Next Article in Journal
Targeting Protein–Protein Interfaces with Peptides: The Contribution of Chemical Combinatorial Peptide Library Approaches
Next Article in Special Issue
C- and N-Phosphorylated Enamines—An Avenue to Heterocycles: NMR Spectroscopy
Previous Article in Journal
Regeneration of Hair Cells from Endogenous Otic Progenitors in the Adult Mammalian Cochlea: Understanding Its Origins and Future Directions
Previous Article in Special Issue
Investigation of Dynamic Behavior of Confined Ionic Liquid [BMIM]+[TCM] in Silica Material SBA-15 Using NMR
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New pecJ-n (n = 1, 2) Basis Sets for Selenium Atom Purposed for the Calculations of NMR Spin–Spin Coupling Constants Involving Selenium

by
Yuriy Yu. Rusakov
and
Irina L. Rusakova
*
A. E. Favorsky Irkutsk Institute of Chemistry, Siberian Branch of the Russian Academy of Sciences, Favorsky St. 1, 664033 Irkutsk, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(9), 7841; https://doi.org/10.3390/ijms24097841
Submission received: 30 March 2023 / Revised: 19 April 2023 / Accepted: 24 April 2023 / Published: 25 April 2023
(This article belongs to the Special Issue NMR Spectroscopy in Materials Chemistry)

Abstract

:
We present new compact pecJ-n (n = 1, 2) basis sets for the selenium atom developed for the quantum–chemical calculations of NMR spin–spin coupling constants (SSCCs) involving selenium nuclei. These basis sets were obtained at the second order polarization propagator approximation with coupled cluster singles and doubles amplitudes (SOPPA(CCSD)) level with the property-energy consistent (PEC) method, which was introduced in our previous papers. The existing SSCC-oriented selenium basis sets are rather large in size, while the PEC method gives more compact basis sets that are capable of providing accuracy comparable to that reached using the property-oriented basis sets of larger sizes generated with a standard even-tempered technique. This is due to the fact that the PEC method is very different in its essence from the even-tempered approaches. It generates new exponents through the total optimization of angular spaces of trial basis sets with respect to the property under consideration and the total molecular energy. New basis sets were tested on the coupled cluster singles and doubles (CCSD) calculations of SSCCs involving selenium in the representative series of molecules, taking into account relativistic, solvent, and vibrational corrections. The comparison with the experiment showed that the accuracy of the results obtained with the pecJ-2 basis set is almost the same as that provided by a significantly larger basis set, aug-cc-pVTZ-J, while that achieved with a very compact pecJ-1 basis set is only slightly inferior to the accuracy provided by the former.

Graphical Abstract

1. Introduction

Selenium chemistry plays an important role in modern materials science and has many applications in modern industries, from photosensitive electronics to pharmaceuticals. For instance, selenium is known to be an essential element in biochemistry [1,2,3,4,5,6,7,8,9]; in particular, it is embedded in selenoproteins as the 21st amino acid selenocysteine [10]. Selenium compounds have also found their applications in photoelectronics [11,12], solar cells technology [13], the semiconductors industry [14], nanotechnology [15,16,17], the synthesis of channelized porous materials [18,19], and in many other fields. Today, any synthesis of novel selenium-containing compounds involves NMR experiments supplemented by high-quality quantum chemical modeling of their NMR spectra [20,21,22,23,24,25,26,27,28]. In this respect, a delicate approach is required for the calculation of the indirect nuclear spin–spin coupling constants (SSCCs) involving selenium [29,30,31,32] because the calculation of the SSCCs with heavy chalcogens from the fourth period and below, in general, represents a challenging task [33,34,35,36,37,38,39,40,41] due to the complexity of the electronic structure of such systems and the need to pay special attention to the relativistic effects [42,43,44,45]. Accurate computation of the selenium SSCCs covers a wide range of applicability, representing, in the first place, a powerful tool of the stereochemical studies of organoselenium compounds [46,47,48,49,50,51,52,53,54]. Indeed, it should be emphasized that Se-H and Se-C SSCCs manifest a pronounced stereochemical behavior basically governed by the orientational lone-pair effect of selenium [48,55]. For example, large and, in some cases, even dramatic differences between geminal and vicinal Se-H SSCCs involving diastereotopic protons allowed making unambiguous diastereotopic assignments of signals in the 77Se-1H HMBC spectra of the series of selenium-containing four-, five-, and six-membered heterocycles, including the derivatives of thiaselenetane, selenasilole, thiaselenole, thiaselenolane, and dihydrothiaselenine [54]. Another example of the usefulness of the accurately computed nJ(77Se,1H) SSCCs was presented in the work [49], where these were used for the correct assignment of the signals in the 1H-77Se HMBC spectra of Se-glycosides and related molecules, allowing the determination of their glycosidic conformation around the C–Se bond. Stereochemical studies of nine Z-2-(vinylsulfanyl)ethenylselanyl organyl sulfides have been carried out by means of NMR experimental measurements and high-quality computations of the 77Se-1H spin–spin coupling constants [50]. Unambiguous resonance assignments of diastereotopic CH2 protons in the anomeric side chain of nine alkyl- and aralkylselenoglycosides have been carried out on the basis of experimental CPMG-HSQMBC measurements and high-quality computations of geminal 77Se-1H spin–spin coupling constants involving diastereotopic pro-R and pro-S protons [51]. Such examples are numerous, inferring the importance of the accurate prediction of the SSCCs involving selenium using high-quality quantum chemical methods and properly chosen basis sets.
In this respect, many accurate quantum–chemical approaches to the calculations of SSCCs have been developed over the past few decades [56,57,58,59,60,61,62]. The most popular ones comprise the multi-configurational self-consistent field approach (MCSCF) [63,64], the density functional theory (DFT) [65,66,67,68,69], the coupled cluster approach (CC) [70,71,72,73,74,75], and the second-order polarization propagator approximation (SOPPA) [76,77] along with its modifications, the SOPPA(CC2) [78] and SOPPA(CCSD) [79,80], obtained using the single and double amplitudes from the coupled-cluster calculations with the CC2 [74] and CCSD [70,75] models, respectively.
Significantly, even if a high-quality ab initio approach has been used when calculating the SSCCs, it does not guarantee the correctness of the result. There are many factors affecting the final values, though the most important one is the quality of the basis set used. Regardless of what high-quality approach was used, a small and inflexible basis set will poorly describe the main contributions to the SSCCs and will spoil the whole calculous. In particular, it is well known that standard energy-optimized basis sets are not effective for the calculation of the second- and higher-order molecular properties, such as NMR SSCCs. This is because such basis sets do not usually provide an appropriate representation of the molecular orbitals in the desired areas and provide slow convergence towards the complete basis set (CBS) limit. As a consequence, to achieve values close to the CBS limit within a particular method, one is forced to resort to rather large nonspecialized basis sets, resulting in high computational demands. This problem is especially pronounced for the SSCCs involving nuclei of fourth and further periods of the periodic table of elements (PTE) because of the complexity of their atomic shells requiring a more flexible description. In this way, specialized J-oriented basis sets are under permanent development. Such basis sets usually reflect the peculiarities of the SSCCs, which follow from the simplest nonrelativistic Ramsey formulation [81]. Each of the four Ramsey contributions to SSCCs, namely, Fermi contact (FC), the spin dipolar (SD), the paramagnetic spin-orbit (PSO), and the diamagnetic spin-orbit (DSO), require basis set flexibility in different exponential regions [82,83]. The most proliferated case is when the FC term predominates over the other three contributions; thus, a majority of J-oriented basis sets were developed so as to improve the description of the FC term only. In that case, the ultimate goal of such basis sets was to enhance the description of the s-electron density at the nuclear positions by extending the standard energy-optimized basis sets with tight s-type functions with very high exponents [80]. This approach can be thought of as the most straightforward one because it is based on the consecutive augmentation of the important angular spaces using a recurrent ratio (geometrical progression, at most). In this respect, it should be mentioned that such an approach was used to obtain the now-popular aug-cc-pVTZ-J series of basis sets for different elements [80,84,85,86,87,88,89,90,91]; the Pople-style basis sets 6-31G-J and 6-311G-J [92]; the acvXz-J for Se, Te, and Sn [93,94]; and the av3z-J for Te [41]. The other approaches involve, in some way or another, the optimization procedure, namely the search for the values of additional exponents is carried out via the variational procedure. In this way, Jensen et al. and Benedikt et al. developed their famous basis sets (aug)pcJ-n (n = 0–4) [82,95,96] and ccJ-pVXZ (X = D, T, Q, 5) [97], which are suitable for the calculation of the SSCCs involving 1–3 and 1–2 row nuclei, respectively.
In this paper, we present new compact J-oriented basis sets for the selenium atom that have been obtained within the property-energy consistent (PEC) method, introduced in our previous studies [98,99,100]. The PEC method is based on the consistent optimization of all exponents using the Monte Carlo (MC) simulations [101,102,103], with respect to the property under consideration and the total molecular energy. It gives compact basis sets that are capable of providing accuracy comparable to that achieved with the other property-oriented basis sets of larger sizes generated with a standard even-tempered technique [104].
This work has been called for by the needs of compact and accurate basis sets for highly computationally demanding calculations of the spin–spin coupling constants involving selenium in large- and medium-sized selenium compounds. Only a few J-oriented basis sets for selenium exist at the moment, and all of them are rather large to be efficiently used in large-scale calculations. These are the acvXz-J (X = 2, 3, 4) [93], the aug-cc-pVTZ-J (the first version) [86], and the aug-cc-pVTZ-J (the second version) [87]. The aug-cc-pVTZ-J (the first version) basis set for selenium was obtained from the energy-optimized uncontracted Dunning’s basis set of triple-zeta quality, aug-cc-pVTZ(uc) (21s14p10d2f) [105], by the addition of two tight s-type functions and the subsequent application of the contraction scheme based on the molecular orbital coefficients of the selenium hydride, (23s14p10d2f) -> [12s9p6d2f]. The f- and g-angular spaces were not considered in the saturation procedure in the mentioned paper. The acvXz-J (X = 2, 3, 4) basis sets [93] were designed from the augmented core-valence Dyall’s basis sets dyall.acvXz (X = 2, 3, 4) [106,107,108] of double- (16s12p8d2f), triple- (24s17p11d5f1g), and quadruple-zeta (31s22p14d6f4g1h) quality by the consecutive even-tempered saturation of the s-, p-, d-, f-, and g-angular functional spaces. As a result, it was found that the double-zeta quality basis set dyall.acv2z is not complete in the f-space for the correct calculation of the selenium SSCCs; i.e., two f-functions are not enough. For example, the addition of only one diffuse f-function to the f-space of dyall.acv2z resulted in the increase of the 1J(77Se,1H) of MeSeH molecule by about 5 Hz. Thus, the proposed acvXz-J basis sets were profoundly expanded in the f-space as compared to the corresponding dyall.acvXz basis sets. Therefore, the acvXz-J basis sets have the following configurations: [19s12p8d4f|11s8p5d4f] for X = 2, [27s17p11d5f1g|14s10p6d5f1g] for X = 3, and [35s22p14d6f4g|18s12p7d6f4g] for X = 4. In this sense, the aug-cc-pVTZ-J for the selenium atom has also been revisited recently [87] because the former version of this basis set contained only two f-functions. Thus, the contemporary configuration of the aug-cc-pVTZ-J for the selenium atom is [26s16p12d5f|17s10p7d5f]. Although the acvXz-J and aug-cc-pVTZ-J (the second version) basis sets are complete enough in all important functional spaces to provide a flexible description of the selenium SSCCs, the problem is that these basis sets are very large. To be more precise, the number of functions for the uncontracted/contracted forms of these basis sets is as follows: 123/88 (acv2z-J), 177/118 (acv3z-J), 249/167 (acv4z-J), and 169/117 (aug-cc-pVTZ-J). On the other hand, we have recently proposed the PEC method [98] for generating property-oriented basis sets that reoptimizes all angular spaces and gives very efficient compact basis sets, providing more accurate results as compared to the property-oriented even-tempered basis sets of similar sizes. Thus, in this paper, we present new pecJ-n (n = 1, 2) basis sets for the calculation of selenium SSCCs that embody both desirable features: they are moderate in size and provide very accurate results.

2. Results and Discussion

2.1. Creation of pecJ-n (n = 1, 2) Basis Sets for Selenium

The PEC method [98] consists of the optimization of basis sets in relation to a certain molecular property, provided that the least possible total molecular energy is achieved. Exponents are randomly generated around the starting basis set via Monte Carlo simulations. The generated arrays are verified whether they give the property under interest within a desired diapason or not. In the end, only one set is selected—the one that provides the property value within the desired range and the lowest energy. For the detailed description of the PEC algorithm and its peculiarities as applied to the generation of the J-oriented basis sets, we refer the reader to our earlier works presenting the basics of the method [98,99]. Although, it should be mentioned that the PEC method is unique in the sense of dealing with two target functions (the property and the energy) simultaneously. The PEC optimization procedure of the exponents {ζi} for the property under consideration with respect to the “ideal” values under the energetic constrain represents a nonlinear problem with multiple solutions. In this case, the PEC optimization procedure can be thought of as approaching the isoline of the “ideal” property value that is formed by the intersection of the “ideal” property plane with the property surface representing the multi-argument function of the varying exponents f (ζ1, ζ2, …, ζn), both determined in the multidimensional exponential space. All points of this isoline give the same “ideal” value of property but different molecular energies. The PEC method is aimed at selecting the basis set that provides the lowest molecular energy among those which belong to the “ideal” isoline, since the lower the energy, the better the description of the molecular wave-function. The optimization problem in the multidimensional exponential space with multiple solutions can hardly be solved using standard procedures based on the directed search, like numerical Newton-like methods [109]. This is connected with the natural limitations of such techniques due to the fact that they are aimed at finding a single extremum of a property in the vicinity of the starting point. In the case of Newton-type optimization, one can imagine the process as a gradual ascending or descending to a certain single point. In that way, such algorithms lapse the other solutions, unless initiated many times from different starting guesses, and, even if they were started from various guesses, they would find only a limited number of the solutions. Based on this reasoning, one can conclude that a conventional directed Newton-like optimization is not fully suitable for the problem with multiple solutions; in this sense, the PEC method is unique and totally justified for generating the J-oriented basis sets.
In this work, all optimizations of exponents were carried out using the SOPPA(CCSD) method for SSCC calculations and the CCSD method for the energy calculations. We used two fitting molecules, SeH2 and Se=C, in which the FC contributions to 1J(77Se,1H) and 1J(77Se,13C) SSCCs were considered as the arguments of the target function. The target function that was minimized represents the mean absolute error of these JFC against their “ideal” values:
Δ ˜ = 1 2 i = 1 2 J ˜ F C , i J F C , i i d e a l min
This minimization was performed under the energetic constrain: n = 1 2 E ˜ n min , which guarantees that the least possible total molecular energy of two molecules is achieved. The energy tolerance threshold was set to 10−4 Hartree. The optimization that involves two fitting molecules provides more robustness of the generated basis sets towards the diversity of the electronic systems as compared to the basis sets obtained with only one fitting molecule.
The “ideal” values were evaluated at the SOPPA(CCSD) level using the extended dyall.aae4z basis set (dyall.aae4z+) on all atoms in both fitting molecules. The dyall.aae4z+ was obtained by adding 3, 2, and 1 tight s-functions to the dyall.aae4z basis set for hydrogen, carbon, and the selenium atom, respectively, in the even-tempered manner. Accordingly, the corresponding configurations of the dyall.aae4z+ basis set for these atoms are as follows: (15s4p3d2f), (21s11p6d4f2g), and (32s22p14d9f5g1h). In this respect, dyall.aae4z+ provides overwhelming flexibility in all angular spaces, including the most important tight s-region to gain the values of SSCCs that are very close to the CBS limit. Thus, J F C , 1 i d e a l = J FC ideal 1 Se 77 , H 1 = 74.65   Hz and J F C , 2 i d e a l = J FC ideal 1 Se 77 , C 13 = 154.25   Hz . It is interesting to note that the ideal values for the FC contributions to one-bond 77Se-1H and 77Se-13C SSCCs are of different signs. The sign of the JFC(A,B) can be deduced from the signs of the nuclear magnetogyric ratios of the coupled nuclei (γA, γB) and that of the reduced SSCC, KFC(A,B), in accordance with the following relationship: JFC(A,B) = (h/4π2)∙(γAγB)∙KFC(A,B). As the magnetogyric ratios of the isotopes 1H, 13C, and 77Se are of the same positive sign, the difference in the sign of the 1JFC(77Se,1H) and 1JFC(77Se,13C) stems from different signs of the FC contributions to the corresponding reduced SSCCs. In this respect, the latter is totally determined by the details of the electronic structure of the considered molecules and can be deduced based on the fundamental rules proposed by Gil and von Philipsborn [110]. The authors proved that, if one of the coupled nuclei has lone electron pair(s) (LEP(s)), they always give the contributions of a negative sign to 1KFC(A,B). As a consequence, the removal of a lone electron pair, namely by protonation, alkylation, oxide formation, or complexation, leads to an increase in 1KFC(A,B). The negative contributions from LEP(s) (1KFCLP) compete with the always positive contributions from the bonding orbital between coupled nuclei (1KFCBD) [111]. In the case of the SeH2 molecule, the 1KFCBD is very large [111] and, apparently, is not quite suppressed by the competing 1KFCLP, giving (in total with minor rest contributions) a relatively small positive 1KFC(Se,H) [43], while for most 1J(Se,C) SSCCs, the FC contribution has a negative sign [32,43], which can be related to the total domination of the negative 1KFCLP over the positive 1KFCBD.
It was shown earlier [87,93] that the expansion of the f-angular space of a standard energy-optimized basis set of double-zeta quality leads to a dramatic change in the one-bond SSCCs involving selenium and other NMR-active half spin nuclei. A weaker but no less important sensitivity of the selenium SSCCs has been found when varying the tight s-region. Therefore, before starting the basis set optimization process, we carried out the investigation of the sensitivity of 1J(77Se,1H) of SeH2 and 1J(77Se,13C) of Se=C to the expansion of different angular spaces of basis sets of double- and triple-zeta quality on the selenium atom. For that purpose, we have set the uncontracted pecJ-1 or pecJ-2 basis sets on hydrogen and carbon atoms [98], while the cc-pVDZ+2f or cc-pVTZ+2f basis sets were set on the selenium atom. The cc-pVDZ+2f represents the uncontracted cc-pVDZ basis set [105] augmented with both f-functions of aug-cc-pVTZ basis set [105], while the cc-pVTZ+2f is the uncontracted cc-pVTZ basis set [105], whose single f-function was replaced with two f-functions of the aug-cc-pVTZ basis set. These basis sets were considered the starting sets for further expansion. The behaviors of the 1J(77Se,1H) and 1J(77Se,13C) SSCCs upon the saturation of the cc-pVDZ+2f or cc-pVTZ+2f basis sets on selenium atom are shown in Figure 1a,b.
The saturation of basis sets on the selenium atom was carried out in the tight region of each angular space by means of applying the geometrical progression or even-tempered recurrent ratio ζi = αβi, with α representing the largest exponent ζn in the original functional set and β being the ratio of two largest exponents, β = ζnn−1. The saturation of basis sets has been made in the consecutive manner, that is, we passed to the saturation of the next angular space only if the current space was totally saturated and provided the converged value.
From Figure 1a,b, one can see that the addition of two tight s-functions to both cc-pVDZ+2f and cc-pVTZ+2f basis sets provides the converged values in both cases. Thus, two s-functions have been added to their initial configurations. The modifications made in the p-space did not cause any effect, while the expansion of the d-space slightly affected both SSCCs only in the case the of cc-pVDZ+2f basis set. Thus, it was decided to add one d-function to the cc-pVDZ+2f basis set and to remove one d-function from the cc-pVTZ+2f basis set, for the sake of lowering the least needed size of the d-space in the latter case. The expansion of the f-region resulted in very substantial changes in both SSCCs calculated with both double- and triple-zeta basis sets. As can be seen from Figure 1, the decrease of the 1J(77Se,1H) and 1J(77Se,13C) upon adding four additional tight functions to the f-space of both basis sets is about 25 and 12–14 Hz, respectively. Figure 1 tells us that the least needed number of additional f-functions to the original two functions is three in all cases, thus giving five f-functions in total. In that way, five f-functions provide sufficiently converged values within the f-space. However, we decided to deal with only four and three f-functions in the f-spaces of the pecJ-2 and pecJ-1 basis sets, respectively. This seemingly controversial decision can be explained by the very essence of our PEC method. This method is not a mere even-tempered augmentation of the basis set but an optimization of the exponents that implies dealing with a fewer number of exponents providing the same or even better accuracy than the larger even-tempered basis set. Thus, leaning on our experience of the PEC method, three additional even-tempered exponents can be replaced with one and two optimized exponents for the double- and triple-zeta levels, respectively, resulting in three and four f-functions in total. That is enough for the correct representation of the first polarization shell. This also can be thought of as an explanation of our decision of removing one d-function in the final configuration of the pecJ-2 basis set. We also decided to introduce one g-function to improve the description of the second polarization shell of the selenium atom. The resulting configurations and the modifications made are represented in Table 1.
Thus, having the resulting configurations for the pecJ-n basis sets, we have set the obtained trial basis sets on the selenium atom and pecJ-1 or pecJ-2 basis sets on the hydrogens and carbon atoms in the fitting molecules SeH2 and Se=C, and commenced the PEC optimization of the exponents. It is worth mentioning once again that, in the current case, we considered only the FC terms of the selenium SSCCs. In this sense, it could be said that we have developed the pecJ-n basis sets for the FC-dominating selenium SSCCs, covering by that the majority of cases [32,47,48,50,112]. It is also very important that all exponents in all shells were varied during the optimization process, which is different from what was conducted in our first work on the PEC method [98], where for the nonhydrogen atoms of the second period, only the s-, d-, and f-shells were varied to converge the target FC term to the ideal value. The final deviations of the FC contributions to the SSCCs of both fitting molecules from their ideal values amounted to ca. 0.01 Hz for both pecJ-1 and pecJ-2 basis sets. The resulting optimized exponents of the pecJ-1 and pecJ-2 basis sets can be found in Tables S2–S5 of Supplementary Materials, where the final selenium pecJ-n basis sets are presented in the format of the Dalton [113] and CFOUR [114] programs.
To reduce the sizes of the obtained uncontracted basis sets (for that we will use the notation “(uc)” throughout the text), pecJ-n(uc), we have applied a general contraction scheme [115]. However, this time, it was not a mere gain of the contraction coefficients from the molecular energy Self-Consistent Field (SCF) calculations of the simplest hydrides but was the application of the PEC algorithm to minimize the contraction error, providing the least possible molecular energy. In more detail, to obtain the contraction coefficients for the selenium pecJ-n basis sets, we consecutively (in relation to shells) minimized the sum of the absolute differences between the values of the 1J(77Se,1H) of SeH2 and 1J(77Se,13C) of Se=C (this time, it was the total values consisting of the four Ramsey contributions) and the corresponding contemporary reference values. The pecJ-1 or pecJ-2 basis sets were set on the hydrogen and carbon atoms, depending on the cardinal number of the basis set used on the selenium atom. This routine can be thought of as the following:
  • The PEC optimization of the contraction coefficients for the s-shell with respect to the target function: Δ ˜ = 1 2 i = 1 2 J ˜ i J i r e f is performed. The J ˜ i are the values obtained with chosen s-contraction pattern with the varying contraction coefficients, and J i r e f are the values obtained with the fully uncontracted pecJ-n(uc) basis set on selenium. When the optimization procedure is completed, the reference values J i r e f are redefined: the new ones are obtained with the modified pecJ-n basis set on selenium, in which the s-shell is contracted with the optimized coefficients, and the remaining shells are uncontracted;
  • The PEC optimization of the contraction coefficients for the p-shell starts; the reference values are those that were recalculated in the previous step. When the optimized coefficients for the p-shell are obtained, the J i r e f are redefined once again: the new ones are calculated with the pecJ-n basis set, in which the s-shell is contracted with the coefficients obtained in the first step and p-shell is contracted with the newly optimized coefficients, while the remaining shells are kept uncontracted;
  • The PEC optimization of the contraction coefficients for the d-shell starts; the reference values are those that were recalculated in the previous step. In the end, we arrive at the final pecJ-n basis set with the contracted s-, p-, and d-shells and uncontracted f-shell, plus 1g function in the pecJ-2 basis set.
During this algorithm, the contraction patterns with different contraction depths were considered for each shell. A decision for the choice of the number of contracted functions was made based on the resulting contraction error Δ for each step. Table 2 shows different contraction patterns with the final averaged absolute errors (Δ) in relation to the current reference values.
The final configurations of the pecJ-n basis sets for the selenium atom together with the mean absolute percentage errors (ε), which they provide at the SOPPA(CCSD) level against the values obtained with the totally uncontracted pecJ-n(uc) basis set used on the selenium atom, are compiled in Table 3. The final contraction coefficients for the pecJ-n basis sets are presented in Tables S2–S5 of Supplementary Materials.
It is worth mentioning that the contracted aug-cc-pVTZ-J basis set for the selenium atom has the configuration [17s10p7d5f], thus providing the Nbas of 117. Our contracted second-level basis set, pecJ-2, has a pronounced benefit in size as compared to the aug-cc-pVTZ-J basis set, being 16 basis set functions fewer than the latter. If we recall the formal computational scaling [62] for the most popular high-quality computational methods applied to the calculation of SSCCs, namely N5 for SOPPA, N5 for SOPPA(CC2), N6 for SOPPA(CCSD), N6 for CCSD, and N4 for DFT, with N being the total number of basis set functions participating in the calculation, the additional 16 functions on selenium result in a significant increasing of the operations needed to evaluate characteristic operators in a molecular orbital basis. To roughly exemplify the operational costs provided by the pecJ-n and aug-cc-pVTZ-J basis sets, we show the assessment of Nk (with k being the scaling factors of different methods) for the simplest selenium hydride, SeH2; see Table 4. We set the basis set of the same type on the hydrogen atom.
From Table 4, it can be seen that the pecJ-n basis sets are beneficial compared to the aug-cc-pVTZ-J basis set. For example, for the most popular SOPPA and CCSD methods, the calculations with the pecJ-1 and pecJ-2 basis sets are, respectively, ten and two times less computationally demanding than that with the aug-cc-pVTZ-J basis set.

2.2. Testing New Basis Sets

In this section, we demonstrate the accuracy provided by our new basis sets against the experiment. All couplings were calculated at the CCSD level of theory, taking into account vibrational, solvent, and relativistic corrections to be properly compared with the experimental data. Overall, we calculated 13 selenium SSCCs of different types in seven representative molecules. The basic CCSD values were calculated with the pecJ-n and aug-cc-pVTZ-J basis sets being set on all atoms with the exclusion of chlorine; for the chlorine, atom we used the Dunning’s correlation consistent basis sets of different qualities depending on the basis sets used on the rest of atoms, namely, the cc-pVDZ on Cl with the pecJ-1 on the rest, and cc-pVTZ on Cl with the pecJ-2 or aug-cc-pVTZ-J on the rest. All three types of corrections were calculated using the SVWN5 exchange-correlation functional [116,117]. The SVWN5 represents the local density approximation (LDA) [118,119], in which the exchange is uniquely defined analytically in the form of the exchange energy of a homogeneous electron gas, while its correlation term is defined through several parameterizations, mostly relying on the highly accurate quantum Monte Carlo simulations. We have chosen this function based on the fact that the LDA model is particularly stable towards triplet instabilities [120], reflecting a balanced description of exchange and correlation. The latter is very important for the calculation of triplet excitation properties such as FC and SD contributions to the spin–spin coupling constants.
The solvent corrections to SSCCs were calculated as the differences between their values obtained in the gas and liquid phases. The IEF-PCM scheme [121,122] for a particular solvent (specified in accordance with the experimental data) was used for each compound in the liquid phase simulation. In the calculations of the solvent corrections, the same basis set schemes as those applied in the CCSD calculations were used.
The vibrational corrections to the SSCCs were calculated at zero temperature (zero-point vibrational corrections, ZPVC) within the vibrational second-order perturbation theory (VPT2) [123] as applied in combination with the effective geometry approach of Ruud et al. [124]. The effective geometry represents the vibrationally averaged molecular geometry to the second-order in perturbation theory (involving the cubic force-field tensor). Thus, by means of calculating the property at the effective geometry, one partially takes into account the contribution to a vibrationally averaged property due to the anharmonicity of the potential, while the inclusion of the contribution from the averaging of the molecular property over the harmonic oscillator requires the additional step where the second derivative of the property surface is calculated at the effective geometry. In all vibrational calculations, we used the pecJ-1 basis set on all atoms except for the chlorine atom, on which the cc-pVDZ basis set was used.
The relativistic corrections to coupling constants were evaluated as the differences between the SSCCs obtained using the relativistic four-component Dirac–Kohn–Sham–Hamiltonian and those obtained within the “10c limit scheme”. The “10c limit scheme” implies the increase of the speed of light by 10 times in the relativistic four-component calculations, resulting in a sufficiently accurate approximation of nonrelativistic values.
The “10c limit scheme” is now a widely used approximation [24,25,93,94,125,126,127,128,129] that is applied to exclude the basis set inequivalence in passing from the four-component relativistic to the one-component nonrelativistic framework. The convergences of the selenium SSCCs with the increasing of the speed of light are shown in Figures S1 and S2 in the Supplementary Materials for the examples of 1J(Se,H) in SeH2 and 1J(Se,C) in Se=C, respectively, calculated at the four-component DFT(SVWN5)//pecJ-2(uc) level. The speed of light was increased from one (totally relativistic calculation) to ten (“10c limit scheme”) times. It follows that, for both molecules, the convergence within ca. 0.1–0.2 Hz already appears at the increasing of the speed of light by seven to eight times. Thus, the “10c limit scheme” can be regarded as appropriate for the calculations of the selenium SSCCs considered in this work.
In the four-component calculations, the same basis sets as in the CCSD calculations were used, though applied in the uncontracted form. The four-component calculations were performed under the unrestricted kinetic balance condition (UKB) [130,131].
The results are compiled in Table 5 and Table 6, with the former presenting the roles of four Ramsey contributions to total CCSD values of SSCCs and the latter showing the comparison of the corrected CCSD values with the experiment.
As can be seen from Table 5, the FC contribution by far dominates the other three contributions for most of the considered SSCCs. Only in the case of Se-P SSCCs are the PSO contributions significant on an absolute scale, though, as compared to the FC terms of the Se-P SSCCs, these are not so significant, being only about 20% of the magnitude of the FC terms. It is also worth mentioning that in some one-bond Se-C and Se-H SSCCs, the SD contribution is noticeable on an absolute scale (for example, in compounds 2, 5, 7), though being substantially inferior to the FC terms.
As can be seen from Table 6, the agreement of total theoretical values obtained with introduced basis sets with the experimental data is rather good. The mean absolute percentage errors (MAPEs) evaluated for all final theoretical selenium SSCCs against the experiment are presented in Figure 2.
From Figure 2, one can see that the total accuracy provided by the pecJ-2 basis set (MAPE = 6.8%) is practically the same as that provided by a rather larger aug-cc-pVTZ-J basis set (MAPE = 6.5%). In terms of formal operational costs, for the most popular SOPPA or CCSD methods, we can arrive at a 50% reduction (as was roughly estimated above in the example of selenium hydride) of the number of operations without any noticeable loss of accuracy by using the pecJ-2 basis set instead of the aug-cc-pVTZ-J basis set. Meanwhile, the calculations with the pecJ-1 basis set are several times less computationally demanding than those with the pecJ-2 basis set, and the accuracy is only slightly (MAPE = 8.9 vs. 6.8%) inferior to that provided by the pecJ-2 basis set. Thus, we can conclude here that our compact pecJ-1 basis set may be very useful in large-scale calculations or in very demanding calculations like those with highly correlated methods involving triple- or higher excitations or in the problem of vibrational averaging.

3. Materials and Methods

Geometry optimizations were carried out at the DFT level of theory with the Minnesota M06-2X exchange-correlation functional [137] using the pc-3 basis set [138,139,140,141]. Media effects were taken into account within the IEF-PCM solvation model when optimizing the geometrical parameters for the testing compounds 17, while the optimization of the geometry of both fitting molecules, SeH2 and Se=C, has been performed in the gas phase. All optimizations were performed in the Gaussian program [142]. Final equilibrium geometries are presented in Table S1 of Supplementary Materials. The SOPPA(CCSD) and CCSD calculations of SSCCs were carried out in the Dalton [113] and CFOUR [114] programs, respectively. Solvent and vibrational corrections to SSCCs were calculated using the Dalton program. Relativistic values were calculated within the DIRAC program [143].

4. Conclusions

We presented new compact and accurate pecJ-n (n = 1, 2) basis sets for the selenium atom purposed for the quantum–chemical calculations of NMR spin–spin coupling constants involving selenium nuclei. These basis sets were obtained with the property-energy consistent method, which is efficient in generating compact property-oriented basis sets due to its peculiarity to fully reoptimize all angular spaces of the trial basis sets. In this way, the final PEC-generated basis sets are significantly smaller than the other specialized basis sets of the same zeta-quality obtained with the even-tempered technique. Thus, new pecJ-1 and pecJ-2 basis sets consist of only 78 and 101 basis functions, respectively. This gives a significant benefit as compared to the other existing J-oriented selenium basis sets, acvXz-J (X = 2, 3, 4) and aug-cc-pVTZ-J, with the total number of functions being as much as 88, 118, 167, and 117, respectively. In particular, for the SeH2 molecule, the operational costs of the SOPPA and CCSD calculations can be said to be significantly reduced by using the pecJ-1 and pecJ-2 basis sets by approximately ten and two times, respectively, as compared to the costs of the calculations with the aug-cc-pVTZ-J basis set.
New basis sets were tested on the CCSD calculations of thirteen SSCCs involving selenium in the representative series of seven molecules carried out with taking into account relativistic, solvent, and vibrational corrections. The comparison with the experiment revealed that the accuracy of the results obtained with a compact pecJ-2 basis set is almost the same as that provided by an essentially larger basis set, aug-cc-pVTZ-J (the MAPEs are 6.8% and 6.5% for the former and the latter, respectively), while the accuracy achieved with a significantly smaller basis set, pecJ-1, is only slightly inferior (MAPE = 8.9%) to the accuracy provided by the pecJ-2 basis set. Overall, we recommend resorting to the pecJ-2 basis set in large-scale calculations of selenium SSCCs; this would provide accuracy comparable to that of the aug-cc-pVTZ-J basis set and give significant CPU time savings. The pecJ-1 can be recommended for highly demanding calculations, such as those involving the coupled cluster methods of higher hierarchy that treat triple- and higher excitations or in the problem of vibrational averaging.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24097841/s1.

Author Contributions

Conceptualization, Y.Y.R. and I.L.R.; methodology, Y.Y.R. and I.L.R.; software, Y.Y.R.; validation, Y.Y.R. and I.L.R.; formal analysis, I.L.R.; investigation, Y.Y.R. and I.L.R.; resources, Y.Y.R. and I.L.R.; data curation, Y.Y.R. and I.L.R.; writing—original draft preparation, I.L.R.; writing—review and editing, Y.Y.R. and I.L.R.; visualization, I.L.R.; supervision, Y.Y.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank the Baikal Analytical Center of SB RAS for providing us with computational facilities, allowing us to perform most parts of the calculations. We are also grateful to the Irkutsk Supercomputer Center of SB RAS for providing computational resources for the computational cluster “Academician V. M. Matrosov” [144]. This work was carried out within the framework of the research project of the Russian Academy of Sciences # 121021000264-1.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Eybl, V.; Kotyzová, D.; Sýkora, J.; Topolčan, O.; Pikner, R.; Mihaljevič, M.; Brtko, J.; Glattre, E. Effects of selenium and tellurium on the activity of selenoenzymes glutathione peroxidase and Type I iodothyronine deiodinase, trace element thyroid level, and thyroid hormone status in rats. Biol. Trace Elem. Res. 2007, 117, 105–114. [Google Scholar] [CrossRef] [PubMed]
  2. Johansson, L.; Gafvelin, G.; Arnér, E.S. Selenocysteine in proteins—Properties and biotechnological use. Biochim. Biophys. Acta 2005, 1726, 1–13. [Google Scholar] [CrossRef] [PubMed]
  3. Schwarz, K.; Foltz, C.M. Selenium as an integral part of factor 3 against dietary necrotic liver degeneration. J. Am. Chem. Soc. 1957, 79, 3292–3293. [Google Scholar] [CrossRef]
  4. Rayman, M.P. Selenium and human health. Lancet 2012, 379, 1256–1268. [Google Scholar] [CrossRef]
  5. Mehdi, Y.; Hornick, J.-L.; Istasse, L.; Dufrasne, I. Selenium in the Environment, Metabolism and Involvement in Body Functions. Molecules 2013, 18, 3292–3311. [Google Scholar] [CrossRef]
  6. Standtman, T.C. Selenium biochemistry. Annu. Rev. Biochem. 1990, 59, 111–127. [Google Scholar] [CrossRef]
  7. Combs, G.F., Jr.; Combs, S.B. The Nutritional Biochemistry of Selenium. Annu. Rev. Nutr. 1984, 4, 257–280. [Google Scholar] [CrossRef]
  8. Ventura, M.; Melo, M.; Carrilho, F. Selenium and Thyroid Disease: From Pathophysiology to Treatment. Int. J. Endocrinol. 2017, 2017, 1297658. [Google Scholar] [CrossRef]
  9. Phiri, F.P.; Ander, E.L.; Lark, R.M.; Bailey, E.H.; Chilima, B.; Gondwe, J.; Joy, E.J.M.; Kalimbira, A.A.; Phuka, J.C.; Suchdev, P.S.; et al. Urine selenium concentration is a useful biomarker for assessing population level selenium status. Environ. Int. 2020, 134, 105218. [Google Scholar] [CrossRef]
  10. Behne, D.; Kyriakopoulos, A. Mammalian selenium-containing proteins. Annu. Rev. Nutr. 2001, 21, 453–473. [Google Scholar] [CrossRef]
  11. Garcia, R.I.; Lozano, R.D.; Martinez, E.A. A transimpedance circuit for use with selenium photoelectric cells. J. Phys. E Sci. Instrum. 1972, 5, 745. [Google Scholar] [CrossRef]
  12. Huang, W.; Zhang, Y.; You, Q.; Huang, P.; Wang, Y.; Huang, Z.N.; Ge, Y.; Wu, L.; Dong, Z.; Dai, X.; et al. Enhanced Photodetection Properties of Tellurium-Selenium Roll-to-Roll Nanotube Heterojunctions. Nano Micro Small 2019, 15, 1900902. [Google Scholar] [CrossRef] [PubMed]
  13. Hadar, I.; Song, T.-B.; Ke, W.; Kanatzidis, M.G. Modern Processing and Insights on Selenium Solar Cells: The World’s First Photovoltaic Device. Adv. Energy Mater. 2019, 9, 1802766. [Google Scholar] [CrossRef]
  14. Malik, M.A.; Ramasamy, K.; O’Brien, P. Selenium and Tellurium Containing Precursors for Semiconducting Materials. In Selenium and Tellurium Chemistry. From Small Molecules to Biomolecules and Materials, 1st ed.; Woollins, J.D., Laitinen, R., Eds.; Springer: Berlin/Heidelberg, Germany, 2011; Chapter 9; pp. 201–238. [Google Scholar] [CrossRef]
  15. Johnson, J.A.; Saboungi, M.-L.; Thiyagarajan, P.; Csencsits, R.; Meisel, D. Selenium Nanoparticles: A Small-Angle Neutron Scattering Study. J. Phys. Chem. B 1999, 103, 59–63. [Google Scholar] [CrossRef]
  16. Piacenza, E.; Presentato, A.; Zonaro, E.; Lampis, S.; Vallini, G.; Turner, R.J. Physical–Chemical Properties of Biogenic Selenium Nanostructures Produced by Stenotrophomonas maltophilia SeITE02 and Ochrobactrum sp. MPV1. Phys. Sci. Rev. 2018, 3, 20170100. [Google Scholar] [CrossRef]
  17. Mayers, B.; Jiang, X.; Sunderland, D.; Cattle, B.; Xia, Y. Hollow Nanostructures of Platinum with Controllable Dimensions Can Be Synthesized by Templating Against Selenium Nanowires and Colloids. J. Am. Chem. Soc. 2003, 125, 13364–13365. [Google Scholar] [CrossRef]
  18. Terasaki, O.; Yamazaki, K.; Thomas, J.M.; Ohsuna, T.; Watanabe, D.; Sanders, J.V.; Barry, J.C. Isolating individual chains of selenium by incorporation into the channels of a zeolite. Nature 1987, 330, 58–60. [Google Scholar] [CrossRef]
  19. Parise, J.B.; MacDougall, J.; Herron, N.; Farlee, R.; Sleight, A.W.; Wang, Y.; Bein, T.; Moller, K.; Moroney, L.M. Characterization of selenium-loaded molecular sieves A, X, Y, AlPO-5, and mordenite. Inorg. Chem. 1988, 27, 221–228. [Google Scholar] [CrossRef]
  20. Rusakova, I.L.; Rusakov, Y.Y. Quantum chemical calculations of 77Se and 125Te nuclear magnetic resonance spectral parameters and their structural applications. Magn. Reson. Chem. 2021, 59, 359–407. [Google Scholar] [CrossRef]
  21. Nakanishi, W.; Hayashi, S. 77Se NMR: Theoretical Aspects and Practical Applications. In Organoselenium Chemistry: Between Synthesis and Biochemistry, 1st ed.; Santi, C., Ed.; Bentham Science Publishers (eBook): Dubai, United Arab Emirates, 2014; Chapter 13; pp. 379–415. [Google Scholar] [CrossRef]
  22. Krivdin, L.B. Recent advances in computational liquid-phase 77Se NMR. Russ. Chem. Rev. 2021, 90, 265–279. [Google Scholar] [CrossRef]
  23. Rusakov, Y.Y.; Rusakova, I.L.; Krivdin, L.B. MP2 calculation of 77Se NMR chemical shifts taking into account relativistic corrections. Magn. Reson. Chem. 2015, 53, 485–492. [Google Scholar] [CrossRef] [PubMed]
  24. Rusakov, Y.Y.; Rusakova, I.L.; Krivdin, L.B. On the significant relativistic heavy atom effect on 13C NMR chemical shifts of β- and γ-carbons in seleno- and telluroketones. Mol. Phys. 2017, 115, 3117–3127. [Google Scholar] [CrossRef]
  25. Rusakov, Y.Y.; Rusakova, I.L. Long-range relativistic heavy atom effect on 1H NMR chemical shifts of selenium- and tellurium-containing compounds. Int. J. Quantum Chem. 2018, 119, e25809. [Google Scholar] [CrossRef]
  26. Vícha, J.; Novotný, J.; Komorovsky, S.; Straka, M.; Kaupp, M.; Marek, R. Relativistic Heavy-Neighbor-Atom Effects on NMR Shifts: Concepts and Trends Across the Periodic Table. Chem. Rev. 2020, 120, 7065–7103. [Google Scholar] [CrossRef]
  27. Vícha, J.; Komorovsky, S.; Repisky, M.; Marek, R.; Straka, M. Relativistic Spin−Orbit Heavy Atom on the Light Atom NMR Chemical Shifts: General Trends Across the Periodic Table Explained. J. Chem. Theory Comput. 2018, 14, 3025–3039. [Google Scholar] [CrossRef]
  28. Rusakova, I.L.; Rusakov, Y.Y. Relativistic Effects from Heavy Main Group p-Elements on the NMR Chemical Shifts of Light Atoms: From Pioneering Studies to Recent Advances. Magnetochemistry 2023, 9, 24. [Google Scholar] [CrossRef]
  29. Wrackmeyer, B. Indirect Nuclear 77Se–77Se Spin–Spin Coupling Constants. Application of Density Functional Theory (DFT) Calculations. Struct. Chem. 2005, 16, 67–71. [Google Scholar] [CrossRef]
  30. Åstrand, P.-O.; Mikkelsen, K.V. Solvent effects on nuclear shieldings and spin–spin couplings of hydrogen selenide. J. Chem. Phys. 1998, 108, 2528–2537. [Google Scholar] [CrossRef]
  31. Rusakova, I.L.; Rusakov, Y.Y. Correlated ab initio calculations of one-bond 31P-77Se and 31P-125Te spin–spin coupling constants in a series of P=Se and P=Te systems accounting for relativistic effects (part 2). Magn. Reson. Chem. 2020, 58, 929–940. [Google Scholar]
  32. Rusakov, Y.Y.; Rusakova, I.L.; Krivdin, L.B. Full four-component relativistic calculations of the one-bond 77Se–13C spin-spin coupling constants in the series of selenium heterocycles and their parent open-chain selenides. Magn. Reson. Chem. 2014, 52, 214–221. [Google Scholar] [CrossRef]
  33. Feindel, K.W.; Wasylishen, R.E. A relativistic DFT study of one-bond fluorine-X indirect spin–spin coupling tensors. Magn. Reson. Chem. 2004, 42, S158–S167. [Google Scholar] [CrossRef] [PubMed]
  34. Tanioku, A.; Hayashi, S.; Nakanishi, W. Analysis of One-Bond Se-Se Nuclear Couplings in Diselenides and 1,2-Diselenoles on the Basis of Molecular Orbital Theory: Torsional Angular Dependence, Electron Density Influence, and Origin in 1J(Se,Se). Bioinorg. Chem. Appl. 2009, 2009, 381925. [Google Scholar] [CrossRef] [PubMed]
  35. Jokisaari, J.; Autschbach, J. 13C–77Se and 77Se–77Se spin–spin coupling tensors in carbon diselenide: NMR experiments and ZORA DFT calculations. Phys. Chem. Chem. Phys. 2003, 5, 4551–4555. [Google Scholar] [CrossRef]
  36. Komorovsky, S.; Jakubowska, K.; Swider, P.; Repisky, M.; Jaszunski, M. NMR Spin–Spin Coupling Constants Derived from Relativistic Four-Component DFT Theory—Analysis and Visualization. J. Phys. Chem. A 2020, 124, 5157–5169. [Google Scholar] [CrossRef] [PubMed]
  37. Pyykkö, P.; Wiesenfeld, L. Relativistically parameterized extended Hückel calculations. IV. Nuclear spin-spin coupling tensors for main group elements. Mol. Phys. 1981, 43, 557–580. [Google Scholar] [CrossRef]
  38. Gomez, S.S.; Romero, R.H.; Aucar, G.A. Fully relativistic calculation of nuclear magnetic shieldings and indirect nuclear spin-spin couplings in group-15 and -16 hydrides. J. Chem. Phys. 2002, 117, 7942–7946. [Google Scholar] [CrossRef]
  39. Aucar, G.A.; Saue, T.; Visscher, L.; Jensen, H.A. On the origin and contribution of the diamagnetic term in four-component relativistic calculations of magnetic properties. J. Chem. Phys. 1999, 110, 6208–6218. [Google Scholar] [CrossRef]
  40. Rusakova, I.L.; Rusakov, Y.Y.; Krivdin, L.B. First example of the correlated calculation of the one-bond tellurium-carbon spin-spin coupling constants: Relativistic effects, vibrational corrections, and solvent effects. J. Comput. Chem. 2016, 37, 1367–1372. [Google Scholar] [CrossRef]
  41. Rusakov, Y.Y.; Krivdin, L.B.; Østerstrøm, F.F.; Sauer, S.P.A.; Potapov, V.A.; Amosova, S.V. First example of a high-level correlated calculation of the indirect spin–spin coupling constants involving tellurium: Tellurophene and divinyl telluride. Phys. Chem. Chem. Phys. 2013, 15, 13101–13107. [Google Scholar] [CrossRef]
  42. Rusakova, I.L.; Krivdin, L.B. Relativistic effects in the NMR spectra of compounds containing heavy chalcogens. Mendeleev Commun. 2018, 28, 1–13. [Google Scholar] [CrossRef]
  43. Rusakova, I.L.; Rusakov, Y.Y.; Krivdin, L.B. Relativistic effects in the one-bond spin–spin coupling constants involving selenium. Magn. Reson. Chem. 2014, 52, 500–510. [Google Scholar] [CrossRef] [PubMed]
  44. Autschbach, J. Perspective: Relativistic effects. J. Chem. Phys. 2012, 136, 150902. [Google Scholar] [CrossRef] [PubMed]
  45. Autschbach, J. Relativistic Effects on NMR Parameters. In High Resolution NMR Spectroscopy. Understanding Molecules and Their Electronic Structures, 1st ed.; Contreras, R.H., Ed.; Elsevier B.V.: London, UK, 2013; Volume 3, Chapter 4; pp. 69–117. [Google Scholar]
  46. Colquhoun, I.J.; McFarlane, H.C.E.; McFarlane, W.; Nash, J.A.; Keat, R.; Rycroft, D.S.; Thompson, D.G. Long-range nuclear spin–spin coupling between selenium-77 and phosphorus-31 in biphosphorus compounds. Org. Magn. Res. 1979, 12, 473–475. [Google Scholar] [CrossRef]
  47. Krivdin, L.B.; Rusakov, Y.Y. Structural and Stereochemical Applications of Computational NMR Using 29Si–1H and 77Se–1H Indirect Spin–Spin Coupling Constants. eMagRes 2014, 3, 87–110. [Google Scholar] [CrossRef]
  48. Rusakov, Y.Y.; Krivdin, L.B. Stereochemical behavior of geminal and vicinal 77Se–13C spin–spin coupling constants studied at the SOPPA(CC2) level taking into account relativistic corrections. Magn. Reson. Chem. 2015, 53, 93–98. [Google Scholar] [CrossRef]
  49. Kövér, K.E.; Kumar, A.A.; Rusakov, Y.Y.; Krivdin, L.B.; Illyés, T.-Z.; Szilágyi, L. Experimental and computational studies of nJ(77Se,1H) selenium–proton couplings in selenoglycosides. Magn. Reson. Chem. 2011, 49, 190–194. [Google Scholar] [CrossRef]
  50. Rusakov, Y.Y.; Krivdin, L.B.; Penzik, M.V.; Potapov, V.A.; Amosova, S.V. Open-chain unsaturated selanyl sulfides: Stereochemical structure and stereochemical behavior of their 77Se–1H spin–spin coupling constants. Magn. Reson. Chem. 2012, 50, 653–658. [Google Scholar] [CrossRef]
  51. Rusakov, Y.Y.; Krivdin, L.B.; Kumar, A.A.; Szilágyi, L.; Kövér, K.E. Resonance assignments of diastereotopic CH2 protons in the anomeric side chain of selenoglycosides by means of 2J(Se,H) spin-spin coupling constants. Magn. Reson. Chem. 2012, 50, 488–495. [Google Scholar] [CrossRef]
  52. Nakanishi, W.; Hayashi, S. Torsional Angular Dependence of 1J(Se,Se) and Fermi Contact Control of 4J(Se,Se): Analysis of nJ(Se,Se) (n = 1–4) Based on Molecular Orbital Theory. Chem. Eur. J. 2008, 14, 5645–5655. [Google Scholar] [CrossRef]
  53. Rusakov, Y.Y.; Krivdin, L.B.; Orlov, N.V.; Ananikov, V.P. Stereochemical study of the sterically crowded phenylselanylalkenes by means of 77Se-1H spin–spin coupling constants. Magn. Reson. Chem. 2011, 49, 570–574. [Google Scholar] [CrossRef]
  54. Rusakov, Y.Y.; Krivdin, L.B.; Potapov, V.A.; Penzik, M.V.; Amosova, S.V. Conformational analysis and diastereotopic assignments in the series of selenium-containing heterocycles by means of 77Se-1H spin-spin coupling constants: A combined theoretical and experimental study. Magn. Reson. Chem. 2011, 49, 389–398. [Google Scholar] [CrossRef]
  55. Rusakov, Y.Y.; Krivdin, L.B. Stereochemical behavior of 2J(Se,H) and 3J(Se,H) spin-spin coupling constants across sp3 carbons: A theoretical scrutiny. Magn. Reson. Chem. 2012, 50, 557–562. [Google Scholar] [CrossRef] [PubMed]
  56. Autschbach, J.; Ziegler, T. Relativistic computation of NMR shieldings and spin-spin coupling constants. In Encyclopedia of Nuclear Magnetic Resonance: Advances in NMR, 1st ed.; Grant, D.M., Harris., R.K., Eds.; John Wiley and Sons: Chichester, UK, 2002; Volume 9, pp. 306–323. [Google Scholar]
  57. Autschbach, J. Relativistic calculations of magnetic resonance parameters: Background and some recent developments. Phil. Trans. R. Soc. A 2014, 372, 20120489. [Google Scholar] [CrossRef] [PubMed]
  58. Facelli, J.C. Chemical shift tensors: Theory and application to molecular structural problems. Prog. Nucl. Magn. Reson. Spectrosc. 2011, 58, 176–201. [Google Scholar] [CrossRef] [PubMed]
  59. Pyykkö, P. Theory of NMR parameters. From Ramsey to Relativity, 1953 to 1983. In Calculation of NMR and EPR Parameters. Theory and Applications, 1st ed.; Kaupp, M., Bühl, M., Malkin, V.G., Eds.; WILEY-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2004; Chapter 2; pp. 7–19. [Google Scholar]
  60. Rusakov, Y.Y.; Krivdin, L.B. Modern quantum chemical methods for calculating spin–spin coupling constants: Theoretical basis and structural applications in chemistry. Russ. Chem. Rev. 2013, 82, 99–130. [Google Scholar] [CrossRef]
  61. Rusakova, I.L.; Rusakov, Y.Y.; Krivdin, L.B. Theoretical grounds of relativistic methods for calculation of spin–spin coupling constants in nuclear magnetic resonance spectra. Russ. Chem. Rev. 2016, 85, 356–426. [Google Scholar] [CrossRef]
  62. Rusakova, I.L. Quantum Chemical Approaches to the Calculation of NMR Parameters: From Fundamentals to Recent Advances. Magnetochemistry 2022, 8, 50. [Google Scholar] [CrossRef]
  63. Vahtras, O.; Ågren, H.; Jørgensen, P.; Helgaker, T.; Jensen, H.J.A. The nuclear spin-spin coupling in N2 and CO. Chem. Phys. Lett. 1993, 209, 201–206. [Google Scholar] [CrossRef]
  64. Barszczewicz, A.; Helgaker, T.; Jaszuński, M.; Jørgensen, P.; Ruud, K. Multiconfigurational self-consistent field calculations of nuclear magnetic resonance indirect spin–spin coupling constants. J. Chem. Phys. 1994, 101, 6822–6828. [Google Scholar] [CrossRef]
  65. Malkin, V.G.; Malkina, O.L.; Salahub, D.R. Calculation of spin-spin coupling constants using density functional theory. Chem. Phys. Lett. 1994, 221, 91–99. [Google Scholar] [CrossRef]
  66. Helgaker, T.; Watson, M.; Handy, N.C. Analytical calculation of nuclear magnetic resonance indirect spin–spin coupling constants at the generalized gradient approximation and hybrid levels of density-functional theory. J. Chem. Phys. 2000, 113, 9402–9409. [Google Scholar] [CrossRef]
  67. Autschbach, J.; Ziegler, T. Nuclear spin–spin coupling constants from regular approximate relativistic density functional calculations. I. Formalism and scalar relativistic results for heavy metal compounds. J. Chem. Phys. 2000, 113, 936–947. [Google Scholar] [CrossRef]
  68. Sychrovský, V.; Gräfenstein, J.; Cremer, D. Nuclear magnetic resonance spin–spin coupling constants from coupled perturbed density functional theory. J. Chem. Phys. 2000, 113, 3530–3547. [Google Scholar] [CrossRef]
  69. Barone, V.; Peralta, J.E.; Contreras, R.H.; Snyder, J.P. DFT Calculation of NMR JFF Spin−Spin Coupling Constants in Fluorinated Pyridines. J. Phys. Chem. A 2002, 106, 5607–5612. [Google Scholar] [CrossRef]
  70. Sekino, H.; Bartlett, R.J. Nuclear spin–spin coupling constants evaluated using many body methods. J. Chem. Phys. 1986, 85, 3945–3949. [Google Scholar] [CrossRef]
  71. Perera, S.A.; Sekino, H.; Bartlett, R.J. Coupled-cluster calculations of indirect nuclear coupling constants: The importance of non-Fermi contact contributions. J. Chem. Phys. 1994, 101, 2186–2191. [Google Scholar] [CrossRef]
  72. Perera, S.A.; Nooijen, M.; Bartlett, R.J. Electron correlation effects on the theoretical calculation of nuclear magnetic resonance spin–spin coupling constants. J. Chem. Phys. 1996, 104, 3290–3305. [Google Scholar] [CrossRef]
  73. Auer, A.; Gauss, J. Triple excitation effects in coupled-cluster calculations of indirect spin–spin coupling constants. J. Chem. Phys. 2001, 115, 1619–1622. [Google Scholar] [CrossRef]
  74. Christiansen, O.; Koch, H.; Jørgensen, P. The second-order approximate coupled cluster singles and doubles model CC2. Chem. Phys. Lett. 1995, 243, 409–418. [Google Scholar] [CrossRef]
  75. Bartlett, R.J. Many-Body Perturbation Theory and Coupled Cluster Theory for Electron Correlation in Molecules. Annu. Rev. Phys. Chem. 1981, 32, 359–401. [Google Scholar] [CrossRef]
  76. Geertsen, J.; Oddershede, J. Second-order polarization propagator calculations of indirect nuclear spin-spin coupling tensors in the water molecule. Chem. Phys. 1984, 90, 301–311. [Google Scholar] [CrossRef]
  77. Geertsen, J.; Oddershede, J. Higher RPA and second-order polarization propagator calculations of coupling constants in acetylene. Chem. Phys. 1986, 104, 67–72. [Google Scholar] [CrossRef]
  78. Kjær, H.; Sauer, S.P.A.; Kongsted, J. Benchmarking NMR indirect nuclear spin-spin coupling constants: SOPPA, SOPPA(CC2), and SOPPA(CCSD) versus CCSD. J. Chem. Phys. 2010, 133, 144106. [Google Scholar] [CrossRef] [PubMed]
  79. Sauer, S.P.A. Second-order polarization propagator approximation with coupled-cluster singles and doubles amplitudes—SOPPA(CCSD): The polarizability and hyperpolarizability of Li. J. Phys. B At. Mol. Opt. Phys. 1997, 30, 3773–3780. [Google Scholar] [CrossRef]
  80. Enevoldsen, T.; Oddershede, J.; Sauer, S.P.A. Correlated calculations of indirect nuclear spin-spin coupling constants using second-order polarization propagator approximations: SOPPA and SOPPA(CCSD). Theor. Chem. Acc. 1998, 100, 275–284. [Google Scholar] [CrossRef]
  81. Ramsey, N.F. Electron Coupled Interactions between Nuclear Spins in Molecules. Phys. Rev. 1953, 91, 303–307. [Google Scholar] [CrossRef]
  82. Jensen, F. The basis set convergence of spin–spin coupling constants calculated by density functional methods. J. Chem. Theory Comput. 2006, 2, 1360–1369. [Google Scholar] [CrossRef]
  83. Helgaker, T.; Jaszuński, M.; Ruud, K.; Górska, A. Basis-set dependence of nuclear spin-spin coupling constants. Theor. Chem. Acc. 1998, 99, 175–182. [Google Scholar] [CrossRef]
  84. Provasi, P.F.; Aucar, G.A.; Sauer, S.P.A. The effect of lone pairs and electronegativity on the indirect nuclear spin–spin coupling constants in CH2X (X=CH2, NH, O, S): Ab initio calculations using optimized contracted basis sets. J. Chem. Phys. 2001, 115, 1324–1334. [Google Scholar] [CrossRef]
  85. Provasi, P.F.; Sauer, S.P.A. Optimized basis sets for the calculation of indirect nuclear spin-spin coupling constants involving the atoms B, Al, Si, P, and Cl. J. Chem. Phys. 2010, 133, 054308. [Google Scholar] [CrossRef]
  86. Rusakov, Y.Y.; Krivdin, L.B.; Sauer, S.P.A.; Levanova, E.P.; Levkovskaya, G.G. Structural trends of 77Se-1H spin-spin coupling constants and conformational behavior of 2-substituted selenophenes. Magn. Reson. Chem. 2010, 48, 44–52. [Google Scholar] [CrossRef] [PubMed]
  87. Steinmann, C.; Sauer, S.P.A. The aug-cc-pVTZ-J basis set for the p-block fourth-row elements Ga, Ge, As, Se, and Br. Magn. Reson. Chem. 2021, 59, 1134–1145. [Google Scholar] [CrossRef] [PubMed]
  88. Barone, V.; Provasi, P.F.; Peralta, J.E.; Snyder, J.P.; Sauer, S.P.A.; Contreras, R.H. Substituent effects on scalar 2J(19F,19F) and 3J(19F,19F) NMR couplings: A comparison of SOPPA and DFT methods. J. Phys. Chem. A 2003, 107, 4748–4754. [Google Scholar] [CrossRef]
  89. Sauer, S.P.A.; Raynes, W.T. Unexpected differential sensitivity of nuclear spin-spin-coupling constants to bond stretching in BH4, NH4+, and SiH4. J. Chem. Phys. 2000, 113, 3121–3129. [Google Scholar] [CrossRef]
  90. Sauer, S.P.A.; Raynes, W.T.; Nicholls, R.A. Nuclear spin-spin coupling in silane and its isotopomers: Ab initio calculation and experimental investigation. J. Chem. Phys. 2001, 115, 5994–6006. [Google Scholar] [CrossRef]
  91. Hedegård, E.D.; Kongsted, J.; Sauer, S.P.A. Optimized basis sets for calculation of electron paramagnetic resonance hyperfine coupling constants: Aug-cc-pVTZ-J for the 3d atoms Sc-Zn. J. Chem. Theory Comput. 2011, 7, 4077–4087. [Google Scholar] [CrossRef]
  92. Kjær, H.; Sauer, S.P.A. Pople Style Basis Sets for the Calculation of NMR Spin–Spin Coupling Constants: The 6-31G-J and 6-311G-J Basis Sets. J. Chem. Theory Comput. 2011, 7, 4070–4076. [Google Scholar] [CrossRef]
  93. Rusakov, Y.Y.; Rusakova, I.L. Hierarchical Basis Sets for the Calculation of Nuclear Magnetic Resonance Spin–Spin Coupling Constants Involving Either Selenium or Tellurium Nuclei. J. Phys. Chem. A 2019, 123, 6564–6571. [Google Scholar] [CrossRef]
  94. Rusakov, Y.Y.; Rusakova, I.L. Efficient J-oriented tin basis sets for the correlated calculations of indirect nuclear spin–spin coupling constants. Magn. Reson. Chem. 2021, 59, 713–722. [Google Scholar] [CrossRef]
  95. Jensen, F. The optimum contraction of basis sets for calculating spin–spin coupling constants. Theor. Chem. Acc. 2010, 126, 371–382. [Google Scholar] [CrossRef]
  96. Aggelund, P.A.; Sauer, S.P.A.; Jensen, F. Development of polarization consistent basis sets for spin-spin coupling constant calculations for the atoms Li, Be, Na, and Mg. J. Chem. Phys. 2018, 149, 044117. [Google Scholar] [CrossRef] [PubMed]
  97. Benedikt, U.; Auer, A.A.; Jensen, F. Optimization of augmentation functions for correlated calculations of spin-spin coupling constants and related properties. J. Chem. Phys. 2008, 129, 064111. [Google Scholar] [CrossRef] [PubMed]
  98. Rusakov, Y.Y.; Rusakova, I.L. An efficient method for generating property-energy consistent basis sets. New pecJ-n (n = 1, 2) basis sets for high-quality calculations of indirect nuclear spin–spin coupling constants involving 1H, 13C, 15N, and 19F nuclei. Phys. Chem. Chem. Phys. 2021, 23, 14925–14939. [Google Scholar] [CrossRef] [PubMed]
  99. Rusakov, Y.Y.; Rusakova, I.L. New pecJ-n (n = 1, 2) Basis Sets for High-Quality Calculations of Indirect Nuclear Spin–Spin Coupling Constants Involving 31P and 29Si: The Advanced PEC Method. Molecules 2022, 27, 6145. [Google Scholar] [CrossRef]
  100. Rusakov, Y.Y.; Rusakova, I.L. New pecS-n (n = 1, 2) basis sets for quantum chemical calculations of the NMR chemical shifts of H, C, N and O nuclei. J. Chem. Phys. 2022, 156, 244112. [Google Scholar] [CrossRef]
  101. Metropolis, N.; Ulam, S. The Monte Carlo Method. J. Am. Stat. Assoc. 1949, 44, 335–341. [Google Scholar] [CrossRef]
  102. Harrison, R.L. Introduction to Monte Carlo Simulation. AIP Conf. Proc. 2010, 1204, 17–21. [Google Scholar] [CrossRef]
  103. Del Moral, P.; Doucet, A.; Jasra, A. Sequential Monte Carlo samplers. J. R. Statist. Soc. B 2006, 68, 411–436. [Google Scholar] [CrossRef]
  104. Reeves, C.M.; Fletcher, R. Use of Gaussian functions in the calculation of wavefunctions for small molecules. III. The orbital basis and its effect on valence. J. Chem. Phys. 1965, 42, 4073–4081. [Google Scholar] [CrossRef]
  105. Wilson, A.K.; Woon, D.E.; Peterson, K.A.; Dunning, T.H. Gaussian basis sets for use in correlated molecular calculations. IX. The atoms gallium through krypton. J. Chem. Phys. 1999, 110, 7667–7676. [Google Scholar] [CrossRef]
  106. Dyall, K.G. Relativistic and nonrelativistic finite nucleus optimized triple-zeta basis sets for the 4p, 5p and 6p elements. Theor. Chem. Acc. 2002, 108, 335–340. [Google Scholar] [CrossRef]
  107. Dyall, K.G. Relativistic quadruple-zeta and revised triple-zeta and double-zeta basis sets for the 4p, 5p and 6p elements. Theor. Chem. Acc. 2006, 115, 441–447. [Google Scholar] [CrossRef]
  108. Dyall, K.G. Relativistic and nonrelativistic finite nucleus optimized double zeta basis sets for the 4p, 5p and 6p elements. Theor. Chem. Acc. 1998, 99, 366–371. [Google Scholar] [CrossRef]
  109. Verbeke, J.; Cools, R. The Newton-Raphson method. Int. J. Math. Educ. Sci. Technol. 1995, 26, 177–193. [Google Scholar] [CrossRef]
  110. Gil, V.M.S.; von Philipsborn, W. Effect of Electron Lone-Pairs on Nuclear Spin-Spin Coupling Constants. Magn. Reson. Chem. 1989, 27, 409–430. [Google Scholar] [CrossRef]
  111. Wu, A.; Gräfenstein, J.; Cremer, D. Analysis of the Transmission Mechanism of NMR Spin-Spin Coupling Constants Using Fermi Contact Spin Density Distribution, Partial Spin Polarization, and Orbital Currents: XHn Molecules. J. Phys. Chem. A 2003, 107, 7043–7056. [Google Scholar] [CrossRef]
  112. Rusakov, Y.Y.; Krivdin, L.B.; Papernaya, L.K.; Shatrova, A.A. Stereochemical behavior of 77Se-1 H spin-spin coupling constants in pyrazolyl-1,3-diselenanes and 1,2-diselenolane. Magn. Reson. Chem. 2012, 50, 169–173. [Google Scholar] [CrossRef]
  113. Aidas, K.; Angeli, C.; Bak, K.L.; Bakken, V.; Bast, R.; Boman, L.; Christiansen, O.; Cimiraglia, R.; Coriani, S.; Dahle, P.; et al. The Dalton quantum chemistry program system. WIREs Comput. Mol. Sci. 2014, 4, 269–284. [Google Scholar] [CrossRef]
  114. Stanton, J.F.; Gauss, J.; Cheng, M.E.; Harding, D.A.; Matthews, P.G.; Asthana, A.; Auer, A.A.; Bartlett, R.J.; Benedikt, U.; Berger, C.; et al. CFOUR, A Quantum Chemical Program Package. Available online: http://www.cfour.de (accessed on 27 March 2023).
  115. Raffenetti, R.C. General contraction of Gaussian atomic orbitals: Core, valence, polarization, and diffuse basis sets; molecular integral evaluation. J. Chem. Phys. 1973, 58, 4452–4458. [Google Scholar] [CrossRef]
  116. Slater, J.C.; Phillips, J.C. Quantum Theory of Molecules and Solids: The Self-Consistent Field for Molecules and Solids, 1st ed.; McGraw-Hill: New York, NY, USA, 1974; Volume 4, pp. 1–583. [Google Scholar] [CrossRef]
  117. Vosko, S.H.; Wilk, L.; Nusair, M. Accurate spin-dependent electron liquid correlation energies for local spin density calculations: A critical analysis. Can. J. Phys. 1980, 59, 1200–1211. [Google Scholar] [CrossRef]
  118. Dirac, P.A.M. Note on exchange phenomena in the Thomas atom. Proc. Camb. Philos. Soc. 1930, 26, 376–385. [Google Scholar] [CrossRef]
  119. Slater, J.C. A simplification of the Hartree-Fock method. Phys. Rev. 1951, 81, 385–390. [Google Scholar] [CrossRef]
  120. Lutnæs, O.B.; Helgaker, T.; Jaszuński, M. Spin-spin coupling constants and triplet instabilities in Kohn-Sham theory. Mol. Phys. 2010, 108, 2579–2590. [Google Scholar] [CrossRef]
  121. Tomasi, J.; Mennucci, B.; Cancès, E. The IEF version of the PCM solvation method: An overview of a new method addressed to study molecular solutes at the QM ab initio level. J. Mol. Struct. THEOCHEM 1999, 464, 211–226. [Google Scholar] [CrossRef]
  122. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3094. [Google Scholar] [CrossRef]
  123. Ruden, T.A.; Lutnæs, O.B.; Helgaker, T. Vibrational corrections to indirect nuclear spin–spin coupling constants calculated by density-functional theory. J. Chem. Phys. 2003, 118, 9572–9581. [Google Scholar] [CrossRef]
  124. Ruud, K.; Åstrand, P.-O.; Taylor, P.R. Zero-point vibrational effects on proton shieldings: Functional-group contributions from ab initio calculations. J. Am. Chem. Soc. 2001, 123, 4826–4833. [Google Scholar] [CrossRef]
  125. Rusakov, Y.Y.; Rusakova, I.L. Relativistic heavy atom effect on 13C NMR chemical shifts initiated by adjacent multiple chalcogens. Magn. Reson. Chem. 2018, 56, 716–726. [Google Scholar] [CrossRef]
  126. Rusakov, Y.Y.; Rusakova, I.L.; Krivdin, L.B. Relativistic heavy atom effect on the 31P NMR parameters of phosphine chalcogenides. Part 1. Chemical shifts. Magn. Reson. Chem. 2018, 56, 1061–1073. [Google Scholar] [CrossRef]
  127. Rusakova, I.L.; Rusakov, Y.Y. On the heavy atom on light atom relativistic effect in the NMR shielding constants of phosphine tellurides. Magn. Reson. Chem. 2019, 57, 1071–1083. [Google Scholar] [CrossRef]
  128. Rusakov, Y.Y.; Rusakova, I.L. What Most Affects the Accuracy of 125Te NMR Chemical Shift Calculations. J. Phys. Chem. A 2020, 124, 6714–6725. [Google Scholar] [CrossRef] [PubMed]
  129. Rusakova, I.L.; Rusakov, Y.Y. On the Utmost Importance of the Basis Set Choice for the Calculations of the Relativistic Corrections to NMR Shielding Constants. Int. J. Mol. Sci. 2023, 24, 6231. [Google Scholar] [CrossRef] [PubMed]
  130. Maldonado, A.F.; Aucar, G.A. The UKB prescription and the heavy atom effects on the nuclear magnetic shielding of vicinal heavy atoms. Phys. Chem. Chem. Phys. 2009, 11, 5615–5627. [Google Scholar] [CrossRef] [PubMed]
  131. Sun, Q.; Liu, W.; Kutzelnigg, W. Comparison of restricted, unrestricted, inverse, and dual kinetic balances for four-component relativistic calculations. Theor. Chem. Acc. 2011, 129, 423–436. [Google Scholar] [CrossRef]
  132. Anderson, J.A.; Odom, J.D. Carbon-13 Chemical Shifts and 77Se-13C Spin-Spin Coupling Constants in Symmetrical Dialkyl Diselenides, Unsymmetrical Alkyl Methyl and Alkyl Phenyl Diselenides, and Related Alkyl Selenols. Organometallics 1988, 7, 267–271. [Google Scholar] [CrossRef]
  133. Poleschner, H.; Radeglia, R.; Kuprat, M.; Richter, A.M.; Fanghänel, E. Organylselenoacetylene und selenocyanate; 77Se- und 13C-NMR-chemische verschiebungen und 77Se-13C-spin-kopplungskonstanten. J. Organomet. Chem. 1987, 327, 7–15. [Google Scholar] [CrossRef]
  134. Schröder, H.F.; Müller, J. Dialkylphosphinsaureazide II. Z. Für Anorg. Allg. Chem. 1979, 451, 158–174. [Google Scholar] [CrossRef]
  135. García-Rodríguez, R.; Liu, H. A Nuclear Magnetic Resonance Study of the Binding of Trimethylphosphine Selenide to Cadmium Oleate. J. Phys. Chem. A 2014, 118, 7314–7319. [Google Scholar] [CrossRef]
  136. Glidewell, C.; Rankin, D.W.H.; Sheldrick, G.M. N.M.R. Studies of Mixed Group IV/Group VI Hydrides. Trans. Faraday Soc. 1969, 65, 1409–1412. [Google Scholar] [CrossRef]
  137. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar] [CrossRef]
  138. Jensen, F. Polarization consistent basis sets: Principles. J. Chem. Phys. 2001, 115, 9113–9125. [Google Scholar] [CrossRef]
  139. Jensen, F. Polarization consistent basis sets. II. Estimating the Kohn-Sham basis set limit. J. Chem. Phys. 2002, 116, 7372–7379. [Google Scholar] [CrossRef]
  140. Jensen, F.; Helgaker, T. Polarization consistent basis sets. V. The elements Si-Cl. J. Chem. Phys. 2004, 121, 3463–3470. [Google Scholar] [CrossRef] [PubMed]
  141. Jensen, F. Polarization consistent basis sets. VII. The elements K, Ca, Ga, Ge, As, Se, Br, and Kr. J. Chem. Phys. 2012, 136, 114107. [Google Scholar] [CrossRef] [PubMed]
  142. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 09, Revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2009. [Google Scholar]
  143. Gomes, A.S.P.; Saue, T.; Visscher, L.; Jensen, H.J.A.; Bast, R.; Aucar, A.; Bakken, V.; Dyall, K.G.; Dubillard, S.; Ekström, U.; et al. DIRAC, a Relativistic Ab Initio Electronic Structure Program, Release DIRAC19. 2019. Available online: http://www.diracprogram.org (accessed on 29 March 2023).
  144. Irkutsk Supercomputer Center of SB RAS. Irkutsk: ISDCT SB RAS. Available online: https://hpc.icc.ru (accessed on 29 March 2023).
Figure 1. Convergence of the one-bond SSCCs of SeH2 and Se=C upon saturation of cc-pVDZ+2f and cc-pVTZ+2f basis sets set on selenium atom: (a) 1J(77Se,1H); (b) 1J(77Se,13C).
Figure 1. Convergence of the one-bond SSCCs of SeH2 and Se=C upon saturation of cc-pVDZ+2f and cc-pVTZ+2f basis sets set on selenium atom: (a) 1J(77Se,1H); (b) 1J(77Se,13C).
Ijms 24 07841 g001
Figure 2. Mean absolute percentage errors (MAPEs) for all computed SSCCs involving selenium against experiment depending on the basis set used.
Figure 2. Mean absolute percentage errors (MAPEs) for all computed SSCCs involving selenium against experiment depending on the basis set used.
Ijms 24 07841 g002
Table 1. Modifications in original configuration of cc-pVXZ+2f (X = D, T).
Table 1. Modifications in original configuration of cc-pVXZ+2f (X = D, T).
Starting ConfigurationTotal ModificationResulting Configuration
cc-pVDZ+2f: (14s11p6d2f)+2s, +0p, +1d, +1fpecJ-1: (16s11p7d3f)
cc-pVTZ+2f: (20s13p9d2f)+2s, +0p, –1d, +2f, +1gpecJ-2: (22s13p8d4f1g)
Table 2. Contraction patterns with varying depths 1.
Table 2. Contraction patterns with varying depths 1.
StepBasis Set Contraction PatternDetailed Contraction PatternFinal Contraction Error Δ (in Relation to Current Reference Values), in Hz
Step 1
(s-shell)
pecJ-1 16 s   10s(8,8,1,1,1,1,1,1,1,1)0.133
16s 11s(7,7,1,1,1,1,1,1,1,1,1)0.024
16 s   12s(6,6,1,1,1,1,1,1,1,1,1,1)0.011
pecJ-2 22 s   14s(10,10,1,1,1,1,1,1,1,1,1,1,1,1)0.133
22s   15s(9,9,1,1,1,1,1,1,1,1,1,1,1,1,1)0.091
22 s   16s(8,8,1,1,1,1,1,1,1,1,1,1,1,1,1,1)0.077
Step 2
(p-shell)
pecJ-1 11 p   6p(7,7,1,1,1,1)1.080
11p   7p(6,6,1,1,1,1,1)0.095
11 p   8p(5,5,1,1,1,1,1,1)0.076
pecJ-2 13 p   7p(8,8,1,1,1,1,1)0.998
13p   8p(7,7,1,1,1,1,1,1)0.047
13 p   9p(6,6,1,1,1,1,1,1,1)0.032
Step 3
(d-shell)
pecJ-1 7 d   4d(4,1,1,1)0.768
7d   5d(3,1,1,1,1)0.123
7 d   6d(2,1,1,1,1,1)0.098
pecJ-2 8 d   4d(5,1,1,1)0.676
8d   5d(4,1,1,1,1)0.127
8 d   6d(3,1,1,1,1,1)0.124
1 Bold face indicates the chosen patterns.
Table 3. Final configurations of pecJ-n (n = 1, 2) for selenium atom.
Table 3. Final configurations of pecJ-n (n = 1, 2) for selenium atom.
Final ConfigurationNbas/Nbas(uc)ε1/2, in % 1
pecJ-1: [16s11p7d3f|11s7p5d3f]78/105ε1 = 0.35; ε2 = 0.08
pecJ-2: [22s13p8d4f1g|15s8p5d4f1g]101/138ε1 = 0.18; ε2 = 0.04
1ε1 and ε2 are the relative contraction errors of the pecJ-n basis sets calculated for 1J(77Se,1H) in SeH2 and 1J(77Se,13C) in Se=C calculated as: ε1 = (|1J(77Se,1H) [pecJ-n] − 1J(77Se,1H) [pecJ-n(uc)] |/|1J(77Se,1H) [pecJ-n(uc)]|)·100%, ε2 =(|1J(77Se,13C) [pecJ-n] − 1J(77Se,13C) [pecJ-n(uc)] |/|1J(77Se,13C) [pecJ-n(uc)]|) × 100%, respectively. Previously developed contracted pecJ-n basis sets were used on the hydrogen and carbon atoms.
Table 4. Formal operational costs of different methods provided by the pecJ-n and aug-cc-pVTZ-J basis sets for SeH2 1.
Table 4. Formal operational costs of different methods provided by the pecJ-n and aug-cc-pVTZ-J basis sets for SeH2 1.
Basis SetNN4
(DFT, RPA)
N5
(SOPPA, SOPPA (CC2))
N6
(SOPPA(CCSD), CCSD)
N8
(CCSDT)
pecJ-178(Se) + 2 × 11(H) = 1001.00 × 108,
R = 0.16
1.00 × 1010,
R = 0.10
1.00 × 1012,
R = 0.07
1.00 × 1016,
R = 0.03
pecJ-2101(Se) + 2 × 20(H) = 1413.95 × 108,
R = 0.65
5.57 × 1010,
R = 0.58
7.86 × 1012,
R = 0.52
1.56 × 1017,
R = 0.42
aug-cc-pVTZ-J117(Se) + 2 × 20(H) = 1576.08 × 108
R = 1.00
9.54 × 1010
R = 1.00
1.50 × 1013
R = 1.00
3.69 × 1017,
R = 1.00
1 R = (Npec-J-n/Naug-cc-pVTZ-J)k represents the ratio of the operational costs of methods with a particular scaling factor k provided by pecJ-n and aug-cc-pVTZ-J basis sets. For the aug-cc-pVTZ-J basis set, this factor is equal to 1.
Table 5. Different contributions to SSCCs 1 involving selenium in compounds 17 calculated at the CCSD level of theory with pecJ-n (n = 1, 2) basis sets.
Table 5. Different contributions to SSCCs 1 involving selenium in compounds 17 calculated at the CCSD level of theory with pecJ-n (n = 1, 2) basis sets.
MoleculeSSCCBasis SetJFCJSDJPSOJDSOTotal JCCSD
Me2Se2 (1)1J(77Se,13C)pecJ-1−77.2111.266.610.09−59.25
pecJ-2−81.2612.326.740.09−62.11
2J(77Se,13C)pecJ-14.131.602.00−0.027.71
pecJ-23.341.571.77−0.026.66
MeSe-C≡N (2)1J(77Se,13Csp3)pecJ-1−57.2312.179.320.05−35.69
pecJ-2−60.9512.539.250.05−39.12
1J(77Se,13Csp)pecJ-1−173.731.96−6.900.10−178.57
pecJ-2−182.112.21−6.170.10−185.97
MeSeH (3)1J(77Se,13C)pecJ-1−60.7212.2511.770.04−36.66
pecJ-2−64.6112.6812.000.04−39.89
Me2P(Se)Cl (4)1J(77Se,31P)pecJ-1−688.03−1.55−152.960.12−842.42
pecJ-2−700.90−0.74−151.210.12−852.73
C4H4Se (5)2J(77Se,1H)pecJ-147.50−0.15−4.61−0.2142.53
pecJ-248.97−0.16−4.92−0.2143.68
3J(77Se,1H)pecJ-18.590.260.56−0.369.05
pecJ-28.740.450.38−0.369.21
1J(77Se,13C)pecJ-1−59.912.31−25.790.09−83.3
pecJ-2−65.342.16−26.530.09−89.62
2J(77Se,13C)pecJ-1−1.832.92−1.20−0.02−0.13
pecJ-2−2.313.7−1.51−0.02−0.14
SePMe3 (6)1J(77Se,31P)pecJ-1−619.154.24−108.920.1−723.73
pecJ-2−621.696.15−104.120.1−719.56
SiH3SeH (7)1J(77Se,1H)pecJ-147.12−0.4823.120.1269.88
pecJ-248.41−0.0422.50.1170.98
2J(77Se,1H)pecJ-115.81−0.250.13−0.1815.51
pecJ-215.91−0.250.13−0.1815.61
1 All values are given in Hz.
Table 6. Testing pecJ-n (n = 1, 2) basis sets on the calculation of SSCCs 1 involving selenium in compounds 17 against experiment 2.
Table 6. Testing pecJ-n (n = 1, 2) basis sets on the calculation of SSCCs 1 involving selenium in compounds 17 against experiment 2.
MoleculeSSCCBasis SetJCCSDΔrel 3Δsolv 3Δvib 3JtotJexp
Me2Se2 (1)1J(77Se,13C)pecJ-1−59.25−7.021.51−0.9−65.66(−)73.87 ± 0.26
pecJ-2−62.11−5.481.56−0.9−66.93
aug-cc-pVTZ-J−63.56−4.511.54−0.9−67.43
2J(77Se,13C)pecJ-17.71−0.290.24−0.117.557.47 ± 0.26
pecJ-26.660.360.29−0.117.20
aug-cc-pVTZ-J6.55−0.640.28−0.116.08
MeSe-C≡N (2)1J(77Se,13Csp3)pecJ-1−35.69−5.760.24−1.00−42.21(−)52.3
pecJ-2−39.12−6.460.24−1.00−46.34
aug-cc-pVTZ-J−40.74−5.490.24−1.00−46.99
1J(77Se,13Csp)pecJ-1−178.57−37.39−5.84−3.22−225.02(−)237.8
pecJ-2−185.97−37.45−5.05−3.22−231.69
aug-cc-pVTZ-J−188.71−34.50−5.28−3.22−231.71
MeSeH (3)1J(77Se,13C)pecJ-1−36.66−5.692.30−1.90−41.95(−)48.26 ± 0.26
pecJ-2−39.89−6.312.39−1.90−45.71
aug-cc-pVTZ-J−41.26−5.192.32−1.90−46.03
Me2P(Se)Cl (4)1J(77Se,31P)pecJ-1−842.42−49.4311.82−4.08−884.11(−)838.4
pecJ-2−852.73−27.0114.27−4.08−869.55
aug-cc-pVTZ-J−861.31−28.4414.12−4.08−879.71
C4H4Se (5)2J(77Se,1H)pecJ-142.533.730.10−0.3546.0147.6
pecJ-243.683.970.21−0.3547.51
aug-cc-pVTZ-J44.463.900.19−0.3548.20
3J(77Se,1H)pecJ-19.05−0.320.23−0.328.649.4
pecJ-29.21−0.320.33−0.328.90
aug-cc-pVTZ-J9.53−0.370.32−0.329.16
1J(77Se,13C)pecJ-1−83.30−16.510.27−2.58−102.12(−)113.5
pecJ-2−89.62−17.610.62−2.58−109.19
aug-cc-pVTZ-J−89.18−17.950.60−2.58−109.11
2J(77Se,13C)pecJ-1−0.13−1.50−0.370.19−1.81(−)2.7
pecJ-2−0.14−1.59−0.370.19−1.91
aug-cc-pVTZ-J−1.21−1.65−0.380.19−3.05
SePMe3 (6)1J(77Se,31P)pecJ-1−723.73−56.5625.26−7.45−762.48(−)720.0
pecJ-2−719.56−35.3627.87−7.45−734.50
aug-cc-pVTZ-J−725.09−44.0127.49−7.45−749.06
SiH3SeH (7)1J(77Se,1H)pecJ-169.88−29.495.954.5950.9351.0 ± 0.1
pecJ-270.98−26.216.964.5956.32
aug-cc-pVTZ-J70.83−26.726.524.5955.22
2J(77Se,1H)pecJ-115.510.10−0.09−0.1615.3615.4 ± 0.2
pecJ-215.61−0.240.01−0.1615.22
aug-cc-pVTZ-J15.59−0.28−0.04−0.1615.11
1 All values are given in Hz. 2 Experimental values were taken from different sources: 1, 3—[132]; 2—[133]; 4—[134]; 5—[93]; 6—[135]; 7—[136]. 3 All corrections were calculated at the DFT(SVWN5) level of theory.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Rusakov, Y.Y.; Rusakova, I.L. New pecJ-n (n = 1, 2) Basis Sets for Selenium Atom Purposed for the Calculations of NMR Spin–Spin Coupling Constants Involving Selenium. Int. J. Mol. Sci. 2023, 24, 7841. https://doi.org/10.3390/ijms24097841

AMA Style

Rusakov YY, Rusakova IL. New pecJ-n (n = 1, 2) Basis Sets for Selenium Atom Purposed for the Calculations of NMR Spin–Spin Coupling Constants Involving Selenium. International Journal of Molecular Sciences. 2023; 24(9):7841. https://doi.org/10.3390/ijms24097841

Chicago/Turabian Style

Rusakov, Yuriy Yu., and Irina L. Rusakova. 2023. "New pecJ-n (n = 1, 2) Basis Sets for Selenium Atom Purposed for the Calculations of NMR Spin–Spin Coupling Constants Involving Selenium" International Journal of Molecular Sciences 24, no. 9: 7841. https://doi.org/10.3390/ijms24097841

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop