Introduction

Noncovalent interactions bind molecules together in condensed phases and play important roles in chemical and biological processes. Rational design of catalysts, pharmaceuticals and supramolecular systems relies heavily on the fine-tuning of noncovalent interactions. Among the noncovalent interactions, the hydrogen-bond forms a special category due to its high strength and directional character, characteristics that are the reasons behind its frequent utilization in natural biological systems as well as in artificially designed systems. In recent years, the undisputed reign of hydrogen-bonding in rational design has become challenged by another form of noncovalent interactions. Halogen bonding shares many of the characteristics of hydrogen bonding, but it is even more highly directional in character and provides a complementary interaction to use in areas, such as drug design and supramolecular design [1, 2]. The natures of hydrogen and halogen bonding have been the focus of numerous studies, but the scientific community is still far from reaching consensus and particularly the importance of charge-transfer and covalent contributions remains under debate [3,4,5,6,7,8,9,10,11,12,13,14,15,16,17].

A large number of methods have been developed for analyzing intermolecular interactions in terms of well-defined and physically significant energy components [18]. This type of analysis is referred to as energy decomposition analysis (EDA). Most EDA methods are based on the supermolecular approach, i.e., variational quantum chemical calculations are performed on both the molecular complex and its isolated fragments, and the interaction energy is decomposed by the use of intermediate wavefunctions [19,20,21,22,23,24,25,26,27]. There are also methods in which the interaction between the fragments is treated as a perturbation to the non-interacting system [5, 22, 28, 29]. The former type of methods originates from the work of Kitaura and Morokuma in the mid-1970s [19], whereas the latter type is dominated by variants of symmetry-adapted perturbation theory (SAPT) with the first practical formulation appearing in the late 1970s [22]. Independent of the EDA type, the interaction energy is typically considered to be dominated by five energy components, i.e., exchange repulsion, electrostatics, polarization, dispersion and charge transfer. However, there are methods, such as those based on the quantum theory of atoms in molecules (QTAIM), that decompose the interaction energy into other contributions [30, 31].

Exchange or Pauli repulsion is a short-range repulsive term that stems from the overlap of the electron densities of the interacting fragments. In EDA methods based on wavefunction theory, this term is a result of the requirement of the supermolecular wavefunction to be antisymmetric with respect to exchange of electrons between the fragments, and an antisymmetry operator is introduced when constructing the first-order wavefunction from the non-interacting fragment wavefunctions. Exchange repulsion is considered a first-order term, but in SAPT it also appears at higher orders, where it is coupled to other interaction terms, such as polarization and dispersion [5, 28].

The electrostatic interaction energy is another first-order term, and it corresponds to the classical Coulomb interaction between the (static) charge distributions of the non-interacting fragments in the geometry of the molecular complex. In some EDA methods, exchange and electrostatics are not separated but considered as a single term [23, 25].

Polarization, or induction, is the lowering of the Coulombic interaction energy due to the polarization of each fragment by the charge distribution of the other. In orbital-based methods, such as Hartree–Fock or Kohn–Sham DFT, polarization results from the excitation of electrons from occupied to virtual orbitals within each fragment.

Dispersion, or London interaction, is often described as a Coulombic interaction arising from the instantaneous and mutual polarization of the charge distributions of the interacting molecules, e.g., induced dipole–induced dipole interactions. In wavefunction theory, dispersion appears as an electron correlation effect due to the contributions of configuration functions with concurrent excitations within the fragments. However, as demonstrated by Feynman, the dispersive force can be calculated from the electron density of the molecular complex and is the result of a polarization of the density of each interacting fragment [32].

Charge transfer is considered to originate from the transfer of electrons from occupied orbitals of one fragment to virtual orbitals of the other. Often it is referred to as a covalent contribution to the interaction, although as we will discuss later in this article, charge transfer and covalency are not necessarily equivalent. In many EDA methods, the charge-transfer term is not computed explicitly but obtained by subtracting out the other energy contributions from the intermolecular interaction energy. A problem with the differentiation between polarization and charge transfer is that it requires the use of an atom-centered basis set, and depends on the size and functional form of the basis set. In the limit of an infinite basis set, the charge-transfer term vanishes as the full density deformation is contained within the polarization term. Similarly, if the basis set is too small or the method is not able to account for polarization fully due to other reasons, charge transfer can be overestimated. Some EDA methods refrain from separating these two effects and combine them into a single term [5, 20, 21, 26].

In this context, it should be emphasized that the charge-transfer term of most EDA methods is different from the classical charge-transfer contribution computed by perturbation theory that is used in natural bond order (NBO) theory, as the latter is not based on the SCF-wavefunction and does not subtract out the electrostatic contribution from the charge-transfer term [33]. This typically results in a much larger magnitude of the charge-transfer energy in NBO compared to EDA analysis.

Summing up the different energy components from the EDA analysis gives the total interaction energy within the molecular complex, and it can be expressed as

$${\Delta E}_{Int}={\Delta E}_{Ex}+{\Delta E}_{Es}+{\Delta E}_{Pol}+{\Delta E}_{Disp}+{\Delta E}_{CT}$$
(1)

However, it should be noted that the different energy terms are generally computed from the geometry of the fragments in the molecular complex. Thus, they do not contain the energy required for deforming the fragments from their geometries in the isolated state. After adding this nuclear deformation energy term (\({\Delta E}_{Nuc}\)) to \({\Delta E}_{Int}\), we obtain the energy for forming the complex from the separated fragments, which we will refer to as the complexation energy (\({\Delta E}_{Cmpl}\)).

$${\Delta E}_{Cmpl}={\Delta E}_{Int}+{\Delta E}_{Nuc}$$
(2)

\({\Delta E}_{Nuc}\) is typically small for weak noncovalent interactions, but generally increases with the strength of the interaction and is often large in interactions that are viewed as having covalent character.

In a recent study, we used a point-charge approach to investigate the importance of electrostatics and polarization for halogen-bond interactions between Br and halogen-bond donors of the types RC \(\equiv\) CBr and R3CBr [15]. Although less elaborate, this point charge (PC) model has some advantages compared to the typical schemes used for EDA. The PC model gives an accurate description of electrostatics and polarization, and the two terms are rigorously defined and separated. In addition, the method is completely free of charge transfer, as the model has no electrons to transfer from the electron donor to the electron acceptor. Our study found that the \({\Delta E}_{Cmpl}\) is accurately reproduced by the PC interaction energy (\({\Delta E}_{Int}^{PC}\)), when the latter is scaled by a factor of 0.9 [15]. Interestingly, this shows that the variation in the \({\Delta E}_{Cmpl}\) over the whole data set is fully accounted for by only considering electrostatics and polarization. In the data set analyzed, \({\Delta E}_{Nuc}\) is generally small and varies between 0.1 and 1.3 kcal mol−1, and thus \({\Delta E}_{Int}\) also correlates well with the scaled \({\Delta E}_{Int}^{PC}\). Furthermore, we found that the polarization energy contributes strongly to \({\Delta E}_{Int}^{PC}\) and varies between -2.8 and -11.5 kcal mol−1. In some of the weakest complexes, the electrostatic interaction energy is positive and the interaction is driven by polarization.

In this study, we have used the PC model to analyze the interactions in a set of HB complexes between Br and halogen, oxygen and sulfur HB donors. We find that \({\Delta E}_{Int}\) is accurately reproduced by the scaled \({\Delta E}_{Int}^{PC}\) with a similar scaling factor as for the XB complexes studied previously. The interactions within the complexes have been analyzed and compared with XB-bonded complexes with the objective of understanding the differences and similarities between HB and XB interactions.

Theoretical background of the PC model

figure a

The scheme above shows the point charge (PC) model for the interaction of Br with a hydrogen-bond donor of the type XH or RXH, exemplified by the Br•••HF complex. The geometry of (R)XH is obtained from the quantum chemical structure optimization of the full complex, and the point charge (qBr) is placed at the position of the Br nucleus (RBr) in that complex. The total interaction energy (ΔEPC) between the point charge and (R)XH is obtained as the difference in the Born–Oppenheimer energy with and without the point charge included in the Hamiltonian.

$${\Delta E}_{Int}^{PC}=E\left[{\widehat{H}}_{q}\left({{\varvec{R}}}_{\mathrm{Cmpl}}\right)\right]-E\left[{\widehat{H}}_{0}\left({{\varvec{R}}}_{\mathrm{Cmpl}}\right)\right]$$
(3)

The \({\Delta E}_{Int}^{PC}\) can be separated into an electrostatic term and a polarization term:

$$\mathrm{}{\Delta E}_{Int}^{PC}={\Delta E}_{ES}^{PC}+{\Delta E}_{Pol}^{PC}$$
(4)

The electrostatic interaction energy (\({\Delta E}_{ES}^{PC}\)) is computed from the electrostatic potential [V(r)] of RXH at the position of the Br in the complex.

$${\Delta E}_{ES}^{PC}={q}_{{Br}^{-}}{V}_{0}\left({{\varvec{r}}}_{q}\right)={-V}_{0}\left({\mathbf{R}}_{{\mathrm{Br}}^{-}}\right)$$
(5)

where we have added the subscript 0 to V(r) to emphasize that V0(r) is computed from the unperturbed (static) charge distribution of the HB donor.

V(r) is rigorously defined

$$V\left({\varvec{r}}\right)=\sum_{A}\frac{{Z}_{A}}{\left|{{\varvec{R}}}_{A}-{\varvec{r}}\right|}-\int \frac{\rho\left({{\varvec{r}}}^{^{\prime}}\right)d{\varvec{r}}}{\left|{{\varvec{r}}}^{^{\prime}}-{\varvec{r}}\right|}$$
(6)

where ZA is the charge on nucleus A located at RA, and ρ(r) is the electron density function. V(r) is a physical observable and can be determined by experiment, but is more commonly computed using wavefunction theory or Kohn–Sham DFT. V(r) is a one-electron property and is, compared to, e.g., the electronic energy, relatively insensitive to the computational method or basis set. \({qV}_{0}\left({{\varvec{r}}}_{q}\right)\) corresponds to the interaction energy between a point charge q at rq and the static (unperturbed) charge distribution of the molecule. It is the exact interaction energy within the limit of an infinitesimal charge, i.e., when polarization is negligible. \(q{V}_{q}\left({{\varvec{r}}}_{q}\right)\) is the interaction energy between q and the perturbed charge distribution of the molecule, i.e., when the electron density is polarized due to q positioned at rq. However, \(q{V}_{q}\left({{\varvec{r}}}_{q}\right)\) is not the complete interaction energy as there is an energy cost of polarizing the electron density, which in the linear response approximation equals minus one half of the gain in interaction energy due to the polarization.

Within the PC model, \({\Delta E}_{Pol}^{PC}\) is simply obtained by

$${\Delta E}_{Pol}^{PC}={\Delta E}_{Int}^{PC}-{\Delta E}_{ES}^{PC}$$
(7)

Assuming linear response, the polarization energy contribution can instead be computed from V0(r) and Vq(r) as

$${\Delta E}_{LinPol}^{PC}=1/2q\left[{V}_{q}\left({{\varvec{r}}}_{q}\right)-{V}_{0}\left({{\varvec{r}}}_{q}\right)\right]$$
(8)

In our previous study on halogen-bonded complexes, we found that the polarization followed linear response closely; \({\Delta E}_{Pol}^{PC}\) and \({\Delta E}_{LinPol}^{PC}\) were nearly identical over the whole data set [15]. Furthermore, Eq. 8 shows that even a large polarization response only generates a smaller change in the polarization energy, e.g., a 100% increase in \(V\left({{\varvec{r}}}_{q}\right)\) upon polarization results in a polarization energy that is only 50% of the electrostatic interaction energy \(\left[\mathrm{if}\;V_q\left({\mathbf r}_q\right)=2V_0\left({\mathbf r}_q\right){\;\Rightarrow\;\Delta E}_{LinPol}=0.5E_{ES}\right]\).

Finally, we should discuss how well an anion such as Br can be represented by a negative point charge of elementary charge (-e = -1 au). Comparing their respective electrostatic potential functions, i.e., \({V}^{Br-}\left({\varvec{r}}\right)\) with \({V}^{-1}\left({\varvec{r}}\right)=-\left(1/r\right)\), we note that they both are spherically symmetric and are identical in magnitude as the radial distance from the nucleus (r) goes towards infinity (Fig. 1). Reducing the distance, the two functions follow each other closely but below 3 Å, \({V}^{Br-}\left(r\right)\) becomes significantly higher than \({V}^{-1}\left(r\right)\) due to an increasing amount of the electronic charge residing outside the limiting distance (r). At 2.5 Å, \({V}^{Br-}\left(r\right)\) is less negative than \({V}^{-1}\left(r\right)\) by 10% and around 1.7 Å, \({V}^{Br-}\left(r\right)\) reaches its minimum, where its value is 80% of \({V}^{-1}\left(r\right)\). It has been shown that this minimum coincides with the radius where exactly one electron resides outside the radius [34]. This so-called charge penetration effect has the result that Br in comparison with a negative point charge has a higher (less negative) interaction energy when interacting with a positive charge or a dipole at short distances [35, 36].

Fig. 1
figure 1

The electrostatic potential [V(r)] of Br (blue line) in kcal mol−1 computed at the M06-2X/6–311 + G(2df,2p) level compared to the electrostatic potential of a negative point charge of elementary charge (-e = -1 au) (red line)

On the basis of the above analysis, it is clear that the PC model overestimates the magnitude of the electrostatic interaction energy, as well the polarization contribution to the total interaction energy due to the polarization of the electron acceptor (hydrogen or halogen bond donor). On the other hand, it does not account for the polarization of the electron donor (here Br), the dispersion energy or any charge-transfer component to the interaction energy.

Methods and procedure

The hydrogen-bonded complexes were analyzed by full structure optimization at the M06-2X/6–311 + G(2df,2p) level. The M06-2X functional is highly accurate for main-group chemistry, including noncovalent interactions, and it explicitly accounts for dispersion interaction [37]. Di Labio et al. evaluated non-counterpoise-corrected DFT interaction energies against a revision of the HB23 data set, and found the M06-2X functional to perform well with a mean absolute deviation of only 0.21 kcal mol−1 [38]. The 6–311 + G(2df,2p) basis set is similar in size to the basis set used by Di Labio, and is sufficiently flexible and diffuse to reduce the basis set superposition error to acceptable levels. Additional computations have been performed at the same level of theory using a point charge to represent the Br anion. These computations used the optimized geometries of the halogen-bonded complexes, but a negative point charge (-1 au) was placed at the position of Br. In addition, the electrostatic potential was computed from the unperturbed density, and from the point-charge-perturbed density, of the halogen-bond donors in the geometries of the complexes. All computations were performed using the Gaussian 16 suite of software [39].

PC model and trends in interaction energies

The different energy components of the PC model as well as the KS-DFT computations for the entire set HB-bonded complexes are listed in Table 1. Because of the anionic HB acceptor (Br), these are strong complexes and \({\Delta E}_{Cmpl}\) varies between -10.7 kcal mol−1 and -31.3 kcal mol−1. Figure 2 shows that there is a good overall linear correlation between \({\Delta E}_{Cmpl}\) and \({\Delta E}_{Int}^{PC}\) with a correlation coefficient (R2) of 0.94 when the (Br-H-Br) complex is excluded from the correlation. The strength of the (Br-H-Br) complex is lower than expected from the relationship. The correlation improves very significantly if the nuclear deformation energy (\({\Delta E}_{Nuc}\)) is subtracted out of \({\Delta E}_{Cmpl}\) to obtain \({\Delta E}_{Int}\). \({\Delta E}_{Int}\) is well reproduced by scaling \({\Delta E}_{Int}^{PC}\) with 0.86 and the correlation coefficient is excellent (R2 =0.999). Again (Br-H-Br) does not follow the general relation relationship, but in contrast to \({\Delta E}_{Cmpl}\), \({\Delta E}_{Int}\) is lower than predicted by the general relationship. We will return to (Br-H-Br) later but for now we will leave it out of the discussion. We note in passing that a very similar relationship was obtained in our previous study of halogen-bonded complexes with \({\Delta E}_{Int}=0.92 {\Delta E}_{Int}^{PC}\), but in that study \({\Delta E}_{Nuc}\) was low (< 1.1 kcal mol−1) for all complexes and there was an excellent correlation with \({\Delta E}_{Cmpl}\), as well. It is not surprising that \({\Delta E}_{Int}^{PC}\) correlates better with \({\Delta E}_{Int}\) than with \({\Delta E}_{Cmpl}\) considering that \({\Delta E}_{Nuc}\) is not included in \({\Delta E}_{Int}^{PC}\). It should also be emphasized that it is common practice to focus on \({\Delta E}_{Int}\) rather than \({\Delta E}_{Cmpl}\) when analyzing noncovalent interactions, and especially in EDA studies. However, we note that \({\Delta E}_{Nuc}\) for many of the HB complexes is relatively large and over the whole data series \({\Delta E}_{Nuc}\) varies between 0.5 and 6.6 kcal mol−1. There are some obvious trends in \({\Delta E}_{Nuc}\). First, \({\Delta E}_{Nuc}\) is generally larger for sulfur compared to oxygen HB donors, and increases when going from the lighter to heavier halogen atoms for halogen HB donors. Secondly, for obvious reasons, \({\Delta E}_{Nuc}\) has a tendency to increase with interaction strength. We will return to \({\Delta E}_{Nuc}\) and discuss it significance later in this article.

Table 1 Equilibrium Br-H distance (in Å) and different energy components (in kcal mol−1) of the point charge interaction energy (\({\Delta E}_{Int}^{PC}\)) and the quantum chemical complexation energy (\({\Delta E}_{Cmpl}\)) for the ionic hydrogen bond complexes with Br as hydrogen-bond acceptor
Fig. 2
figure 2

The graph to the left shows the linear correlation between \({\Delta E}_{Cmpl}\) and \({\Delta E}_{Int}^{PC}\) for the entire data set of HB complexes with Br. The graph to the right shows the corresponding correlation between ∆EInt and \({\Delta E}_{Int}^{PC}\). The complex between HBr and Br is an outlier in both correlations

The excellent agreement between \({\Delta E}_{Int}\) and the scaled \({\Delta E}_{Int}^{PC}\) is remarkable considering that the data set consists of a diverse group of halogen, oxygen and sulfur HB, donors and considering the simplicity of the PC model, which only accounts for electrostatics and the polarization of the HB donor. There is only a fair correlation between the electrostatic interaction energy (\({\Delta E}_{ES}^{PC}\)) and \({\Delta E}_{Int}\) (R2 =0.89), but much better correlations with \({\Delta E}_{ES}^{PC}\) are obtained if groups of HB donors of similar type, e.g., alcohols, are considered separately. This shows that caution should be used when interpreting the character of intermolecular interactions based on the analysis of a group of congeneric molecules. The polarization energy is always smaller in magnitude than the electrostatic energy, but \({\Delta E}_{Pol}^{PC}\) varies considerably, both in value (− 2.8 to − 11.5 kcal mol−1) and in terms of its relative contribution (11- 42%) to \({\Delta E}_{Int}^{PC}\). It should be noted that the large importance of polarization is a consequence of the anionic HB acceptor; in complexes with neutral HB acceptors, we see much smaller relative contributions from polarization and the interactions are clearly dominated by electrostatics [40]. There are similar but more distinct trends in the variation of \({\Delta E}_{Pol}^{PC}\) compared to the variation of \({\Delta E}_{Nuc}\). \({\Delta E}_{Pol}^{PC}\) is lower for the sulfur HB donors when compared to the corresponding oxygen donors, as well as for HCl compared to the oxygen donors. It also decreases with the polarizability and the conjugation of the R group. Furthermore, \({\Delta E}_{Pol}^{PC}\) generally decreases with the strength of the interaction.

Energy decomposition along the potential-energy surface

In order to understand the relevance of the different energy components and the overall correlation between \({\Delta E}_{Int}\) and \({\Delta E}_{Int}^{PC}\), we will analyze the potential-energy surfaces of some of the complexes by studying the variation of the energy components with varying distance to the Br. We will begin with the HF complex, and then continue with HCl, which has a similar \({\Delta E}_{Cmpl}\) but \({\Delta E}_{Pol}^{PC}\) and \({\Delta E}_{Nuc}\) of larger magnitudes.

When analyzing the HF complex in Fig. 3, we note that at R distances greater than 4 Å \({\Delta E}_{Cmpl}\), \({\Delta E}_{Int}\), \({\Delta E}_{Int}^{PC}\) and \({\Delta E}_{ES}^{PC}\) follow each other closely, whereas \({\Delta E}_{Pol}^{PC}\) and \({\Delta E}_{Nuc}\) are negligible. This shows that the interaction is almost entirely electrostatic at these longer distances. At distances shorter than 4 Å, \({\Delta E}_{Pol}^{PC}\) starts to decrease, leading to an increasing separation between \({\Delta E}_{Int}^{PC}\) and \({\Delta E}_{ES}^{PC}\), but \({\Delta E}_{Pol}^{PC}\) remains above -1.0 kcal mol−1 down to R = 2.8 Å. \({\Delta E}_{Cmpl}\) and \({\Delta E}_{Int}\) also become slightly lower than \({\Delta E}_{Int}^{PC}\) below 4 Å, but the difference never exceeds 1.0 kcal mol−1. This difference can be attributed to polarization of Br and dispersion, interactions that are not included in the PC model. \({\Delta E}_{Nuc}\) remains negligible until below 2.5 Å and reaches a value of 1.0 kcal mol−1 at the potential energy minimum (R = 2.13 Å); Consequently, \({\Delta E}_{Int}\) is 1.0 kcal mol−1 lower than \({\Delta E}_{Cmpl}\) at the minimum. After the minimum, \({\Delta E}_{Nuc}\) increases steadily, and this is the main reason for the increase in \({\Delta E}_{Cmpl}\) at short distances. Just before the minimum at 2.2 Å, \({\Delta E}_{Int}^{PC}\) becomes lower than \({\Delta E}_{Int}\), mainly due to increasing charge penetration, and at the minimum, \({\Delta E}_{Int}^{PC}\) (-29.4 kcal mol−1) is 3.8 kcal mol−1 lower than \({\Delta E}_{Int}\). At this point, \({\Delta E}_{ES}^{PC}\) is -23.2 kcal mol−1 and \({\Delta E}_{Pol}^{PC}\) is -2.8 kcal mol−1, clearly showing that this interaction is dominated by electrostatics and has only a minor contribution from polarization.

Fig. 3
figure 3

The different energy components from the full quantum chemical model and the PC model as functions of the Br-H distance (R) for complexes of HF, HCl and HBr with Br. The vertical dotted line marks the Br···HX distance (R0) at the complex minimum. The corresponding structure is shown as an inset, and the X–H distance in italics is for the free hydrogen bond donor

When analyzing the HCl complex at distances beyond 4 Å, we see similar behavior as for HF, with \({\Delta E}_{Int}\) (and \({\Delta E}_{Int}^{PC}\)) completely dominated by the electrostatic term (\({\Delta E}_{ES}^{PC}\)). However, below 4 Å polarization starts to build up and \({\Delta E}_{Pol}^{PC}\) decreases with decreasing R more strongly than for HF. At the minimum (R = 1.93 Å), \({\Delta E}_{Pol}^{PC}\)= -7.8 kcal mol−1 and constitutes 30% of \({\Delta E}_{Int}^{PC}\); the corresponding contribution for HF is only 10%. However, when we analyze the behavior of the sum of electrostatic and polarization terms (\({\Delta E}_{Int}^{PC}\)) compared to \({\Delta E}_{Int}\) it is almost identical to that in HF; below 4 Å, \({\Delta E}_{Int}\) becomes slightly more negative than \({\Delta E}_{Int}\) (maximum difference 1.0 kcal mol−1), and below 2.2 Å, the curves cross due to increasing charge penetration. \({\Delta E}_{Nuc}\) varies in similar manner with distance as for HF, but its magnitude is much larger at all distances, and at the complex minimum it reaches a value of 5.7 kcal mol−1. \({\Delta E}_{Nuc}\) alone is the reason for the minimum in \({\Delta E}_{Cmpl}\), as the gradient of \({\Delta E}_{Int}\) is everywhere positive in the range of distances investigated. In other words, the \({\Delta E}_{Int}\) curve is attractive at all distances. The large value of \({\Delta E}_{Nuc}\) is a consequence of that the H-Cl bond length increases from 1.28 Å in the free HCl molecule to 1.42 Å in the complex. Comparing the HCl complex with the HF complex, the main differences are lower \({\Delta E}_{Pol}^{PC}\) and higher \({\Delta E}_{Nuc}\) in the former.

$$\mathrm{Br}^\circleddash\;\mathrm H-\mathrm{Br}\leftrightarrow\mathrm{Br}-\mathrm H\;\mathrm{Br}^\circleddash$$

As already mentioned, (Br-H-Br) is the only complex that does not follow the scaling relationship between \({\Delta E}_{Int}\) and \({\Delta E}_{Int}^{PC}\). This is not surprising, considering that it is very far from a typical hydrogen-bonded complex, with the Lewis structure representation consisting of two equivalent resonance structures, each with a -1 charge on opposite bromines. The optimized structure with two equal bond lengths of 1.715 Å (compared to 1.421 Å in HBr) confirms the Lewis structure picture of approximately a 0.5 covalent bond order for each H-Br bond and the partial electron transfer from Br to the other Br upon complex formation. In order to better understand the potential effect of charge transfer and change in covalency, we have analyzed the potential-energy surface in a similar manner as for the HF and HCl complexes. Going from HCl to HBr follows a very similar trend as going from HF to HCl in terms of the evolution of the different energy components with decreasing R distance. In particular, there is a consistent increase in magnitude of \({\Delta E}_{Nuc}\) and \({\Delta E}_{Pol}^{PC}\) going from HF via HCl to HBr. However, it should be noted that for HBr, \({\Delta E}_{Nuc}\) is very high, 16.9 kcal mol−1, more than 2.5 times higher than for any of the other hydrogen-bonded complexes in Table 1. In addition, HBr stands out from HF and HCl in that \({\Delta E}_{Int}^{PC}\) does not become lower than \({\Delta E}_{Int}\) at distances below 2.2 Å, but instead the two energies are nearly identical at the shorter distances. This could be interpreted as an additional energy contribution from charge transfer to \({\Delta E}_{Int}\) that is building up at the shorter distances, a contribution that is missing from \({\Delta E}_{Int}^{PC}\). Alternatively, it could be viewed as that the repulsive contributions to \({\Delta E}_{Int}\) are weakened compared to HCl, following the same trend as going from HF to HCl.

We have also analyzed the potential-energy surfaces for the Br complexes with CF3OH and CF3SH. First, it can be noted that these surfaces and the variations of the different energy components with R are very similar in appearance as for the surfaces of HF and HCl. This clearly indicates that the character of the hydrogen bonding in oxygen and sulfur HB donors is not fundamentally different from that of the halogen HB donors. Analyzing CF3OH first, we find that it forms a stronger complex (\({\Delta E}_{Cmpl}\)= -25.5 kcal mol−1) than HF and HCl, and the interaction is dominated by electrostatics. At the minimum, \({\Delta E}_{Pol}^{PC}\)= -7.8 kcal mol−1 and constitutes 15% of \({\Delta E}_{Int}^{PC}\). This can be compared to 10% and 30%, respectively in the HF and HCl complexes. \({\Delta E}_{Nuc}\) is 3.6 kcal mol−1, which also is intermediate between HF and HCl. Thus, it is clear that the higher strength of the CF3OH complex is an effect of a much stronger electrostatic interaction.

Turning to CF3SH, it forms a weaker complex than not only CF3OH but also HF and HCl. However, \({\Delta E}_{Pol}^{PC}\) (-8.5 kcal mol−1) is larger in magnitude than for CF3OH and HF, and the relative contribution, 41% of \({\Delta E}_{Int}^{PC}\), is even larger than for HCl. Thus, the weaker interaction of CF3SH compared to the other is the effect of a weaker electrostatic interaction. The value of \({\Delta E}_{Nuc}\) at the minimum is nearly identical to that in CF3OH, but it plays a larger relative role for \({\Delta E}_{Cmpl}\) in CF3SH compared to CF3OH.

Polarization and charge transfer

The excellent agreement between the supermolecular interaction energy and the scaled point-charge interaction energy shows that the variation of the interaction energy within the data set is fully reproduced by considering only electrostatics and polarization (remember that the PC description is completely free of charge transfer, as there are no electrons that can be transferred to the HB donor). This result may seem surprising considering the character of these interactions, and particularly in some complexes, such as Br···HCl, a significant charge-transfer contribution to the interaction energy could have been anticipated. There is no doubt that any EDA method that differentiates between charge transfer and polarization will indicate a charge-transfer component to the Br···HCl complex, although the actual size of that component will depend upon the choice of method and basis set.

The results of the PC model could be used to argue that charge transfer plays a very minor role for the strength of these complexes. However, we rather view it as a manifestation of that these interaction terms are difficult to separate and that any division between them is arbitrary. In the PC model, the interactions must be described as electrostatics and polarization, because the model does not allow for charge transfer. On the other hand, some EDA methods will find a very significant contribution from electron transfer between the fragments as a consequence of the functional form of the atom-centered basis set. In this context, one could ask whether the failure of the PC model to describe (Br-H-Br) should be taken as evidence that a separate charge-transfer term is needed to reproduce the interaction energy of this complex. It should first be remembered that the PC model does not allow for any polarization of Br. In addition, the polarization of the H-Br unit is limited by the flexibility of the basis set used in the calculation. Still, the question remains whether a model that allows for full polarization of both Br and H-Br in the geometry of the (Br-H-Br) complex would reproduce the charge distribution of the (Br-H-Br) complex as well as its interaction energy. However, such a model would require an extremely large and diffuse basis set for each fragment, which would make it impossible to distinguish polarization from charge transfer. In this context, we like to emphasize that the charge-transfer character of an interaction should not be equated with the covalent character of an individual bond. There is no doubt that the two Br-H bonds in (Br-H-Br) have an equally strong covalent character, and this results from one covalent bond being weakened and another partly covalent bond being formed upon the interaction between Br and H-Br. In the same manner, we would argue that the Br-H bond in the Br···HCl complex has a partial covalent character, even though we find that the interaction energy is equally well described by the scaled PC-energy as interactions with a much smaller covalent character.

The nuclear deformation energy as an indicator of charge transfer

Another energy term that is interesting to analyze is the nuclear deformation energy \({\Delta E}_{Nuc}\). It plays a very significant role for the potential-energy surfaces presented in Figs. 4 and 5, and with the exception of the HF complex, the \({\Delta E}_{Nuc}\) contribution is solely responsible for the minimum in the \({\Delta E}_{Cmpl}\) and the repulsive character of \({\Delta E}_{Cmpl}\) at shorter distances. This is different from the typical behavior of noncovalent interactions, where \({\Delta E}_{Nuc}\) generally is of minor importance at the minimum. It should be remembered that the driving force for the nuclear deformation is to lower the total energy, and the increase in \({\Delta E}_{Nuc}\) at short distances is compensated by a larger decrease in \({\Delta E}_{Int}\). Thus, it may be more appropriate to use the term nuclear relaxation rather than nuclear deformation. The PC model takes the nuclear relaxation into account as the interaction energy is calculated using the geometry of the supramolecular complex. This is the same approach that is commonly used in supramolecular EDA methods as well as in SAPT. However, the PC model fails to reproduce the nuclear relaxation in the sense that optimizing the HB donor in the presence of the point charge will generate a much smaller deformation of the HB donor than the deformation that is induced by the presence of Br. It could be argued that the nuclear relaxation of the complex results from an electron donation from Br into virtual orbitals of the HB donor, followed by a rehybridization of the occupied orbitals. Such a behavior would also imply that \({\Delta E}_{Nuc}\) could be used as an indicator of the size of the charge-transfer contribution to an intermolecular interaction. This interpretation finds partial support from the very large magnitude of \({\Delta E}_{Nuc}\) in the (Br-H-Br) complex, which clearly fulfils the criteria for a strong charge-transfer complex with a significant covalent contribution to the bonding. However, it cannot, for example, explain the large variation in \({\Delta E}_{Nuc}\) of the complexes with sulfur HB donors, as it seems unlikely that the much larger \({\Delta E}_{Nuc}\) of HSCF2CN (6.6 kcal mol−1) compared to HSPhNO2 (1.7 kcal mol−1) reflects a much larger charge-transfer contribution in the former complex. Furthermore, in our earlier study of XB bond complexes involving C–Br groups, \({\Delta E}_{Nuc}\) was generally much smaller than for the HB complexes of this study. The small \({\Delta E}_{Nuc}\) values seem to mainly reflect the character of the C − Br bond, and should not be taken to indicate a lower contribution from charge transfer in XB complexes with Br compared to HB complexes with Br. In line with this conclusion, we note that XB complexes involving Br and dihalogens, such as Br2 or BrF, have rather large \({\Delta E}_{Nuc}\) values.

Fig. 4
figure 4

The different energy components from the full quantum chemical model and the PC model as functions of the Br-H distance (R) for complexes of CF3OH and CF3SH. The vertical dotted line marks the Br···H distance (R0) of the lowest energy structure. The lowest energy structure is shown as an inset, and the X–H distance in italics refers to the free hydrogen bond donor

Fig. 5
figure 5

The graph shows the effect of charge penetration on the different energy components of the PC as functions of the Br···H distance (R) for the complexes with HF. Charge penetration corrected energies are marked with superscript V and compared to the PC energies and to \({\Delta E}_{Int}\). Similarly to \({\Delta E}_{Int}\), the corrected energy curves, \({\Delta E}_{ES}^{V}\) and \({\Delta E}_{Int}^{V}\), are repulsive at short distances

Effect of charge penetration on the potential-energy surface

In connection to the discussion about the importance of electrostatic and polarization versus charge transfer, it should be emphasized that the PC model overestimates the magnitudes of the former contributions at shorter distances due to the neglect of charge penetration. This is the main reason that \({\Delta E}_{Int}^{PC}\) is lower than \({\Delta E}_{Int}\) at distances below around 2.2 Å, as is found for all complexes in Figs. 3 and 4. However, other energy contributions, such as exchange repulsion, could also potentially contribute to \({\Delta E}_{Int}\) and increase its value at short distances. In the case of the HF complex, \({\Delta E}_{Int}\) has repulsive character with a negative gradient at distances below 2.0 Å; a behavior that is distinctly different from \({\Delta E}_{Int}^{PC}\), which has an increasingly positive gradient with decreasing distance. Thus, it is interesting to investigate how the neglect of charge penetration affects the electrostatic and polarization contributions of \({\Delta E}_{Int}^{PC}\) along potential-energy surface of the HF complex. The HF molecule is well suited for investigating these effects, as its charge distribution can be well approximated by an atomic monopole expansion, i.e., a positive charge on H and negative charge of the same magnitude on F. On the basis of this approximation, we can estimate the charge penetration corrected energies, \({\Delta E}_{ES}^{V}\) and \({\Delta E}_{Pol}^{V}\), by

$${\Delta E}_{ES}^{V}={k}_{HF}^{V}{\Delta E}_{ES}^{PC}$$
(9)

and

$${\Delta E}_{Pol}^{V}={k}_{HF}^{V}{\Delta E}_{Pol}^{PC}$$
(10)

where

$${k}_{HF}^{V}=\frac{{V}^{Br-}\left({\mathbf{r}}_{\mathrm{H}}\right)-{V}^{Br-}({\mathbf{r}}_{\mathrm{F}})}{{V}^{-1}\left({\mathbf{r}}_{\mathrm{H}}\right)-{V}^{-1}({\mathbf{r}}_{\mathrm{F}})}$$
(11)

In Eq. 11, \({V}^{Br-}\left({\mathbf{r}}_{\mathrm{H}}\right)\) is the electrostatic potential of Br at the position of the H nucleus in the complex and \({V}^{Br-}\left({\mathbf{r}}_{\mathrm{H}}\right)\) is the corresponding value at the position of the F nucleus. The \({V}^{-1}\left({\mathbf{r}}_{\mathrm{H}/F}\right)\) terms refer to the electrostatic potential of a point charge (-e = -1 au) placed at the position of Br. Equation 9 is exact within the approximation that the charge distribution of HF can be approximated by a monopole expansion. Equation 10 is based on the additional approximation that the exact charge distribution of Br polarizes the charge distribution of HF to the same extent as a point-charge representation. This additional approximation is likely to lead to an underestimation of \({\Delta E}_{Pol}^{V}\) and thus also the predicted \({\Delta E}_{Int}^{V}\) should be too low. However, the magnitude of the polarization component is smaller than that of the electrostatic component for the HF complex, and thus, the resulting error should have a relatively small impact on the size of \({\Delta E}_{Int}^{V}\). In Fig. 5, we compare \({\Delta E}_{ES}^{V}\), \({\Delta E}_{Pol}^{V}\) and \({\Delta E}_{Int}^{V}\) with the corresponding PC energies as well as \({\Delta E}_{Int}^{V}\) as functions of the intermolecular distance Br-H. First, we note that \({\Delta E}_{ES}^{V}\) and \({\Delta E}_{Int}^{V}\) already begin to deviate significantly from the corresponding PC values at about 3.5 Å, whereas for \({\Delta E}_{Pol}^{V}\) it is not until around 2.5 Å that the deviation becomes significant. \({\Delta E}_{Int}^{V}\) consistently lies above \({\Delta E}_{Int}\) down to a distance of about 1.8 Å, where the two curves approach each other. This energy difference can mainly be attributed to dispersion and polarization of Br, energy components that are not included in \({\Delta E}_{Int}^{V}\). At the complex minimum (R = 2.13 Å), \({\Delta E}_{Int}\) is 1.6 kcal mol−1 lower than \({\Delta E}_{Int}^{V}\), which is in line with the expected energy contribution from dispersion and polarization of Br. \({\Delta E}_{Int}\) also continues to decrease after the minimum in \({\Delta E}_{Cmpl}\) and reaches its minimum at 1.98 Å, after which the energy increases. The general increase in \({\Delta E}_{Int}\) at shorter distances for noncovalent interactions is attributed to exchange repulsion in most EDA methods. Therefore, it is interesting to note that \({\Delta E}_{ES}^{V}\) behaves similarly to \({\Delta E}_{Int}\) and has a minimum at a similar distance, despite that \({\Delta E}_{ES}^{V}\) lacks a contribution from exchange repulsion; the increase energy at shorter distances is instead a result of charge penetration. \({\Delta E}_{Int}^{V}\), which also lacks exchange repulsion, behaves similarly to \({\Delta E}_{Int}\) and \({\Delta E}_{ES}^{V}\), but reaches its minimum at a slightly shorter distance.

Difference between hydrogen and halogen bonding

Finally, we discuss the difference in character of HB versus XB bonding based on the analysis with the PC model. First of all, it should be noted that we have found very similar relationships between \({\Delta E}_{Int}\) and \({\Delta E}_{Int}^{PC}\) for the two types of bonding in the data sets that we have investigated, i.e., \({\Delta E}_{Int}\) ≈ 0.9 \({\Delta E}_{Int}^{PC}\). This agreement is encouraging, as it indicates that the high correlations are not fortuitous but rather that the PC model reflects the physics of the interactions. Secondly, as already noted, we find a much larger variation in \({\Delta E}_{Nuc}\) for the HB data set, where it varies between 0.5 and 6.6 kcal mol−1. In the XB data set, the variation is only between 0.1 and 1.3 kcal mol−1, which we attribute to the rigidity of the C–Br bond. Instead, we find that the most significant difference between the HB and XB complexes is in the relative contributions of polarization and electrostatics. In the HB data set, the contribution of \({\Delta E}_{Pol}^{PC}\) to \({\Delta E}_{Int}^{PC}\) varies between − 2.8 and − 11.5 kcal mol−1 in absolute terms and between 11 and 42% in relative terms, whereas for the XB data set the corresponding numbers are -4.0 to -9.9 kcal mol−1 and 29% to 880%. The reason for the extremely large relative contributions of \({\Delta E}_{Pol}^{PC}\) in some XB complexes is that these are weakly bonded XB complexes where the electrostatic contribution is positive and thus is counteracted by a strongly negative \({\Delta E}_{Pol}^{PC}\). Thus, overall, we find that polarization has much higher importance for XB compared to HB, and this is the main difference in the character of the two types of bonding. We have found no indications that a separate charge-transfer term is needed to describe the difference between halogen and hydrogen bonding.

Summary and conclusion

In this study, we have shown that the PC model describes the variation of the halogen bond interaction energy accurately within a diverse group of hydrogen-bond donors and their complexes with Br, as indicated by the relationship \({\Delta E}_{Int}\)=0.86 \({\Delta E}_{Int}^{PC}\) with R2 =0.999. The excellent correlation is remarkable considering that the PC model only accounts for electrostatics and polarization, and considering the large variation in chemical structure; the data set includes halogen, oxygen and sulfur HB donors, and the two latter feature both electron donating and accepting substituents. The only complex that does not follow the general correlation is (Br-H-Br). This is not surprising considering that this complex is classified as a strong charge-transfer complex with a large covalent character rather than an HB complex. However, the failure of the PC model to reproduce the interaction energy of this complex can partly be ascribed to incomplete description of polarization.

The different energy components for the PC model and the full quantum chemical model have been investigated along the potential-energy surface (PES) for the complexes of five HB donors. In all complexes, we find that the long-range interaction is dominated by electrostatics, and that \({\Delta E}_{Int}^{PC}\) approximately follows \({\Delta E}_{Int}\) down to around 2.2 Å. With the exception of the (Br-H-Br) complex, \({\Delta E}_{Int}^{PC}\) becomes lower than \({\Delta E}_{Int}\) below 2.2 Å, due to increasing charge penetration at shorter distances. In all complexes, except for the complex with HF, \({\Delta E}_{Int}\) has a positive gradient at all distances, and \({\Delta E}_{Nuc}\) defines the repulsive character of \({\Delta E}_{Cmpl}\) at short distances. An investigation of the charge-penetration effect on the electrostatic and polarization energies in the HF complex indicates that the repulsive component to \({\Delta E}_{Int}\) of this complex can to a great extent be attributed to charge penetration and that the contribution from exchange repulsion is relatively minor.

The results of the current study have been compared to our previous study on halogen bonding complexes between Br and XB bond donors of the types RC \(\equiv\) CBr and R3CBr [15]. In that study, we found a very similar relationship between \({\Delta E}_{Int}\) and \({\Delta E}_{Int}^{PC}\), i.e., \({\Delta E}_{Int}\)=0.92 \({\Delta E}_{Int}^{PC}\). The similar scaling factor of the two studies supports the conclusion that the PC model is able to to describe the physics behind the interactions. The main difference in the potential-energy surfaces of the two data sets is in the \({\Delta E}_{Nuc}\) term, and it generally is smaller and varies less in the XB bonding data set. Our interpretation is that the difference is a reflection of the character of the C–Br bond rather than a consequence of a larger charge-transfer contribution in the hydrogen-bond interactions. The main differences in the character of the bonding between the HB- and XB-bonded complexes are instead a larger influence of electrostatics and smaller contribution from polarization in the former type. However, we do not find it necessary to invoke charge transfer in order to understand the difference in character between ionic hydrogen and halogen bonding.

It may seem that the results of this and our previous study using the PC model are to some extent at odds with the conclusions typically drawn from supermolecular EDA and SAPT. However, there is a lack of studies where those type of methods have been used to analyze strong ionic hydrogen and halogen bonds of the types studied here. We hope that someone will pick up the baton, and employ such methods to help us increase the understanding of these important interactions on the border between noncovalent interactions and covalent bonding.