3.1. Relation of Electronic Levels of Ions to Adsorption and Binding Energy
Let us evaluate how the energy level of an ion's orbitals can affect binding into an adsorption complex with solvent molecules or an electrode surface.
To consider the simplest scenario of the interaction of an ion with a solvent molecule through a one-electron operator, let us represent their wave functions in terms of Gaussian orbitals for a solvent molecule (assuming that the solvent consists of sufficiently light atoms) ψ
solv=A
solv∙r
n∙exp(-α
solv∙r [
2]), where n=0 for the s-orbital and n=1 for the p-orbital and ψ
ion=A
ion∙r
n∙exp(-α
ion∙r [
2]) for the s-, p- or d-orbital for the ion (n=2 for the d-orbital). A
solv and A
ion are normalizing coefficients, α
solv=0.1 a.u.
- [
2] and α
ion are exponential coefficients, and all the wave functions are normalized
. The overlap of the wave functions is given by the expression
.
The eigenvalues of the orbitals of the solvent molecule and ion [
71,
72] will look as follows:
where
is the one-electron perturbation operator, which is the sum of the electron kinetic energy (Hkin) and electron-nuclear interaction (Hen); is the Laplace operator; is the distance between the electron density (product of the wave functions) at an arbitrary point x, y, z and a point charge qi with xi, yi, and zi coordinates of the i-atom; h is the Planck constant; me and e0 are the mass and charge of the electron; ε0 is the vacuum dielectric constant; i is the serial number of an atom in the solvent molecule.
The matrix element (V) for the orbitals is expressed through operator
:
The binding energy between the ψ
solv and ψ
ion orbitals is described by the expression: [
4]
where
is the solution to the Schrödinger equation for the ground state,
The energy level of the ψ
ion orbital can be changed by varying the α
ion parameter in the range of 0.001 – 0.5 a.u.
- [
2].
Now let us consider the adsorption of ions on the electrode using the Newnes-Andersen (NA) model. [
3,
73] According to the NA model, E
ads is determined by the sum of the contributions: [
4,
5,
74,
75]
The first term represents the hybridization energy due to the overlap between the adsorbent orbitals and the electronic states of the electrode, the second term is the orthogonalization energy (Pauli repulsion),
where f(ε) is the Fermi function, εa is the energy level of the adsorbent valence orbital, na is the filling of the adsorbent valence orbital determined by the expression , and is the degree of filling of the electrode electronic level.
For simplicity, we assume that V is proportional to S=-α∙|V|, where α is a coefficient, and thus arrive at .
The density of states on the adsorbent is given by the expression
where Δ and Λ are the chemisorption functions described by the expressions [
76]
, and Λ is the Hilbert transform of Δ:
, where
P denotes the principal part. For simplicity, instead of the sum of values for all the k-vectors, we represent Δ(ε) in the form
. [
5,
72,
77]
The adsorbent nature in the NA model, eq. (
6−
7), is given directly only by parameters ε
a and V [
2,
5] S is indirectly related to the matrix element. In case of the ions, as the data show (Supplementary Material of ref. [
72]) for O
2 and intermediates of oxygen reduction OH and OOH, the matrix element increases significantly due to the contribution of the electron-nuclear interaction. That is why we will evaluate the influence of, first of all, V [
2] and then ε
a on E
ads.
Another way to estimate adsorption is to represent the effective orbital of an electrode as an s-function distributed along the z-axis perpendicular to the electrode, as in the jellium model: [
78,
79]
ψel=Ael∙exp(-αel∙z[2]) (8)
Such approach has been used in numerous works [
71,
72,
80,
81,
82,
83,
84] to estimate the matrix element of the overlap between the electrode and redox system orbitals. In the following calculations, α
el was set at 0.01 a.u.
- [
2] and the normalization factor A
el in (
8) was calculated at the electrode surface (z=0) and at some distance from it using the expression
.
In this method, the ψ
el function substitutes for ψ
solv in expressions (
1−
3). After that, all further calculations are made using equations (
4,
5), and the final expression for the adsorption energy can be rewritten as
where
,
,
,
and the ground state energy
.
The results of the calculations of binding of the orbitals of cations and anions with the solvent orbital using the model of eq. (
1−
5) give a set of graphs that fully or partially form two volcano-shaped dependencies (
Figure 1 A,B). The graphs show that the orbital energies for the anions are mostly lower than those for the cations. It means that the centers of the volcano-shaped dependencies on the energy scale are shifted to the left for the anions and to the right for the cations.
In the regions of decreasing or increasing Ebind values, the graphs are quite close to linear or exponential dependencies. In case of s/s (schematic notation: solvent molecule orbital/ion orbital), s/pz and cation orbital interactions, the maximum region becomes sloped, which is caused by a large overlap of the orbitals due to their shape (σ-bond formation), lowering the binding energy value. When the solvent s-orbitals are combined with the and orbitals of the ions, rather weak π-bonds are formed. In case of the solvent p-orbitals, similar trends are observed for px/pz and pz/ and, to a lesser extent, for pz/dxz and px/px. The steeper slope of pz/dxz and px/px compared to s/px, etc. is explained by the larger extent of the p-orbital compared to the s-orbital, which results in a strong overlap and a decrease in the Ebind value. The specific shape of pz/ is caused by the orientation of pz-orbital directly towards the point charge of the ion. As a result, the contribution of the Hne operator makes the bond stronger.
The most interesting trend is observed when the p
z-orbital of the anion participates in the binding, which causes a split in the left side of the branch (
Figure 1 A,B). A special case are the results for the combination of s/s, and s/
orbitals at q=-1; p
x/p
x and p
x/p
y at q= -1 and p
x/p
x, p
x/p
y, and p
z/d
xy at q= +1, for which stable solutions are obtained only for one of the branches, the shape of which is close to a linear function, and in some cases (s/
) to an exponential one.
The dependence of E
ads(ε
a) according to the NA model, eq. (
6,
7), is almost linear and does not depend significantly on coefficient α (
Figure 1C), while in case of E
ads(V [
2]) an exponential dependence is observed (
Figure 1D). In the region of moderately high values V [
2]> 0.5 eV [
2], the dependence of E
ads(V [
2]) is almost linear. Notably, the approximation of specific results of DFT calculations of DOS adsorbents, for example, [
74,
75] shows that each bond type or band involved in the adsorption can have its own set of parameters ε
a and ε
c (band center). Therefore, for sufficiently high adsorption energies and, respectively, high V [
2], the shape of the E
ads(V [
2]) or E
ads(ε
a) dependencies may become more complicated due to the contribution of a larger number of electrode bands and adsorbent orbitals involved in the adsorption process.
The calculations for the ions carried out by eq. (
9) for the
,
,
orbitals lead to two volcano-shaped dependencies (
Figure 1E). Their side branches are close to linear dependencies and the interaction of the orbitals with the electrode is rather weak in this case. In case of the
and
orbitals, a quickly decreasing exponential dependence is observed. For the cations involving these orbitals, the E
ads dependence represents a function that decreases as the ε
ion value becomes higher. In general, this result is very close to the data obtained by the first model by eq. (
1−
5).
Since the HOMO or LUMO orbitals of real ions represent a superposition of a number of MOs, the total contribution of such composite orbitals to the final adsorption energy or energy of binding can become more complicated. If we consider the interaction between the solvent molecules and the ions, the dependence of the binding energy on the ion energy level becomes volcano-shaped. When the p-orbitals of the ion are oriented along the z-axis (perpendicular to the MO of the solvent molecule), the left side of the volcano-shaped Ebind dependence splits into two branches. In addition to the splitting, a scenario of parallel dependencies caused by the simultaneous contribution of multiple MOs of the ion is also possible. Within the NA model, the adsorption dependencies can be linear or exponential. However, if we consider the adsorption on the electrode through the effective wave function by eq. (8, 9), we can expect the same trends as those obtained in the first model by eq. (1−5), i.e. formation of volcano-like dependencies or an exponential or linear decrease/increase in the adsorption energy.
3.2. Trends in Ion Adsorption on α-Al2O3 (0001), Graphene and Au (111)
Summarizing the data on the adsorption of ions (
Figure S1.1−
S1.12,
Table S2.3 in Supplementary Information) and their binding into complexes (
Figure S2.1−
S2.4,
Table S3.3), we can conclude that it is most convenient to evaluate the energy levels of the HOMO (ε
HOMO) and LUMO (ε
LUMO) orbitals (
Table S4.1,
S4.2). In some cases, second highest (HOMO-1) and second lowest (LUMO+1) molecular orbitals can also be used. In addition, there are descriptors that combine HOMO and LUMO with HOMO-1 and LUMO+1 orbitals, e.g. hyperhardness or hypersoftness. [
21] In case of strong correlations, combining parameters into descriptors may well make practical sense in some problems. Working directly with ε
HOMO and ε
LUMO levels allows us to describe the adsorption and distance from the electrode to the ion, avoiding additional calculations of descriptors made up of number of parameters. Therefore, we will mainly focus on the analysis of trends in adsorption and binding depending on one orbital only.
In case of Al
2O
3, quantum chemical calculations show that the value of the cation adsorption energy decreases exponentially (
Figure 2A) as the ε
HOMO value becomes lower and almost linearly when the ε
LUMO value goes down (
Figure 2B). The strongest binding to the surface is observed for alkali metal cations and H
3O
+ and the weakest binding for EMI
+. The distance from the electrode surface to the cation also follows the E
ads trends, where z
0 becomes lower as the ε
HOMO (
Figure 2A) and ε
LUMO (
Figure 2B) values decrease. Only H
3O
+ is slightly out of the general E
ads(ε
HOMO) trend, which may be caused by the formation of strong adsorption bonds between the hydrogen of H
3O
+ and the oxygen of the Al
2O
3 surface.
In case of z0(εLUMO), only PyMe+ does not follow the linear trend. Since the z0 distance is determined at the ion center of mass, the elongated shape of the PyMe molecule and, consequently, the orbitals may cause overestimation of the z0 value.
No correlation is observed between the anion adsorption and ε
HOMO (
Figure 2E), while in case of E
ads(ε
LUMO), there is a volcano-like dependence with a maximum at ε
LUMO = 0.245 a.u. (
Figure 2F). The strongest binding to the surface is observed for the RCOO
-, OH
-, and F
- systems, while that for the PF
6- anion is the weakest. OH
- and F
- are out of the E
ads(ε
LUMO) trend, with the adsorption energies for the above ions getting as high as -4.0 and -4.3 eV, respectively; the adsorption energy in these cases is close to that of a strong chemical bond with the Al atoms.
The shape of z
0(ε
HOMO) and z
0(ε
LUMO) becomes more complicated (
Figure 2G,H), with the dependencies splitting into two curves that are reasonably well approximated by linear and exponential functions. In case of z
0(ε
HOMO), the linear dependence is formed by PF
6-, BF
4- and F
-, while the exponential one is represented by the other anions (TFSI
-, TfO
-; HSO
4-, MeSO
3-; ClO
4-, NO
3-, OH
-; HCOO
-, AcO
-, EtCOO
- and Cl
-, Br
-). z
0(ε
LUMO) also splits into two dependencies, with the first linear one consisting of OH
-; HSO
4-, MeSO
3-; HCOO
-, AcO
-, and EtCOO
- and the second exponential one represented by the TFSI
-, TfO
-, BF
4-, PF
6-; ClO
4-, and NO
3- anions and halides. In all the cases, z
0 decreases smoothly as the HOMO and LUMO energy values increase.
As the analysis (Figure S1.13) shows, the distance from the ions to the electrode surface for a number of systems correlates with the adsorption energy. For the anions, there is no Eads(z0) correlation, while for the cations a linear dependence is observed (Figure S1.13 A, B).
It can be concluded that the adsorption of cations on aluminum oxide is the result of contributions of both the HOMO and the LUMO orbitals. The adsorption of the anions is determined by the LUMO, HOMO-1, and LUMO+1 levels (B, D in Figure S1.1, S1.2) and is independent of the HOMO one.
Let us continue our consideration of adsorption by studying a semiconductor as an example. As in the previous system, the alkali metal cations are strongly adsorbed on the graphene surface, the PyMe
+ cation is weakly bound to the surface. The E
ads value for the cations on graphene decreases linearly as the ε
HOMO value becomes lower (
Figure 3A) and in case of E
ads(ε
LUMO), the dependence (
Figure 3B) is a volcano-shaped plot with an energy maximum in the region where ε
LUMO ~ -0.10 a.u. At the minimum distance from the electrode are the compact alkali metal cations, whereas the farthest from the surface are the voluminous TPA
+ and TBA
+ ions. As the ε
HOMO value becomes higher, the z
0 distance from the electrode increases exponentially (
Figure 3C), and at an increase in the ε
LUMO value, z
0 grows linearly (
Figure 3D).
The Eads(εHOMO) dependence for the anions is almost linear; in case of Eads(εLUMO) two branches are observed, the first linear one is represented by the ClO4-, NO3-, OH-; RSO3-; TFSI-, TfO-, BF4-, PF6-; and RCOO- anions and the second one consisting of halides is close to an exponential dependence. Only i-PrCOO- is out of the trend. The presence of an i-Pr- voluminous group probably leads to a rather strong repulsion due to the van der Waals interactions, making the adsorption energy value for this ion positive. The lowest adsorption energy values are observed for BF4-, PF6- and NO3-, the highest energies are registered for some of the halides (Cl-, Br-), EtSO3- and i-PrCOO-.
z
0(ε
HOMO) for the anions splits into two linear dependencies, the first one represented by the BF
4-, PF
6-, F
- and OH
- ions, the second one – by the other ions; i-PrCOO
- and p-TsO
- are noticeably out of the trend. In case of p-TsO
-, such spread of values may also be caused by the total volume and the associated asymmetry of the system due to the Ph- group, leading to an overestimation of the z
0 distances relative to the trend. The i-PrCOO
- distance from the electrode is also associated with the voluminous i-Pr group. The z
0(ε
LUMO) plot, as the one for the ions on Al
2O
3 (
Figure 2H), is formed by a linear dependence consisting of the OH
-; HSO
4-, MeSO
3-, EtSO
3-, NH
2SO
3-, p-TsO
-; HCOO
-, AcO
-, EtCOO
-, PrCOO
-, and i-PrCOO
- ions and an exponential dependence consisting of TFSI
-, TfO
-, BF
4-, PF
6-; ClO
4-, NO
3- and halides.
As in case of Al
2O
3, the E
ads(z
0) dependence on graphene is observed only for the cations (
Figure S1.13 C, D) and has a volcano shape with a maximum at z
0 ~ 4 Å and E
ads ~ -0.5 eV. According to the above dependencies (
Figure 3) and the data of (
Figure S1.5 B, D), the adsorption of the cations is determined by the HOMO orbital (except HOMO-1), as well as LUMO and LUMO+1. In case of the anions (
Figure 3,
Figure S1.7 B, D), the HOMO, HOMO-1 and LUMO orbitals are involved in the adsorption.
The strongest binding to the gold surface is observed for hydroxonium, which is followed by alkali metal cations and EMPyr
+. PyMe
+, as in case of graphene, is weakly bound to the surface. For the Au (111) surface, the voluminous organic PyMe
+ and EMPyr
+ cations drop out of the exponential dependence of E
ads(ε
HOMO) and E
ads(ε
LUMO) (
Figure 4A,B). In addition, a significant shift towards negative energies is observed for H
3O
+. In case of the alkali metal, TMA
+ and EMI
+ cations, the dependencies are almost exponential, the E
ads value in both cases decreases with an increase in ε
HOMO and ε
LUMO.
z
0 grows exponentially with an increase in ε
HOMO and ε
LUMO (
Figure 4С,D). Quite far from the z
0(ε
HOMO) dependence are only TMA
+, EMPyr
+ and PyMe
+. The alkali metal cations, H
3O
+, and EMI
+ fit quite well into the exponential dependence (
Figure 4 С), and all the ions considered fit into the linear z
0(ε
LUMO) dependence (
Figure 4 D).
Among the anions, the strongest interaction with the surface is observed for halides, which are followed by OH
- and HCOO
-. The weakest interaction with the (111) Au surface is observed for the nitrate ion, followed by TfO
- and MeSO
3-. The anions form a characteristic volcano-shaped E
ads(ε
HOMO) dependence (
Figure 4E) with a maximum at ε
HOMO ~ -0.25 a.u., while no correlation is found for E
ads(ε
LUMO) (
Figure 4F). The z
0(ε
HOMO) dependence is represented by a linear relationship consisting of BF
4-, PF
6- and OH
- ions and an exponential relationship with the other systems (
Figure 4G). Two dependencies are observed for z
0(ε
LUMO) (
Figure 4H), with the first linear one formed by OH
-; HSO
4-, MeSO
3-, EtSO
3-, NH
2SO
3-, p-TsO
-; HCOO
-, AcO
-, EtCOO
-, PrCOO
-, and i-PrCOO
- ions and the second exponential one made up by TFSI
-, TfO
-, BF
4-, PF
6-; ClO
4-, NO
3- and halides.
Some degradation of the correlations on the metal and deviation of a number of cations (on average, cations have stronger binding to Au (111) than anions) can be caused by the increased complexity of the adsorption process of bulk systems on the metal. In addition to a bigger population of electronic states in the metal and, consequently, a bigger number of electrode wave functions involved in the adsorption, there are also local perturbations of the electron density in the vicinity of the ions. In particular, we can talk about image charges, [
79,
85,
86] which can make a significant contribution to the adsorption energy [
87] when the adsorbent approaches the metal electrode at a sufficiently close distance. [
85]
Unlike the previous model electrodes, on Au (111) the E
ads(z
0) dependence is observed for both the cations (an exponential one,
Figure S1.13 E, except for PyMe
+, EMPyr
+ and H
3O
+) and the anions (a linear one,
Figure S1.13 F).
To sum it up (
Figure 4,
Figure S1.9 B, D and
Figure S1.11 B, D), the adsorption of the cations, except for some systems, and the anions on Au (111) is formed by both HOMO, HOMO-1 and LUMO, LUMO-1. Such conclusion is expectable considering the high density of electronic levels in the metal involved in the adsorption process.
The proposed descriptors characterize the adsorption parameters of cations on Al
2O
3 and graphene quite well (R [
2]=0.64−0.86 for E
ads and R [
2]=0.87−0.96 for z
0). On Au (111), the correlations for the cations look slightly worse because some organic cations fall out of the dependencies. The correlation coefficients of adsorption and distances for the cations increase as follows: Au (111) < graphene < Al
2O
3 (0001). In case of the anions, the correlations for E
ads are somewhat worse R [
2]=0.35−0.55, but still acceptable for qualitative and semi-quantitative evaluations. And the correlations for the z
0 parameter are approximately at the same level R [
2]=0.64−0.95 as those for the cations.
Most of the dependencies on the orbital energy are exponential, in some cases they are close to the linear type, as the theoretical estimates suggest (
Figure 1). Linear dependencies are also reported for the adsorption of neutral molecules on metal oxides in refs. [
32,
33,
34] Volcano-shaped dependencies are observed for the anions on Al
2O
3 (
Figure 2F) and cations on graphene (
Figure 3B), which is also in agreement with the theory. The splitting of the dependencies for E
ads into two branches is observed only for the anions on graphene (
Figure 3F). There is a split in all the z
0 anion dependencies on ε
HOMO and ε
LUMO. One cause of the division of the anions into two groups may be the presence of several MOs with the same energy (
Table S6.1). This effect will be discussed in more detail in the next paragraph.
Some limitations on the applicability of εHOMO and εLUMO as adsorption descriptors may arise in case of voluminous organic ions whose interaction with the electrode is poorly described by the MO energy level due to steric perturbations. Another limiting factor may be strong chemisorption involving the formation of a chemical bond with the surface, comparable in energy to the interatomic bond in a solid. In general, the descriptors provided in the manuscript are best suited for characterizing semiconductors and dielectrics.
3.3. Ion Binding in Complexes with Water and Dimethyl Carbonate
The free energy of cation binding into an aqueous complex is described well by an exponential function of ε
HOMO (
Figure 5A), the TBA
+ cation falls out of the trend, and a polynomial dependence is observed for G
bind(ε
LUMO) (
Figure 5B). The distance from the center of mass of the cation and water is reasonably well described by the exponential r
0(ε
HOMO) and linear r
0(ε
LUMO) functions (
Figure 5C,D). Only the pyridinium derivatives (RPy
+) fall out of the r
0(ε
LUMO) trend; the distance between the centers of mass for RPy
+ according to quantum chemical calculations is almost 1 Å longer than the obtained trend. The cause of this discrepancy may be greater asymmetry of RPy
+ compared to the other systems. The lowest binding energies and shortest distances are registered for the alkali metal cations, while the highest values are observed for the organic cations.
For the anions, the dependence of G
bind on ε
HOMO is divided into linear and exponential curves (
Figure 5E), while there is no correlation between G
bind and ε
LUMO (
Figure 5F). The lower linear G
bind(ε
HOMO) correlation (
Figure 5E) is formed by BF
4-, PF
6-, F
- and OH
- ions, the upper one is made up by ClO
4-, NO
3-, DCA
-; TFSI
-, TfO
-; Cl
-, Br
-; RSO
3-; and RCOO
- ions. The ion-water distances (
Figure 5G, H) depend only on ε
LUMO, and no correlations are observed for ε
HOMO. The linear dependence of r
o(ε
LUMO) is represented by OH
-; HSO
4-, MeSO
3-, EtSO
3-, NH
2SO
3-, p-TsO
-; HCOO
-, AcO
-, EtCOO
-, PrCOO
-, and i-PrCOO
- ions, the exponential one – by TFSI
-, TfO
-, BF
4-, PF
6-; ClO
4-, NO
3-; F
-, Cl
-, and Br
- ions. Only DCA
- deviates slightly from the exponential trend. The lowest values of binding energies and distances are observed for F
- and OH
-, while the highest values are observed for the organic cations. The highest values of binding energies are observed for fluorine-containing anions TFSI
-, TfO
-, BF
4-, and PF
6-, as well as DCA
- and RSO
3-. The longest r
0 distances are found for BF
4-, PF
6- and DCA
-.
For the cations, G
bind is exponentially related to r
0 (
Figure S2.9 A), while for the anions there is no such correlation (
Figure S2.9 B).
It can be concluded that the HOMO, LUMO (
Figure 5A,B) and HOMO-1, LUMO+1 orbitals (
Figure S2.1. B, D) are actively involved in the cation binding to water. The anionic aqueous complex is only made up of HOMO and HOMO-1 orbitals (
Figure 5E,
Figure S2.3 C), but not the LUMO orbitals.
The dependence of the free energy of cation binding to DMC on ε
HOMO is approximated by an exponential function (
Figure 6A) and on ε
LUMO by a polynomial function with a maximum at -0.09 a.u. (
Figure 6B). The distances from the center of mass of the cation and DMC grow exponentially depending on ε
HOMO and inversely exponentially depending on ε
LUMO (
Figure 6C,D). Only PyMe
+ and EMI
+ drop out of the r
0(ε
LUMO) dependence, with the predicted r
0 values tending to be almost 1 Å lower than the calculated r
0 ones. The strongest binding to DMC is observed for the alkali metal cations, while the weakest interaction is registered for TEA
+ and EMPyr
+.
The plot of the anion binding energy vs ε
HOMO represents two linear dependencies (
Figure 6E), the first one consisting of PF
6-, BF
4- and F
-, the second one — of ClO
4-, NO
3-; MeSO
3-; TfO
-, TFSI
-, halides and RCOO
- ions. G
bind(ε
LUMO) is formed (
Figure 6F) by a linear relationship composed of MeSO
3- anions, halides and RCOO
- and an exponential one consisting of ClO
4-, NO
3-; PF
6-, BF
4-, TfO
-, and TFSI
-. Like G
bind(ε
HOMO), r
0(ε
HOMO) for the anions also represents two dependencies (
Figure 6G), while r
0(ε
LUMO) is a linear relationship (
Figure 6H). The strongest interaction with the solvent is observed for F
- and the weakest one — for the TFSI
- and PF
6- anions.
For the cations, the G
bind(r
0) data are approximated by an exponential function (
Figure S2.9 C) and for the anions — by a linear function (
Figure S2.9 D).
The HOMO, LUMO (
Figure 6A,B), HOMO-1, and LUMO+1 orbitals (
Figure S2.6B,D) are actively involved in the binding of cations to DMC. However, the formation of the anionic complex with DMC involves only the HOMO, LUMO (
Figure 6E,B), and HOMO-1 (
Figure S2.7 B) orbitals.
In addition to describing the adsorption on the electrode, the energy level of the HOMO and LUMO is an excellent descriptor of the energy of binding into a complex (R [
2]=0.70−0.95) and the distance between the solvent molecule and the ions (R [
2]=0.36−0.98). In almost all the cases, the correlation coefficients are much better for the complexes than for the electrodes. Reducing the number of orbitals involved in the binding makes the complex formation easier and closer to theoretical models (paragraph
1).
The splitting of the dependencies is mainly due to the orbital type of the ion involved in the complex formation. For example, the dependence of the energy of solvent binding to anions on the ε
HOMO parameter (
Figure 5E,
Figure 6E) is split into two curves: one consisting of OH
-, F
-, BF
4-, and PF
6- and the second one comprised of the other ions. The MO analysis shows (
Table S6.1) that the ions from the first dependence possess multiple HOMOs of the same energy. A similar division into two dependencies can be seen in z
0(ε
HOMO), for example, for the complexes with DMC (
Figure 6G) and for the adsorption of anions on Al
2O
3 (0001) (
Figure 2G), graphene (
Figure 3G), and Au (111) (
Figure 4G). The z
0(ε
LUMO) plots are also represented by two dependencies: a linear one consisting of OH
-; RSO
3-; and RCOO
- and an exponential one composed of ClO
4-, NO
3-; BF
4-, PF
6-, TfO
-, TFSI
-, and halides. There is a similar trend for the adsorption on electrodes (Figures 2−4H) and for complexation with water (
Figure 5H), while it is not observed for DMC (
Figure 6H). Theoretical evaluations can shed light on the nature of such dependencies by establishing the distance between the ion and the electrode or solvent molecule at which the total energy of the system would be minimized. Taking into account the available data, we can assume that the division is caused by the specific composition of the LUMO orbital and does not depend on the number of LUMO orbitals with the same energy. Taking into account the data on DMC, it is quite natural to conclude that the molecular orbitals of the solvent molecule influence the complex formation.
3.4. Correlations with Physicochemical and Electrochemical Parameters of Ions
Judging by the dependence for dipole moment [
88] and the more complicated one for polarization [
89] (since the perturbation of the system by the external field is calculated), the wave functions are explicitly included in all the expressions of these parameters. In several models, the wave functions are also included in some terms of the solvation energy expression. [
63] Notably, in addition to a number of summands responsible for the environment polarization, [
63,
64] cavity formation and other contributions, [
90] the solvation energy is determined by the first solvation shell, [
91,
92] which is formed by the solvent directly bound to the ion. As paragraph
3 shows, the energy of binding to water and the intermolecular distances are related to the HOMO and LUMO. The redox potential, in turn, contains the solvation energies of the reactants [
93,
94] in the original entry and, therefore, depends on the internal rearrangement of the bonds and the environment response to the charge. Such facts and some literature data [
35] suggest a possible correlation of the mentioned parameters with the orbital energy level. In general, the mentioned physicochemical parameters, as in works, [
26,
27] are likely to depend on a whole range of descriptors, but for simplicity of the analysis we will focus on the correlations with MO levels only..
According to the calculations (
Figure 7A), the dependence of the dipole moment for the cations on ε
HOMO is nearly sigmoidal, for compact one-atom ions (alkali metals) and symmetric ions (TMA
+) in the range ε
HOMO = -3 – -0.7 a.u. the dipole moment is equal to zero. Starting from ε
HOMO > -0.62 a.u. a sharp increase in the dipole moment value is observed, with the maximum values corresponding to the voluminous and asymmetric CP
+, CTA
+ and DMI
+ cations. The dependence of the dipole moment of anions on ε
LUMO (
Figure 7B) decreases exponentially in the range ε
LUMO = 0.08 − 0.25 a.u., the maximum dipole moment is observed for voluminous DS
-, followed by BuCOO
-, p-TsO
-. At ε
LUMO > 0.25 a.u. the dipole moment of the anions becomes close to zero (for halides and symmetric NO
3-, ClO
4-, etc.).
The polarization dependence on ε
HOMO for the cations (
Figure 7C) is split into two curves: an exponential one represented by H
3O
+, alkali metal cations, RMPyr
+, and R
4N
+, and a linear one represented by CTA
+, CP
+, RMI
+ (except PhMI
+, which is out of the trend), and RPy
+. The dependence of P on ε
LUMO, as the one for the dipole moment, decreases exponentially (
Figure 7D), with TFSI
- falling out of the trend. The maximum polarization values are obtained for the voluminous and heavy DS
- and p-TsO
- cations, the lowest ones for the light F
-, OH
- and BF
4- ions.
The value of the solvation energy for the cations decreases exponentially with an increase in the ε
HOMO or ε
LUMO values (
Figure 7E,F). The lowest solvation values are observed for Li
+, which is followed by the other alkali metals, with the maximum values obtained for the organic cations. The G
solv(ε
HOMO-1) dependence for the anions has a characteristic volcano shape with a maximum at ε
HOMO-1 = -0.42 a.u. (
Figure 7G). The left part of the dependence is formed by the halides, the right part – by the other ions. A significant deviation from the trend is observed for OH
-, the energy value of which is significantly underestimated. TFSI
-, the solvation energy of which is slightly overestimated, is out of the trend. The lowest values of G
solv are observed for F
-, OH
- and the highest values for TFSI
-.
A dependence of the oxidation potentials on ε
HOMO is only observed for the anions (
Figure 7H). E
SHE decreases exponentially as ε
HOMO goes up.
The present paragraph clearly shows that in addition to describing the adsorption of ions or their binding to the solvent, the HOMO and LUMO parameters are also well suited for characterizing a number of physicochemical and electrochemical parameters. The latter, first of all, include dipole moment and polarization, whose correlation coefficients of the approximations reach R [
2]=0.8−0.94 and the cation solvation energy R [
2]=0.95. Slightly worse are R [
2]=0.74−0.78, but there are clear dependencies on the solvation energy and oxidation potential of anions.
3.5. Practical Evaluations
In this paragraph, we will show how orbital values can be applied to solving a practical problem. For this purpose, in continuation of the study of aqueous electrolytes, [
28,
44,
45,
46,
47,
50,
51] let us consider the ability of cations and anions to displace water from the inner layer of the electric double layer (EDL). The point of this task is to displace the electroactive component of the electrolyte water as far away from the electrode surface as possible [
28,
44,
45,
46] and, as a result, to increase the electrochemical potential window and energy capacity of the batteries. Since such problem considers a potential electrolyte for a battery, we will use a Li
+ ion as the cation. The parameters of the anions will be calculated using the approximations obtained previously (
Table S2.3,
S3.3,
S5.3). We will use graphene as the electrode because it can serve as a simplified model system representing different carbon materials for supercapacitors [
95,
96] and batteries. [
97,
98] By varying the ε
HOMO and ε
LUMO parameters in the range characteristic of anions, we will find systems with the minimal water concentration in the inner layer of EDL.
Under equilibrium conditions, the reversible process of displacement of ions (i) by water molecules (w) in the inner layer from the surface into the bulk
is described by a multicomponent adsorption isotherm: [
99,
100,
101]
and parameters
where θ
i and θ
w are the degrees of surface coverage by ions and water molecules,
is the potential drop in the inner layer of the EDL, σ
0 is the surface charge density on the electrode,
is the surface charge density of adsorbed ions in the inner layer,
is the potential drop in the diffuse layer according to the Gui-Chapman model, [
102] ε
b is the dielectric permittivity of the solution in the bulk (since dilute aqueous solutions are considered, the value is approximated as 78 [
103]), c
0 is the concentration of the supporting electrolyte in mol/L, q
i is the i ion charge,
is the ratio of areas of the i ion and water molecule, r
i and r
w are the effective radius of the ion or water molecule. [
28]
The parameters included in coefficients b
j and c
j (
11,
12) have the following physical meaning: μ
j is the dipole moment perpendicular to the electrode surface;
is the dielectric constant of component j, calculated according to the model described in ref. [
104] (P
j is the polarizability);
is the area occupied by component j;
and
are the effective values of the dielectric constant and thickness of the inner layer.
The
ratio of the activity coefficients of the electrolyte components in the inner layer according to the mean field approximation [
99,
100] is generally determined by the expression
For a 1:1 electrolyte (A stands for cations, B designates anions) and a solvent, equation (
13) looks as follows:
If k≠m, the A
km coefficients between components k and m are written as
where z
CN is the effective coordination number in the two-dimensional lattice, E is the energy of the interaction between the electrolyte components, when k=i and m=w: E
iw is the energy of binding of the ion and the water molecule (taken equal to the free binding energy of the components in the following estimations), E
ww is the energy of binding of the water molecules into a dimer, E
ii is the energy of binding of the ions with each other.
For simplicity, Eii is represented as a Coulomb interaction of charges qi separated from each other by the distance of their own effective ionic radii: for two cations, for two anions, and for one cation and one anion.
β
i is the parameter which determines the difference between the chemical potentials of the reaction components on the surface and in the bulk.
To simplify the problem, let us write this parameter as a function of the adsorption energies of the components on the electrode
As work [
28] suggests, the effective radius of cations should be estimated taking into account the solvation water. Keeping that in mind, we used the following parameters as the initial data for the lithium cation with a correction for solvation water: E
ads = -1.27 eV, E
bind,w = -0.33 eV,
= 3.00 Å (including the solvation water (
Table S7.1)), z
0 = 3.45 Å, μ and P are equal to zero; the parameters for the water molecule are E
ads = -0.11 eV, r
w = 1.41 Å, z
0 = 3.29 Å, μ = 2.19 D, P =7.21 Å [
3], E
ww = -0.13 eV. We described the dipole moment and polarizability of all the anions by the expressions µ
D=0.55598+353.79894∙exp(-31.72148∙ε
LUMO) and P=194.4018∙exp(-ε
LUMO/0.03634) 12.32756∙ε
LUMO+7.63034. The effective coordination number was taken to be z
CN = 6.
In accordance with the specific features of the dependencies of adsorption of the anions and their interaction with water on ε
HOMO and ε
LUMO, we divided all the systems into three sets (
Figure 8A–C).
The first set of systems (
A) consists of F
-, BF
4- and PF
6-; the adsorption of these ions is described by the expressions: E
ads=1.93766∙ε
HOMO+0.27664 and z
0=1.54334-5.459∙ε
HOMO, the aqueous complex parameters – by the expressions G
bind=-3.43293∙ε
HOMO-1.47558 and r
0=6.13597-16.63311∙ε
LUMO. According to the obtained solutions (
Figure 8A) to equations (
10−18), the lowest water coverage degree (θ
w < 0.08) is observed for the BF
4- and PF
6- ions. The only practical limitation on using BF
4- and PF
6- as aqueous electrolytes is their slow hydrolysis, [
105,
106,
107,
108,
109] which is why they have not been mentioned as salts for aqueous systems.
Figure 8А does not show F
-, which lies much higher at ε
LUMO= 0.54 a.u., but this system also extends to the region of low values θ
w < 0.08. In contrast to BF
4- and PF
6-, the fluoride appears in studies [
97,
110] as a component of an aqueous electrolyte with a wide potential window.
The second set of systems (
B) is represented by ions whose adsorption is described in a way similar to the systems in (
Figure 8A), except for the parameter z
0=1.92102-7.83216∙ε
HOMO, and the aqueous complex parameters are expressed as G
bind=0.03092∙exp(-9.00895∙ε
HOMO)-0.46743 and r
0=6.13597-16.63311∙ε
LUMO. All the systems (
Figure 8B) lie in the region θ
w < 0.08. Almost all of the above systems are reported in the literature as aqueous electrolytes with a wide potential window: TFSI
-, [
44,
45,
46] TfO
-, [
111] ClO
4-, [
112,
113] and NO
3-. [
114,
115] The exceptions are Cl
- and Br
- halides, whose electroactivity is quite high, especially that of Br
-, due to the relatively low oxidation potential, which limits their use in high voltage electrolytes. Only KCl is employed in supercapacitors but its use is limited too. [
116,
117,
118]
The third set of systems (
C) is represented by OH
-, HSO
4-, CF
3COO
-, RSO
3-, and RCOO
- anions whose adsorption is described by expressions similar to those in (
Figure 8B), the parameters of interaction with water are described by the equations G
bind=0.03092∙exp(-9.00895∙ε
HOMO)-0.46743 and r
0=6.37246∙exp(-6.99633∙ε
LUMO)+2.3211. The lowest water coverage θ
w < 0.08 (
Figure 8C) is observed for the CF
3COO
- system as well as for HSO
4- and NH
2SO
3-. CF
3COO
- is mentioned in refs. [
119,
120] In general, the application of such system is limited by their cost and toxicity. HSO
4- in an aqueous medium is characterized by dissociation with a marked pH decrease, [
121] which significantly narrows the potential window by shifting the cathodic region in the acidic medium. NH
2SO
3- is electrochemically unstable as the experiments and calculations show. [
28] The MeSO
3-, EtSO
3-, and HCOO
- systems lie in the moderate water coverage degree region 0.08 < θ
w < 0.16, AcO
- lies at the boundary of the region θ
w ~ 0.16. OH
- itself is quite electroactive and shifts the potential window; however, in concentrated alkalis the potential window slightly increases to ~1.6 V, [
116,
117,
122,
123] which is higher than in pure water (1.23 V [
115]). Such systems are used in supercapacitors. [
116,
117,
123] In practice, sulfonate is more often used in its acid form as an additive to electrolytes in electroplating and hydrometallurgy, [
124] and as an anion in redox flow batteries. [
125] EtSO
3- has hardly ever been mentioned as an aqueous electrolyte in the literature, except for some references where it was mentioned as an electrolyte for zinc-ion batteries. [
126] HCOO
- as an electrolyte was mentioned in ref. [
127]; in general, alkaline salts of carboxylic acids are often reported as potential electrolytes. [
28,
127,
128,
129,
130]
The low values of the electrode water coverage in electrolytes are in most cases consistent with the electrochemical stability of the system, confirming that the coverage is one of the main parameters [
28] determining the potential window. In addition to the known systems, electrolytes with anions represented by sulfonic acid derivatives are potentially interesting as objects of further investigation. Of course, a comprehensive analysis of the ion suitability for electrolytes requires, in addition to the water coverage, evaluation of the electrochemical stability of the anion itself and its potential intermediates during electrochemical decomposition, as it was done in case of NH
2SO
3-. [
28]
The use of the εHOMO and εLUMO parameters makes it possible to quickly evaluate the water coverage of the electrode and select the optimal anions, as it has been shown by a particular example. Due to the dependencies of the other electrochemical parameters (solvation energy, redox potential, etc.) on MO energy levels, such analysis can be performed in a comprehensive way, taking into account all the necessary parameters of cations and anions at the same time.