Skip to main content
Log in

Physical mechanisms of intermolecular interactions from symmetry-adapted perturbation theory

  • Original Paper
  • Published:
Journal of Molecular Modeling Aims and scope Submit manuscript

Abstract

Symmetry-adapted perturbation theory (SAPT) is a method for computational studies of noncovalent interactions between molecules. This method will be discussed here from the perspective of establishing the paradigm for understanding mechanisms of intermolecular interactions. SAPT interaction energies are obtained as sums of several contributions. Each contribution possesses a clear physical interpretation as it results from some specific physical process. It also exhibits a specific dependence on the intermolecular separation R. The four major contributions are the electrostatic, induction, dispersion, and exchange energies, each due to a different mechanism, valid at any R. In addition, at large R, SAPT interaction energies are seamlessly connected with the corresponding terms in the asymptotic multipole expansion of interaction energy in inverse powers of R. Since such expansion explicitly depends on monomers’ multipole moments and polarizabilities, this connection provides additional insights by rigorously relating interaction energies to monomers’ properties.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Data availability

Not applicable.

Code availability

Not applicable.

References

  1. Hirschfelder JO (1967) Perturbation theory for exchange forces, I. Chem Phys Lett 1:325

    Article  CAS  Google Scholar 

  2. Jeziorski B, Chałasiński G, Szalewicz K (1978) Symmetry forcing and convergence properties of perturbation expansions for molecular interaction energies. Int J Quantum Chem 14:271–287. https://doi.org/10.1002/qua.560140306

    Article  CAS  Google Scholar 

  3. Hirschfelder JO (1967) Perturbation theory for exchange forces, II. Chem Phys Lett 1:363–368

    Article  Google Scholar 

  4. van der Avoird A (1967) Perturbation theory for intermolecular interactions in the wave-operator formalism. J Chem Phys 47:3649

    Article  Google Scholar 

  5. Musher JI, Amos AT (1967) Theory of weak atomic and molecular interactions. Phys Rev 164:31

    Article  CAS  Google Scholar 

  6. Murrell JN, Shaw G (1967) Intermolecular forces in region of small orbital overlap. J Chem Phys 46:1768

    Article  CAS  Google Scholar 

  7. Chipman DM, Bowman JD, Hirschfelder JO (1973) Perturbation theories for the calculations of molecular interaction energies. I. General formulation. J Chem Phys 59:2830–2837

    Article  CAS  Google Scholar 

  8. Chipman DM, Hirschfelder JO (1973) Perturbation theories for the calculations of molecular interaction energies. II. Application to \(\mathrm H_2^+\). J Chem Phys 59:2838–2857

  9. Chałasiński G, Jeziorski B, Szalewicz K (1977) On the convergence properties of the Rayleigh-Schrödinger and Hirschfelder-Silbey perturbation expansions for molecular interaction energies. Int J Quantum Chem 11:247–257. https://doi.org/10.1002/qua.560110205

    Article  Google Scholar 

  10. Adams WH, Polymeropoulos EE (1978) Exchange perturbation theory. I. General definitions and relations. Phys Rev A 17:11–17. https://doi.org/10.1103/PhysRevA.17.11

    Article  CAS  Google Scholar 

  11. Chałasiński G, Szalewicz K (1980) Degenerate symmetry-adapted perturbation theory. Convergence properties of perturbation expansions for excited states of \(\mathrm H_2^+\) ion. Int J Quantum Chem 18:1071–1089. https://doi.org/10.1002/qua.560180414

  12. Jeziorski B, Schwalm WA, Szalewicz K (1980) Analytic continuation in exchange perturbation theory. J Chem Phys 73:6215–6222. https://doi.org/10.1063/1.440116

    Article  CAS  Google Scholar 

  13. Adams WH (1990) Perturbation-theory of intermolecular interactions - what is the problem, are there solutions? Int J Quantum Chem S24:531–547. https://doi.org/10.1002/qua.560382452

    Article  Google Scholar 

  14. Adams WH (1992) The problem of unphysical states in the theory of intermolecular interactions. J Math Chem 10:1–23. https://doi.org/10.1007/BF01169168

    Article  CAS  Google Scholar 

  15. Ćwiok T, Jeziorski B, Kołos W, Moszyński R, Szalewicz K (1992) On the convergence of the symmetrized Rayleigh-Schrödinger perturbation theory for molecular interaction energies. J Chem Phys 97:7555–7559

    Article  Google Scholar 

  16. Patkowski K, Jeziorski B, Korona T, Szalewicz K (2002) Symmetry-forcing procedure and convergence behavior of perturbation expansions for molecular interaction energies. J Chem Phys 117:5124–5134. https://doi.org/10.1063/1.1499488

    Article  CAS  Google Scholar 

  17. Adams WH (2002) Two new symmetry-adapted perturbation theories for the calculation of intermolecular interaction energies. Theor Chem Acc 108:225–231. https://doi.org/10.1007/s00214-002-0377-3

    Article  CAS  Google Scholar 

  18. Patkowski K, Jeziorski B, Szalewicz K (2004) Unified treatment of chemical and van der Waals forces via symmetry-adapted perturbation expansion. J Chem Phys 120:6849–6862. https://doi.org/10.1063/1.1676119

    Article  CAS  PubMed  Google Scholar 

  19. Jeziorski B, Bulski M, Piela L (1976) First-order perturbation treatment of the short-range repulsion in a system of many closed-shell atoms or molecules. Int J Quantum Chem 10:281–297. https://doi.org/10.1002/qua.560100208

    Article  CAS  Google Scholar 

  20. Jeziorski B, van Hemert M (1976) Variation-perturbation treatment of hydrogen bond between water molecules. Mol Phys 31:713–729. https://doi.org/10.1080/00268977600100551

    Article  CAS  Google Scholar 

  21. Szalewicz K, Jeziorski B (1979) Symmetry-adapted double-perturbation analysis of intramolecular correlation effects in weak intermolecular interactions. Mol Phys 38:191–208. https://doi.org/10.1080/00268977900101601

    Article  CAS  Google Scholar 

  22. Rybak S, Jeziorski B, Szalewicz K (1991) Many-body symmetry-adapted perturbation theory of intermolecular interactions - H2O and HF dimers. J Chem Phys 95:6576–6601. https://doi.org/10.1063/1.461528

    Article  CAS  Google Scholar 

  23. Lotrich VF, Szalewicz K (1997) Symmetry-adapted perturbation theory of three-body nonadditivity of intermolecular interaction energy. J Chem Phys 106:9668–9687. https://doi.org/10.1063/1.473831

    Article  CAS  Google Scholar 

  24. Misquitta AJ, Jeziorski B, Szalewicz K (2003) Dispersion energy from density-functional theory description of monomers. Phys Rev Lett 91:033201. https://doi.org/10.1103/PhysRevLett.91.033201

    Article  CAS  PubMed  Google Scholar 

  25. Heßelmann A, Jansen G (2003) Intermolecular dispersion energies from time-dependent density functional theory. Chem Phys Lett 367:778–784. https://doi.org/10.1016/S0009-2614(02)01796-7

    Article  Google Scholar 

  26. Misquitta AJ, Podeszwa R, Jeziorski B, Szalewicz K (2005) Intermolecular potentials based on symmetry-adapted perturbation theory including dispersion energies from time-dependent density functional calculations. J Chem Phys 123:214103. https://doi.org/10.1063/1.2135288

    Article  CAS  PubMed  Google Scholar 

  27. Heßelmann A, Jansen G, Schütz M (2005) Density-functional theory-symmetry-adapted intermolecular perturbation theory with density fitting: a new efficient method to study intermolecular interaction energies. J Chem Phys 122:014103. https://doi.org/10.1063/1.1824898

    Article  CAS  Google Scholar 

  28. Podeszwa R, Szalewicz K (2007) Three-body symmetry-adapted perturbation theory based on Kohn-Sham description of the monomers. J Chem Phys 126:194101. https://doi.org/10.1063/1.2733648

    Article  CAS  PubMed  Google Scholar 

  29. Zuchowski PS, Podeszwa R, Moszyński R, Jeziorski B, Szalewicz K (2008) Symmetry-adapted perturbation theory utilizing density functional description of monomers for high-spin open-shell complexes. J Chem Phys 129:084101. https://doi.org/10.1063/1.296855

    Article  PubMed  Google Scholar 

  30. Pernal K, Szalewicz K (2009) Third-order dispersion energy from response functions. J Chem Phys 130:034103–(1-5). https://doi.org/10.1063/1.3058477

    Article  CAS  Google Scholar 

  31. Patkowski K, Zuchowski PS, Smith DGA (2018) First-order symmetry-adapted perturbation theory for multiplet splittings. J Chem Phys 148:164110. https://doi.org/10.1063/1.5021891

    Article  PubMed  CAS  Google Scholar 

  32. Hapka M, Przybytek M, Pernal K (2019) Second-order dispersion energy based on multireference description of monomers. J Chem Theory Comput 15:1016–1027. https://doi.org/10.1021/acs.jctc.8b01058

    Article  CAS  PubMed  Google Scholar 

  33. Hapka M, Przybytek M, Pernal K (2019) Second-order exchange-dispersion energy based on multireference description of monomers. J Chem Theory Comput 15:6712–6723. https://doi.org/10.1021/acs.jctc.9b00925

    Article  CAS  PubMed  Google Scholar 

  34. Hohenstein EG, Sherrill CD (2010) Density fitting and Cholesky decomposition approximations in symmetry-adapted perturbation theory: implementation and application to probe the nature of pi-pi interactions in linear acenes. J Chem Phys 132:184111–(1-10)

    Article  CAS  Google Scholar 

  35. Hohenstein EG, Sherrill CD (2010) Density fitting of intramonomer correlation effects in symmetry-adapted perturbation theory. J Chem Phys 133:014101–(1-12)

    Google Scholar 

  36. Hohenstein EG, Sherrill CD (2010) Efficient evaluation of triple excitations in symmetry-adapted perturbation theory via second-order Møller-Plesset perturbation theory natural orbitals. J Chem Phys 133:104107–(1-7)

    Google Scholar 

  37. Hohenstein EG, Parrish RM, Sherrill CD, Turney JM, Schaefer HF (2011) Large-scale symmetry-adapted perturbation theory computations via density fitting and Laplace transformation techniques: investigating the fundamental forces of DNA-intercalator interactions. J Chem Phys 135:174107

    Article  PubMed  CAS  Google Scholar 

  38. Parker TM, Burns LA, Parrish RM, Ryno AG, Sherrill CD (2014) Levels of symmetry adapted perturbation theory (SAPT). I. Efficiency and performance for interaction energies. J Chem Phys 140:094106

    Article  PubMed  CAS  Google Scholar 

  39. Gonthier JF, Sherrill CD (2016) Density-fitted open-shell symmetry-adapted perturbation theory and application to pi-stacking in benzene dimer cation and ionized DNA base pair steps. J Chem Phys 145:134106

    Article  PubMed  CAS  Google Scholar 

  40. Maurer SA, Beer M, Lambrecht DS, Ochsenfeld C (2013) Linear-scaling symmetry-adapted perturbation theory with scaled dispersion. J Chem Phys 139:184104–(1-7)

    Google Scholar 

  41. Parrish RM, Thompson KC, Martínez TJ (2018) Large-scale functional group symmetry-adapted perturbation theory on graphical processing units. J Chem Theory Comput 14:1737–1753. https://doi.org/10.1021/acs.jctc.7b01053

    Article  CAS  PubMed  Google Scholar 

  42. Garcia J, Szalewicz K (2022) Efficient calculations of interaction energies using symmetry-adapted perturbation theory based on density functional theory description of monomers. J Chem Theory Comput (Manuscript in preparation)

  43. Xie W, Gao J (2007) Design of a next generation force field: the X-POL potential. J Chem Theory Comput 3:1890–1900

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Gordon MS, Slipchenko LV, Li H, Jensen JH (2007) The effective fragment potential: a general method for predicting intermolecular interactions. Ann Rep Comp Chem 3:177–193

    CAS  Google Scholar 

  45. Jacobson LD, Herbert JM (2011) An efficient, fragment-based electronic structure method for molecular systems: self-consistent polarization with perturbative two-body exchange and dispersion. J Chem Phys 134:094118–(1-17)

    Article  CAS  Google Scholar 

  46. Flick JC, Kosenkov D, Hohenstein EG, Sherrill CD, Slipchenko LV (2012) Accurate prediction of noncovalent interaction energies with the effective fragment potential method: comparison of energy components to symmetry-adapted perturbation theory for the S22 test set. J Chem Theory Comput 8:2835–2843

    Article  CAS  PubMed  Google Scholar 

  47. Lao KU, Herbert JM (2012) Accurate intermolecular interactions at dramatically reduced cost: XPol+SAPT with empirical dispersion. J Chem Phys Lett 3:3241–3249

    Article  CAS  Google Scholar 

  48. Lao KU, Herbert JM (2013) An improved treatment of empirical dispersion and a many-body energy decomposition scheme for the explicit polarization plus symmetry-adapted perturbation theory (XSAPT) method. J Chem Phys 139:034107

    Article  PubMed  CAS  Google Scholar 

  49. Lao KU, Herbert JM (2015) Accurate and efficient quantum chemistry calculations for noncovalent interactions in many-body systems: the XSAPT family of methods. J Phys Chem A 119:235–252

    Article  CAS  PubMed  Google Scholar 

  50. Lao KU, Herbert JM (2018) Atomic orbital implementation of extended symmetry-adapted perturbation theory (XSAPT) and benchmark calculations for large supramolecular complexes. J Chem Theory Comput 14:2955–2978

    Article  CAS  PubMed  Google Scholar 

  51. Kim Y, Bui Y, Tazhigulov RN, Bravaya KB, Slipchenko VL (2020) Effective fragment potentials for flexible molecules: transferability of parameters and amino acid database. J Chem Theory Comput 16:7735–7747

    Article  CAS  PubMed  Google Scholar 

  52. Carter-Fenk K, Lao KU, Herbert JM (2021) Predicting and understanding non-covalent interactions using novel forms of symmetry-adapted perturbation theory. Acc Chem Res 54:3679–3690

    Article  CAS  PubMed  Google Scholar 

  53. Jeziorski B, Moszyński R, Szalewicz K (1994) Perturbation theory approach to intermolecular potential energy surfaces of van der Waals complexes. Chem Rev 94:1887–1930. https://doi.org/10.1021/cr00031a008

    Article  CAS  Google Scholar 

  54. Szalewicz K, Patkowski K, Jeziorski B (2005) Intermolecular interactions via perturbation theory: from diatoms to biomolecules. Struct & Bond 116:43–117. https://doi.org/10.1007/430_004

    Article  CAS  Google Scholar 

  55. Szalewicz K (2012) Symmetry-adapted perturbation theory of intermolecular forces. Wiley Interdisc Rev–Comp Mol Sci 2:254–272. https://doi.org/10.1002/wcms.86

    Article  CAS  Google Scholar 

  56. Hohenstein EG, Sherrill CD (2012) Wavefunction methods for noncovalent interactions. Wiley Interdisc Rev–Comp Mol Sci 2:304–326

    Article  CAS  Google Scholar 

  57. Jansen G (2014) Symmetry-adapted perturbation theory based on density functional theory for noncovalent interactions. Wiley Interdisc Rev–Comp Mol Sci 4:127–144. https://doi.org/10.1002/wcms.1164

    Article  CAS  Google Scholar 

  58. Patkowski K (2019) Recent developments in symmetry-adapted perturbation theory. Wiley Interdisc Rev–Comp Mol Sci 10:e1452. https://doi.org/10.1002/wcms.1452

    Article  CAS  Google Scholar 

  59. Garcia J, Podeszwa R, Szalewicz K (2020) SAPT Codes for calculations of intermolecular interaction energies. J Chem Phys 152:184109–(1-23). https://doi.org/10.1063/5.0005093

    Article  CAS  Google Scholar 

  60. Kaplan IG (1986) Theory of molecular interactions. Elsevier, Amsterdam, ISBN: 978-0444426963

  61. Arrighini P (1981) Intermolecular forces and their evaluation by perturbation theory, Lecture Notes in Chemistry, vol 25. Springer, Berlin. https://doi.org/10.1007/978-3-642-93182-6

    Book  Google Scholar 

  62. Stone AJ (2013) The theory of intermolecular forces, 2nd edn. Clarendon Press, Oxford. https://doi.org/10.1093/acprof:oso/9780199672394.001.0001

    Book  Google Scholar 

  63. Micha DA (2020) Molecular interactions: concepts and methods. Cambridge University Press, ISBN: 978-1-119-31907-8

  64. Ángyán J, Dobson J, Jansen G, Gould T (2020) London dispersion forces in molecules, solids and nano-structures: an introduction to physical models and computational methods. Royal Society of Chemistry, Cambridge, UK. https://doi.org/10.1039/9781782623861

    Google Scholar 

  65. Boys SF, Bernardi F (1970) The calculation of small molecular interactions by the differences of separate total energies. Some procedures with reduced errors. Mol Phys 19:553–566

    Article  CAS  Google Scholar 

  66. Williams HL, Korona T, Bukowski R, Jeziorski B, Szalewicz K (1996) Helium dimer potential from symmetry-adapted perturbation theory. Chem Phys Lett 262:431–436. https://doi.org/10.1063/1.473556

    Article  CAS  Google Scholar 

  67. Bukowski R, Jeziorski B, Szalewicz K (1996) Basis set superposition problem in interaction energy calculations with explicitly correlated bases. Saturated second- and third-order energies for He2. J Chem Phys 104:3306

    Article  CAS  Google Scholar 

  68. Lesiuk M, Jeziorski B (2019) Size consistency and counterpoise correction in explicitly correlated calculations of interaction energies and interaction-induced properties. J Chem Theory Comput 15:5398–5403

    Article  CAS  PubMed  Google Scholar 

  69. Kitaura K, Morokuma K (1976) A new energy decomposition scheme for molecular interactions within the Hartree-Fock approximation. Int J Quantum Chem 10:325–340

    Article  CAS  Google Scholar 

  70. Mo Y, Gao J, Peyerimhoff S (2000) Energy decomposition analysis of intermolecular interactions using a block-localized wave function approach. J Chem Phys 112:5530–5538

    Article  CAS  Google Scholar 

  71. Stevens WJ, Fink WH (1987) Frozen fragment reduced variational space analysis of hydrogen-bonding interactions - application to the water dimer. Chem Phys Lett 139:15–22. https://doi.org/10.1016/0009-2614(87)80143-4

    Article  CAS  Google Scholar 

  72. Khaliullin RZ, Cobar EA, Lochan RC, Bell AT, Head-Gordon M (2007) Unravelling the origin of intermolecular interactions using absolutely localized molecular orbitals. J Phys Chem A 111:8753–8765. https://doi.org/10.1021/jp073685z

    Article  CAS  PubMed  Google Scholar 

  73. Horn PR, Mao Y, Head-Gordon M (2016) Defining the contributions of permanent electrostatics, Pauli repulsion, and dispersion in density functional theory calculations of intermolecular interaction energies. J Chem Phys 144:114107

    Article  PubMed  CAS  Google Scholar 

  74. Horn PR, Mao Y, Head-Gordon M (2016) Probing non-covalent interactions with a second generation energy decomposition analysis using absolutely localized molecular orbitals. Phys Chem Chem Phys 18:23067–23079

    Article  CAS  PubMed  Google Scholar 

  75. Andres J, Ayers PW, Boto RA, Carbo-Dorca R, Chermette H, Cioslowski J, Contreras-Garcia J, Cooper DL, Frenking G, Gatti C, Heidar-Zadeh F, Joubert L, Martin Pendas A, Matito E, Mayer I, Misquitta AJ, Mo Y, Pilme J, Popelier PLA, Rahm M, RamosCordoba E, Salvador P, Schwarz WHE, Shahbazian S, Silvi B, Sola M, Szalewicz K, Tognetti V, Weinhold F, Zins EL (2019) Nine questions on energy decomposition analysis. J Comp Chem 40:2248–2283. https://doi.org/10.1002/jcc.26003

    Article  CAS  Google Scholar 

  76. Veccham SP, Lee J, Mao Y, Horn PR, Head-Gordon M (2021) A non-perturbative pairwise-additive analysis of charge transfer contributions to intermolecular interaction energies. Phys Chem Chem Phys 23:928–943

    Article  CAS  PubMed  Google Scholar 

  77. Naseem-Khan S, Gresh N, Misquitta AJ, Piquemal JP (2021) Assessment of SAPT and supermolecular EDA approaches for the development of separable and polarizable force fields. J Chem Theory Comput 17:2759–2774

    Article  CAS  PubMed  Google Scholar 

  78. Ahlrichs R (1976) Convergence properties of the intermolecular force series (1/R-expansion). Theor Chim Acta 41:7–15

    Article  CAS  Google Scholar 

  79. Kreek H, Meath WJ (1969) Charge-overlap effects. Dispersion and induction forces. J Chem Phys 50:2289

    Article  CAS  Google Scholar 

  80. Knowles PJ, Meath WJ (1986) Non-expanded dispersion and damping functions for Ar2 and Li2. Chem Phys Lett 124:164–171

    Article  CAS  Google Scholar 

  81. Knowles PJ, Meath WJ (1986) Non-expanded dispersion and induction energies, and damping functions, for molecular-interactions with application to HF-He. Mol Phys 59:965–984

    Article  CAS  Google Scholar 

  82. Knowles PJ, Meath WJ (1987) A separable method for the calculation of dispersion and induction energy damping functions with applications to the dimers arising from He, Ne and HF. Mol Phys 60:1143–1158

    Article  CAS  Google Scholar 

  83. Tang KT, Toennies JP (1984) An improved simple-model for the van der Waals potential based on universal damping functions for the dispersion coefficients. J Chem Phys 80:3726–3741

    Article  CAS  Google Scholar 

  84. Lotrich VF, Williams HL, Szalewicz K, Jeziorski B, Moszyński R, Wormer PES, van der Avoird A (1995) Intermolecular potential and rovibrational levels of Ar-HF from symmetry-adapted perturbation theory. J Chem Phys 103:6076–6092. https://doi.org/10.1063/1.470436

    Article  CAS  Google Scholar 

  85. Stone AJ (2017) Natural bond orbitals and the nature of the hydrogen bond. J Phys Chem A 121:1531–1534. https://doi.org/10.1021/acs.jpca.6b12930

    Article  CAS  PubMed  Google Scholar 

  86. Stone AJ, Szalewicz K (2018) Reply to “Comment on Natural bond orbitals and the nature of the hydrogen bond”. J Phys Chem A 122:733–736. https://doi.org/10.1021/acs.jpca.7b09307

    Article  CAS  PubMed  Google Scholar 

  87. Ćwiok T, Jeziorski B, Kołos W, Moszyński R, Rychlewski J, Szalewicz K (1992) Convergence properties and large-order behavior of the polarization expansion for the interaction energy of hydrogen atoms. Chem Phys Lett 195:67–76. https://doi.org/10.1063/1.463475

    Article  Google Scholar 

  88. Patkowski K, Jeziorski B, Szalewicz K (2001) Symmetry-adapted perturbation theory with regularized Coulomb potential. J Mol Struct (Theochem) 547:293–307. https://doi.org/10.1016/S0166-1280(01)00478-X

    Article  CAS  Google Scholar 

  89. Patkowski K, Korona T, Jeziorski B (2001) Convergence behavior of the symmetry-adapted perturbation theory for states submerged in Pauli forbidden continuum. J Chem Phys 115:1137–1152. https://doi.org/10.1063/1.1379330

    Article  CAS  Google Scholar 

  90. Szalewicz K, Bukowski R, Jeziorski B (2005) On the importance of many-body forces in clusters and condensed phase. In: Dykstra CE, Frenking G, Kim KS, Scuseria GE (eds) Theory and applications of computational chemistry: the first fourty years, Elsevier, Amsterdam, chap. 33, pp 919–962. https://doi.org/10.1016/B978-044451719-7/50076-7

  91. Szalewicz K, Leforestier C, van der Avoird A (2009) Towards complete understanding of water by first-principle computational approach. Chem Phys Lett 482:1–14. https://doi.org/10.1016/j.cplett.2009.09.029

    Article  CAS  Google Scholar 

  92. Góra U, Podeszwa R, Cencek W, Szalewicz K (2011) Interaction energies of large clusters from many-body expansion: water clusters. J Chem Phys 135:224102–(1-19). https://doi.org/10.1063/1.3664730

    Article  CAS  Google Scholar 

  93. Góra U, Cencek W, Podeszwa R, van der Avoird A, Szalewicz K (2014) Predictions for water clusters from a first-principles two- and three-body force field. J Chem Phys 140:194101–(1-20). https://doi.org/10.1063/1.4875097

    Article  CAS  Google Scholar 

  94. Chałasiński G, Szczęśniak MM (1994) Origins of structure and energetics of van der Waals clusters from ab-initio calculations. Chem Rev 94:1723–1765. https://doi.org/10.1021/cr00031a001

    Article  Google Scholar 

  95. Moszyński R, Wormer PES, Jeziorski B, van der Avoird A (1995) Symmetry-adapted perturbation-theory of nonadditive 3-body interactions in van-der-Waals molecules. 1. General theory. J Chem Phys 103:8058–8074. https://doi.org/10.1063/1.470171, Erratum: 107, 672 (1997), https://doi.org/10.1063/1.47532

    Article  Google Scholar 

  96. McLachlan AD, Ball MA (1964) Time-dependent Hartree-Fock theory for molecules. Rev Mod Phys 36:844–855. https://doi.org/10.1103/RevModPhys.36.844

    Article  CAS  Google Scholar 

  97. Moszyński R, Jeziorski B, Szalewicz K (1993) Møller-Plesset expansion of the dispersion energy in the ring approximation. Int J Quantum Chem 45:409–432. https://doi.org/10.1002/qua.560450502

    Article  Google Scholar 

  98. Bohm D, Pines D (1953) A collective description of electron interactions. 3. Coulomb interactions in a degenerate electron gas. Phys Rev 92:609–625

    Article  CAS  Google Scholar 

  99. Moszyński R, Jeziorski B, Rybak S, Szalewicz K, Williams HL (1994) Many-body theory of exchange effects in intermolecular interactions – density-matrix approach and applications to He-F, He-HF, H2-HF, and Ar-H2 dimers. J Chem Phys 100:5080–5092. https://doi.org/10.1063/1.467225

    Article  Google Scholar 

  100. Williams HL, Szalewicz K, Moszyński R, Jeziorski B (1995) Dispersion energy in the coupled pair approximation with noniterative inclusion of single and triple excitations. J Chem Phys 103:4586–4599. https://doi.org/10.1063/1.470646

    Article  CAS  Google Scholar 

  101. Korona T, Moszyński R, Jeziorski B (2002) Electrostatic interactions between molecules from relaxed one-electron density matrices of the coupled cluster singles and doubles model. Mol Phys 100:1723–1734. https://doi.org/10.1080/00268970110105424

    Article  CAS  Google Scholar 

  102. Korona T, Jeziorski B (2006) One-electron properties and electrostatic interaction energies from the expectation value expression and wave function of singles and doubles coupled cluster theory. J Chem Phys 125:184109–(1-13). https://doi.org/10.1063/1.2364489

    Article  CAS  Google Scholar 

  103. Korona T, Jeziorski B (2008) Dispersion energy from density-fitted density susceptibilities of singles and doubles coupled cluster theory. J Chem Phys 128:144107–(1-10). https://doi.org/10.1063/1.2889006

    Article  CAS  Google Scholar 

  104. Korona T (2009) Exchange-dispersion energy: a formulation in terms of monomer properties and coupled cluster treatment of intramonomer correlation. J Chem Theory Comput 5:2663–2678. https://doi.org/10.1021/ct900232j

    Article  CAS  PubMed  Google Scholar 

  105. Korona T (2010) Coupled cluster treatment of intramonomer correlation effects in intermolecular interactions. In: Carsky P, Paldus J, Pittner J (eds) Recent progress in coupled cluster methods. Springer, Dordrecht, pp 267–298. https://doi.org/10.1007/978-90-481-2885-3_11.

  106. Chałasiński G, Szczęśniak MM (1988) On the connection between the supermolecular Møller-Plesset treatment of the interaction energy and the perturbation theory of intermolecular forces. Mol Phys 63:205–224. https://doi.org/10.1080/00268978800100171

    Article  Google Scholar 

  107. Patkowski K, Szalewicz K, Jeziorski B (2006) Third-order interactions in symmetry-adapted perturbation theory. J Chem Phys 125:154107. https://doi.org/10.1063/1.2358353

    Article  CAS  PubMed  Google Scholar 

  108. Patkowski K, Szalewicz K, Jeziorski B (2010) Orbital relaxation and the third-order induction energy in symmetry-adapted perturbation theory. Theor Chem Acc 127:211–221

    Article  CAS  Google Scholar 

  109. Williams HL, Chabalowski CF (2001) Using Kohn-Sham orbitals in symmetry-adapted perturbation theory to investigate intermolecular interactions. J Phys Chem A 105:646–659. https://doi.org/10.1021/jp003883p

    Article  CAS  Google Scholar 

  110. Misquitta AJ, Szalewicz K (2002) Intermolecular forces from asymptotically corrected density functional description of monomers. Chem Phys Lett 357:301–306. https://doi.org/10.1016/S0009-2614(02)00533-X

    Article  CAS  Google Scholar 

  111. Heßelmann A, Jansen G (2002) First-order intermolecular interaction energies from Kohn-Sham orbitals. Chem Phys Lett 357:464–470. https://doi.org/10.1016/S0009-2614(02)00538-9

    Article  Google Scholar 

  112. Heßelmann A, Jansen G (2002) Intermolecular induction and exchange-induction energies from coupled-perturbed Kohn-Sham density functional theory. Chem Phys Lett 362:319–325

    Article  Google Scholar 

  113. Misquitta AJ, Szalewicz K (2005) Symmetry-adapted perturbation theory calculations of intermolecular forces employing density functional description of monomers. J Chem Phys 122:214109. https://doi.org/10.1063/1.1924593

    Article  CAS  PubMed  Google Scholar 

  114. Bukowski R, Podeszwa R, Szalewicz K (2005) Efficient calculations of coupled Kohn-Sham dynamic susceptibility functions and dispersion energies with density fitting. Chem Phys Lett 414:111–116. https://doi.org/10.1016/j.cplett.2005.08.048

    Article  CAS  Google Scholar 

  115. Taylor DC, Angyan JG, Galli G, Zhang C, Gygi F, Hirao K, Song JW, Rahul K, von Lilienfeld OA, Podeszwa R, Bulik IW, Henderson TM, Scuseria GE, Toulouse J, Peverati R, Truhlar DG, Szalewicz K (2016) Blind test of density-functional-based methods on intermolecular interaction energies. J Chem Phys 145:124105. https://doi.org/10.1063/1.4961095

    Article  CAS  PubMed  Google Scholar 

  116. Tozer DJ, Handy NC (1998) Improving virtual Kohn-Sham orbitals and eigenvalues: application to excitation energies and static polarizabilities. J Chem Phys 109:10180–10189. https://doi.org/10.1063/1.477711

    Article  CAS  Google Scholar 

  117. Grüning M, Gritsenko OV, van Gisbergen SJA, Baerends EJ (2001) Shape corrections to exchange-correlation potentials by gradient-regulated seamless connection of model potentials for inner and outer region. J Chem Phys 114:652–660. https://doi.org/10.1063/1.1327260

    Article  Google Scholar 

  118. Shahbaz M, Szalewicz K (2018) Do semilocal density-functional approximations recover dispersion energies at small intermonomer separations? Phys Rev Lett 121:113402

    Article  CAS  PubMed  Google Scholar 

  119. Jaszuński M, McWeeny R (1985) Time-dependent Hartree-Fock calculations of dispersion energy. Mol Phys 55:1275–1286. https://doi.org/10.1080/00268978500102021

    Article  Google Scholar 

  120. Heßelmann A, Jansen G (2003) The helium dimer potential from a combined density functional theory and symmetry-adapted perturbation theory approach using an exact exchange-correlation potential. Phys Chem Chem Phys 5:5010–5014

    Article  Google Scholar 

  121. Boese AD, Jansen G (2019) ZMP-SAPT: DFT-SAPT Using ab initio densities. J Chem Phys 150:154101

    Article  PubMed  CAS  Google Scholar 

  122. Podeszwa R, Bukowski R, Szalewicz K (2006) Density fitting methods in symmetry-adapted perturbation theory based on Kohn-Sham description of monomers. J Chem Theory Comput 2:400–412. https://doi.org/10.1021/ct050304h

    Article  CAS  PubMed  Google Scholar 

  123. Metz MP, Piszczatowski K, Szalewicz K (2016) Automatic generation of intermolecular potential energy surfaces. J Chem Theory Comput 12:5895–5919. https://doi.org/10.1021/acs.jctc.6b00913

    Article  CAS  PubMed  Google Scholar 

  124. Schäffer R, Jansen G (2012) Intermolecular exchange-induction energies without overlap expansion. Theor Chim Acta 131:1235. https://doi.org/10.1007/s00214-012-1235-6

    Article  CAS  Google Scholar 

  125. Schäffer R, Jansen G (2013) Single-determinant-based symmetry-adapted perturbation theory without single-exchange approximation. Mol Phys 111:2570–2584

    Article  CAS  Google Scholar 

  126. Waldrop J, Patkowski K (2021) Nonapproximated third-order exchange induction energy in symmetry-adapted perturbation theory. J Chem Phys 154:024103

    Article  CAS  PubMed  Google Scholar 

  127. Korona T, Williams HL, Bukowski R, Jeziorski B, Szalewicz K (1997) Symmetry-adapted perturbation theory calculation of He–He interaction energy. J Chem Phys 106:5109–5122. https://doi.org/10.1063/1.473556

    Article  CAS  Google Scholar 

  128. Burcl R, Chałasiński G, Bukowski R, Szczęśniak MM (1995) On the role of bond functions in interaction energy calculations: Ar⋯HCl, Ar⋯H2O, (HF)2. J Chem Phys 103:1498–1507. https://doi.org/10.1063/1.469771

    Article  CAS  Google Scholar 

  129. Williams HL, Mas EM, Szalewicz K, Jeziorski B (1995) On the effectiveness of monomer-, dimer-, and bond-centered basis functions in calculations of intermolecular interaction energies. J Chem Phys 103:7374–7391. https://doi.org/10.1063/1.470309

    Article  CAS  Google Scholar 

  130. Tao FM, Pan YK (1992) Møller-Plesset perturbation investigation of the He2 potential and the role of midbond basis functions. J Chem Phys 97:4989–4995

    Article  CAS  Google Scholar 

  131. Przybytek M (2018) Dispersion energy of symmetry-adapted perturbation theory from the explicitly correlated F12 approach. J Chem Theory Comput 14:5105–5117

    Article  CAS  PubMed  Google Scholar 

  132. Kodrycka M, Holzer C, Klopper W, Patkowski K (2019) Explicitly correlated dispersion and exchange dispersion energies in symmetry-adapted perturbation theory. J Chem Theory Comput 15:5965–5986

    Article  CAS  PubMed  Google Scholar 

  133. Kodrycka Patkowski K (2021) Efficient density-fitted explicitly correlated dispersion and exchange dispersion energies. J Chem Theory Comput 17:1435–1456

    Article  CAS  PubMed  Google Scholar 

  134. Herring C, Flicker M (1964) Asymptotic exchange coupling of two hydrogen atoms. Phys Rev 134:A362–A366

    Article  Google Scholar 

  135. Silkowski M, Pachucki K (2020) Long-range asymptotics of exchange energy in the hydrogen molecule. J Chem Phys 152:174308

    Article  CAS  PubMed  Google Scholar 

  136. Morgan JD III, Simon B (1980) Behavior of molecular potential energy curves for large nuclear separations. Int J Quantum Chem 17:1143–1166

    Article  CAS  Google Scholar 

  137. Tafipolsky M (2016) Challenging dogmas: hydrogen bond revisited. J Phys Chem A 120:4550–4559. https://doi.org/10.1021/acs.jpca.6b04861

    Article  CAS  PubMed  Google Scholar 

  138. Totton TS, Misquitta AJ, Kraft M (2010) A first principles development of a general anisotropic potential for polycyclic aromatic hydrocarbons. J Chem Theory Comput 6:683–695. https://doi.org/10.1021/ct9004883

    Article  CAS  PubMed  Google Scholar 

  139. Stone AJ, Misquitta AJ (2007) Atom-atom potentials from ab initio calculations. Int Rev Phys Chem 26:193–222. https://doi.org/10.1080/01442350601081931

    Article  CAS  Google Scholar 

  140. Misquitta AJ, Stone AJ (2016) Ab initio atom-atom potentials using CamCASP: theory and application to many-body models for the pyridine dimer. J Chem Theory Comput 12:4184–4208

    Article  CAS  PubMed  Google Scholar 

  141. Metz MP, Szalewicz K (2020) Automatic generation of flexible-monomer intermolecular potential energy surfaces. J Chem Theory Comput 16:2317–2339. https://doi.org/10.1021/acs.jctc.9b01241

    Article  CAS  PubMed  Google Scholar 

  142. Stone AJ, Misquitta AJ (2009) Charge-transfer in symmetry-adapted perturbation theory. Chem Phys Lett 473:201–205. https://doi.org/10.1016/j.cplett.2009.03.073

    Article  CAS  Google Scholar 

  143. Misquitta AJ (2013) Charge transfer from regularized symmetry-adapted perturbation theory. J Chem Theory Comput 9:5313–5326. https://doi.org/10.1021/ct400704a

    Article  CAS  PubMed  Google Scholar 

  144. Dereka B, Yu Q, Lewis NHC, Carpenter WB, Bowman JM, Tokmakoff A (2021) Crossover from hydrogen to chemical bonding. Science 371:160–164

    Article  CAS  PubMed  Google Scholar 

  145. Mas EM, Szalewicz K (1996) Effects of monomer geometry and basis set saturation on depth of water dimer potential. J Chem Phys 104:7606–7614

    Article  CAS  Google Scholar 

  146. Mas EM, Szalewicz K, Bukowski R, Jeziorski B (1997) Pair potential for water from symmetry-adapted perturbation theory. J Chem Phys 107:4207–4218

    Article  CAS  Google Scholar 

  147. Jankowski P, Jeziorski B, Rybak S, Szalewicz K (1990) Perturbation analysis of the first-order exchange energy for the helium dimer. J Chem Phys 92:7441–7447

    Article  CAS  Google Scholar 

  148. Rybak S, Szalewicz K, Jeziorski B, Corongiu G (1992) Symmetry-adapted perturbation theory calculations of uracil-water interaction energy. Chem Phys Lett 199:567–573

    Article  CAS  Google Scholar 

  149. Moszyński R, Jeziorski B, Szalewicz K (1992) Many-body symmetry-adapted perturbation theory study of the He⋯F- interaction. Chem Phys Lett 166:329–339

    Google Scholar 

  150. Williams HL, Szalewicz K, Jeziorski B, Moszyński R, Rybak S (1993) Symmetry-adapted perturbation theory calculation of the Ar-H2 intermolecular potential energy surface. J Chem Phys 98:1279–1292

    Article  CAS  Google Scholar 

  151. Patkowski K, Korona T, Moszyński R, Jeziorski B, Szalewicz K (2002) Ab initio potential energy surface and second virial coefficient for He-H2O complex. J Mol Struct (Theochem) 591:231–243

    Article  CAS  Google Scholar 

  152. Mas EM, Bukowski R, Szalewicz K (2003) Ab initio three-body interactions for water. I. Potential and structure of water trimer. J Chem Phys 118:4386–4403

    Article  CAS  Google Scholar 

  153. Podeszwa R, Bukowski R, Szalewicz K (2006) Potential energy surface for the benzene dimer and perturbational analysis of ππ interactions. J Phys Chem A 110:10345–10354

    Article  CAS  PubMed  Google Scholar 

  154. van der Avoird A, Podeszwa R, Szalewicz K, Leforestier C, van Harrevelt R, Bunker PR, Schnell M, von Helden G, Meijer G (2010) Vibration-rotation-tunneling states of the benzene dimer: an ab initio study. Phys Chem Chem Phys 12:8219–8240

    Article  PubMed  CAS  Google Scholar 

  155. Groenenboom GC, Mas EM, Bukowski R, Szalewicz K, Wormer PES, van der Avoird A (2000) The pair and three-body potential of water. Phys Rev Lett 84:4072–4075

    Article  CAS  PubMed  Google Scholar 

  156. Groenenboom GC, Wormer PES, van der Avoird A, Mas EM, Bukowski R, Szalewicz K (2000) Water pair potential of near spectroscopic accuracy: II. Vibration-rotation-tunneling levels of the water dimer. J Chem Phys 113:6702–6715

    Article  CAS  Google Scholar 

  157. van der Avoird A, Szalewicz K (2008) Water trimer torsional spectrum from accurate ab initio and semi-empirical potentials. J Chem Phys 128:014302–(1-8)

    Google Scholar 

  158. Podeszwa R, Bukowski R, Rice BM, Szalewicz K (2007) Potential energy surface for cyclotrimethylene trinitramine dimer from symmetry-adapted perturbation theory. Phys Chem Chem Phys 9:5561–5569

    Article  CAS  PubMed  Google Scholar 

  159. Podeszwa R, Rice BM, Szalewicz K (2008) On predicting structure of molecular crystals from first principles. Phys Rev Lett 101:115503

    Article  PubMed  CAS  Google Scholar 

  160. Podeszwa R, Rice BM, Szalewicz K (2009) Crystal structure prediction for cyclotrimethylene trinitramine (RDX) from first principles. Phys Chem Chem Phys 11:5512–5518

    Article  CAS  PubMed  Google Scholar 

  161. Taylor D, Rice BM, Podeszwa R, Rob F, Szalewicz K (2011) A molecular dynamics study of 1,1-diamino-2,2-dinitroethylene (FOX-7) crystal using a symmetry adapted perturbation theory-based intermolecular force field. Phys Chem Chem Phys 13:16629–16636

    Article  CAS  PubMed  Google Scholar 

  162. Reilly AM, Cooper RI, Adjiman CS, Bhattacharya S, Boese AD, Brandenburg JG, Bygrave PJ, Bylsma R, Campbell JE, Car R, Case DH, Chadha R, Cole JC, Cosburn K, Cuppen HM, Curtis F, Day GM, DiStasio Jr RA, Dzyabchenko A, van Eijck BP, Elking DM, van den Ende JA, Facelli JC, Ferraro MB, Fusti-Molnar L, Gatsiou CA, Gee TS, de Gelder R, Ghiringhelli LM, Goto H, Grimme S, Guo R, Hofmann DWM, Hoja J, Hylton RK, Iuzzolino L, Jankiewicz W, de Jong DT, Kendrick J, de Klerk NJJ, Ko HY, Kuleshova LN, Li X, Lohani S, Leusen FJJ, Lund AM, Lv J, Ma Y, Marom N, Masunov AE, McCabe P, McMahon DP, Meekes H, Metz MP, Misquitta AJ, Mohamed S, Monserrat B, Needs RJ, Neumann MA, Nyman J, Obata S, Oberhofer H, Oganov AR, Orendt AM, Pagola GI, Pantelides CC, Pickard CJ, Podeszwa R, Price LS, Price SL, Pulido A, Read MG, Reuter K, Schneider E, Schober C, Shields GP, Singh P, Sugden IJ, Szalewicz K, Taylor CR, Tkatchenko A, Tuckerman ME, Vacarro F, Vasileiadis M, Vazquez-Mayagoitia A, Vogt L, Wang Y, Watson RE, de Wijs GA, Yang J, Zhu Q, Groom CR (2016) Report on the sixth blind test of organic crystal-structure prediction methods. Acta Cryst B 72:439–459

    Article  CAS  Google Scholar 

  163. Metz MP, Shahbaz M, Song H, Vogt-Maranto L, Tuckerman ME, Szalewicz K (2022) Crystal structure predictions for 4-amino-2,3,6-trinitrophenol using a tailor-made first-principles-based force field. Cryst Growth Des 22:1182–1195. https://doi.org/10.1021/acs.cgd.1c01117

    Article  CAS  Google Scholar 

  164. Shahbaz M, Szalewicz K (2019) Evaluation of methods for obtaining dispersion energies used in density-functional calculations of intermolecular interactions. Theor Chem Acc 138:25–(1-17)

    Article  CAS  Google Scholar 

  165. Ćwiok T, Jeziorski B, Kołos W, Moszyński R, Szalewicz K (1994) Symmetry-adapted perturbation theory of potential energy surfaces for weakly bound molecular complexes. J Mol Struct (Theochem) 307:135

    Article  Google Scholar 

  166. Cencek W, Szalewicz K (2013) On asymptotic behavior of density functional theory. J Chem Phys 139:024104–(1:27). Erratum: 140, 149902–(1:4) (2014)

    Article  PubMed  CAS  Google Scholar 

  167. Hapka M, Rajchel Ł, Modrzejewski M, Chałasiński G, Szczęśniak MM (2014) Tuned range-separated hybrid functionals in the symmetry-adapted perturbation theory. J Chem Phys 141:134120

    Article  PubMed  CAS  Google Scholar 

  168. Lao KU, Herbert JM (2014) Symmetry-adapted perturbation theory with Kohn-Sham orbitals using non-empirically tuned, long-range-corrected density functionals. J Chem Phys 140:044108

    Article  PubMed  CAS  Google Scholar 

  169. Hapka M, Modrzejewski M, Chałasiński G, Szczęśniak MM (2020) Assessment of SAPT(DFT) with meta-GGA, functionals. J Mol Mod 26:102

    Article  CAS  Google Scholar 

  170. Perdew JP, Burke K, Ernzerhof M (1996) Generalized gradient approximation made simple. Phys Rev Lett 77:3865–3868

    Article  CAS  PubMed  Google Scholar 

  171. Adamo C, Barone V (1999) Toward reliable density functional methods without adjustable parameters: the PBE0 model. J Chem Phys 110:6158–6170

    Article  CAS  Google Scholar 

  172. Leforestier C (2014) Water dimer equilibrium constant calculation: a quantum formulation including metastable states. J Phys Chem 140:074106

    Article  CAS  Google Scholar 

  173. Parrish RM, Gonthier JF, Corminboeuf C, Sherrill CD (2015) Communication: practical intramolecular symmetry-adapted perturbation theory via Hartree-Fock embedding. J Chem Phys 143:051103

    Article  PubMed  CAS  Google Scholar 

  174. Meitei OR, Hesselmann A (2017) Intramolecular interactions in sterically crowded hydrocarbon molecules. J Comp Chem 38:2500–2508

    Article  CAS  Google Scholar 

  175. Kita S, Noda K, Inouye H (1976) Repulsive potentials for Cl-R and Br-R (R=He, Ne, and Ar) derived from beam experiments. J Chem Phys 64:3446–3449

    Article  CAS  Google Scholar 

  176. Wheatley RJ, Price SL (1990) An overlap model for estimating the anisotropy of repulsion. Mol Phys 69:507–533

    Article  CAS  Google Scholar 

  177. Hodges MP, Wheatley RJ (2000) Application of the overlap model to calculating correlated exchange energies. Chem Phys Lett 326:263–268

    Article  CAS  Google Scholar 

  178. Jing A, Szalewicz K, van der Avoird A (2022) Ammonia dimer: extremely fluxional but still hydrogen bonded. Nature Comm 13:1470–(1:8)

    Article  CAS  Google Scholar 

Download references

Funding

This work was supported by the NSF grant CHE-1900551 to KS. BJ thanks the support from the National Science Center, Poland, Project No. 2017/27/B/ST4/02739.

Author information

Authors and Affiliations

Authors

Contributions

Both authors contributed equally.

Corresponding author

Correspondence to Krzysztof Szalewicz.

Ethics declarations

Conflict of interest

The authors declare no competing interests.

Additional information

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

This paper belongs to the Topical Collection Conversation on Non-Covalent Interactions

Appendices

Appendix: Third-order induction energy

In contrast to the second-order case, separation of the induction energy in higher orders is not straightforward. A definition of the infinite-order induction energy was proposed in ref. [53] (see also a more extensive discussion of this approach in ref. [107]). This definition leads to the formula for \(E^{(3)}_{\text {ind}}\) given by Eq. (60) in ref. [107] (or equivalently by Eq. (20) in ref. [53] where the reduced-resolvent notation is used). In both cases, \(E^{(3)}_{\text {ind}}\) is expressed via the zeroth- and first-order functions only. To get an alternative physical interpretation of this quantity, we derive here an expression for \(E^{(3)}_{\text {ind}}\) involving only the second-order induction functions. The infinite-order induction energy can be defined [53] as the minimum of the expectation value of H0 + V with the trial function of the form \(\tilde {\Psi }^{A}_{\text {ind}} \tilde {\Psi }^{B}_{\text {ind}}\)

$$J_{\text{ind}}\left[\tilde{\Psi}^{A}_{\text{ind}},\tilde{\Psi}^{B}_{\text{ind}}\right] \equiv \frac{\langle \tilde{\Psi}^{A}_{\text{ind}} \tilde{\Psi}^{B}_{\text{ind}}| H_{0} +V| \tilde{\Psi}^{A}_{\text{ind}} \tilde{\Psi}^{B}_{\text{ind}} \rangle} {\langle \tilde{\Psi}^{A}_{\text{ind}} \tilde{\Psi}^{B}_{\text{ind}}| \tilde{\Psi}^{A}_{\text{ind}} \tilde{\Psi}^{B}_{\text{ind}} \rangle} .$$
(A.1)

A simple way to derive the equations for the induction functions is to first assume that the exact induction wave function for monomer B is known. Equation (A.1) can then be written as

$$J_{\text{ind}}\left[\tilde{\Psi}^{A}_{\text{ind}},{\Psi}^{B}_{\text{ind}}\right] = \frac{\langle \tilde{\Psi}^{A}_{\text{ind}} | H_{A} + \bar{\Omega}_{B}| \tilde{\Psi}^{A}_{\text{ind}} \rangle} {\langle \tilde{\Psi}^{A}_{\text{ind}}| \tilde{\Psi}^{A}_{\text{ind}} \rangle} ,$$
(A.2)

where

$$\begin{aligned} \bar{\Omega}_{B} = \frac{\langle {\Psi}^{B}_{\text{ind}}| V {\Psi}^{B}_{\text{ind}} \rangle + \langle {\Psi}^{B}_{\text{ind}}|H_{B} |{\Psi}^{B}_{\text{ind}} \rangle} {\langle {\Psi}^{B}_{\text{ind}}|{\Psi}^{B}_{\text{ind}} \rangle} . \end{aligned}$$
(A.3)

The standard application of the variational principle leads then to the following equation for the induction wave function

$$\begin{array}{@{}rcl@{}} \left[ H_{A} + \bar{\Omega}_{B} \right] {\Psi}^{A}_{\text{ind}} = {\mathcal E} {\Psi}^{A}_{\text{ind}} , \end{array}$$
(A.4)

where \({\mathcal E} = J_{\text {ind}}[{\Psi }^{A}_{\text {ind}}, {\Psi }^{B}_{\text {ind}}]\). The analogous equation for monomer B is

$$\begin{array}{@{}rcl@{}} \left[ H_{B} + \bar{\Omega}_{A} \right] {\Psi}^{B}_{\text{ind}} = {\mathcal E} {\Psi}^{B}_{\text{ind}} . \end{array}$$
(A.5)

If all quantities in Eqs. (A.4) and (A.5) are expanded in powers of V, one gets (using a self-explanatory short-hand notation)

$$\begin{array}{@{}rcl@{}} {\mathcal E} = {\mathcal E}^{0} + {\mathcal E}^{1} + {\mathcal E}^{2} + {\mathcal E}^{3} + {\dots} = {E_{0}^{A}} + {E_{0}^{B}} + E^{(1)}_{\text{elst}} + E^{(2)}_{\text{ind}} + E^{(3)}_{\text{ind}} + {\dots} , \end{array}$$
(A.6)
$$\begin{array}{@{}rcl@{}} |A\rangle = |A^{0}\rangle + |A^{1}\rangle + |A^{2}\rangle + {\dots} , \end{array}$$
(A.7)

and similarly for monomer B, where all the terms in the sums except for the last ones have been defined before. To derive formulas for the latter quantities, we start from the following expansion of \(\bar {\Omega }_{B}\):

$$\begin{array}{@{}rcl@{}} \bar{\Omega}_{B} = {{\Omega}^{0}_{B}} + {{\Omega}^{1}_{B}} + {{\Omega}^{2}_{B}} + {{\Omega}^{3}_{B}} + {\dots} , \end{array}$$
(A.8)

where

$$\begin{array}{@{}rcl@{}} {{\Omega}^{0}_{B}} &=& {E_{0}^{B}}; \quad {{\Omega}^{1}_{B}} = \langle B^{0}|V|B^{0} \rangle = {\Omega}_{B} + \langle B^{0}|V^{A}|B^{0} \rangle + V_{0}; \end{array}$$
(A.9)
$$\begin{array}{@{}rcl@{}} {{\Omega}^{2}_{B}} &= &2\langle B^{0}|V|B^{1} \rangle - \langle B^{0}|{\Omega}_{A}|B^{1} \rangle \end{array}$$
(A.10)

and

$$\begin{aligned} {{\Omega}^{3}_{B}} &= \langle B^{1}|V|B^{1} \rangle + 2 \langle B^{0}|V|B^{2} \rangle\\& - 2 \langle B^{0}|{\Omega}_{A}|B^{2} \rangle - \langle B^{0}|V|B^{0} \rangle\langle B^{1}|B^{1} \rangle . \end{aligned}$$
(A.11)

The second-order equation defining \(|B^{2} \rangle\) is

$$\begin{aligned} (H_{B} - {E_{0}^{B}}) |B^{2}\rangle& = \left(\langle B^{0}|{\Omega}_{A}|B^{0} \rangle-{\Omega}_{A}\right) |B^{1}\rangle\\ &+ \left(\langle B^{0}|{\Omega}_{A}|B^{1} \rangle -2 \langle A^{0}|V|A^{1} \rangle \right) |B^{0}\rangle , \end{aligned}$$
(A.12)

with an analogous equation for \(|A^{1} \rangle\). Finally, the third-order equation is

$$\begin{array}{@{}rcl@{}} (H_{A} - {E_{0}^{A}}) \big|A^{3}\rangle = \sum\limits_{i=1}^{3} \left({\mathcal E}^{i} - {{\Omega}_{B}^{i}} \right) |A^{3-i}\rangle , \end{array}$$
(A.13)

resulting in the following expression for the third-order induction energy

$$\begin{aligned} {\mathcal E}^{3} &= \langle A^{0}|{\Omega}_{B}|A^{2} \rangle + 2 \langle A^{0} B^{0} | V | A^{1} B^{1} \rangle \\&+ \langle B^{1}|{\Omega}_{A}|B^{1} \rangle - \langle B^{0}|{\Omega}_{A}|B^{0} \rangle\langle B^{1}|B^{1} \rangle . \end{aligned}$$
(A.14)

Since

$$\begin{aligned}\langle B^{1}|{\Omega}_{A}|B^{1} \rangle & = \langle B^{0}|{\Omega}_{A}|B^{2} \rangle + \langle B^{0}|{\Omega}_{A}|B^{0} \rangle\langle B^{1}|B^{1} \rangle\\ & - 2 \langle A^{0} B^{0} | V | A^{1} B^{1} \rangle , \end{aligned}$$
(A.15)

we finally obtain

$$E^{(3)}_{\text{ind}} \equiv {\mathcal E}^{3} = \langle A^{0}|{\Omega}_{B}|A^{2} \rangle + \langle B^{0}|{\Omega}_{A}|B^{2} \rangle .$$
(A.16)

Returning to the explicit notation, Eq. (A.16) becomes

$$E^{(3)}_{\text{ind}} = \langle {{\Psi}_{0}^{A}} |{\Omega}_{B}| {\Psi}_{\text{ind}}^{(2)A} \rangle + \langle {{\Psi}_{0}^{B}} |{\Omega}_{A}| {\Psi}_{\text{ind}}^{(2)B} \rangle .$$
(A.17)

To find an explicit expression for \({\Psi }_{\text {ind}}^{(2)A}\), let us first write the equivalent of Eq. (A.12) for monomer A:

$$\begin{aligned} (H_{A} - {E_{0}^{A}}) |A^{2}\rangle& = \left(\langle A^{0}|{\Omega}_{B}|A^{0} \rangle-{\Omega}_{B}\right) |A^{1}\rangle\\ & + \left(\langle A^{0}|{\Omega}_{B}|A^{1} \rangle -2 \langle B^{0}|V|B^{1} \rangle \right) |A^{0}\rangle , \end{aligned}$$
(A.18)

which leads to the following formula for the second-order induction function of monomer A expressed as a spectral sum

$$\begin{aligned} {\Psi}_{\text{ind}}^{(2)A}& = \sum\limits_{k\ne 0} \frac{\langle {{\Psi}_{k}^{A}} |\langle {{\Psi}_{0}^{A}} | {\Omega}_{B} {{\Psi}_{0}^{A}} \rangle -{\Omega}_{B}| {\Psi}_{\text{ind}}^{(1)A} \rangle} {{E_{k}^{A}} - {E_{0}^{A}}} {{\Psi}_{k}^{A}}\\ &- 2 \sum\limits_{k\ne 0} \frac{ \langle {{\Psi}_{k}^{A}} {{\Psi}_{0}^{B}} | V_{\text{ee}} | {{\Psi}_{0}^{A}} {\Psi}_{\text{ind}}^{(1)B} \rangle } {{E_{k}^{A}} - {E_{0}^{A}}} {{\Psi}_{k}^{A}} . \end{aligned}$$
(A.19)

In practice, \(E^{(3)}_{\text {ind}}\) is much easier to compute using the expression depending only on the zeroth- and first-order functions, see Eq. (A.22) below, but the use of the second-order function allows a transparent interpretation of the third-order induction energy. Of course, the importance of this function stems also from the fact that it can be used to compute the fourth- and fifth-order induction energy contributions.

Expression (A.17) can be written in terms of densities as

$$\begin{array}{@{}rcl@{}} E^{(3)}_{\text{ind}} &= \frac{1}{2}\int \rho^{(02)A}_{\text{ind}}({\boldsymbol r}) \omega^{B}({\boldsymbol r}) d^{3}\boldsymbol{r} + \frac{1}{2}\int \rho^{(02)B}_{\text{ind}}({\boldsymbol r}) \omega^{A}({\boldsymbol r}) d^{3}\boldsymbol{r} \end{array}$$
(A.20)

where \(\rho ^{(02)A}_{\text {ind}}({\boldsymbol r})\), a component of the total second-order induction density \(\rho ^{(2)A}_{\text {ind}}({\boldsymbol r})\), is defined by a procedure similar to that leading to Eq. (21), but with \({\Psi }_{\text {ind}}^{(2)A}\) added to ΨA

$$\begin{aligned}\rho^{(2)A}_{\text{ind}}({\boldsymbol r}) &= 2 N_{A} \sum\limits_{s,s_{2},\dots,s_{N}}\int {{\Psi}_{0}^{A}}(\boldsymbol{x},\boldsymbol{x}_{2}, \dots, \boldsymbol{x}_{N_{A}}) {\Psi}_{\text{ind}}^{(2)A}(\boldsymbol{x},\boldsymbol{x}_{2}, \dots, \boldsymbol{x}_{N_{A}}) d^{3}\boldsymbol{r}_{2} {\dots} d^{3}\boldsymbol{r}_{N_{A}} \\ &\quad +N_{A} \sum\limits_{s,s_{2},\dots,s_{N}}\int {\Psi}_{\text{ind}}^{(1)A}(\boldsymbol{x},\boldsymbol{x}_{2}, \dots, \boldsymbol{x}_{N_{A}}) {\Psi}_{\text{ind}}^{(1)A}(\boldsymbol{x},\boldsymbol{x}_{2}, \dots, \boldsymbol{x}_{N_{A}}) d^{3}\boldsymbol{r}_{2} {\dots} d^{3}\boldsymbol{r}_{N_{A}} \\ & = \rho^{(02)A}_{\text{ind}}({\boldsymbol r}) + \rho^{(11)A}_{\text{ind}}({\boldsymbol r}), \end{aligned}$$
(A.21)

with the second-order induction function given by Eq. (A.19) and the two terms in the last expression corresponding to the consecutive terms in the preceding expression. Thus, the second-order induction densities have two components: one originating from the second-order induction function and one originating from the product of first-order induction functions, and only the former term enters the expression for \(E^{(3)}_{\text {ind}}\). Equation (A.20) shows that the third-order induction interaction can be interpreted as the interaction of the second-order induction density components resulting from \({\Psi }_{\text {ind}}^{(2)X}\) with the unperturbed electrostatic potentials. Note again the factors of 1/2 multiplying the Coulomb interactions of \(\rho ^{(02)A}_{\text {ind}}({\boldsymbol r})\) with ωB(r) in the first term, and analogously in the second.

Using Eq. (A.15), formula (A.16), expressed in terms of the unperturbed functions and of the second-order induction functions, can be transformed into formula (60) from ref. [107] expressed in terms of the unperturbed functions and of the first-order induction functions only:

$$\begin{aligned} E^{(3)}_{\text{ind}} &= \langle A^{1}|{\Omega}_{B}|A^{1} \rangle + \langle B^{1}|{\Omega}_{A}|B^{1} \rangle - \langle A^{0}|{\Omega}_{B}|A^{0} \rangle\langle A^{1}|A^{1} \rangle\\ &- \langle B^{0}|{\Omega}_{A}|B^{0} \rangle\langle B^{1}|B^{1} \rangle + 4 \langle A^{0} B^{0} | V | A^{1} B^{1} \rangle , \end{aligned}$$
(A.22)

which written in explicit notation is the same as Eq. (60) in ref. [107]. This formula can be expressed in terms of densities as

$$\begin{aligned} E^{(3)}_{\text{ind}}& = \int \rho^{(11)A}_{\text{ind}}({\boldsymbol r}) \tilde{\omega}^{B}({\boldsymbol r}) d^{3}\boldsymbol{r} + \int \rho^{(11)B}_{\text{ind}}({\boldsymbol r}) \tilde{\omega}^{A}({\boldsymbol r}) d^{3}\boldsymbol{r} \\ & +\iint \rho^{(1)A}_{\text{ind}}({\boldsymbol r}_{1}) \frac{1}{r_{12}} \rho^{(1)B}_{\text{ind}}({\boldsymbol r}_{2}) d^{3}\boldsymbol{r}_{1}d^{3}\boldsymbol{r}_{2} ,\end{aligned}$$
(A.23)

where

$$\tilde{\omega}^{B}({\boldsymbol r}) = \omega^{B}({\boldsymbol r}) - \frac{1}{N_{A}} \int {\rho^{A}_{0}}({\boldsymbol r}) \omega^{B}({\boldsymbol r}) d^{3}\boldsymbol{r} .$$

The third term is the same as in Eq. (25). The first term is numerically the same as the first term in Eq. (25), but the density is now the component of the second-order induction density resulting from the product of the first-order induction functions and it interacts with the shifted electric potential of molecule B, similarly for the second term.

One more interpretation of \(E^{(3)}_{\text {ind}}\) can be obtained via regrouping the terms in formula (A.17) and defining an alternative second-order induction function, different from the one of Eq. (A.19). The resulting expression for the third-order induction energy is algorithmically different from Eq. (A.17), but gives the same numerical value of this quantity. The former definition just neglects the second expression on the right-hand side of Eq. (A.19)

$$\begin{aligned}\bar{\Psi}_{\text{ind}}^{(2)A} = \sum\limits_{k\ne 0} \frac{\langle {{\Psi}_{k}^{A}} |\langle {{\Psi}_{0}^{A}} | {\Omega}^{B} {{\Psi}_{0}^{A}} \rangle -{\Omega}^{B}| {\Psi}_{\text{ind}}^{(1)A} \rangle} {{E_{k}^{A}} - {E_{0}^{A}}} {{\Psi}_{k}^{A}} . \end{aligned}$$
(A.24)

The alternative second-order equation is

$$\begin{array}{@{}rcl@{}} (H_{A} - {E_{0}^{A}}) |\bar{A}^{2}\rangle = \left(\langle A^{0}|{\Omega}_{B}|A^{0} \rangle-{\Omega}_{B}\right) |A^{1}\rangle . \end{array}$$
(A.25)

Notice that \(|\bar {A}^{2}\rangle\) cannot be called the second-order induction function since the function of Eq. (A.18) is the unique and only such function. We can now design an energy expression

$$E^{(3)}_{\text{ind}} = \langle A^{0}|{\Omega}_{B}|\bar{A}^{2} \rangle + \langle B^{0}|{\Omega}_{A}|\bar{B}^{2} \rangle + 4 \langle A^{0} B^{0} | V | A^{1} B^{1} \rangle$$
(A.26)

or in explicit notation

$$E^{(3)}_{\text{ind}} = \langle {{\Psi}_{0}^{A}} |{\Omega}_{B}| \bar{\Psi}_{\text{ind}}^{(2)A} \rangle + \langle {{\Psi}_{0}^{B}} |{\Omega}_{A}| \bar{\Psi}_{\text{ind}}^{(2)B} \rangle + 4\langle {{\Psi}_{0}^{A}} {{\Psi}_{0}^{B}} |V| {\Psi}_{\text{ind}}^{(1)A} {\Psi}_{\text{ind}}^{(1)B} \rangle ,$$
(A.27)

which produces the same numerical values as given by Eq. (A.17). To show this, transform Eq. (A.26) into Eq. (A.22). To this end, write Eq. (A.25) as

$$|\bar{A}^{2}\rangle = {R_{0}^{A}} \left(\langle A^{0}|{\Omega}_{B}|A^{0} \rangle-{\Omega}_{B}\right) |A^{1}\rangle ,$$
(A.28)

where \({R_{0}^{A}}\) is the reduced resolvent of monomer A, and use it in Eq. (A.26). We get

$$\begin{aligned}\langle A^{0}|{\Omega}_{B}|\bar{A}^{2} \rangle &= \langle A^{0}|{\Omega}_{B} {R_{0}^{A}} \left(\langle A^{0}|{\Omega}_{B}|A^{0} \rangle-{\Omega}_{B}\right) |A^{1}\rangle \\&= \langle A^{1}|{\Omega}_{B}|A^{1} \rangle - \langle A^{0}|{\Omega}_{B}|A^{0} \rangle\langle A^{1}|A^{1} \rangle \end{aligned}$$
(A.29)

and similarly for the second term, which indeed gives the desired formula. Equation (A.27) expressed in terms of densities yields Eq. (25).

Appendix: Conversations

Mo et al. commented:

The authors focus on the elegance of the SRS formulation of SAPT, but perhaps ignore chemistry, and at the same time they are dismissing all other EDA methods as nonunique. They make statements about the uniqueness and precision of SAPT in providing energy partition that sums to the total interaction energy between the molecules/fragments A and B, and dismisses effects like charge transfer, covalency, and charge delocalization which arise from other EDA methods (and also from NBO, VB, and BLW). Our general comment is that there are multiple perspectives in a matter of fact, which are in no way unphysical. What is unscientific is to claim uniqueness and truth for one of these choices, namely SAPT, and to dismiss on this ground all other approaches. This is done without providing the reader with a single example that compares SAPT (e.g., what about BrHBr?) to other EDA methods. In a nutshell the paper is a blizzard of equations without any example. This is a major problem for most chemists, who would like to see examples with numerical data, as proofs of correctness of statements.

Reply:

We do not believe we ignore chemistry, but it depends what one has in mind by “chemistry”. Noncovalent intermolecular interactions are a part of chemistry and our whole paper is devoted to such interactions. Thus, in this sense we cannot agree that we ignore chemistry. On the other hand, we do not discuss making and breaking of chemical bonds since these are not processes that SAPT was designed for (although please see a discussion of this issue later on). In fact, the first sentence in the abstract states “Symmetry-adapted perturbation theory (SAPT) is a method for computational studies of noncovalent interactions between molecules.”

We also do not believe we are dismissing all other EDA methods as nonunique. While this is a plain fact that the EDA methods are highly nonunique, the physical components in some of these methods come reasonably close to the corresponding SAPT components. Such methods are, in our opinion, important since their application on top of some supermolecular calculations of interaction energies does give sufficiently precise physical insights.

Concerning the dismissal of “effects like charge-transfer, covalency, charge delocalization,” indeed, SAPT is not designed to investigate covalency effects as it is practically limited to noncovalent interactions. We make no statements on EDA methods applied to chemically reactive systems. One the other hand, we do not dismiss charge-delocalization effects (which in our terminology are equivalent to charge-transfer effects) and we discuss these effects in “SAPT contributions and EDA methods.” To summarize this discussion: SAPT does include charge-delocalization effects, but it appears there is no unique ways to separate them. However, very reasonable ways to perform such separation approximately have been designed by Misquitta and Stone [140, 142, 143] and analyzed from the point of view of applications in developments of force fields [77].

While we agree that different viewpoints are useful in science, it does not mean that all viewpoints are correct. In particular, if method X states that the dispersion interaction in a given dimer is zero, while method Y states that it is one of major attractive forces, only one of these viewpoints can be correct. Thus, if the authors of method X believe their results is correct, they should explain why method Y makes wrong predictions.

There are so many examples of SAPT interaction energy decompositions in literature that we did not think examples are needed in the present paper. Nevertheless, in the revised version we added Table 1 with such examples. The geometries of the dimers included are available in literature, so these data can perhaps serve as a useful reference point for authors of EDA approaches. We have not included the BrHBr complex in Table 1. While one can trivially compute SAPT components for this system, it is a system which is to a large effect covalently bound, see the recent work on FHF dimer [144]. While both systems would be interesting cases to study by SAPT, their special character does not make them appropriate as examples of SAPT analysis of noncovalent interactions. One may add that a somewhat similar system, H2O-F, is included in ref. [77].

Most published SAPT calculations listed the components of interaction energies, at least for some selected geometries. Here is a selection of such papers for readers who would like to study more examples: water dimer (refs. [20, 22, 145, 146]), helium dimer (refs. [21, 147]), water-uracil (ref. [148]), He-F (ref. [149]), Ar–H2 (ref. [150]), Ar–HF (ref. [84]), He-H2O (ref. [151]), water trimer (ref. [152]), and many other.

Mo et al. commented:

On pages 13–14 the authors state that “one cannot define uniquely the charge transferred from one monomer.” But the fact is that charge transfer accompanies nevertheless many reactions. How does SAPT handles the CT in e.g., SN2 reaction? As the authors say, CT is handled as damping of other terms of SAPT. It is hard to buy such a statement.

Reply:

We never say that there is no charge delocalization (or transfer) in noncovalent intermolecular interactions, we only say that the amount of charge that was delocalized cannot be uniquely determined. These are two different statements. Again, we make no statements about chemical reactions.

We never say that charge-delocalization energy is due to damping of asymptotic expansion. We just point out that in the method of determining this energy developed by Misquitta and Stone in ref. [140], the numerical value of this quantity depends on assumptions concerning damping.

Mo et al. commented:

On page 14, the authors speak about CT and ask: “what would be the use of such information besides a physical insight?” Let us ask the authors: is there any science without insight?

Reply:

We were not trying to belittle the importance of physical insight, but the major goal of science are predictions about nature. We added an additional sentence at this place to make our views clear.

Mo et al. commented:

On page 14, the authors do recognize charge-transfer-delocalization “and one may assume that this is due to the difficulties of such a basis set to model charge delocalization”. It is good that Stone found a way to add CT to SAPT. There is not much chemistry without CT.

Reply:

We recognize the existence of charge delocalization all the time. In the sentence quoted, we discuss the assumptions of the method proposed by Stone and Misquitta [142] in 2009 and this sentence is not a statement of recognition of charge delocalization. The work of these authors shows how difficult it is to define charge-delocalization energies within SAPT as the three proposed approaches [140, 142, 143] produce quite different numerical values of these quantities. This is one more confirmation of the point we make: charge delocalization is a fact, a unique determination of its energetic effect is not possible, but reasonable approximate definitions of this effect may be useful.

Mo et al. commented:

In fact, MO (or MO-CI – any level) theory tells us that when two molecules approach, there are orbital interactions dominated by one HOMO in one side and one LUMO on the other side. These orbital interactions stabilize the complex by CT interaction that creates covalency. Where is this in SAPT?

Reply:

Again, SAPT is not designed to treat covalent bonds.

Mo et al. commented:

In addition, when two molecules approach one another, their individual molecular orbitals will be perturbed and thus reshuffled. This physical effect is the polarization effect. Thus, the obvious question would be how the SAPT method quantifies the electron (or charge) transfer and polarization energies. The author did address this question in passing in the end of the paper by mentioning the approaches by Stone and Misquitta.

Reply:

The polarization effect, which we call the induction effect, is discussed in detail in “Induction interaction”. We can only repeat one more time that a unique separation of induction energies into parts due to “fixed” (but deformed) and delocalized charge is not possible. Nonunique, approximate separations such as those proposed by Misquitta and Stone [140, 142, 143] can be useful.

Mo et al. commented:

In chemistry, not only inter- but also intramolecule charge transfer plays a significant role and has been well recognized. For example, in benzene (C6H6), the delocalization of the six p electrons has a profound influence on molecular structures and properties. It will be helpful for readers if some data from the SAPT computations in this regard can be presented.

Reply:

In SAPT, monomers can be described at various levels of theory, see “Levels of intramonomer electron correlation in SAPT”. Even the lowest possible level, the HF method, should take account of these effects. This is demonstrated by the fact that the SAPT potential for the benzene dimer [153] is the most accurate published one for this system, capable to produce predictions of spectroscopic accuracy [154].

Mo et al. commented:

We are also curious about the basis set dependency of the SAPT method. Taking the simple example of a two-body complex H3N⋯BH3, can the author show the SAPT results with the basis sets from 6-31G to 6-311+G(d) to 6-311++G(d,p) to aug-cc-pvtz for this very simple complex? Of course, the comparison of different correlated methods would also be helpful. We believe that the case of H3N⋯BH3 with various basis sets would be illuminating if the author is willing to share the computation results.

Reply:

Multiple basis set convergence studies for SAPT components including the basis sets listed above and several other basis sets have been published, see in particular refs. [22] and [129]. We believe no further studies of basis set convergence are needed.

Mo et al. commented:

Getting back to the “Introduction,” the author wrote that “there is no place for terms not present in SAPT since SAPT’s contributions sum up to an accurate value of the interaction energy”. This is quite confusing as all other EDA approaches also sum up all terms to an accurate value of the interaction energy. We do not see disagreements with the interaction energy values, and all controversies come from the interpretation of the energy terms. The accuracy of the SAPT towards the final interaction energies cannot be used as evidence for the correctness of its physical interpretations or energy partition schemes. Again, data can speak better.

Reply:

This statement means the following: virtually all EDA methods identify terms labeled in the same way as in SAPT: electrostatic, first-order exchange, induction, and dispersion energies. In SAPT, these terms sum up to an accurate total interaction energies. Thus, if these four components are defined in an EDA in such a way that they are close to the SAPT values, they also sum up to total interaction energies. Then, no other terms with significant magnitude can be added. We have modified the quoted text to make our point more clear.

Mo et al. commented:

Besides, any theoretical results need be justified by experimental evidence, directly or indirectly. On page 5, the author wrote “In fact, there is no resemblance between SAPT(DFT) and supermolecular DFT interaction energies for majority of dimers”. This is again quite confusing. DFT interaction energies rely on the DFT methods themselves not any particular EDA method. It seems that the author is comparing orange with apple here as SAPT(DFT) and DFT are not at the same theoretical levels.

Reply:

SAPT results are fully confirmed by experimental evidence. The most convincing confirmation are vibration-rotation-tunneling (VRT) spectra of dimers and trimers: spectra computed from SAPT potentials agree very well with experiment [84, 154,155,156,157]. Another example are crystal-structure predictions from SAPT-based force fields, which correctly identify the experimental crystal as one of the top-ranked polymorphs [158,159,160,161,162,163].

A different question is if the individual energy components predicted by SAPT can be related to experimental data. The evidence is less direct here, but it does exist. In the long-range region, the interaction energy of a dimer made from polar monomers is dominated by the electrostatic energy. For such systems, scattering experiments can sometimes identify the so-called long-range entrance-channel states which are located in the regions of strongest electrostatic interactions. Furthermore, the electrostatic energies are guiding crystallographers in designing crystals of polar molecules. The induction energy dominates the long-range total potential in interactions of ions with rare-gas atoms. Therefore scattering experiments on such systems directly probe the induction components of PESs. Similarly, interactions of rare-gas atoms are dominated at long range by the dispersion energies. Thus, the measured s-wave scattering lengths probe dispersion interactions. Finally, the exchange component is related to van der Waals radii of elements. This component determines the repulsive wall of potentials and the repulsive wall in turn determines the van der Waals radii.

The second question is orthogonal to the first. Yes, SAPT(DFT) and supermolecular DFT are different levels of theory. Still, different approaches can produce similarly accurate interaction energies. For example, interaction energies from high-order SAPT based on the FCI description of monomers agree to several digits with supermolecular FCI energies [16, 18, 66, 127]. On the other hand, while SAPT(DFT) gives accurate interaction energies, those from supermolecular DFT, as it is well known, have in general dramatically large errors, mainly due to the fact that semilocal DFT methods do not reproduce dispersion energies at physically relevant intermolecular separations [118, 164]. The underlying reason for this problem is the “shortsightedness” of interelectron interactions in semilocal DFT approximations. However, DFT describes monomers reasonably well. This is why Kohn-Sham monomer determinants and TD-DFT monomer density-density response functions can be used to construct SAPT(DFT) components. The dispersion energy is obtained in SAPT(DFT) from wave-function-type expressions and therefore the DFT shortsightedness problem does not matter. References [26, 113] provide theoretical justifications for high accuracy of SAPT(DFT). It is easy to understand this in the case of electrostatic energy. Most variants of semilocal generalized-gradient approximation (GGA) DFTs give quite accurate electron densities, except at large separations from nuclei. Because of the latter problem, the first SAPT(KS) calculations [109] gave poor electrostatic energies. This problem can be fixed by applying the asymptotic correction as done in refs. [110, 111], leading to densities accurate at all separations. Since the electrostatic energy is just an integral of electron densities of unperturbed monomers, if the densities are accurate, so is the electrostatic energy.

Mo et al. commented:

On Page 2 the authors state that SAPT in higher order is accurate even for diatomics, e.g., LiH. Can he show an example or two? There is no chemistry without delocalization, there is no chemistry without covalency. What about H2? Where is the covalency in SAPT? What about resonance?

Reply:

The question presumably concerns chemically bound diatomics (for NCIs in diatomics, SAPT is accurate already in the low order). For two diatomics, H2 and LiH, SAPT was applied to chemically bonded ground states of these systems [15, 16, 18, 54, 87, 165]. Table I in ref. [15] shows that in 60th order SRS recovers the energy of the singlet state of H2 at R = 2.0 bohr to within 0.0005%. Table III in ref. [18] shows that the best working variant of SAPT recovers the LiH binding energy at the chemical minimum to within 0.0001%. Figure 8 in the same paper shows potential energy curves in the region of chemical minimum computed using 4th-order SAPT. So clearly, SAPT can recover covalent interactions. Yet, because of the necessity to apply a high-order treatment, we do not recommend SAPT to study strong covalent interactions. However, this excellent convergence says nothing about presence or absence of charge delocalization. As already stated several times, SAPT does account for delocalization effects, but cannot separate them from polarization effects. Finally, resonances do appear in intermolecular interactions when the interacting systems are degenerate. It is possible to construct SAPT applicable to such systems [11].

Mo et al. commented:

Some technical questions:

Are the orbitals in A and B orthogonal? Presumably they are not since Pauli repulsion is accounted for during the anti-symmetrization procedure. But Pauli repulsion necessarily bring about electronic effects like CT. Where are these in the SAPT picture? Any example?

Reply:

Orbitals within system A are orthogonal to each other, and similarly for systems B. However, orbitals of A are not orthogonal to orbitals of B. This is independent of antisymmetrization as such non-orthogonality exists already at the level of RSPT. The non-orthogonality is treated in SAPT exactly, i.e., proper orbital overlap integrals appear in all formulas. Indeed, antisymmetrization leads to a distortion of charge density, so the exchange components of SAPT do contain some delocalization, it is not only the induction energy which contributes to charge delocalization. Since delocalization effects are not separable from charge distortion effects that do not involve any shift of charge, no examples can be given.

Mo et al. commented:

Doesn’t SAPT miss one electronic effect like CT because its perturbation Hamiltonian includes only bielectronic Coulombic terms [\(H_{0} = H_{A}+H_{B}+ H_{C} + V_{AB}+V_{AC}+V_{BC} + \dots\)]?

Reply:

The terms on the right-hand side define the exact Hamiltonian of Schrödinger’s quantum mechanics for atoms and molecules. No three-electrons interactions are present in such Hamiltonians. [BTW, since this is the total Hamiltonian of a cluster, it should be denoted as H rather than H0 since in the customary notation H0 = HA + HB + HC.]

Brink and Borrfors commented:

It is very reassuring that SAPT to high orders is formally an exact theory and can handle covalent bond formation. However, the most common SAPT variants are truncated at the second order or possibly the third order and better suited for weaker interactions. In addition, there are other approximations that commonly are employed, such as \(\delta E_{\text {int, resp}}^{\text {HF}}\) (eq. 7) and the S2 approximation. Is there any approach for estimating the accuracy of the employed SAPT level for a given problem? Can the \(\delta E_{\text {int, resp}}^{\text {HF}}\) value be used as such an indicator? For example, does a value lower than a certain number or lower than a certain fraction of the total SAPT energy indicate that the SAPT level is sufficiently accurate for the problem at hand?

Reply:

As discussed above, SAPT is not designed for interactions leading to formation of covalent bonds. For NCIs, second-order treatment works well in practice. In fact, in most SAPT calculations basis set incompleteness errors are larger than SAPT theory-level errors. In a number of papers, SAPT interaction energies were compared to CCSD(T) energies at complete basis set (CBS) limits. Perhaps the most thorough comparison was performed in ref. [115] on 10 dimers and about 100 configurations total. The median unsigned percentage error computed for all dimers in an augmented triple-zeta basis relative to CCSD(T)/CBS was 2.6%. This should be compared to the same error for CCSD(T) in the same basis set amounting to 1.2%.

Estimates of SAPT accuracy by comparisons to other accurate methods are the only reliable ones. The magnitude of \(\delta E_{\text {int, resp}}^{\text {HF}}\) is not an indication of the overall error of SAPT. In fact, SAPT performs very well for interactions of strongly polar systems, while \(\delta E_{\text {int, resp}}^{\text {HF}}\) is always large for such systems. Thus, other than the average errors such as those found in ref. [115], there are no a priori estimates of the size of SAPT error for a given dimer. In practice, one usually performs CCSD(T) calculation for a couple of points on a potential energy surface to estimates the uncertainties of SAPT, as well as performs a few calculations at the CBS limits to estimate basis set incompleteness errors.

It appears that the errors due to the use of the S2 approximation and due to the addition of \(\delta E_{\text {int, resp}}^{\text {HF}}\) are smaller than the errors due to the truncation of SAPT expansion at the second order, although there is no study showing this unequivocally. The S2 approximation can now be eliminated [124,125,126]; however, it affects the results significantly only at very short separations and the S2 errors are mostly removed by the use of \(\delta E_{\text {int, resp}}^{\text {HF}}\). The physical reason for including \(\delta E_{\text {int, resp}}^{\text {HF}}\) is to account for the induction and exchange-induction effects of the third (or fourth) and higher orders. For polar systems, the advantages of adding the third and higher-order induction effects much outweigh the small inaccuracies [94, 106] that this addition introduces in the first and second order.

Brink and Borrfors commented:

What are the main reasons for the higher accuracy of SAPT(DFT) compared to supermolecular DFT? Is the difference in interaction energy dominated by the more accurate description of the dispersion energy in the former approach or is SAPT(DFT) able to describe other energy contributions more accurately, as well? Is it still possible to determine third-order and higher terms by a similar equation to eq. 7 (\(\delta E_{\text {int, resp}}^{\text {HF}}\) approximation) in SAPT(DFT)? A related question concerns the functional dependence of SAPT(DFT). Is SAPT(DFT) less dependent than supermolecular DFT on the choice of DFT functional? Furthermore, which energy term in SAPT is most functional dependent?

Reply:

The reasons that SAPT(DFT) interaction energies are more accurate than the supermolecular DFT ones have already been discussed in the reply to one of the Mo et al. questions. Indeed, the inability of semilocal GGA approaches to recover dispersion interactions is one of the reasons. However, it is not the only reason. An extensive discussion [118] of the other reasons based on analysis of numerical results for several dimers led to the conclusion that inaccuracies originating from DFT components unrelated to the dispersion energy are of similar magnitude. This work analyzed the quantity

$$E_{\text{int}}^{\text{extra}} = E_{\text{int}}^{\text{DFT}} - E_{\text{int}}^{\text{dispersionless}},$$

where \(E_{\text {int}}^{\text {DFT}}\) is the supermolecular DFT interaction energy and \(E_{\text {int}}^{\text {dispersionless}}\) is a near-exact interaction energy minus the dispersion and exchange-dispersion contribution. If the dispersion energy was the only problem of DFT, it should recover \(E_{\text {int}}^{\text {dispersionless}}\) well, i.e., \(E_{\text {int}}^{\text {extra}}\) should be small (except possibly at very small intermonomer separations where the electrons of the interacting monomers get into the “visibility” region of DFT). Figures 4 and 6 in ref. [118] show that this is not the case, in fact, the recovery of \(E_{\text {int}}^{\text {dispersionless}}\) is poor. So the answer is confirmative: the accurate description of the dispersion interaction in SAPT(DFT) compared to essentially no description in supermolecular DFT is one reason for SAPT(DFT) being so much more accurate, but SAPT(DFT)’s ability to describe the other interaction energy contributions more accurately than does supermolecular DFT is another, perhaps equally important reason.

Yes, the addition of \(\delta E_{\text {int, resp}}^{\text {HF}}\) is as rigorous in SAPT(DFT) as in SAPT based on wave-function description of monomers. The reason is that the orders in V in each version of SAPT are rigorously separated from each other. In particular, SAPT(DFT) in its current version includes only first- and second-order terms, while \(\delta E_{\text {int, resp}}^{\text {HF}}\) includes only the third- and higher-order terms (plus a small “contamination” in lower orders [94, 106] which it the reason the addition of \(\delta E_{\text {int, resp}}^{\text {HF}}\) is an approximation, as discussed above).

Yes, SAPT(DFT) interaction energies change insignificantly when different variants of GGAs are used (provided an asymptotic correction is applied) compared to dramatically different interaction energies from different variants of supermolecular DFT. This issue was investigated in a number of papers [26, 113, 166,167,168,169]. Interestingly, the PBE0 functional [170, 171] shows consistently the best performance in SAPT(DFT) calculations.

The SAPT component most dependent on the choice of the density functional depends on type of interactions. For dispersion-bonded systems like rare-gas dimers, the effect is the largest in absolute terms for the dispersion energies, see for example Table IV in ref. [26]. For dimers of polar monomers, the largest effects come from the first-order and induction energies, see Table V in ref. [26] and Table IV in ref. [113].

Brink and Borrfors commented:

A limitation of SAPT for the analysis of larger systems seems to be the lack of an efficient procedure for structure optimization of molecular complexes. In particular, strong interactions often lead to conformational changes and changes to intramolecular geometry parameters, e.g., intramolecular bond lengths. How are the structure optimizations of such systems best handled? When employing SAPT(DFT) it does not seem advisable to use a supermolecular approach for structure optimization as supermolecular DFT is much less accurate than SAPT(DFT) for intermolecular interactions. Would it be possible to use a mixed approach where supermolecular calculations are used to determine binding conformations and intramolecular parameters and where SAPT is used for refining intermolecular distances? Would such a procedure be sufficiently accurate and can it be automated? A related question concerns the best approach for computing vibrational corrections (zero point and thermal corrections) to complexation enthalpies and free energies?

Reply:

Actually, our programs provide one of the most efficient approaches to structure optimization for molecular complexes. Let us focus first on rigid monomers. In this case, one needs monomers’ structures. To get them, one can use standard electronic structure programs. For not too large monomers (containing up to couple dozen atoms), one can usually determine a small number of starting monomer’s configurations based on chemical intuition. Then, local optimization algorithms (i.e., algorithms finding the minimum closest from the starting point) will reliably find global and local minima (conformers) of each monomer. If one wanted to proceed in this way to find minimum structures of the dimer, this approach would frequently fail since the locations of minima on the potential energy surface of the dimer are often in very nonintuitive places. Thus, many starting points would have to be tried, which makes such optimizations very expensive even at the DFT level. Our approach is to first fit a potential energy surface and then use the fit function to search for minima. Both tasks are performed completely automatically by the autoPES programs [123, 141]. This protocol is the mixed approach mentioned in the question (and yes, the procedure is accurate, robust, and fully automated). For optimization of monomer geometries any method can be used, e.g., MP2, not necessarily DFT (and if DFT is used, it should be a dispersion-corrected DFT approach). Since the latest version of autoPES can develop flexible-monomer potential energy surfaces, one can now optimize full-dimensional dimer structures. This allows investigations of effects of intermolecular interactions on monomer conformations.

As mentioned earlier, SAPT potentials for smaller clusters have been often used to compute VRT spectra of these clusters. Such calculations produce very accurate zero-point energies and give energy levels allowing computations of thermodynamic quantities (see for example ref. [172]). For larger clusters, the potential energy surface can be used to compute the Hessian and proceed in the standard way to obtain thermodynamic quantities in the harmonic approximation.

Popelier commented:

Question 1: Let us take a single molecule, such as Br(CH2)10Br, and curl back its chain so that the two Br atoms end up in close contact but without being bonded. Surely there is a dispersion-like interaction between the two Br atoms but can SAPT, as presented in this article, calculate its energy value? SAPT’s basic assumption is the partitioning of the total Hamiltonian into a sum of Hamiltonians of separated monomers. However, Br(CH2)10Br is a single molecule and cannot be separated into monomers. Is there a conceptual challenge in partitioning the Hamiltonian for this covalently bound system, in the absence of monomers? Is a way forward the “atomic SAPT partition” or A-SAPT (J. Chem. Phys. 2014, 141, 044115)? However, A-SAPT struggles to produce chemically useful partitions of the electrostatic energy, caused by the buildup of oscillating partial charges on adjacent functional groups. This is why “functional-group SAPT” or F-SAPT (J. Chem. Theor. Comp. 2014, 10, 4417) was proposed. But then F-SAPT is formulated entirely in terms of fragments with integral charge (including zero), which may suit this molecule but which is not realistic in general.

Reply:

Yes, SAPT is a theory starting with the assumption that the system separates into a set of monomers at infinite distances from each other and these monomers are well-defined molecules (not necessarily closed-shell). As discussed above, this separation should not involve any breaking of chemical bonds. Thus, standard SAPT cannot be applied to Br(CH2)10Br.

A-SAPT and F-SAPT have been proposed with the goal to approximately assign interaction energy contributions to atoms or groups of atoms in the standard SAPT approach involving dimers of two closed-shell monomers. An extension of SAPT to intramonomer NCIs, named ISAPT, was proposed later [173]. Another approach of this type was developed in ref. [174]. One should emphasize that all intramolecular SAPT applications require one to make several assumptions and that different but equally reasonable assumptions can lead to very different predictions for a given system.

Popelier commented:

Question 2: It is stated that SAPT defines energy contributions each of which results from a differential equation that has an exact solution. Please give examples of such differential equations as they do not seem to appear in the standard SAPT literature.

Reply:

The differential equations for the wave function corrections are the foundations of RSPT. Such equations appear in many papers developing SAPT, see for example Eqs. (5), (10), and (27) in ref. [21] and Eqs. (18), (29), (A.4), and (A.12) in the present work.

Popelier commented:

Question 3: Does the author agree with the opinion of Konrad Patkowski who writes in reference [58] that “my personal least favorite SAPT term is the \(\delta E^{(2)}_{HF}\) correction of Equation (20) (...). I consider its presence as an admission that pure SAPT has a difficulty that cannot be fully resolved from within, and it requires outside help in a form of supermolecular HF”? Note that his Eq. (20) is the same as Eq. (7) in the current article and hence \(\delta E^{(2)}_{HF} = \delta E_{\text {int, resp}}^{\text {HF}}\).

Reply:

As already discussed above, from the practical point of view the addition of \(\delta E_{\text {int, resp}}^{\text {HF}}\) poses no problems. It does account for higher-order induction and exchange-induction effects and the errors introduced by this addition are very small. Furthermore, for nonpolar and mildly polar systems the addition of \(\delta E_{\text {int, resp}}^{\text {HF}}\) is not needed if the third-order SAPT interaction energies are computed [108]. Thus, the issue is more of aesthetic than practical nature. It may be possible to increase the range of systems that do not need \(\delta E_{\text {int, resp}}^{\text {HF}}\) by removing some approximations in the present set of third-order terms. Another possible step in this direction is to apply the formula for the second-order induction wave function derived in the present work in computations of the fourth- and fifth-order induction energy corrections. One more possible avenue is to make the regularized SAPT [18, 88] applicable to general monomers (the regularized SAPT exhibits a faster convergence of induction energies).

Popelier commented:

Question 4: It is stated that “One cannot uniquely determine the total charge transferred from monomer A to monomer B since this requires choosing an arbitrary boundary between the monomers.” The claim that charge cannot be uniquely assigned to atoms or even molecules is typically perpetuated, yet there is a great need to do so both in terms of interpretative (quantum) chemistry and force field construction. Which criterion (or criteria) give(s) rise to uniqueness if it is not experimental arbitration? According to this article, present-day SAPT (“the theory of intermolecular forces (...) providing the ‘standard model’ for EDA methods”) is declared unique because the symmetrized Rayleigh-Schrödinger method is the only one used in practice. With such perhaps relaxed uniqueness criterion, can Occam’s razor not be used to propose the topological partitioning as a satisfactory method to settle the debate on how to quantify charge transfer (even at the level of tens of millielectrons)? Moreover, an extensive and thorough comparison between fuzzy (interpenetrating) and non-fuzzy (space-filling, e.g., QTAIM) partitioning methods (J. Comp. Chem. 2007, 28, 161) showed that the latter “may be preferred from the chemical consistency point of view” as they also ”better preserve the atomic or fragment identity from the energetic point of view”.

Reply:

First, we changed the terminology to “reference model”. Nevertheless, we maintain that SAPT components are the quantities that EDA methods should compare to (and mostly do). As discussed in the reply to Mo et al., there is, actually, reasonably convincing experimental evidence for the physical character of SAPT components.

Also in the reply to Mo et al., we have extensively discussed our position on charge delocalization. Of course, AIM-atoms give a possible definition of charge delocalization, but still do not allow to determine its energetic effect using SAPT.

Popelier commented:

Question 5: Can SAPT match the quantification of steric effects that the Interacting Quantum Atoms (IQA) method is able to achieve (J. Phys. Chem. A 2016, 120, 9647; ChemPhysChem 2021, 22, 775; Chemistry Open 2019, 8, 560)?

Reply:

As stated in the quote from Richard Feynman in the “Introduction,” molecules are “repelling upon being squeezed into one another.” A simple interpretation relates this repulsion to steric effects: since atoms are approximately spherical, a molecule can be viewed as a shape resulting from superposition of such spheres, and if two interacting molecules come close enough to each other, the penetration of the spheres results in a repulsion. The more the spheres overlap, the larger the repulsive component becomes. In force fields, the repulsive effects are approximated by exponential terms or by large inverse powers of R (mostly 1/R12) with positive coefficients. SAPT, of course, accounts for steric effects, as demonstrated by the fact that SAPT PESs are accurate in the repulsive regions. The steric effects computed using SAPT cannot be compared with those computed using IQA as the two quantities are defined in different and unrelated ways. We would like to reiterate here that, as it is extensively discussed in “Exchange interactions”, SAPT provides physical interpretation of the steric interactions grounded in quantum mechanics. The main repulsive effect comes from the first-order exchange energy, Eq. (43). Although, as discussed in “Exchange interactions”, the exchange interactions are due to electron tunneling between monomers and not to overlap of wave functions, this tunneling is proportional to such overlap and can be modelled using overlap integrals [175,176,177]. The exchange interactions appear also in the second-order of SAPT, see Eqs. (46) and (47). These exchange interactions are repulsive as well, but significantly smaller than the first-order ones. They are also proportional to overlap integrals. Another group of effects related to the wave-function overlap are damping effects [53, 79,80,81,82]. In most cases, damping leads to positive contributions to interaction energies; however, as discussed briefly in the “Introduction,” it can also produce negative contributions, which then violate the simple steric interactions picture. Since the first-order exchange energy dominates the repulsive interactions, the values of this component presented in Table 1 provide a reasonable approximation of the total steric effect.

An example of unusual steric effects on shapes of PESs is provided in a recent study of the ammonia dimer [178]. This PES is very unusual in that it has a very narrow canyon-like valley where two equivalent minima are located, with a very small barrier between them at the lowest-energy saddle point. An overlap-driven variation of the components in the direction perpendicular to the interconversion path through this saddle, i.e., across the valley, explains the narrowness of this valley. One has to consider both the exchange and damping effects to get a complete explanation.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Szalewicz, K., Jeziorski, B. Physical mechanisms of intermolecular interactions from symmetry-adapted perturbation theory. J Mol Model 28, 273 (2022). https://doi.org/10.1007/s00894-022-05190-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00894-022-05190-z

Keywords

Navigation