We present a new, novel implementation of the Many-Body Expansion (MBE) to account for the breaking of covalent bonds, thus extending the range of applications from its previous popular usage in the breaking of hydrogen bonds in clusters to molecules. A central concept of the new implementation is the in situ atomic electronic state of an atom in a molecule that casts the one-body term as the energy required to promote it to that state from its ground state. The rest of the terms correspond to the individual diatomic, triatomic, etc., fragments. Its application to the atomization energies of the XHn series, X = C, Si, Ge, Sn and n = 1–4, suggests that the (negative, stabilizing) 2-B is by far the largest term in the MBE with the higher order terms oscillating between positive and negative values and decreasing dramatically in size with increasing rank of the expansion. The analysis offers an alternative explanation for the purported “first row anomaly” in the incremental Hn−1X–H bond energies seen when these energies are evaluated with respect to the lowest energy among the states of the XHn molecules. Due to the “flipping” of the ground/first excited state between CH2 (3B1 ground state, 1A1 first excited state) and XH2, X = Si, Ge, Sn (1A1 ground state, 3B1 first excited state), the overall picture does not exhibit a “first row anomaly” when the incremental bond energies are evaluated with respect to the molecular states having the same in situ atomic states.

The Many-Body Expansion (MBE) is a concept based on combinatorial mathematics that was first introduced over 300 years ago and is usually employed to count the number of elements in the union of finite sets.1 Its first application to chemical physics problems considered individual water molecules as “bodies” connected via hydrogen bonds as it was used to estimate the non-additive three-body term by partitioning the energy of a water trimer.2 Since then the MBE analysis based on the energies of distinct, non-overlapping sub-fragments has been applied to hydrogen bonded clusters by some of us3–9 and others10–27 to quantify the importance of non-additive terms in the binding energies of aqueous clusters. Recently, the details of the MBE based on high level electronic structure calculations related to the size of the orbital basis set and the level of electron correlation used in the expansion28–30 as well as a molecular dynamics protocol based on the MBE (MBE-MD)31 were reported. This type of MBE for hydrogen bonded molecular systems (including ions), where the definition of a “body” is straightforward and the system is partitioned in non-overlapping sub-fragments (monomers, dimers, trimers, etc.) by just breaking hydrogen bonds, has laid the foundation for the development of accurate, ab initio based, many-body interaction potentials for water.32–53 This is to be contrasted to different partitions that are, for instance, based on overlapping sub-fragments, such as the Molecular Tailoring Approach (MTA),54–59 or other fragmentation based approaches.60–64 Note that the MBE has also been applied to incorporate molecular orbitals as “bodies” in order to extrapolate the total correlation energy of molecules65–72 and solids.73–75 

In this paper, we extend the general idea of the MBE to the breaking of covalent bonds, that is, considering a polyatomic molecule as consisting of a collection of atoms, diatomics, triatomics, etc. A recent study76 of the carbon, silicon, and germanium hydrides, XHn (n = 1–4) based on the spin-coupled generalized valence bond theory has examined the qualitative changes between carbon and silicon/germanium in the Hn−1X–H bond energies (De) on the context of the “first raw anomaly”77 by expanding on Kutzelnigg’s argument78 based on the hybridization of the X atom bond orbitals due to the increase in the spatial separation of the ns and np orbitals between atoms in the first row and the following rows of the Periodic Table. The results presented in our study offer (vide infra) an alternative explanation for this result.

Our analysis is based on casting the atomization energy (ΔEatomiz⋅) of a molecule of N atoms computed with respect to its constituent atoms i (i = 1, …, N) in their respective ground states (Ei0) in the usual way4 as

ΔEatomiz⋅=ΔE1B+ΔE2B+ΔE3B++ΔEnB,
(1)

where

ΔE1B=iNEipEi0,
(2)
ΔE2B=i,jNΔ2Eij=i,jN(EijpqEipEjq),
(3)
ΔE3B=i,j,kNΔ3Eijk=i,j,kNEijkpqrΔ2EijΔ2EikΔ2EjkEipEjqEkr,
(4)

and Eip, Eijpq, Eijkpqr, Eijklpqrs, etc., are the energies of the in situ (ground or excited) states of atom (i), dimer (ij), trimer (ijk), tetramer (ijkl), etc., in the molecule. It is obvious that if the in situ state of an atom in the molecule (Eip) is the atom’s ground state (Ei0), the 1-B term for that atom, given by Eq. (2), is zero.

Since the concept of the in situ state of an atom in the molecule is central to our analysis, we will further discuss it using Fig. 1, as discussed by Heitler.79 Consider a carbon atom in its ground (3P) atomic state. The pairing of the four (2s22p2) valence electrons in this electronic state is not appropriate to accommodate bonding with four hydrogen atoms to form CH4. They should rather be promoted to the 5S excited electronic state lying 4.18 eV above80 the 3P ground state in order to form the 4 equivalent C–H bonds. In order words, the 2s and the three 2p orbitals of the carbon atom should be hybridized into four equivalent sp3 orbitals forming the atomic 5S state, which is the in situ electronic state of the carbon atom in CH4. Examples of the concept of in situ atomic states have been previously discussed for molecules81,82 and metal aqueous clusters.83,84 As it is evident from Eqs. (1)(4), the choice of the in situ electronic state of an atom will affect the MBE.

FIG. 1.

The in situ electronic structure of the carbon atom in CH4.

FIG. 1.

The in situ electronic structure of the carbon atom in CH4.

Close modal

The calculations for the ground states of the XHn hydrides, X = C, Si, Ge, Sn and n = 1–4, were performed at the Coupled Cluster Singles, Doubles and perturbative Triples [CCSD(T) and RCCSD(T)]85,86 levels of theory. Additionally, two excited states for CH and one for the remaining diatomic and all triatomic molecules were computed. For the CHn series, we employed Dunning’s augmented correlation consistent basis sets, aug-cc-pVxZ, x = D, T, Q.87,88 For the SiHn, GeHn, and SnHn species, we employed the aug-cc-pVQZ basis sets for H,87,88 Si,89 and Ge,90 and for Sn Peterson’s91 aug-cc-pVQZ-PP augmented correlation consistent basis sets, which employ accurate small-core (1s22s22p63s23p63d10; 28 electrons) relativistic pseudopotentials.

The atomization energies, the XHn−1–H dissociation energies, and the individual many body terms were corrected for basis set superposition error (BSSE)92,93 as described in Ref. 94. Note that the deformation (or relaxation) energy in that reference that arises from geometrical distortions due to the interaction is zero in the present case since the atomization energy is computed with respect to the atoms. The BSSE-corrected interaction energies (ΔEatomiz.) and the BSSE-corrected many-body terms, ΔE′(n-B), e given in Eqs. (5)(10),4 

ΔEatomiz⋅=EXiXjxixj(XiXj)iEXiXjxixj(Xi),
(5)
ΔE(1B)=ΔE(1B),
(6)
ΔE(2B)=i,jΔ2Eij,
(7)
Δ2Eij=EX1X2x1x2(XiXj)EX1X2x1x2(Xi)+EX1X2x1x2(Xj),
(8)
ΔE(3B)=i,j,kΔ3Eijk,
(9)
Δ3Eijk=EX1X2x1x2(XiXjXk)EX1X2x1x2(Xi)+EX1X2x1x2(Xj)+EX1X2x1x2(Xk)Δ2EX1X2.x1x2(XiXj)+Δ2EX1X2x1x2(XiXk)+Δ2EX1X2x1x2(XjXk),
(10)

etc., where EGs(M) refers to the total energy of the molecule M computed at the geometry G with basis set s.

In order to evaluate the appropriateness of RCCSD(T), which is a single-reference method, we checked the single (t1) and the double (t2) amplitudes95 as well as the T1 and D1 diagnostics.96,97 For all calculations of the present study, the t1 and t2 amplitudes were small and the T1 and D1 diagnostics were T1 < 0.02 and D1 < 0.04, expect for the 4Σ state of XH, for which the diagnostics were T1 ∼ 0.03 and D1 ∼ 0.07. We can, therefore, conclude that the single reference RCCSD(T) method is an appropriate methodology to be employed for the systems in this study. All calculations were carried out with the MOLPRO suite of codes.98 

An important detail of the above analysis is that, for each electronic state of the molecule, the two- and higher-body terms are not necessarily computed at the diatomic, triatomic, etc., ground electronic states but at the respective electronic states of these fragments that are formed from the in situ electronic states of the constituent atoms in the full molecule. We will further elaborate on this important detail in Sec. III when we present the case of XH2.

The results in this subsection are presented in Figs. 213 and Tables IV. Below we will discuss the individual members of the XHn series for each n as well as the trends with the atom identity.

FIG. 2.

Many-body decomposition of ΔEatomiz. for the 2Π, 4Σ, and 2Δ states of CH at the RCCSD(T)/aug-cc-pVxZ, x = D, T, Q, level of theory.

FIG. 2.

Many-body decomposition of ΔEatomiz. for the 2Π, 4Σ, and 2Δ states of CH at the RCCSD(T)/aug-cc-pVxZ, x = D, T, Q, level of theory.

Close modal
FIG. 3.

Many-body decomposition of ΔEatomiz. for the 2Π and 4Σ states of the XH series (X = C, Si, Ge, Sn) at the RCCSD(T)/aug-cc-pVQZ(-PP)Sn level of theory.

FIG. 3.

Many-body decomposition of ΔEatomiz. for the 2Π and 4Σ states of the XH series (X = C, Si, Ge, Sn) at the RCCSD(T)/aug-cc-pVQZ(-PP)Sn level of theory.

Close modal
FIG. 4.

Many-body decomposition of the ΔEatomiz. for the X3B1 and a1A1 states of CH2 at the RCCSD(T)/aug-cc-pVxZ, x = D, T, Q, level of theory.

FIG. 4.

Many-body decomposition of the ΔEatomiz. for the X3B1 and a1A1 states of CH2 at the RCCSD(T)/aug-cc-pVxZ, x = D, T, Q, level of theory.

Close modal
FIG. 5.

Many-body decomposition of the ΔEatomiz. for the 3B1 and 1A1 states of the XH2 series (X = C, Si, Ge, Sn) at RCCSD(T)/aug-cc-pVQZ(-PP)Sn level of theory.

FIG. 5.

Many-body decomposition of the ΔEatomiz. for the 3B1 and 1A1 states of the XH2 series (X = C, Si, Ge, Sn) at RCCSD(T)/aug-cc-pVQZ(-PP)Sn level of theory.

Close modal
FIG. 6.

Many-body decomposition of the ΔEatomiz. for the X2A1 state of CH3 at the RCCSD(T)/aug-cc-pVxZ, x = D, T, Q, level of theory.

FIG. 6.

Many-body decomposition of the ΔEatomiz. for the X2A1 state of CH3 at the RCCSD(T)/aug-cc-pVxZ, x = D, T, Q, level of theory.

Close modal
FIG. 7.

Many-body decomposition of the ΔEatomiz. for the X2A1 state of the XH3 series (X = C, Si, Ge, Sn) at the RCCSD(T)/aug-cc-pVQZ(-PP)Sn level of theory.

FIG. 7.

Many-body decomposition of the ΔEatomiz. for the X2A1 state of the XH3 series (X = C, Si, Ge, Sn) at the RCCSD(T)/aug-cc-pVQZ(-PP)Sn level of theory.

Close modal
FIG. 8.

Many-body decomposition of the ΔEatomiz. for the X1A1 state of CH4 at the RCCSD(T)/aug-cc-pVxZ, x = D, T, Q, level of theory.

FIG. 8.

Many-body decomposition of the ΔEatomiz. for the X1A1 state of CH4 at the RCCSD(T)/aug-cc-pVxZ, x = D, T, Q, level of theory.

Close modal
FIG. 9.

Many-body decomposition of the ΔEatomiz. for the X1A1 state of the XH4 series (X = C, Si, Ge, Sn) at the RCCSD(T)/aug-cc-pVQZ(-PP)Sn level of theory.

FIG. 9.

Many-body decomposition of the ΔEatomiz. for the X1A1 state of the XH4 series (X = C, Si, Ge, Sn) at the RCCSD(T)/aug-cc-pVQZ(-PP)Sn level of theory.

Close modal
FIG. 10.

Summary of the MBE for the ΔEatomiz. of the CHn species, n = 1–4, at the RCCSD(T)/aug-cc-pVQZ level of theory.

FIG. 10.

Summary of the MBE for the ΔEatomiz. of the CHn species, n = 1–4, at the RCCSD(T)/aug-cc-pVQZ level of theory.

Close modal
FIG. 11.

Many-body decomposition for the ΔEatomiz. of the SiHn species, n = 1–4, at RCCSD(T)/aug-cc-pVQZ level of theory.

FIG. 11.

Many-body decomposition for the ΔEatomiz. of the SiHn species, n = 1–4, at RCCSD(T)/aug-cc-pVQZ level of theory.

Close modal
FIG. 12.

Many-body decomposition for the ΔEatomiz. of the GeHn species, n = 1–4, at RCCSD(T)/aug-cc-pVQZ level of theory.

FIG. 12.

Many-body decomposition for the ΔEatomiz. of the GeHn species, n = 1–4, at RCCSD(T)/aug-cc-pVQZ level of theory.

Close modal
FIG. 13.

Many-body decomposition for the ΔEatomiz. of the SnHn species, n = 1–4, at RCCSD(T)/aug-cc-pVQZ(-PP)Sn level of theory.

FIG. 13.

Many-body decomposition for the ΔEatomiz. of the SnHn species, n = 1–4, at RCCSD(T)/aug-cc-pVQZ(-PP)Sn level of theory.

Close modal
TABLE I.

1-B term (eV), 2-B term (eV), atomization energy ΔEatomiz. (eV) with respect to ground state products, and dissociation energy ΔE (eV) with respect to the in situ atomic state without and with BSSE correction (second lines) of the XH molecules, X = C, Si, Ge, and Sn, at the RCCSD(T) level of theory.

Basis setMolecule1-B2-BΔEatomiz.ΔE
AVDZ CH (X2Π) 0.000 −3.350 −3.350 −3.350 
0.000 −3.309 −3.309 −3.309 
CH (A4Σ4.000 −6.707 −2.707 −6.707 
4.016 −6.679 −2.663 −6.679 
CH (α2Δ) 1.557 −1.730 −0.173 −1.729 
1.570 −1.713 −0.143 −1.713 
AVTZ CH (X2Π) 0.000 −3.571 −3.571 −3.571 
0.000 −3.548 −3.548 −3.548 
CH (A4Σ4.096 −6.960 −2.864 −6.960 
4.111 −6.950 −2.839 −6.950 
CH (α2Δ) 1.419 −1.871 −0.452 −1.871 
1.430 −1.864 −0.434 −1.864 
AVQZ CH (X2Π) 0.000 −3.616 −3.616 −3.616 
0.000 −3.606 −3.606 −3.606 
CH (A4Σ4.139 −7.025 −2.886 −7.025 
4.146 −7.022 −2.876 −7.022 
CH (α2Δ) 1.391 −1.895 −0.504 −1.895 
1.397 −1.891 −0.494 −1.891 
SiH (X2Π) 0.000 −3.172 −3.172 −3.172 
0.000 −3.165 −3.165 −3.165 
SiH (A4Σ3.907 −5.404 −1.497 −5.404 
3.912 −5.402 −1.490 −5.402 
GeH (X2Π) 0.000 −2.931 −2.931 −2.931 
0.000 −2.925 −2.925 −2.925 
GeH (A4Σ4.394 −5.505 −1.111 −5.506 
4.398 −5.503 −1.105 −5.504 
AVQZ(-PP)Sn SnH (X2Π) 0.000 −2.662 −2.662 −2.662 
0.000 −2.658 −2.658 −2.658 
SnH (A4Σ4.061 −4.896 −0.835 −4.896 
4.064 −4.894 −0.830 −4.894 
Basis setMolecule1-B2-BΔEatomiz.ΔE
AVDZ CH (X2Π) 0.000 −3.350 −3.350 −3.350 
0.000 −3.309 −3.309 −3.309 
CH (A4Σ4.000 −6.707 −2.707 −6.707 
4.016 −6.679 −2.663 −6.679 
CH (α2Δ) 1.557 −1.730 −0.173 −1.729 
1.570 −1.713 −0.143 −1.713 
AVTZ CH (X2Π) 0.000 −3.571 −3.571 −3.571 
0.000 −3.548 −3.548 −3.548 
CH (A4Σ4.096 −6.960 −2.864 −6.960 
4.111 −6.950 −2.839 −6.950 
CH (α2Δ) 1.419 −1.871 −0.452 −1.871 
1.430 −1.864 −0.434 −1.864 
AVQZ CH (X2Π) 0.000 −3.616 −3.616 −3.616 
0.000 −3.606 −3.606 −3.606 
CH (A4Σ4.139 −7.025 −2.886 −7.025 
4.146 −7.022 −2.876 −7.022 
CH (α2Δ) 1.391 −1.895 −0.504 −1.895 
1.397 −1.891 −0.494 −1.891 
SiH (X2Π) 0.000 −3.172 −3.172 −3.172 
0.000 −3.165 −3.165 −3.165 
SiH (A4Σ3.907 −5.404 −1.497 −5.404 
3.912 −5.402 −1.490 −5.402 
GeH (X2Π) 0.000 −2.931 −2.931 −2.931 
0.000 −2.925 −2.925 −2.925 
GeH (A4Σ4.394 −5.505 −1.111 −5.506 
4.398 −5.503 −1.105 −5.504 
AVQZ(-PP)Sn SnH (X2Π) 0.000 −2.662 −2.662 −2.662 
0.000 −2.658 −2.658 −2.658 
SnH (A4Σ4.061 −4.896 −0.835 −4.896 
4.064 −4.894 −0.830 −4.894 
TABLE II.

1-B term (eV), 2-B term (eV), 3-B term (eV), atomization energy ΔEatomiz. (eV) with respect to ground state products, and dissociation energy ΔE (eV) with respect to the in situ atomic state without and with BSSE correction (second lines) of the XH2 molecules, X = C, Si, Ge, and Sn at the RCCSD(T) level of theory. States are listed in decreasing magnitude of ΔEatomiz.

Basis setMolecule1-B2-BXH2-BHH2-B3-BΔEatomiz.ΔE
AVDZ CH2 (X3B14.000 −6.705 0.259 −13.151 1.404 −7.747 −11.747 
4.031 −6.696 0.249 −13.142 1.449 −7.661 −11.692 
CH2 (a1A10.000 −3.348 0.527 −6.169 −1.127 −7.296 −7.296 
0.000 −3.317 0.516 −6.117 −1.098 −7.214 −7.214 
AVTZ CH2 (X3B14.096 −6.957 0.265 −13.650 1.429 −8.125 −12.221 
4.122 −6.957 0.263 −13.650 1.446 −8.082 −12.204 
CH2 (a1A10.000 −3.570 0.534 −6.605 −1.105 −7.710 −7.710 
0.000 −3.550 0.533 −6.567 −1.101 −7.668 −7.668 
AVQZ CH2 (X3B14.139 −7.023 0.265 −13.780 1.443 −8.198 −12.337 
4.152 −7.024 0.265 −13.782 1.451 −8.180 −12.332 
CH2 (a1A10.000 −3.614 0.534 −6.694 −1.102 −7.795 −7.795 
0.000 −3.606 0.534 −6.678 −1.100 −7.778 −7.778 
SiH2 (X1A10.000 −3.171 0.146 −6.196 −0.425 −6.621 −6.621 
0.000 −3.166 0.146 −6.185 −0.424 −6.609 −6.609 
SiH2 (a3B13.907 −5.403 0.050 −10.755 1.111 −5.737 −9.644 
3.916 −5.403 0.050 −10.756 1.116 −5.725 −9.640 
GeH2 (X1A10.000 −2.930 0.109 −5.752 −0.359 −6.111 −6.111 
0.000 −2.925 0.109 −5.742 −0.358 −6.100 −6.100 
GeH2 (a3B14.394 −5.502 0.034 −10.970 1.481 −5.094 −9.488 
4.402 −5.502 0.033 −10.972 1.487 −5.083 −9.485 
AVQZ(-PP)Sn SnH2 (X1A10.000 −2.662 0.049 −5.275 −0.262 −5.537 −5.537 
0.000 −2.658 0.049 −5.267 −0.261 −5.528 −5.528 
SnH2 (a3B14.061 −4.892 0.012 −9.772 1.315 −4.397 −8.457 
4.068 −4.893 0.011 −9.775 1.320 −4.387 −8.455 
Basis setMolecule1-B2-BXH2-BHH2-B3-BΔEatomiz.ΔE
AVDZ CH2 (X3B14.000 −6.705 0.259 −13.151 1.404 −7.747 −11.747 
4.031 −6.696 0.249 −13.142 1.449 −7.661 −11.692 
CH2 (a1A10.000 −3.348 0.527 −6.169 −1.127 −7.296 −7.296 
0.000 −3.317 0.516 −6.117 −1.098 −7.214 −7.214 
AVTZ CH2 (X3B14.096 −6.957 0.265 −13.650 1.429 −8.125 −12.221 
4.122 −6.957 0.263 −13.650 1.446 −8.082 −12.204 
CH2 (a1A10.000 −3.570 0.534 −6.605 −1.105 −7.710 −7.710 
0.000 −3.550 0.533 −6.567 −1.101 −7.668 −7.668 
AVQZ CH2 (X3B14.139 −7.023 0.265 −13.780 1.443 −8.198 −12.337 
4.152 −7.024 0.265 −13.782 1.451 −8.180 −12.332 
CH2 (a1A10.000 −3.614 0.534 −6.694 −1.102 −7.795 −7.795 
0.000 −3.606 0.534 −6.678 −1.100 −7.778 −7.778 
SiH2 (X1A10.000 −3.171 0.146 −6.196 −0.425 −6.621 −6.621 
0.000 −3.166 0.146 −6.185 −0.424 −6.609 −6.609 
SiH2 (a3B13.907 −5.403 0.050 −10.755 1.111 −5.737 −9.644 
3.916 −5.403 0.050 −10.756 1.116 −5.725 −9.640 
GeH2 (X1A10.000 −2.930 0.109 −5.752 −0.359 −6.111 −6.111 
0.000 −2.925 0.109 −5.742 −0.358 −6.100 −6.100 
GeH2 (a3B14.394 −5.502 0.034 −10.970 1.481 −5.094 −9.488 
4.402 −5.502 0.033 −10.972 1.487 −5.083 −9.485 
AVQZ(-PP)Sn SnH2 (X1A10.000 −2.662 0.049 −5.275 −0.262 −5.537 −5.537 
0.000 −2.658 0.049 −5.267 −0.261 −5.528 −5.528 
SnH2 (a3B14.061 −4.892 0.012 −9.772 1.315 −4.397 −8.457 
4.068 −4.893 0.011 −9.775 1.320 −4.387 −8.455 
TABLE III.

1-B term (eV), 2-B term (eV), 3-B term (eV), 4-B term (eV), atomization energy ΔEatomiz. (eV) with respect to ground state products, and dissociation energy ΔE (eV) with respect to the in situ atomic state without and with BSSE correction (second lines) of the XH3 molecules, X = C, Si, Ge, and Sn at the RCCSD(T) level of theory.

Basis setMolecule1-B2-BXH2-BHH2-B3-BXHH3-BHHH3-B4-BΔEatomiz.ΔE
AVDZ CH3 (2A14.000 −6.704 0.352 −19.056 1.380 −0.199 3.942 −1.453 −12.567 −16.567 
4.045 −6.715 0.341 −19.120 1.435 −0.197 4.108 −1.475 −12.441 −16.486 
AVTZ CH3 (2A14.096 −6.958 0.362 −19.786 1.406 −0.212 4.006 −1.443 −13.128 −17.224 
4.130 −6.965 0.360 −19.814 1.428 −0.212 4.073 −1.460 −13.071 −17.201 
AVQZ CH3 (2A14.139 −7.023 0.363 −19.980 1.420 −0.213 4.048 −1.451 −13.244 −17.383 
4.156 −7.027 0.362 −19.995 1.430 −0.213 4.078 −1.459 −13.220 −17.376 
SiH3 (2A13.907 −5.402 0.068 −16.003 1.117 −0.028 3.322 −1.052 −9.826 −13.733 
3.918 −5.405 0.068 −16.013 1.124 −0.028 3.343 −1.058 −9.809 −13.727 
GeH3 (2A14.394 −5.501 0.049 −16.356 1.495 −0.019 4.467 −1.400 −8.895 −13.289 
4.405 −5.505 0.049 −16.367 1.503 −0.019 4.490 −1.407 −8.879 −13.284 
AVQZ(-PP)Sn SnH3 (2A14.061 −4.892 0.020 −14.618 1.337 −0.006 4.006 −1.260 −7.811 −11.872 
4.070 −4.896 0.019 −14.629 1.344 −0.006 4.026 −1.266 −7.799 −11.869 
Basis setMolecule1-B2-BXH2-BHH2-B3-BXHH3-BHHH3-B4-BΔEatomiz.ΔE
AVDZ CH3 (2A14.000 −6.704 0.352 −19.056 1.380 −0.199 3.942 −1.453 −12.567 −16.567 
4.045 −6.715 0.341 −19.120 1.435 −0.197 4.108 −1.475 −12.441 −16.486 
AVTZ CH3 (2A14.096 −6.958 0.362 −19.786 1.406 −0.212 4.006 −1.443 −13.128 −17.224 
4.130 −6.965 0.360 −19.814 1.428 −0.212 4.073 −1.460 −13.071 −17.201 
AVQZ CH3 (2A14.139 −7.023 0.363 −19.980 1.420 −0.213 4.048 −1.451 −13.244 −17.383 
4.156 −7.027 0.362 −19.995 1.430 −0.213 4.078 −1.459 −13.220 −17.376 
SiH3 (2A13.907 −5.402 0.068 −16.003 1.117 −0.028 3.322 −1.052 −9.826 −13.733 
3.918 −5.405 0.068 −16.013 1.124 −0.028 3.343 −1.058 −9.809 −13.727 
GeH3 (2A14.394 −5.501 0.049 −16.356 1.495 −0.019 4.467 −1.400 −8.895 −13.289 
4.405 −5.505 0.049 −16.367 1.503 −0.019 4.490 −1.407 −8.879 −13.284 
AVQZ(-PP)Sn SnH3 (2A14.061 −4.892 0.020 −14.618 1.337 −0.006 4.006 −1.260 −7.811 −11.872 
4.070 −4.896 0.019 −14.629 1.344 −0.006 4.026 −1.266 −7.799 −11.869 
TABLE IV.

1-B term (eV), 2-B term (eV), 3-B term (eV), 4-B term (eV), 5-B term (eV), atomization energy ΔEatomiz. (eV) with respect to ground state products, and dissociation energy ΔE (eV) with respect to the in situ atomic state without and with BSSE correction (second lines) of the XH4 molecules, X = C, Si, Ge, and Sn at the RCCSD(T) level of theory.

Basis setMolecule1-B2-BXH2-BHH2-B3-BXHH3-BHHH3-B4-BXHHH4-BHHHH4-B5-BΔEatomiz.ΔE
AVDZ CH4 (1A14.000 −6.707 0.451 −24.124 1.463 −0.267 7.711 −1.625 0.157 −6.343 1.511 −17.244 −21.244 
4.036 −6.729 0.439 −24.283 1.522 −0.264 8.074 −1.658 0.160 −6.471 1.537 −17.108 −21.143 
AVTZ CH4 (1A14.096 −6.960 0.461 −25.070 1.488 −0.282 7.804 −1.615 0.171 −6.290 1.500 −17.961 −22.057 
4.131 −6.976 0.460 −25.147 1.515 −0.281 7.963 −1.634 0.171 −6.363 1.519 −17.897 −22.028 
AVQZ CH4 (1A14.139 −7.025 0.463 −25.322 1.502 −0.284 7.880 −1.620 0.173 −6.307 1.507 −18.104 −22.243 
4.157 −7.033 0.462 −25.357 1.515 −0.284 7.953 −1.630 0.173 −6.346 1.516 −18.078 −22.234 
SiH4 (1A13.907 −5.402 0.074 −21.164 1.124 −0.032 6.620 −1.083 0.013 −4.318 0.961 −13.994 −17.901 
3.919 −5.407 0.074 −21.185 1.133 −0.031 6.671 −1.090 0.013 −4.347 0.968 −13.974 −17.893 
GeH4 (1A14.394 −5.500 0.054 −21.675 1.503 −0.021 8.933 −1.432 0.008 −5.721 1.311 −12.758 −17.152 
4.406 −5.506 0.054 −21.699 1.512 −0.021 8.989 −1.441 0.008 −5.755 1.319 −12.740 −17.146 
AVQZ(-PP)Sn SnH4 (1A14.061 −4.890 0.021 −19.434 1.334 −0.007 7.979 −1.260 0.002 −5.039 1.177 −11.255 −15.316 
4.071 −4.895 0.021 −19.456 1.343 −0.007 8.029 −1.268 0.002 −5.069 1.185 −11.240 −15.312 
Basis setMolecule1-B2-BXH2-BHH2-B3-BXHH3-BHHH3-B4-BXHHH4-BHHHH4-B5-BΔEatomiz.ΔE
AVDZ CH4 (1A14.000 −6.707 0.451 −24.124 1.463 −0.267 7.711 −1.625 0.157 −6.343 1.511 −17.244 −21.244 
4.036 −6.729 0.439 −24.283 1.522 −0.264 8.074 −1.658 0.160 −6.471 1.537 −17.108 −21.143 
AVTZ CH4 (1A14.096 −6.960 0.461 −25.070 1.488 −0.282 7.804 −1.615 0.171 −6.290 1.500 −17.961 −22.057 
4.131 −6.976 0.460 −25.147 1.515 −0.281 7.963 −1.634 0.171 −6.363 1.519 −17.897 −22.028 
AVQZ CH4 (1A14.139 −7.025 0.463 −25.322 1.502 −0.284 7.880 −1.620 0.173 −6.307 1.507 −18.104 −22.243 
4.157 −7.033 0.462 −25.357 1.515 −0.284 7.953 −1.630 0.173 −6.346 1.516 −18.078 −22.234 
SiH4 (1A13.907 −5.402 0.074 −21.164 1.124 −0.032 6.620 −1.083 0.013 −4.318 0.961 −13.994 −17.901 
3.919 −5.407 0.074 −21.185 1.133 −0.031 6.671 −1.090 0.013 −4.347 0.968 −13.974 −17.893 
GeH4 (1A14.394 −5.500 0.054 −21.675 1.503 −0.021 8.933 −1.432 0.008 −5.721 1.311 −12.758 −17.152 
4.406 −5.506 0.054 −21.699 1.512 −0.021 8.989 −1.441 0.008 −5.755 1.319 −12.740 −17.146 
AVQZ(-PP)Sn SnH4 (1A14.061 −4.890 0.021 −19.434 1.334 −0.007 7.979 −1.260 0.002 −5.039 1.177 −11.255 −15.316 
4.071 −4.895 0.021 −19.456 1.343 −0.007 8.029 −1.268 0.002 −5.069 1.185 −11.240 −15.312 
TABLE V.

Bond distances RXH (Å) and angles φHΧH (degrees) for the various states of XHn, (X = C, Si, Ge, Sn; n = 1–4) at the RCCSD(T) level of theory.

CHnSiHnGeHnSnHn
AVDZAVTZAVQZAVQZAVQZAVQZ
XH(Χ2Π) RXH 1.1400 1.1219 1.1203 1.5242 1.6028 1.7901 
XH(A4ΣRXH 1.1066 1.0908 1.0894 1.4969 1.5694 1.7550 
XH(α2Δ) RXH 1.1229 1.1069 1.1056    
XH2(3B1RXH 1.0943 1.0791 1.0775 1.4815 1.5449 1.7281 
φHΧH 66.53 66.78 66.81 59.19 59.70 59.31 
XH2(1A1RXH 1.1271 1.1107 1.1088 1.5184 1.5963 1.7841 
φHΧH 50.57 50.94 51.00 46.14 45.82 45.60 
XH3(X2A1RXH 1.0932 1.0795 1.0780 1.4810 1.5443 1.7289 
φHΧH 120.00 120.00 120.00 107.59 107.68 109.21 
XH4(X1A1RXH 1.1027 1.0899 1.0883 1.4803 1.5414 1.7289 
φHΧH 109.47 109.47 109.47 109.49 109.47 109.21 
CHnSiHnGeHnSnHn
AVDZAVTZAVQZAVQZAVQZAVQZ
XH(Χ2Π) RXH 1.1400 1.1219 1.1203 1.5242 1.6028 1.7901 
XH(A4ΣRXH 1.1066 1.0908 1.0894 1.4969 1.5694 1.7550 
XH(α2Δ) RXH 1.1229 1.1069 1.1056    
XH2(3B1RXH 1.0943 1.0791 1.0775 1.4815 1.5449 1.7281 
φHΧH 66.53 66.78 66.81 59.19 59.70 59.31 
XH2(1A1RXH 1.1271 1.1107 1.1088 1.5184 1.5963 1.7841 
φHΧH 50.57 50.94 51.00 46.14 45.82 45.60 
XH3(X2A1RXH 1.0932 1.0795 1.0780 1.4810 1.5443 1.7289 
φHΧH 120.00 120.00 120.00 107.59 107.68 109.21 
XH4(X1A1RXH 1.1027 1.0899 1.0883 1.4803 1.5414 1.7289 
φHΧH 109.47 109.47 109.47 109.49 109.47 109.21 

XH (X = C, Si, Ge, Sn): The analysis for the ground (2Π), first (4Σ), and second (2Δ) excited states of CH is shown in Fig. 2. Energy differences are taken with respect to the ground state of the two atoms, viz., C(3P) + H(2S). The 2Π ground state of CH correlates with the corresponding ground state atomic states, i.e., the in situ atomic state of C in the ground state of CH is the atomic ground 3P state. In other words, the 1-B term is zero since there is no need to promote the ground 3P state of the C atom to form the C–H bond in its ground 2Π state. In this case, the 2-B term is identical with the atomization energy ΔEatomiz⋅ However, the situation is different for the first and second excited states of the CH molecule. The in situ electronic state of the C atom in the first 4Σ excited state of CH is the atomic 5S state, so the (positive) 1-B term is the energy required to promote the C atom from 3P → 5S. Accordingly, the 2-B term is the energy difference between the C(5S) + H(2S) and CH(4Σ) energy levels, which is also the difference between the total atomization energy (ΔEatomiz⋅) and the 1-B term. Similarly, the in situ electronic state of the C atom in the second 2Δ state if CH is the atomic 2D state and the 1-B and 2-B terms are computed accordingly. The results with the various basis sets (AVDZ, AVTZ, and AVQZ) are denoted with different colors for the different energy levels obtained with the three basis sets in Fig. 2. The individual 2-B terms for the ground (2Π), first (4Σ), and second (2Δ) excited states of CH are decreasing in that order (cf. Table I), and the presence of the repulsive 1-B term for the last two determines the overall order in the atomization energies with respect to the atomic ground states, as shown in Fig. 2. The σ bond length of the 4Σ state is shorter than the corresponding value of the X2Π state by 0.03 Å, showing that the bonding of the 4Σ state is stronger than in the X2Π state. The 2-B term of the 4Σ state is significantly larger (double) than the 2-B term of X2Π. This happens because the formation of a σ bond with a highly open shell system, such as C(5S), significantly stabilizes the carbon atom and consequently the whole C–H system. On the contrary, the C(3P) is lower in energy than C(5S), it has only two unpaired electrons, and, thus, even though the σ bond stabilize the C–H molecule, this stabilization is not dramatically large.

The situation for the ground (2Π) and the first (4Σ) excited states of SiH, GeH, and SnH is similar to that for CH; the results for the XH series (X = C, Si, Ge, Sn) are graphically summarized in Fig. 3, and the individual numbers (including the ones corrected for BSSE) are listed in Table I. The decrease of the individual 2-B terms for the respective ground (2Π) and excited (A4Σ) states in the XH series monotonically follows the trends of increasing X–H separations (cf. Table V) in an almost linear fashion.

XH2(X = C, Si, Ge, Sn): The MBE analysis for the ground (X3B1) and the first excited (a1A1) states of CH2 is schematically shown in Fig. 4, where the zero in the energy scale is taken with respect to the C(3P) + 2 × H(2S) asymptote. The ground (X3B1) state of CH2 correlates with the C(5S) + H(2S) + H(2S) atomic states so the 1-B term is positive and corresponds to the 3P → 5S promotion energy. The 2-B term is negative, whereas there is a smaller positive 3-B term. The MBE based on the various fragments allows for the further attribution of the individual terms to specific interactions between the constituent atoms. As can be seen from Table II, the main contribution to the total 2-B term (being the sum of the individual interactions between C–H and H–H) comes from the attractive (stabilizing) 2-BCH term with the remaining 2-BHH term being quite small and repulsive (destabilizing), amounting to just ∼4% of the total 2-B interaction. In contrast, for the first (a1A1) excited state of CH2, also shown in Fig. 4, the 1-B term is zero (since the in situ electronic state of C in the (a1A1) state of CH2 is the atomic 3P state), and, in this case, both the 2- and 3-B terms are negative. This is expected for the 3-B term for the 1A1 state of CH2 and for the rest of the XH2 molecules (see below) because this state corresponds to a closed shell system, and, thus, the simultaneous existence of the three atoms further stabilize it. On the contrary, this not the case for the X3B1 state in which the system is not closed shell and, thus, it follows the usual trend, i.e., the many-body terms oscillate between the positive and negative values.

The analysis for the combined results for the XH2 series, X = C, Si, Ge, Sn, are schematically shown in Fig. 5 and listed in Table II. A notable difference between CH2 and the rest of the XH2 series (X = Si, Ge, Sn) is that the order of the ground/first excited state is reversed. Indeed, as can be seen from Figs. 4 and 5 and the results of Table I, for CH2, the ground state is the 3B1 and the first excited the 1A1 state, whereas, for X = Si, Ge, Sn, the ground state of XH2 is the 1A1 and the first excited the 3B1 state. Therefore, the in situ electronic state of the carbon atom in the ground 3B1 state of CH2 is C(5S), whereas the in situ electronic state of the Si/Ge/Sn in the ground state of SiH2/GeH2/SnH2 is Si/Ge/Sn (3P) (see Figs. 4 and 5). The difference on the type of the ground state results from the fact that the Si, Ge, and Sn atoms are larger than C, forming significant larger X–H bond lengths and this favors the formation of two X–H bonds from the X(3P) rather than from the X(5S) states for X = Si, Ge, Sn. Note that the bond arising from the atomic 5S state is stabilized via shorter X–H bond lengths than the one arising from the X(3P) atomic state. Additionally, the 2-BXH and 2-BHH terms are decreasing in the series for both the ground and the excited states, in line with the corresponding monotonic increase of the X–H bond distance (cf. Table V). Note that the 3-B term for the ground (X3B1) state is positive (destabilizing), whereas the one for the first excited (a1A1) state is negative (stabilizing) in the series. For both states, the 3-B is much smaller than the 2-B term, as seen from Fig. 5 and Table II.

In the following, we elaborate on an important detail of the MBE analysis related to the electronic states used to compute the energies of the fragments. As it can be seen from Figs. 2 and 3, the 2Π ground state of XH (X = C, Si, Ge, Sn) is formed from the respective X(3P) + H(2S) states, whereas the first excited (4Σ) state from the X(5S) + H(2S) states. The ground 3B1 state of CH2 is formed (cf. Fig. 4) from the C(5S) + 2H(2S) states and, when evaluating the MBE for that state, the two-body “CH” term is computed with CH at its excited 4Σ state [also formed from C(5S) + H(2S)]. In contrast, the first excited a1A1 state of CH2 is formed (cf. Fig. 4) from C(3P) + 2H(2S) and the two-body “CH” term for that state is computed with CH at its ground 2Π state [also formed from C(3P) + H(2S)]. The calculation of the two-body “XH” term in XH2 for X = Si, Ge, Sn is evaluated in a similar manner while also noting that for these molecules the order of the ground and first excited states flip with respect to CH2 (see Fig. 5 and Table II). This protocol is also followed in the evaluation of the MBE for the larger molecules.

XH3(X = C, Si, Ge, Sn): The MBE for the ground (X2A1) state of CH3 is schematically shown in Fig. 6. The ground state of CH3 correlates with the C(5S) + H(2S) + H(2S) + H(2S) atomic states, viz., there is a positive 1-B term corresponding to the promotion 3P → 5S for the carbon atom. The 3-B term is smaller than the 2-B and positive, whereas the 4-B is even smaller than the 3-B and negative. The trends for the MBE of the XH3 series, X = C, Si, Ge, Sn, schematically shown in Fig. 7 and listed in Table III, are similar to the ones for CH3 with the terms oscillating between the positive (destabilizing) and negative (stabilizing) values, whereas they overall decrease in magnitude with rank k after the 2-B term, as can also be seen from Table III. The MBE, therefore, seems to be “converging” albeit in a slow, oscillating manner. Again, the total 2-BXH (sum of three identical components) is the major contributor to the 2-B term with the 2-BHH term being quite small and repulsive as in the XH2 series; they both decrease in magnitude in the series. With regard to the 3-B term, the signs of the individual components are reversed when compared to the 2-B term: the 3-BXHH is positive, whereas the 3-BHHH is very small and negative.

XH4(X = C, Si, Ge, Sn): The in situ atomic state of carbon in the ground (X1A1) state of CH4 is 5S, giving rise to a positive 1-B energy. The negative (stabilizing) 2-B term is the largest one among the higher order terms, which are oscillating between the positive and negative values and are decreasing in size after the 2-B term as can be seen from Fig. 8. The same trends are observed for the rest of the XH4 series (X = C, Si, Ge, Sn) as shown in Fig. 9 and Table IV. The 3-BXHH and 3-BHHH terms are similar in sign, magnitude, and trend across the XH4 series as in the XH3 series. The same behavior is seen for the 4-BXHHH and 4-BHHHH terms (cf. Table IV). A notable difference is that the total (positive) 3-B term almost cancels the total (negative) 4-B term. The 5-B term is small (∼10% of the atomization energy).

Trends in the MBE of XHn(X = C, Si, Ge, Sn): The MBE for the XHn series is summarized in Figs. 1013 for X = C, Si, Ge, and Sn, respectively. The trends in both the sign and magnitude of the individual MBE terms are similar across the series, namely, that the 1-B term is zero for n = 1 and positive (energy for the 3P → 5S promotion of the carbon atom) for n = 2–4 with the 2-B term being the largest one in the MBE and the rest of the terms oscillating in sign and diminishing after that. The behavior of the MBE is similar across the series. The correlation between the X–H distances and the magnitude of the individual 2-BXH terms, discussed earlier, is shown in Fig. 14 for the XHn series. The linear correlation across the XHn series for each n confirms that the X–H distance is a descriptor of the magnitude of the individual 2-BXH terms. Note the difference in the magnitude of the 2-B terms for the XH series due to the absence of a 1-B term for this case. This is because the individual 2-B terms are determined with respect to the C(3P) + H(2S) for XH and with respect to the C(5S) + n × H(2S) for the XHn, n = 2, 3, 4 series. It is interesting that this descriptor is independent of n, i.e., the size of the XHn molecules. The linear fits corresponding to the same in situ atoms are almost identical as shown in Fig. 14 

FIG. 14.

Correlation between the individual 2-BXH term and the X–H distance for the XHn series, X = C, Si, Ge, Se and n = 1–4. Note the different magnitude for the XH species due to the fact that the energy difference is taken with respect to the ground state atoms (consistent with the absence of the 1-B term) for that case.

FIG. 14.

Correlation between the individual 2-BXH term and the X–H distance for the XHn series, X = C, Si, Ge, Se and n = 1–4. Note the different magnitude for the XH species due to the fact that the energy difference is taken with respect to the ground state atoms (consistent with the absence of the 1-B term) for that case.

Close modal

The variation of the individual 2-BHH and 2-BXH terms for the XHn series, X = C, Si, Ge, Se and n = 2–4 is shown in Fig. 15. In the left panel, which shows the variation of the 2-BHH across the series, the filled symbols trace states in which X is at the in situ5S atomic state, whereas the open symbols denote states in which X is at the in situ3P atomic state. The larger relative increase in the 2-BHH term across the CHn series is due to the fact that the hydrogen atoms are closer to one another with increasing n as the result of the shorter C–H distances being in the 1.08–1.12 Å range (cf. Table V). In contrast, the 2-BHH terms for XHn, where X = Si, Ge, and Sn, are almost the same for n = 2–4, a result of the longer X–H distances (1.50–1.80 Å range, cf. Table V) that bring the hydrogen atoms further apart. The variation of the 2-BXH term across the series is shown on the right panel of Fig. 15, where the two different curves refer to the calculation when the heavy atom is at the in situ5S and 3P atomic states in the molecule. As a general trend, when X is at the atomic 3P state in the molecule, we observe a monotonic decrease of the 2-BXH term with the size of the X atom. This is consistent with the fact that the X–H bond distance, already being identified as a descriptor of the 2-BXH term (cf. Fig. 14), increases across the series. When X is at the in situ atomic 5S state, the 2-BXH is also decreasing across the series, except for Ge. This “anomaly” is attributed to the fact that the 3P → 5S energy is the largest one for Ge (cf. Tables IIV). It is 6.1% larger than C, 12.5% larger than Si, and 8.2% larger than Sn, resulting in the 2-BGe-H term of Ge(5S) to be further stabilized.

FIG. 15.

Variation of the individual 2-BHH and 2-BXH terms for the XHn series, X = C, Si, Ge, Se and n = 2–4 at the RCCSD(T)/AVQZ(-PP)Sn level of theory. In the left panel, solid/open symbols denote the states in which X is at the 5S/3P atomic states.

FIG. 15.

Variation of the individual 2-BHH and 2-BXH terms for the XHn series, X = C, Si, Ge, Se and n = 2–4 at the RCCSD(T)/AVQZ(-PP)Sn level of theory. In the left panel, solid/open symbols denote the states in which X is at the 5S/3P atomic states.

Close modal

The incremental Hn−1X–H bond energies for X = C, Si, Ge, Sn are shown in Fig. 16. In the left panel of that figure, the energies are computed with respect to the lowest in energy (ground states) of the Hn−1X molecule. Following our previous discussion in Sec. III A, these are 3B1 for CH2 and 1A1 for SiH2, GeH2, and SnH2 (indicated on the left panel of Fig. 16 for n = 2). The left panel of Fig. 16 is the same as the right panel of Fig. 9 of Ref. 76, but extended for X = Sn. As discussed in this earlier publication and reproduced by our results, the trend in the incremental bond energies with respect to the lowest in energy states is qualitatively different between the CHn and the rest of XHn series: indeed, there exists a “peak” for CH3 (the bond energy for CH3 is larger than the ones for CH2 and CH5), whereas there is a “dip” for the XHn, X = Si, Ge, Sn, series (the bond energy for XH3 is lower than that for XH2 and XH5 when X = Si, Ge, Sn).

FIG. 16.

The incremental Hn−1X–H bond energy, X = C, Si, Ge, Sn. Left panel: energies with respect to the lowest in energy of the Hn−1X molecule. Right panel: energies with respect to the Hn−1X molecule, where X is at the 5S state (solid line) and at the 3P state (dotted line). All energies are computed at the RCCSD(T)/AVQZ(-PP)Sn level of theory.

FIG. 16.

The incremental Hn−1X–H bond energy, X = C, Si, Ge, Sn. Left panel: energies with respect to the lowest in energy of the Hn−1X molecule. Right panel: energies with respect to the Hn−1X molecule, where X is at the 5S state (solid line) and at the 3P state (dotted line). All energies are computed at the RCCSD(T)/AVQZ(-PP)Sn level of theory.

Close modal

However, when the bond energies are computed with respect to the molecular states having the same in situ atomic state (i.e., not necessarily the lowest energy ones but the ones correlating with the same atomic states), the situation is different as shown in the right panel of Fig. 16. In that panel, the various states, with respect to which the individual incremental energies were computed, are noted. The Hn−1X–H incremental bond energies are obtained with respect to the Hn−1X molecule, where X is at either the 5S (solid line) or at the 3P (dotted line) atomic states. When this protocol is followed, the variation of the bond energies with n is qualitatively similar across the series and the concept of the “first row anomaly” is not justified. One can nevertheless argue that the (first row) carbon atom can be still considered different than the (second and subsequent rows) Si, Ge, and Sn atoms in lieu of the “flipping” of the ground/excited states for n = 2 in the series (CH2 vs SiH2/GeH2/SnH2). However, this qualitative change is not a property of the individual atoms per se but rather their different behavior in just the XH2, but not the rest of the hydrides.

We have described a novel decomposition scheme for the MBE in molecules and applied it to the XHn series (X = C, Si, Ge, Sn; n = 1–4). The protocol allows for the decomposition of the atomization energy in terms of atoms, dimers, trimers, etc. In the present implementation of the MBE that is based on the breaking of covalent bonds to define subsystems, the 1-B term is both qualitative and quantitatively different than the analogous 1-B term that has been popularized in the MBE of aqueous ionic systems. When breaking hydrogen bonds to define the various subsystems, the 1-B term represents the geometrical distortion94 of the individual fragments from their gas phase geometries due to the interaction with other neighboring molecules or ions. It amounts to a few kcal/mol and it can be indirectly probed experimentally by infrared (IR) spectroscopy,99,100 which records the change in the position of the individual vibrational bands from the gas phase as a result of the change in the fragment’s geometry from the isolated species. In contrast, when breaking covalent bonds, the 1-B represents the electronic excitation of an individual atom101 to bring it to its in situ electronic state in the molecule, and it amounts to a few eV (i.e., it is about two orders of magnitude larger) and can be probed by Ultraviolet–Visible (UV) spectroscopy.

We have found that the MBE in the XHn series is oscillating between positive and negative values, and it is converging with increasing rank of the expansion. Among the individual terms, the 2-B is by far the largest one in the MBE. The X–H distance is a good descriptor of the strength of the 2-BXH term, and its variation across the series can be attributed to the respective geometrical changes. The analysis can offer an alternative explanation for the purported “first row anomaly” that is based upon the different variation of the incremental bond energies for CHn compared to XHn, X = Si, Ge, Sn, when these bond energies are evaluated with respect to the lowest energies (ground states) of the members in the series. However, there is a “flipping” between the ground and the first excited state in CH2 (ground is the 3B1 and excited is the 1A1 states) compared to XH2, where X = Si, Ge, and Sn (ground is the 1A1 and excited is the 3B1 state). When these incremental binding energies are evaluated with respect to the molecular states that have the same in situ atomic state, their variation is the same with n for all X (C, Si, Ge, Sn) in the XHn series and the concept of the “first row anomaly” is, thus, not justified.

Note that the present analysis is simplified due to the presence of just one heavy atom (C, Si, Ge, Sn) in each of the systems studied and can be much more complex when more than one heavy atom is involved. Nevertheless, it provides a straightforward extension of the popular MBE for hydrogen bonded systems to incorporate the breaking of covalent bonds and offers valuable insights into the chemical bonding of chemical systems.

S.S.X. was supported from the Center for Scalable Predictive methods for Excitations and Correlated phenomena (SPEC), which is funded by the U.S. Department of Energy, Office of Science, Basic Energy Sciences, Chemical Sciences, Geosciences and Biosciences Division as part of the Computational Chemical Sciences (CCS) program at Pacific Northwest National Laboratory. Battelle operates the Pacific Northwest National Laboratory for the U.S. Department of Energy. This research used computer resources provided by the National Energy Research Scientific Computing Center, which is supported by the Office of Science of the U.S. Department of Energy under Contract No. DE-AC02-05CH11231.

The authors have no conflicts to disclose.

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
F.
Roberts
and
B.
Tesman
,
Applied Combinatorics
, 2nd ed. (
CRC Press
,
2009
).
2.
D.
Hankins
,
J. W.
Moskowitz
, and
F. H.
Stillinger
, Jr.
,
J. Chem. Phys.
53
,
4544
4554
(
1970
).
3.
S. S.
Xantheas
and
T. H.
Dunning
, “
The structure of the water trimer from ab initio calculations
,”
J. Chem. Phys.
98
,
8037
8040
(
1993
).
4.
S. S.
Xantheas
, “
Ab initio studies of cyclic water clusters (H2O)n, n = 1–6. II. Analysis of many-body interactions
,”
J. Chem. Phys.
100
,
7523
7534
(
1994
).
5.
S. S.
Xantheas
and
T. H.
Dunning
, “
Structures and energetics of F-(H2O)n, n = 1–3, clusters from ab initio calculations
,”
J. Phys. Chem.
98
,
13489
13497
(
1994
).
6.
S. S.
Xantheas
, “
Quantitative description of hydrogen bonding in chloride-water clusters
,”
J. Phys. Chem.
100
,
9703
9713
(
1996
).
7.
S. S.
Xantheas
, “
Significance of higher-order many-body interaction energy terms in water clusters and bulk water
,”
Philos. Mag. B
73
,
107
115
(
1996
).
8.
M. P.
Hodges
,
A. J.
Stone
, and
S. S.
Xantheas
, “
Contribution of many-body terms to the energy for small water clusters: A comparison of ab initio calculations and accurate model potentials
,”
J. Phys. Chem. A
101
,
9163
9168
(
1997
).
9.
S. S.
Xantheas
, “
Cooperativity and hydrogen bonding network in water clusters
,”
Chem. Phys.
258
,
225
231
(
2000
).
10.
R. M.
Richard
and
J. M.
Herbert
, “
A generalized many-body expansion and a unified view of fragment-based methods in electronic structure theory
,”
J. Chem. Phys.
137
,
064113
(
2012
).
11.
R. M.
Richard
and
J. M.
Herbert
, “
Many-body expansion with overlapping fragments: Analysis of two approaches
,”
J. Chem. Theory Comput.
9
,
1408
1416
(
2013
).
12.
R. M.
Richard
,
K. U.
Lao
, and
J. M.
Herbert
, “
Approaching the complete-basis limit with a truncated many-body expansion
,”
J. Chem. Phys.
139
,
224102
(
2013
).
13.
R. M.
Richard
,
K. U.
Lao
, and
J. M.
Herbert
, “
Understanding the many-body expansion for large systems. I. Precision considerations
,”
J. Chem. Phys.
141
,
014108
(
2014
).
14.
K. U.
Lao
,
K.-Y.
Liu
,
R. M.
Richard
, and
J. M.
Herbert
, “
Understanding the many-body expansion for large systems. II. Accuracy considerations
,”
J. Chem. Phys.
144
,
164105
(
2016
).
15.
J.
Liu
and
J. M.
Herbert
, “
Pair–pair approximation to the generalized many-body expansion: An alternative to the four-body expansion for ab initio prediction of protein energetics via molecular fragmentation
,”
J. Chem. Theory Comput.
12
,
572
584
(
2016
).
16.
K.-Y.
Liu
and
J. M.
Herbert
, “
Understanding the many-body expansion for large systems. III. Critical role of four-body terms, counterpoise corrections, and cutoffs
,”
J. Chem. Phys.
147
,
161729
(
2017
).
17.
K.-Y.
Liu
and
J. M.
Herbert
, “
Energy-screened many-body expansion: A practical yet accurate fragmentation method for quantum chemistry
,”
J. Chem. Theory Comput.
16
,
475
487
(
2020
).
18.
R. A.
Christie
and
K. D.
Jordan
, in
Intermolecular Forces and Clusters II. Structure and Bonding
, edited by
D. J.
Wales
(
Springer
,
2005
), Vol. 116, pp.
27
41
.
19.
E. E.
Dahlke
and
D. G.
Truhlar
, “
Electrostatically embedded many-body expansion for simulations
,”
J. Chem. Theory Comput.
4
,
1
6
(
2008
).
20.
J. F.
Ouyang
,
M. W.
Cvitkovic
, and
R. P. A.
Bettens
, “
Trouble with the many-body expansion
,”
J. Chem. Theory Comput.
10
,
3699
3707
(
2014
).
21.
S.
Prasad
,
J.
Lee
, and
M.
Head-Gordon
, “
Polarized many-body expansion: A perfect marriage between embedded mean-field theory and variational many-body expansion
,” in
Abstracts of Papers of the American Chemical Society
(
American Chemical Society
,
2019
), Vol. 257.
22.
M. A.
Collins
,
M. W.
Cvitkovic
, and
R. P. A.
Bettens
, “
The combined fragmentation and systematic molecular fragmentation methods
,”
Acc. Chem. Res.
47
,
2776
2785
(
2014
).
23.
M. A.
Collins
and
R. P. A.
Bettens
, “
Energy-based molecular fragmentation methods
,”
Chem. Rev.
115
,
5607
5642
(
2015
).
24.
R. M.
Richard
,
K. U.
Lao
, and
J. M.
Herbert
, “
Aiming for benchmark accuracy with the many-body expansion
,”
Acc. Chem. Res.
47
,
2828
2836
(
2014
).
25.
W.
Chen
and
M. S.
Gordon
, “
Energy decomposition analyses for many-body interaction and applications to water complexes
,”
J. Phys. Chem.
100
,
14316
14328
(
1996
).
26.
M. S.
Gordon
 et al, “
Accurate methods for large molecular systems
,”
J. Phys. Chem. B
113
,
9646
9663
(
2009
).
27.
M. S.
Gordon
,
D. G.
Fedorov
,
S. R.
Pruitt
, and
L. V.
Slipchenko
, “
Fragmentation methods: A route to accurate calculations on large systems
,”
Chem. Rev.
112
,
632
672
(
2012
).
28.
J. P.
Heindel
and
S. S.
Xantheas
, “
The many-body expansion for aqueous systems revisited: I. Water–water interactions
,”
J. Chem. Theory Comput.
16
,
6843
6855
(
2020
).
29.
J. P.
Heindel
and
S. S.
Xantheas
, “
The many-body expansion for aqueous systems revisited: II. Alkali metal and halide ion–water interactions
,”
J. Chem. Theory Comput.
17
,
2200
2216
(
2021
).
30.
K. M.
Herman
,
J. P.
Heindel
, and
S. S.
Xantheas
, “
The many-body expansion for aqueous systems revisited: III. Hofmeister ion–water interactions
,”
Phys. Chem. Chem. Phys.
23
,
11196
11210
(
2021
).
31.
J. P.
Heindel
and
S. S.
Xantheas
, “
Molecular dynamics driven by the many-body expansion (MBE-MD)
,”
J. Chem. Theory Comput.
17
,
7341
7352
(
2021
).
32.
C. J.
Burnham
,
J.
Li
,
S. S.
Xantheas
, and
M.
Leslie
, “
The parametrization of a Thole-type all-atom polarizable water model from first principles and its application to the study of water clusters (n = 2–21) and the phonon spectrum of ice Ih
,”
J. Chem. Phys.
110
,
4566
4581
(
1999
).
33.
S. S.
Xantheas
,
C. J.
Burnham
, and
R. J.
Harrison
, “
Development of transferable interaction models for water. II. Accurate energetics of the first few water clusters from first principles
,”
J. Chem. Phys.
116
,
1493
1499
(
2002
)..
34.
C. J.
Burnham
and
S. S.
Xantheas
, “
Development of transferable interaction models for water. III. Reparametrization of an all-atom polarizable rigid model (TTM2-R) from first principles
,”
J. Chem. Phys.
116
,
1500
1510
(
2002
).
35.
C. J.
Burnham
and
S. S.
Xantheas
, “
Development of transferable interaction models for water. IV. A flexible, all-atom polarizable potential (TTM2-F) based on geometry dependent charges derived from an ab initio monomer dipole moment surface
,”
J. Chem. Phys.
116
,
5115
5124
(
2002
).
36.
G. S.
Fanourgakis
and
S. S.
Xantheas
, “
The flexible, polarizable, thole-type interaction potential for water (TTM2-F) revisited
,”
J. Phys. Chem. A
110
,
4100
4106
(
2006
).
37.
G. S.
Fanourgakis
and
S. S.
Xantheas
, “
The bend angle of water in ice Ih and liquid water: The significance of implementing the nonlinear monomer dipole moment surface in classical interaction potentials
,”
J. Chem. Phys.
124
,
174504
(
2006
).
38.
G. S.
Fanourgakis
and
S. S.
Xantheas
, “
Development of transferable interaction potentials for water. V. Extension of the flexible, polarizable, Thole-type model potential (TTM3-F, v. 3.0) to describe the vibrational spectra of water clusters and liquid water
,”
J. Chem. Phys.
128
,
074506
(
2008
).
39.
G. S.
Fanourgakis
,
G. K.
Schenter
, and
S. S.
Xantheas
, “
A quantitative account of quantum effects in liquid water
,”
J. Chem. Phys.
125
,
141102
(
2006
).
40.
V.
Babin
,
C.
Leforestier
, and
F.
Paesani
, “
Development of a ‘first principles’ water potential with flexible monomers: Dimer potential energy surface, VRT spectrum, and second virial coefficient
,”
J. Chem. Theory Comput.
9
,
5395
5403
(
2013
).
41.
V.
Babin
,
G. R.
Medders
, and
F.
Paesani
, “
Development of a ‘first principles’ water potential with flexible monomers. II: Trimer potential energy surface, third virial coefficient, and small clusters
,”
J. Chem. Theory Comput.
10
,
1599
1607
(
2014
).
42.
G. R.
Medders
,
V.
Babin
, and
F.
Paesani
, “
Development of a ‘first-principles’ water potential with flexible monomers. III. Liquid phase properties
,”
J. Chem. Theory Comput.
10
,
2906
2910
(
2014
).
43.
G. R.
Medders
,
A. W.
Götz
,
M. A.
Morales
,
P.
Bajaj
, and
F.
Paesani
, “
On the representation of many-body interactions in water
,”
J. Chem. Phys.
143
,
104102
(
2015
).
44.
D. J.
Arismendi-Arrieta
,
M.
Riera
,
P.
Bajaj
,
R.
Prosmiti
, and
F.
Paesani
, “
i-TTM model for ab initio-based ion–water interaction potentials. 1. Halide–water potential energy functions
,”
J. Phys. Chem. B
120
,
1822
1832
(
2016
).
45.
S. K.
Reddy
 et al, “
On the accuracy of the MB-pol many-body potential for water: Interaction energies, vibrational frequencies, and classical thermodynamic and dynamical properties from clusters to liquid water and ice
,”
J. Chem. Phys.
145
,
194504
(
2016
).
46.
M.
Riera
,
A. W.
Götz
, and
F.
Paesani
, “
The i-TTM model for ab initio-based ion–water interaction potentials. II. Alkali metal ion–water potential energy functions
,”
Phys. Chem. Chem. Phys.
18
,
30334
30343
(
2016
).
47.
Y.
Wang
and
J. M.
Bowman
, “
Towards an ab initio flexible potential for water, and post-harmonic quantum vibrational analysis of water clusters
,”
Chem. Phys. Lett.
491
,
1
10
(
2010
).
48.
Y.
Wang
and
J. M.
Bowman
, “
Ab initio potential and dipole moment surfaces for water. II. Local-monomer calculations of the infrared spectra of water clusters
,”
J. Chem. Phys.
134
,
154510
(
2011
).
49.
Y.
Wang
,
X.
Huang
,
B. C.
Shepler
,
B. J.
Braams
, and
J. M.
Bowman
, “
Flexible, ab initio potential, and dipole moment surfaces for water. I. Tests and applications for clusters up to the 22-mer
,”
J. Chem. Phys.
134
,
094509
(
2011
).
50.
R.
Conte
,
C.
Qu
, and
J. M.
Bowman
, “
Permutationally invariant fitting of many-body, non-covalent interactions with application to three-body methane–water–water
,”
J. Chem. Theory Comput.
11
,
1631
1638
(
2015
).
51.
A. K.
Samanta
,
Y.
Wang
,
J. S.
Mancini
,
J. M.
Bowman
, and
H.
Reisler
, “
Energetics and predissociation dynamics of small water, HCl, and mixed HCl-water clusters
,”
Chem. Rev.
116
,
4913
4936
(
2016
).
52.
R.
Schwan
 et al, “
Observation of the low-frequency spectrum of the water trimer as a sensitive test of the water-trimer potential and the dipole-moment surface
,”
Angew. Chem., Int. Ed.
59
,
11399
11407
(
2020
).
53.
A.
Nandi
 et al, “
A CCSD(T)-based 4-body potential for water
,”
J. Phys. Chem. Lett.
12
,
10318
10324
(
2021
).
54.
S. R.
Gadre
,
R. N.
Shirsat
, and
A. C.
Limaye
, “
Molecular tailoring approach for simulation of electrostatic properties
,”
J. Phys. Chem.
98
,
9165
9169
(
1994
).
55.
S. D.
Yeole
and
S. R.
Gadre
, “
Molecular cluster building algorithm: Electrostatic guidelines and molecular tailoring approach
,”
J. Chem. Phys.
134
,
084111
(
2011
).
56.
N.
Sahu
,
S. D.
Yeole
, and
S. R.
Gadre
, “
Appraisal of molecular tailoring approach for large clusters
,”
J. Chem. Phys.
138
,
104101
(
2013
).
57.
N.
Sahu
and
S. R.
Gadre
, “
Molecular tailoring approach: A route for ab initio treatment of large clusters
,”
Acc. Chem. Res.
47
,
2739
2747
(
2014
).
58.
M. M.
Deshmukh
and
S. R.
Gadre
, “
Molecular tailoring approach for the estimation of intramolecular hydrogen bond energy
,”
Molecules
26
,
2928
(
2021
).
59.
N.
Sahu
 et al, “
Low energy isomers of (H2O)25 from a hierarchical method based on Monte Carlo temperature basin paving and molecular tailoring approaches benchmarked by MP2 calculations
,”
J. Chem. Phys.
141
,
164304
(
2014
).
60.
J.
Liu
and
X.
He
, “
Fragment-based quantum mechanical approach to biomolecules, molecular clusters, molecular crystals and liquids
,”
Phys. Chem. Chem. Phys.
22
,
12341
12367
(
2020
).
61.
Z.
Wang
,
Y.
Han
,
J.
Li
, and
X.
He
, “
Combining the fragmentation approach and neural network potential energy surfaces of fragments for accurate calculation of protein energy
,”
J. Phys. Chem. B
124
,
3027
3035
(
2020
).
62.
C.
Shen
,
X.
Jin
,
W. J.
Glover
, and
X.
He
, “
Accurate prediction of absorption spectral shifts of proteorhodopsin using a fragment-based quantum mechanical method
,”
Molecules
26
,
4486
(
2021
).
63.
S.
Hirata
 et al, “
Fast electron correlation methods for molecular clusters in the ground and excited states
,”
Mol. Phys.
103
,
2255
2265
(
2005
).
64.
X.
He
,
O.
Sode
,
S. S.
Xantheas
, and
S.
Hirata
, “
Second-order many-body perturbation study of ice Ih
,”
J. Chem. Phys.
137
,
204505
(
2012
).
65.
J. S.
Boschen
,
D.
Theis
,
K.
Ruedenberg
, and
T. L.
Windus
, “
Accurate ab initio potential energy curves and spectroscopic properties of the four lowest singlet states of C2
,”
Theor. Chem. Acc.
133
,
1425
(
2013
).
66.
J. S.
Boschen
,
D.
Theis
,
K.
Ruedenberg
, and
T. L.
Windus
, “
Correlation energy extrapolation by many-body expansion
,”
J. Phys. Chem. A
121
,
836
844
(
2017
).
67.
L.
Bytautas
and
K.
Ruedenberg
, “
Correlation energy extrapolation by intrinsic scaling. III. Compact wave functions
,”
J. Chem. Phys.
121
,
10852
10862
(
2004
).
68.
L.
Bytautas
and
K.
Ruedenberg
, “
Correlation energy extrapolation by intrinsic scaling. I. Method and application to the neon atom
,”
J. Chem. Phys.
121
,
10905
10918
(
2004
).
69.
L.
Bytautas
and
K.
Ruedenberg
, “
Correlation energy extrapolation by intrinsic scaling. II. The water and the nitrogen molecule
,”
J. Chem. Phys.
121
,
10919
10934
(
2004
).
70.
L.
Bytautas
and
K.
Ruedenberg
, “
Correlation energy extrapolation by intrinsic scaling. IV. Accurate binding energies of the homonuclear diatomic molecules carbon, nitrogen, oxygen and fluorine
,”
J. Chem. Phys.
122
,
154110
(
2005
).
71.
L.
Bytautas
and
K.
Ruedenberg
, “
Correlation energy extrapolation by intrinsic scaling. V. Electronic energy, atomization energy, and enthalpy of formation of water
,”
J. Chem. Phys.
124
,
174304
(
2006
).
72.
L.
Bytautas
and
K.
Ruedenberg
, “
Accurate ab initio potential energy curve of O2. I. Nonrelativistic full configuration interaction valence correlation by the correlation energy extrapolation by intrinsic scaling method
,”
J. Chem. Phys.
132
,
074109
(
2010
).
73.
H.
Stoll
, “
Correlation-energy of diamond
,”
Phys. Rev. B
46
,
6700
6704
(
1992
).
74.
H.
Stoll
,
B.
Paulus
, and
P.
Fulde
, “
On the accuracy of correlation-energy expansions in terms of local increments
,”
J. Chem. Phys.
123
,
144108
(
2005
).
75.
B.
Paulus
, “
The method of increments—A wavefunction-based ab initio correlation method for solids
,”
Phys. Rep.
428
,
1
52
(
2006
).
76.
L. T.
Xu
,
J. V. K.
Thompson
, and
T. H.
Dunning
, “
Spin-coupled generalized valence bond description of group 14 species: The carbon, silicon and germanium hydrides, XHn (n = 1–4)
,”
J. Phys. Chem. A
123
,
2401
2419
(
2019
).
77.
G. L.
Miessler
and
D. A.
Tarr
,
Inorganic Chemistry
, 3rd ed. (
Pearson Prentice Hall
,
1991
).
78.
W.
Kutzelnigg
, “
Chemical bonding in higher main group elements
,”
Angew. Chem., Int. Ed.
23
,
272
295
(
1984
).
79.
W.
Heitler
, “
Quantum chemistry: The early period
,”
Int. J. Quantum Chem.
1
,
13
36
(
1967
).
80.
N.
Takeda
 et al, “
Gas phase protonated nicotine is a mixture of pyridine- and pyrrolidine-protonated conformers: Implications for its native structure in the nicotinic acetylcholine receptor
,”
Phys. Chem. Chem. Phys.
24
,
5786
5793
(
2022
).
81.
A.
Kalemos
and
A.
Mavridis
, “
Bonding elucidation of the three common acids H2SO4, HNO3, and HClO4
,”
J. Phys. Chem. A
113
,
13972
13975
(
2009
).
82.
A.
Kalemos
,
I. R.
Ariyarathna
,
S. N.
Khan
,
E.
Miliordos
, and
A.
Mavridis
, “‘
Hypervalency’ and the chemical bond
,”
Comput. Theor. Chem.
1153
,
65
74
(
2019
).
83.
E.
Miliordos
and
S. S.
Xantheas
, “
Ground and excited states of the [Fe(H2O)6]2+ and [Fe(H2O)6]3+ clusters: Insight into the electronic structure of the [Fe(H2O)6]2+–[Fe(H2O)6]3+ complex
,”
J. Chem. Theory Comput.
11
,
1549
1563
(
2015
).
84.
E.
Miliordos
and
S. S.
Xantheas
, “
Elucidating the mechanism behind the stabilization of multi-charged metal cations in water: A case study of the electronic states of microhydrated Mg2+, Ca2+ and Al3+
,”
Phys. Chem. Chem. Phys.
16
,
6886
6892
(
2014
).
85.
G. D.
Purvis
and
R. J.
Bartlett
, “
A full coupled-cluster singles and doubles model: The inclusion of disconnected triples
,”
J. Chem. Phys.
76
,
1910
1918
(
1982
).
86.
M. J. O.
Deegan
and
P. J.
Knowles
, “
Perturbative corrections to account for triple excitations in closed and open-shell coupled-cluster theories
,”
Chem. Phys. Lett.
227
,
321
326
(
1994
).
87.
T. H.
Dunning
, “
Gaussian-basis sets for use in correlated molecular calculations .I. The atoms boron through neon and hydrogen
,”
J. Chem. Phys.
90
,
1007
1023
(
1989
).
88.
R. A.
Kendall
,
T. H.
Dunning
, and
R. J.
Harrison
, “
Electron-affinities of the first-row atoms revisited. Systematic basis-sets and wave-functions
,”
J. Chem. Phys.
96
,
6796
6806
(
1992
).
89.
D. E.
Woon
and
T. H.
Dunning
, “
Gaussian-basis sets for use in correlated molecular calculations. III. The atoms aluminum through argon
,”
J. Chem. Phys.
98
,
1358
1371
(
1993
).
90.
A. K.
Wilson
,
D. E.
Woon
,
K. A.
Peterson
, and
T. H.
Dunning
, “
Gaussian basis sets for use in correlated molecular calculations. IX. The atoms gallium through krypton
,”
J. Chem. Phys.
110
,
7667
7676
(
1999
).
91.
K. A.
Peterson
, “
Systematically convergent basis sets with relativistic pseudopotentials. I. Correlation consistent basis sets for the post-d group 13–15 elements
,”
J. Chem. Phys.
119
,
11099
11112
(
2003
).
92.
S. F.
Boys
and
F.
Bernardi
, “
Calculation of small molecular interactions by differences of separate total energies. Some procedures with reduced errors
,”
Mol. Phys.
19
,
553
566
(
1970
).
93.
B.
Liu
and
A. D.
McLean
, “
Accurate calculation of attractive interaction of two ground-state helium-atoms
,”
J. Chem. Phys.
59
,
4557
4558
(
1973
).
94.
S. S.
Xantheas
, “
On the importance of the fragment relaxation energy terms in the estimation of the basis set superposition error correction to the intermolecular interaction energy
,”
J. Chem. Phys.
104
,
8821
8824
(
1996
).
95.
I. N.
Levine
,
Quantum Chemistry
(
Prentice Hall
,
1991
).
96.
T. J.
Lee
and
P. R.
Taylor
, “
A diagnostic for determining the quality of single-reference electron correlation methods
,”
Int. J. Quantum Chem.
36
,
199
207
(
1989
).
97.
T. J.
Lee
,
J. E.
Rice
,
G. E.
Scuseria
, and
H. F.
Schaefer
, “
Theoretical investigations of molecules composed only of fluorine, oxygen and nitrogen: determination of the equilibrium structures of FOOF, (NO)2 and FNNF and the transition-state structure for FNNF cis-trans isomerization
,”
Theor. Chim. Acta
75
,
81
98
(
1989
).
98.
H.-J.
Werner
,
P. J.
Knowles
,
G.
Knizia
,
F. R.
Manby
, and
M.
Schütz
, “
Molpro: a general-purpose quantum chemistry program package
,”
Wiley Interdisciplinary Rev. Comput. Mol. Sci.
2
,
242
253
(
2012
).
99.
J. A.
Fournier
 et al, “
Snapshots of proton accommodation at a microscopic water surface: Understanding the vibrational spectral signatures of the charge defect in cryogenically cooled H+(H2O)n=2–28 clusters
,”
J. Phys. Chem. A
119
,
9425
9440
(
2015
).
100.
N.
Yang
 et al, “
Mapping the temperature-dependent and network site-specific onset of spectral diffusion at the surface of a water cluster cage
,”
Proc. Natl. Acad. Sci. U. S. A.
117
,
26047
26052
(
2020
).
101.
C. E.
Moore
,
National Standard Reference Data Series
(
US National Bureau of Standards, Office of Standard Reference Data
,
Washington DC
,
1972
), Vols. I–III.