and 489; 89, 437; 8_l, 109 (1926) L. Pauling, E. B. Wilson, Jr. “Introduction to Quantum Mechanics With Applications to Chemistry ” Dover (1985) I. N. Levine “Quantum Chemistry ” Prentice Hall (1991) D. R. Bates, K. Ledsham, A. L. Stewart, Philosophical Transactions of the Royal Society of London A246, 215 (1953) W. Bian, C. Deng Theor. Chim. Acta. 22, 135 (1995) R. G. Parr, W. Yang “Density-Functional Theory of Atoms and Molecules” Oxford (1989) D. R. Hartree Proc. Cambridge Phil. Soc. 251, 89, 111, 426 (1928) E. A. Hylleraas Z. Physik _SA, 347 (1929) Conference proceedings in Rev. Mod. Phys. volume _3_2, nmnber 2 (1960) T. L. Gilbert Rev. Mod. Phys. 3_5, 491 (1963) W. Kolos, C. C. J. Roothaan Rev. Mod. Phys. g, 205, 219, (1960) N. Metropolis, A. W. Rosenbluth, M. N. Rosenbluth, A. H. Teller, E. Teller J. Chem. Phys. 2_1, 1087 (1953) H. ConroyJ. Chem. Phys. 4_1_, 1327, 1331, 1336, 1341 (1964) 181 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25 . 26. 27. 28. 29. 30. 182 H. Conroy J. Chem. Phys. 41, 912, 930, 5307 (1967) H. Conroy, B. L. Bnmer J. Chem. Phys. _4_7, 921 (1967) D. Ceperley, G. V. Chester, M. H. Kalos Phys. Rev. B. 16, 3081 (1977) D. E. Knuth “Seminumerical Algorithms ” Addison Wesley (1969) J. Gleick “Chaos ” Viking (1987) w. H. Press, B. P. Flannery, s. A. Teukolsky, w. T. Vetterling “Numerical Recipies ” Cambridge (1989) I. Vattulainen, T. Ala-Nissila, K. Kankaala Phys. Rev. Lett. ‘_7_3_, 2513 (1994) B. L. Hammond, W. A. Lester, Jr., P. J. Reynolds “Monte Carlo Methods in Ab Initio Quantum Chemistry” World Scientific (1994) W. Hierse, P. M. Oppeneer Int. J. Quantum Chem. fl, 1249 (1994) H. W. Jones Int. J. Quantum Chem. 5_1_, 417 (1994) O. Treutler, R. Ahlrichs J. Chem. Phys. _lfl, 346 (1995) I. I. Guscinov, R. F. Yassen Int. J. Quantum Chem. 3, 197 (1995) D. R. Hartree, A. L. Ingman Mem. Proc. Manchester Lit & Phil. Soc. _71, 79 (1933) P. Pluvinage Ann. Phys. 5, 145 (1950) C. C. J. Roothaan, A. W. Weiss Rev. Mod. Phys. 32, 194 (1960) T. Kinoshita Phys. Rev. 10_5_, 1490 (1957) P. PulayJ. Comp. Chem. 3, 556 (1982) 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 183 M. H. Kalos, P. A. Witlock “Monte Carlo Methods Volume 1: Basics ” John Wiley & Sons (1986) J. B. Anderson, C. A. Traynor, B. M. Boghosian J. Chem. Phys. 92, 345 (1993) P. Belohorec, S. M. Rothstein, J. Vrbik J. Chem. Phys. E, 6401 (1993) P. J. Hay, W. R. Wadt J. Chem. Phys. 82, 270, 299 (1985) M. Krauss, W. J. Stevens Ann. Rev. Phys. Chem. 3_5_, 357 (1984) J. F. Harrison, K. L. Krmze “Organometallic Ion Chemistry” 89, Kluwer (1996) C. L. Pekeris Phys. Rev. _11_5_, 5 (1959) A. J. Thakkar, T. Koga Phys. Rev. A _S_Q, 845 (1994) D. E. Frermd, B. D. Huxtable, J. D. Morgan, III Phys. Rev. A Q, 980 (1984) J. Midtdal Phys. Math. Univ. Osloensis 32 (1966) T. Koga, Y. Kasai, A. J. Thakkar Int. J. Quantum Chem. 44, 689 (1993) H. M. Schwartz Phys. Rev. m3, 110 (1956) H. M. Schwartz Phys. Rev. m, 483 (1960) H. M. Schwartz Phys. Rev. m, 1029 (1963) C. Schwartz Phys. Rev. _12_8, 1146 (1962) r. P. Tsien, R. r. Pack J. Chem. Phys. 49, 4247 (1968) 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 184 H. O. Pritchard J. Chem. Phys. 49, 1496 (1964) W. A. Lester, Jr., M. Krauss J. Chem. Phys. fl, 1407 (1964) R. Jastrow Phys. Rev. 2_8_, 1479 (1955) S. E. Koonin “Computational Physics ” Benjamin/Cummings (1986) T. D. H. Baber, H. R. Hasse' Proc. Cambridge Phil. Soc. 33, 253 (1937) L. C. Green, C. Stephens, E. K. Kolchin, C. C. Chen, P. P. Rush J. Chem. Phys. fl, 1061 (1959) A. Bijl Physica 1, 869 (1940) J. w. Moskowitz, M. H. Kalos Int. J. Quantum Chem. 24, 1107 (1981) H.-J. Flad, A. Savin J. Chem. Phys. 143, 691 (1995) E. Clementi, “Ab Initio Computations in Atoms and Molecules”, IBM Journal of Research and Development 2, 2 (1965) P. A. Christianst Chem. Phys. _9_5_, 361 (1991) L. Mitds “Computer Simulation Studies in Condensed-Matter Physics V” Eds. D. P. Landau, K. K. Mon, H. B. Schr’ittler; Springer-Veflag (1993) L. Mitas Phys. Rev. A 42, 4411 (1994) M. M. Hurley, L. F. Pacios, P. A. Christiansen, R. B. Ross, W. C. Ermler J. Chem. Phys. 8_4, 6840 (1986) L. A. LaJohn, P. A. Christiansen, R. B. Ross, T. Atashroo, W. C. Ermler J. Chem. Phys. 8_7, 2812 (1987) Chapter IV PROPERTIES OF TOTAL ELECTRON DENSITY A. Introduction In the attempt to determine the physical nature of lone pair electrons (Chapter 11), one common misconception formd was that these electrons were visible as a maximum in the total electron density distribution. Orbital descriptions of bonding, although a useful description of chemistry, have also led to confusion between orbital densities and the total electron density. These misconceptions prompted the writing of a paper on electrdn density for the Journal of Chemical Education [1]. The focus of this work is a clarification of what theorems describing the electron density exist, what is typically observed and what exceptions have been formd. B. Misconceptions about electron density While students are taught orbital descriptions of bonding and geometry, they often have misconceptions about the name of the total electron density. For example, many students and experienced researchers expect to see large 185 186 bumps in the electron density where the lone pair electrons are located which simply don’t exist. Some misconceptions are starting to be eliminated through the availability of programs to graphically display the electron density. However the most general, and thus most powerful, statement of the form of the electron density is given by the associated theorems. Theorems are generalized descriptions of the behavior of all atoms or molecules. As such, theorems will describe trends which are true for many systems, but will not predict exact values for specific molecules. Although some scientists may be well aquainted with these theorems, many misconceptions remain among chemists at all levels. The remainder of this chapter will attempt to srunmarize the existing theorems describing the behavior of the electron density. For cases, where theorems have not been constructed, the normal behavior and anomalous cases will be described. We will also give a very brief description of some of the other properties which are often confused with the electron density. 187 EMILE The electron density, p, is the number of electrons expected to be found per 1mit volume of space, usually expressed in units of electrons/A3 . This is the number of electrons times the probability of finding an electron. As such, we do not examine the trajectory of a single electron as we would the planets in the solar system. In an orbital description this is the smn of the densities of the occupied orbitals. The electron density is sometimes referred to as the total electron density in order to difierentiate it from orbital densities. For the hydrogen atom, the density is known exactly [2]. For the ground state of the hydrogen atom p0) = ire-2' (1) 7: Thus the density is a maximum on the nucleus and decays exponentially and monotonically (continuously) to zero at infinite distance as shown in figure 4.1 . The electron density is spread away from the proton in order to obtain the lowest energy balance between potential energy which is lowest near the nucleus and kinetic energy which is highest near the nucleus. For excited states, the decay to infinity is still exponential, but there is not necessarily a 188 Figure 4.1 - Electron density for the grolmd state of the H atom 13 14 15 1.64 174» 1a 1.9 1) 189 peak on the nucleus, nor is the decay monotonic, as evidenced by the 2p density shown in Figure 4.2 . D. Theorems The wave frmction, and thus p, must decay to zero at long distances from an atom or molecule in order to satisfy the normalizability condition [2]. An upper bormd for the decay of the electron density to zero at infinite distance was derived by Hofl'rnann-Ostenhof & Hofinann-Ostenhof [3]. Their proof utilized properties of differential equations, the variational principle and the Cauchy-Schwartz inequality. For molecules, they found that p(r) s k r‘2 (r — p)"’“ e”“"”’ for large values of r (2) a = (28)"2 (3) where k is a constant, a is the ionization potential of the molecule, Z is the sum of nuclear charges, r is the distance from the center of mass and p is the distance fi'om the center of mass to the nucleus farthest fi'om the center of mass. For atoms this reduces to p(r) s k r” e’“ (4) fl=Z/a-1 (5) 190 Figure 4.2 - Electron density extending away from the nucleus of a hydrogen atom in the excited state with the electron in a 2p orbital. 191 This work is based upon the analysis of wave frmction behavior by Ahhiches [4-6]. A more detailed mathematical wave frmction analysis has been published by Agrnon [7]. Ta] [8] increased the accmacy of the previous expression for atoms by deriving fl=Z-n+l_1 (6) a where n is the number of electrons. His derivation utilized a density matrix theory formulation. This is in agreement with the wave frmction analysis of Katriel and Davidson [9]. For the ground states of atoms and atomic ions, Weinstein, Politzer & Srebrenik [10] have shown that the density is a maximum at the nucleus and decays monotonically. This is proven as a statement which they did not put in equation form. Their analysis shows this to be a direct consequence of Poisson’s equation. These findings are further supported by the approximate analysis of Angulo & Dehesa [l 1] based on the Stieltjes-moment-problem technique. ‘ The preceding results show that the electron density of the grormd state of all atoms is expected to look qualitatively like the hydrogen atom ls 192 density plotted in Figure 4.1 . For excited states of atoms, the density need not decay monotonically or be a maximum on the nucleus [10]. Hofinann-Ostenhof & Morgan [12] have analyzed the behavior of one electron homonuclear diatomic molecules (H; , He;3 , etc.). They have shown that there can be maxima in the electron density only on the nuclei. They have also shown that there are no local minima. The density is a minimmn at the center of the bond along the line connecting the nuclei and a maxirmun along lines perpendicular to the bond. They derived these results by analyzing the wave frmction with differential inequality techniques in several coordinate systems. E. Cases nqt covered by Theorems Although no formal proofs have been presented regarding maxima and monotonicity in molecules, many studies have been done [13-16]. For the grormd state electron distribution in almost all of the molecules ever studied the only maxima in the electron density are on the nuclei. This is illustrated by Figure 4.3 in which the electron density in the plane of the Czl-Ia molecule is plotted as a displacement in the third dimension. The peaks on the carbon atoms have been tnmcated in order to make the hydrogen atom peaks visible. 193 Figure 4.3 - Total electron density in the plane of the C211. molecule plotted as the height above the plane. The density was obtained from a CISD calculation with a 6-311G*"‘-H- basis set. J/\/ (7\ 194 The density around each atom appears nearly spherical when plotted as shown for the CO molecule in Figure 4.4 . The bonding and lone pair regions show some elongation. Note that there are no pronounced tear-drop shaped regions associated with individual bonds or lone pair electrons. Thus the electron density for a molecule is very similar to the sum of the densities of the component atoms. The shift in electron density upon forming a molecule from the separated atoms is typically less than or equal to 10% of the atomic densities [10]. In a few rare cases, a peak in the electron density is found in the middle of the bond for the grormd state of a molecule. This maximum in the bond region is referred to as a nonnuclear attractor. The most pronomrced and well studied case is the Li; molecule [17]. The magnitude of this peak is three orders of magnitude smaller than the peaks on the nuclei as shown in Figure 4.5 . This behavior has been shown to be expected only in very polarizable systems such as Li and Na clusters. F. Other molecular progrties The remaining question is “Where does the chemistry come fi'om?” Chemistry is energetically a very subtle effect as evidenced by a carbon- 195 Figure 4.4 - contour diagram of the electron density of the CO molecule. The carbon atom is on the. left. Contours are plotted for density values of 0.1, 0.2, 0.3, 0.4, 0.5, 1.0 and 5.0 electrons per cubic bohr. The density was obtained from a configuration interaction calculation with a 6-311G*+ basis set. The computed bond length was 1.125 Angstroms. /\ (a V/ 196 Figure 4.5 - Plot of the electron density along the internuclear line of the Li; molecule illustrating the magnitude of the maximrun in the middle of the bond. The density was obtained from a CISD calculation with a 6-311G“+ basis set and a bond length of 2.692 A L 1 . fi' T ‘ _ ‘T...’ ____,.__.——-—— “.- - 537:» density (olAngstrom"3) '8 8 8 S 8 3‘ 8 8 ‘ C G L T //’ / / t\ i i. ( AAAAAA -25 .23 I. -21 i -1 9 J -17 -15 -13 -1 1 . -03 Y -07 i .95 -03 ........................... ~ 0 ....................................................................................................................... a a oooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooooo 197 carbon single bond having an energy of arormd 4.3 eV while 1029.8 eV are required to remove all of the electrons fi'om a carbon atom. There are a number of molecular properties, other than the electron density, which illustrate chemical structure much more graphically. The following are short descriptions of a number of such properties and their relationship to the electron density. These are mentioned here only in order to decrease confusion. It is not the intention of this paper to give a detailed description of these properties. Electron density difference plots show the result of subtracting the atom densities fi'om the molecular densities. This gives an indication of how the density shifts on bond formation rather than the absolute density [2, 16]. In radical compormds, the spins of all of the electrons do not cancel to a net zero spin. Spin densities are plots of the net electron spin in radical compounds [13, l4, 16]. The electrons in an atom are often assigned to K, L and M shells. This atomic shell structure is seen in the total radial distribution fimction but not in the electron density [2]. The radial distribution function gives the probability of finding an electron at a given distance from the nucleus of an atom. The radial distribution frmction is obtained by integrating over the angular 198 variables in order to obtain an angular averaged frmction, which is a function only of the radial distance from the nucleus. Much of our understanding of bonding comes from molecular orbitals [2, 13, 14]. Orbitals, such as those obtained from Hartree-Fock calculations, are one electron frmctions, meaning that they show the behavior of one electron in the presence of all of the other electrons. Orbitals, which are similar but not identical, can be generated from other calculations such as configruation interaction and density frmctional theory calculations. Molecular orbitals can show quite a bit of feature such as sigma and pi bonding characters or electron lone pairs. The square of an orbital gives the contribution to the electron density from that orbital. The total electron density is the sum of the electron densities from all of the orbitals. The energy of an infinitesimally small charge placed at a point in space is the charge times the molecular electrostatic potential evaluated at that point [16]. The electrostatic potential at a point in space is calculatedby summing the repulsions with all of the nuclei and adding to that the integral of the attractions to the electron density at all points in space. This is sometimes used as an indication of hydrophobic vs. hydrophilic regions or equated to the behavior of lone pair electrons. 199 The gradient of the electron density is a vector quantity. This shows how electrons would move most rapidly from regions of high electron density to regions of low electron density [15]. The Laplacian of the electron density, V2 p , shows the behavior of the electron density relative to an exponential decay. There will be a negative minimum in the Laplacian of the electron density where there is the maximum positive deviation from exponential decay. Bader and coworkers have utilized this property to interpret electron lone pairs as maxima in plots of —V2 p [15]. Note that this is a maximum relative to an exponential decay, but is still a region of monotonically decaying electron density. G. Conclusions A nmnber of theorems have been constructed to describe the behavior of the electron density for simple systems. However, existing theorems do not describe all aspects of the electron density for all systems. Descriptions of the expected behavior of the electron density in molecules are supported by a large volume of theoretical and experimental evidence. The electron density is for the most part featureless. The true nature of the total electron density has been shown to be devoid of large peaks denoting bonds and lone pair electrons. 'Featm’es which are sometimes 200 attributed to the electron density are often those of other properties which focus on the more subtle effects which dictate the chemistry. Most of these properties are related to the electron density in some way. Much insight can be gained from these related techniques. However, their behavior should not be confused with that of the total electron density. LIST OF REFERENCES 10. ll. 12. 13. LIST OF REFERENCES D. C. Young submitted to J. Chem. Educ. I. N. Levine “Quantum Chemist ” Prentice Hall (1991) M. Hoflinann-Ostenhof, T. Hofirnmn-Ostenhof Phys. Rev. A ,1_6, 1782 (1977) R. Ahlrichs Chem. Phys. Lett. ii, 609 (1972) R. Ahlrichs J. Math. Phys. 1;, 1860 (1973) R. Ahlrichs Chem. Phys. Lett. _1_8, 521 (1973) S. Agrnon “Lectures on Exponential Decay of Solutions of Second- Order Elliptic Equations: Bounds on E igenfimctions of N-Body Schrodinger Operators” Princeton Univerisity Press (1982) Y. Tal Phys. Rev. A _1__8_, 1781 (1978) J. Katriel, E. R. Davidson Proc. Natl. Acad. Sci. USA _71, 4403 (1980) H. Weinstein, P. Politzer, S. Srebrenik Theoret. Chim. Acta (Berl.) 38, 159 (197 5) J. C. Angulo, J. S. Dehesa Phys. Rev. A 44, 1516 (1991) T. Hofi'mann-Ostenhof, J. D. Morgan III J. Chem. Phys. 2;, 843 (1981) E. R. Davidson “Reduced Density Matrices in Quantum Chemistry” Academic Press (1976) 201 14. 15. 16. 17. 202 R. G. Parr, W. Yang “Density-Functional Theory of Atoms and Molecules” Oxford University Press (1989) R. F. W. Bader “Atoms in Molecules A Quantum Theory” Clarendon Press (1994) P. Coppens, M. B. Hall, Eds. “Electron Distributions and the Chemical Bond” Plenum Press (1982) G. I. Bersuker, C. Peng, J. E. Boggs J. Phys. Chem. 91, 9323 (1993) "tannin)“