Skip to main content
Molecules logoLink to Molecules
. 2014 Dec 17;19(12):21253–21275. doi: 10.3390/molecules191221253

Largely Reduced Grid Densities in a Vibrational Self-Consistent Field Treatment Do Not Significantly Impact the Resulting Wavenumbers

Oliver M D Lutz 1,*, Bernd M Rode 2, Günther K Bonn 1, Christian W Huck 1
Editor: Derek J McPhee
PMCID: PMC6270979  PMID: 25525825

Abstract

Especially for larger molecules relevant to life sciences, vibrational self-consistent field (VSCF) calculations can become unmanageably demanding even when only first and second order potential coupling terms are considered. This paper investigates to what extent the grid density of the VSCF’s underlying potential energy surface can be reduced without sacrificing accuracy of the resulting wavenumbers. Including single-mode and pair contributions, a reduction to eight points per mode did not introduce a significant deviation but improved the computational efficiency by a factor of four. A mean unsigned deviation of 1.3% from the experiment could be maintained for the fifteen molecules under investigation and the approach was found to be applicable to rigid, semi-rigid and soft vibrational problems likewise. Deprotonated phosphoserine, stabilized by two intramolecular hydrogen bonds, was investigated as an exemplary application.

Keywords: VSCF, grid density, spectroscopy, anharmonicity, MP2, infrared, vibrational self-consistent field, phosphoserine

1. Introduction

The continuously increasing availability of computational resources as well as the development of accurate and efficient quantum chemical approaches have made computational vibrational spectroscopy an indispensable field complementing experimental techniques. Nowadays, almost every quantum chemical software package enables the analysis of second-order properties of the energy (e.g., IR and Raman absorptions, IR intensities, electric dipole polarizabilities, nuclear magnetic resonance chemical shifts, spin-spin coupling constants). Arguably, the prediction of IR and Raman absorptions is among the most important applications considering the prevalence of these technique. While most fundamental IR absorptions can be assigned quite satisfyingly on an empirical basis or via normal coordinate analysis, especially the near-infrared region is cluttered by overtones and combination excitations that are cumbersome to assign.

With regard to computational approaches, the harmonic oscillator approximation (HOA) is the most fundamental technique for obtaining vibrational spectroscopic data. However, due to the rigorous assumption introduced, in that the bond potential exhibits a harmonic shape, significant deviation from experiment is frequently observed. The most simple solution accounting for the lack of anharmonicities is to introduce empirical scaling factors that are multiplied with the harmonically approximated absorptions. Scott and Radom derived scaling factors for a vast number of ab initio and semi-empirical methods and a large number of basis sets [1,2]. It has to be stressed, however, that an empirical scaling may not be applied on a system-independent basis, even though the scaling factors have been derived for a rather large set of molecules. Noteworthy, a harmonic bond potential cannot be assumed a proper basis for considerations towards excitations involving more than one quantum of energy.

Especially due to these fundamental deficiencies, further corrective techniques have been developed that account for anharmonicities in an explicit manner. The two most prominent approaches are the vibrational self-consistent field (VSCF) method [3,4,5,6,7,8] and the vibrational second order perturbative ansatz (VPT2) [9,10]. VPT2 is a technique that relies on the computation of higher-order derivatives of the energy. Gaussian [11] is probably the most prominent commercial software package incorporating a VPT2 algorithm which relies on third and semi-diagonal fourth derivatives of the energy with respect to the nuclear coordinates. This method is applied regularly [12,13] but it may only be employed safely to vibrational problems where the harmonic part of the potential is dominant. Moreover, the computationally demanding generation of third and fourth derivatives limits the approach to rather small systems. VSCF on the other hand utilizes a separability ansatz which enables the anharmonicity of each mode to be accounted for by screening the respective potential in a point-wise manner. Mode interactions may be evaluated by computing d dimensional grids and the maximum possible dimensionality is determined by the number of vibrational degrees of freedom (N ) featured by the molecule under investigation. In a conventional VSCF calculation, each mode’s potential is characterized by 16 grid points [14]. In the past, scientific papers have been reported that employed grids of higher (r = 32) [15] and lower resolutions (r = 8) [16,17,18]. Roy et al. [19] recently presented data based on a variety of grid densities ranging from 8 to 16 points. Since they reported overall deviations summed over a set of reference molecules, one could conclude that this rather diverse choice of grid resolutions does not impose a significant error on the wavenumbers. However, to date there still seem to exist ambiguities with regard to a generally applicable grid density that is computationally feasible. Hence, and considering the massive computational effort involved in high resolution VSCF calculations, it seemed promising to conduct a systematic study addressing these issues. 15 reference molecules are investigated in detail and an application to the lowest energy conformer of phosphoserine is presented. At this point it seems worth mentioning that while many quantum chemical software packages such as GAMESS [20], NWChem [21] and MOLPRO [22] incorporate VSCF implementations, GAMESS is the program employed in this work due to the fact that it is freely available and compatible with the most prominent computer operating systems.

2. Methods

In this section, a brief overview describing the procedures underlying a VSCF evaluation is given. Particularities of the involved techniques are presented while reference to the original literature is given for further details.

2.1. Energy Minimization

A prerequisite for every computational spectroscopic analysis is a proper structural ensemble. For the majority of cases, the user would want to obtain absorption data corresponding to an equilibrium structure and thus, an energetically favorable reference geometry is required. This implies that at least a local energy minimum (or the global energy minimum) is obtained and this is reflected by a nearly zero gradient of the energy g(q)∆q with respect to the nuclear coordinates q. While the program’s default criteria suffice for energetical and structural considerations in many cases, a VSCF calculation requires a thoroughly minimized geometry. This condition is realized by setting up rigorous cutoff tolerances for g(q)∆q. We considered a molecular geometry properly minimized when each gradient contribution is smaller than 0.000001 Eh· Bohr−1. Every calculation reported herein was performed at the Møller-Plesset level of theory, accounting for electron correlation effects via a perturbative ansatz [23,24,25]. The frozen-core approximation [26] was employed, explicitly correlating all but the core electrons. Dunning’s correlation-consistent polarized basis set of triple-ζ quality has been utilized since it can be considered a routinely employed set of functions that is known to deliver results of acceptable accuracy [27]. The appropriate point group was imposed onto the molecular geometry where applicable in order to make use of the efficient symmetry optimized SCF algorithm in GAMESS.

2.2. Second Derivatives of the Energy

A potential V(q) may be expressed as a Taylor series

V(q)=V0(q)+g(q)Δq+12ΔqTΔq+ (1)

and for the case that the reference geometry resembles an energy minimum and assuming a harmonic potential shape, all but the first and third terms vanish. This simplified approach enables the evaluation of the elements of ℍr,s by analytical or numerical analysis:

r,s=2V(q)qrqs (2)

A transformation into mass-weighted coordinates and diagonalization of the Hessian matrix ℍr,s yields the eigenvectors corresponding to the normal modes. At this point, computational spectroscopic data at the HOA level is available. It seems worth mentioning that the further description of extended vibrational analysis is based on mass-weighted normal coordinates or other non-redundant coordinates for the sake of clarity, whereas the expansion of the potential energy surfaces was realized in rectilinear coordinates (Section 3.1). VSCF calculations require a Hessian corresponding to the equilibrium geometry, as will be discussed in the next paragraph. The Hessian matrices in this work have been obtained semi-numerically involving two displacements of ±0.01 Bohr about the reference geometry for each atom.

2.3. The VSCF Routine

Even though VSCF accounts for anharmonicities explicitly, it still requires ℍr,s as a reference state. This is owed to the fact, that VSCF assumes a molecule’s vibrational wave function Ψ(Q1,..,N) to be separable into single mode wave functions ψi(n):

Ψ(Q1,,N)=i=1Nψi(n)(Qi) (3)

While this product ansatz can be implemented rather efficiently, it inherently limits the accuracy of the VSCF technique since it represents decoupled vibrational modes. There are, besides the methods mentioned in Section 2.4, techniques available which are able to circumvent this limitation [28,29]. Introducing the variational principle [30,31], the single mode VSCF equation [6,32] is formulated as:

[122Qi2+Vi(n)¯(Qi)]ψi(n)(Qi)=εi(n)ψi(n)(Qi) (4)

where Vi(n)¯(Qi) is mode Qi’s effective potential. Equation (4) neglects parts of the full Watson Hamiltonian for non-linear molecules [33] which accounts for N vibrational degrees of freedom:

Ĥ=12i=1N2Qi2+V(Q1,,N)+12αβπˆαμαβπˆβ18αμαα (5)

The second and third part of the Hamiltonian represent the generalized inverse of the effective moment of inertia µαβ and the vibrational angular momentum operator πˆα. Given that the full Watson Hamiltonian is cumbersome to implement computationally, most implementations resort to neglecting the last two terms of Ĥ. The error introduced is small especially for fundamental excitations as has been shown by Bowman et al. [34].

The VSCF equations are solved in an iterative manner until self-consistency is achieved. Illustratively, VSCF may be seen as an approach where each vibration is influenced by the mean field of the surrounding modes. However, especially when certain mode interactions exhibit unique properties, this mean field ansatz does no longer properly describe the vibrational problem sufficiently well. A number of post-VSCF techniques accounting for this deficiency have been presented and they are discussed in the subsequent section.

Since the VSCF technique involves a grid-based screening of bond potentials and their interactions, a discussion of the so-called hierarchical expansion is justified. For a system exhibiting N normal modes, the VSCF expansion is denoted as:

V(Q1,,N)=i=1NVidiag(Qi)+i<jNVi,jpairs(Qi,Qj)+i<j<kNVi,j,ktriples(Qi,Qj,Qk)+ (6)

A complete interaction scheme is realized by imposing N -dimensionality onto Equation (6) but obviously, such a treatment scales unfavorably with the studied system size. Noteworthy, most applications of VSCF theory truncate the expansion largely by including only i=1NVidiag(Qi) and i<jNVi,jpairs(Qi,Qj) contributions [35,36,37,38]. This simplification is valid for a large number of applications and the inclusion of higher-order interaction terms is hardly justifiable for other than the smallest molecules or for cases, where convergence is only achieved when including such terms. Programs such as MULTIMODE [34,39], Molpro [22] and MIDAScpp enable coupling orders of d > 3 yielding IR bands of exceptional accuracy [7,40]. In this article, we confine our discussion to i=1NVidiag(Qi) and i<jNVi,jpairs(Qi,Qj) contributions since this study aims at larger molecules where an incorporation of higher coupling terms would be prohibitive. The number of grid points due (Np) is calculated via Equation (7) .

Np=r×N+N×(N1)2×r2 (7)

Assuming r = 16 and taking glycine as an example, the pair-wise approximation would require 71,040 points to be computed while a VSCF calculation involving also three-mode interactions already gives rise to 8,361,344 energy evaluations, a nearly 120 fold increase in computational burden. Herein, we will examine the impact of the grid density on the quality of the absorption data by conducting VSCF calculations involving between 6 and 16 grid points per mode. For the displacements underlying the PES scan, a symmetric grid range of [−4ωi0.5, +4ωi0.5], with ωi being the harmonic frequency of the i-th normal mode, was chosen. The grid points within these boundaries have been set-up in an equidistant manner which implies that for even grid densities (i.e., 6, 8, 10, 12, 14 and 16 points), equilibrium is located between the two innermost grid points. For odd grid densities (i.e., 7, 9, 11, 13 and 15 points), equilibrium is described by one distinct grid point. For every resolution r < 16, the grid points are interpolated to the original resolution of r = 16 by a polynomial fit.

2.4. Extensions of the VSCF Approach

As mentioned earlier, the VSCF assumption does not hold for many applications. Therefore, a perturbative approach has been suggested, accounting for the error introduced by the mean-field VSCF ansatz [15,41,42]. A prerequisite is that the difference between the "true" energy and the mean-field VSCF energy is small. An expansion known from electronic structure theory [23] is then formulated for the energy and wave function of a state n that is truncated at second order:

EnEn0+λEn1+λ2En2 (8)
ΨnΨn0+λΨn1+λ2Ψn2 (9)

with H = Hn0 + λ∆V and an insertion of En and Ψn in H Ψn = EnΨn, the energy contribution at second order may be formulated as:

En2=mnΨn0|ΔV|Ψm0Ψm0|ΔV|Ψn0En0Em0 (10)

Herein, Ψn0 and Ψm0 are the unperturbed vibrational product wave functions and En0 and Em0 are the unperturbed VSCF energies. This technique, which is commonly abbreviated as second order perturbation theory augmented (PT2)-VSCF or correlation-corrected (CC)-VSCF, has proven useful for many applications. However, since obtaining correlation-corrected VSCF energies requires significant computational resources, the basic PT2-VSCF approach is limited to small systems. Due to the inherent need for an efficient solution of Equation (10), Gerber’s work group developed a solution based on the assumption of orthogonal vibrational single mode wave functions (Equation (3)) which leads to the annihilation of diagonal elements in Equation (10) [43]. Tests involving Glyn peptides showed that the runtime improvement is greater for larger molecules: for monomeric Gly, a speedup factor close to 6 was observed while the correlation-corrected wavenumbers of tetraglycine have been evaluated more than 16 times faster as with the conventional PT2-VSCF ansatz. This acceleration technique, and also the fact that this accelerated PT2-VSCF technique is readily implemented in GAMESS [20], are probably the main reasons for PT2-VSCF being a routinely employed VSCF correction. It has to be stressed, however, that the evaluation of the VSCF equations with its further corrections may not be confused with the preceding and highly demanding evaluation of the potential energy grid (vide supra).

Problems may arise when degenerate vibrational states are present. PT2-VSCF can fail here due to a close to zero denominator in Equation (4) which can lead to a largely overestimated perturbative correction. Hence, the degenerate PT2-VSCF method has been developed [44] but to date it has been implemented in GAMESS exclusively for degeneracies arising from fundamental excitations. A more generally available simple solution available in the GAMESS program code is that contributions involving denominators falling below a critical value are excluded from the treatment. Most applications reporting PT2-VSCF derived data rely on this simplification and therefore, we will also confine our calculations to this simplified technique.

Besides perturbation theory augmented VSCF, significant effort is put into post-VSCF techniques involving configuration interaction [45,46,47,48], coupled-cluster [49,50,51] and multi-configurational SCF theory [52,53]. While such methods yield highly accurate data, they are to date only applicable to small vibrational problems with less than 20 atoms.

Due to the popularity of the PT2-VSCF method and the fact that results of good quality at manageable computational cost are available also for larger molecules, all data presented in this paper is corrected exclusively with this perturbative ansatz.

3. Results and Discussion

The fifteen molecules under investigation gave rise to 176 distinct vibrational degrees of freedom, considering all fundamental stretching, deformation and torsional vibrations. PT2-VSCF can yield questionable or sometimes even divergent results for very low-lying and floppy torsions, which is due to the fact that the PES expansion is conventionally carried out in Cartesian coordinates [18,19,54]. The error induced through an anharmonic correction of such a vibration can exceed the boundaries of accuracy known from the HOA [54]. Hence, the usual procedure is to either treat such vibrations harmonically or to describe the underlying displacements in internal coordinates. A substitution for physically more meaningful internal coordinates was proposed by Njegic and Gordon [54] and could be shown to yield good results for formamide and thioformamide [55] as well as for H2O2 but by introducing a new expansion technique for the kinetic energy operator [56]. Nonetheless, setting up internal coordinates that properly describe the displacements underlying a VSCF treatment is by no means a trivial task. The GAMESS code is able to identify each normal mode’s contributions to particular internal coordinates [57], but the user still has to input a balanced description of each vibrational degree of freedom which can become unmanageably difficult for larger systems. Hence, and especially when large molecules are investigated, the majority of users of VSCF theory resort to a PES expansion in Cartesian coordinates and the contributions of critical torsions are omitted when self-consistency is not achieved.

For glycine (C2H5NO2), three normal modes (i.e., the N-Cα-Ccarb-O torsion, the NH2 group torsion and δCcarbO2,oop) had to be excluded from the VSCF treatment since they are known to lead to divergence during a perturbation theory corrected VSCF evaluation [19]. Similarly, for methanol (CH3OH) the CO axis torsion has been omitted from the PT2-VSCF treatment due to an inadequate description of this particularly floppy torsional mode within the VSCF framework [58]. Dimethylether also exhibits two floppy torsions involving the C-O axes that have been excluded likewise. For ethane, one normal mode near 290 cm−1 [59,60,61,62] was omitted due to its floppy character but the other six missing modes did arise from degenerate states in νCH3,as, δCH3,as and ρCH3. Experimentally, these degeneracies cannot be distinguished and since the PT2-VSCF derived values did not exhibit significant numerical discrepancies, each pair of degeneracies is presented as a single mean value for the sake of visibility.

3.1. Performance of the PT2-VSCF Approach

The computationally obtained absorption data are compared to experimental data in Table 1. The column headers indicate the employed number of grid points during the VSCF evaluations. As a measure of quality, the mean absolute percentage error µ for each molecule and each grid density is calculated according to Equation (11):

μ=100Ni=1N|ExptiVSCFiExpti| (11)

Table 1.

Computed and experimental absorption data and the calculated values for µ in %. n.a. means “not applicable” and n.o. means “not observed”.

H2O (C2v) Sym. 6 7 8 9 10 11 12 13 14 15 16 Expt. [59,71,72]
νOH2 ,as B2 2875 3997 3731 3772 3766 3766 3765 3765 3765 3765 3765 3756
νOH2 ,s A1 3041 3785 3657 3691 3684 3683 3683 3682 3682 3682 3682 3652
δOH2 A1 1216 1688 1570 1590 1586 1586 1586 1586 1586 1586 1586 1595
μ 21.3 5.3 0.8 0.6 0.6 0.6 0.5 0.5 0.5 0.5 0.5 n.a.
CO2 (Dh) Sym. 6 7 8 9 10 11 12 13 14 15 16 Expt. [59,73,74]
νCO2,as g+ 1791 2547 2359 2391 2386 2386 2385 2385 2385 2385 2385 2349
νCO2,s u+ 993 1408 1255 1319 1316 1315 1315 1315 1315 1315 1315 1285
δCO2 μ 487 699 645 654 653 653 653 653 653 653 653 667
δCO2 μ 487 699 645 654 653 653 653 653 653 653 653 667
μ 25.1 6.9 2.3 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0 n.a.
CH2O (C2v) Sym. 6 7 8 9 10 11 12 13 14 15 16 Expt. [59,75,76,77]
νCH2,as B1 2184 3000 2809 2837 2833 2833 2833 2832 2832 2832 2832 2843
νCH2,s A1 2250 2905 2801 2823 2818 2818 2818 2817 2817 2817 2817 2782
νCO A1 1312 1855 1719 1745 1741 1740 1740 1740 1740 1740 1740 1746
δCH2 A1 1157 1614 1502 1521 1518 1518 1518 1518 1518 1518 1518 1500
ρCH2 B1 945 1340 1240 1257 1254 1254 1254 1254 1254 1254 1254 1249
ωCH2 B2 891 1250 1162 1176 1174 1173 1173 1173 1173 1173 1173 1167
μ 23.0 6.4 0.8 0.8 0.7 0.7 0.7 0.7 0.7 0.7 0.7 n.a.
C2H2 (Dh) Sym. 6 7 8 9 10 11 12 13 14 15 16 Expt. [59,60,78,79,80]
νC2H2,s g+ 2699 3490 3375 3390 3384 3380 3384 3384 3384 3384 3384 3374
νC2H2,as u+ 2523 3405 3199 3233 3228 3227 3227 3227 3227 3227 3227 3289
νCC g+ 1469 2046 1912 1937 1932 1932 1932 1932 1932 1932 1932 1974
δC2H2,s u 557 792 733 746 743 743 743 743 743 743 743 730
δC2H2,s u 557 792 733 746 743 743 743 743 743 743 743 730
δC2H2,as g 443 638 589 602 598 598 598 598 598 598 598 612
δC2H2,as g 443 638 589 602 598 598 598 598 598 598 598 612
μ 24.5 5.2 2.0 1.7 1.8 1.8 1.8 1.8 1.8 1.8 1.8 n.a.
HCOOH (Cs) Sym. 6 7 8 9 10 11 12 13 14 15 16 Expt. [59,81,82,83,84]
νOH A' 2841 3682 3530 3559 3550 3557 3552 3552 3552 3552 3552 3570
νCH A' 2246 3134 2959 2983 2978 2977 2977 2977 2977 2977 2977 2943
νC=O A' 1337 1910 1764 1790 1786 1786 1786 1785 1785 1785 1785 1770
δCH A' 1043 1480 1367 1384 1381 1381 1381 1381 1381 1381 1381 1387
δOH A' 982 1367 1265 1280 1277 1277 1277 1277 1277 1277 1277 1387
νC-O A' 848 1148 1077 1089 1086 1086 1086 1086 1086 1086 1086 1105
δCH A" 786 1114 1029 1042 1040 1040 1040 1040 1040 1040 1040 1033
δOH,oop A" 515 631 588 623 609 608 609 609 609 609 609 638
δOCO A' 469 659 611 619 617 617 617 617 617 617 617 625
μ 22.8 6.0 2.1 1.4 1.7 1.7 1.7 1.7 1.7 1.7 1.7 n.a.
CH4 (Td) Sym. 6 7 8 9 10 11 12 13 14 15 16 Expt. [59,60,85,86]
νCH3,as F2 2333 3210 3000 3035 3029 3029 3028 3028 3028 3028 3028 3019
νCH3,as F2 2299 3177 2990 3020 3015 3015 3014 3014 3014 3014 3014 3019
νCH3,as F2 2330 3209 2999 3033 3028 3027 3027 3027 3027 3027 3027 3019
νCH4,s A1 2588 2968 2948 2943 2945 2944 2943 2943 2943 2943 2943 2917
δCH4,as E 1166 1641 1523 1542 1539 1539 1539 1539 1539 1539 1539 1534
δCH4,as E 1166 1641 1523 1542 1539 1539 1539 1539 1539 1539 1539 1534
δCH3,s F2 993 1390 1293 1309 1306 1306 1306 1306 1306 1306 1306 1306
δCH3,s F2 992 1390 1293 1309 1306 1306 1306 1306 1306 1306 1306 1306
δCH3,s F2 993 1390 1294 1309 1307 1307 1307 1306 1306 1306 1306 1306
µ 22.3 5.9 0.9 0.4 0.3 0.3 0.3 0.3 0.3 0.3 0.3 n.a.
CH3Cl (C3v) Sym. 6 7 8 9 10 11 12 13 14 15 16 Expt. [59,87,88,89,90]
νCH3,as E 2338 3203 3002 3035 3030 3029 3029 3029 3029 3029 3029 3039
νCH3,as E 2306 3192 2974 3019 3016 3015 3015 3015 3015 3015 3014 3039
νCH3,s A1 2339 3074 2964 2989 2987 2986 2986 2985 2985 2985 2985 2937
δCH3,as E 1106 1555 1445 1463 1460 1460 1460 1460 1460 1460 1460 1452
δCH3,as E 1106 1556 1444 1462 1460 1459 1459 1459 1459 1459 1459 1452
δCH3,s A1 1036 1464 1357 1375 1372 1372 1371 1371 1371 1371 1371 1355
ρCH3 E 773 1099 1017 1031 1028 1028 1028 1028 1028 1028 1028 1017
ρCH3 E 776 1099 1020 1033 1031 1031 1031 1031 1031 1031 1031 1017
νCCl A1 585 802 749 760 758 758 758 758 758 758 758 732
µ 23.0 7.0 0.9 1.4 1.2 1.2 1.2 1.2 1.2 1.2 1.2 n.a.
CH3OH (Cs) Sym. 6 7 8 9 10 11 12 13 14 15 16 Expt. [59,60,91,92,93]
νOH A’ 2812 3900 3646 3694 3687 3686 3686 3686 3686 3686 3686 3681
νCH3,as A’ 2301 3198 2984 3035 3030 3029 3028 3028 3028 3028 3028 3000
νCH3,as A” 2230 3091 2881 2912 2908 2908 2907 2907 2907 2907 2907 2960
νCH3,s A’ 2363 2951 2849 2957 2947 2942 2941 2941 2941 2941 2941 2844
δCH3,as A’ 1130 1587 1475 1493 1490 1490 1490 1490 1490 1490 1490 1477
δCH3,as A” 1116 1574 1459 1478 1475 1475 1475 1475 1474 1474 1474 1477
δCH3,s A’ 1104 1563 1447 1466 1463 1463 1463 1463 1463 1463 1463 1455
δOH A’ 1020 1458 1342 1362 1359 1359 1358 1358 1358 1358 1358 1345
ρCH3 A” 843 1178 1069 1083 1081 1081 1080 1080 1080 1080 1080 1060
νCO A’ 789 1060 1015 1028 1026 1026 1025 1025 1025 1025 1025 1033
µ 22.9 6.4 0.9 1.3 1.2 1.2 1.2 1.2 1.1 1.1 1.1 n.a.
CH3CN (C3v) Sym. 6 7 8 9 10 11 12 13 14 15 16 Expt. [59,94,95,96,97,98,99,100]
νCH3,as E 2281 3153 2940 2972 2967 2967 2967 2966 2966 2966 2966 3009
νCH3,as E 2322 3172 2976 2994 2990 2989 2989 2988 2988 2988 2988 3009
νCH3,s A1 2322 3059 2951 2970 2968 2967 2967 2967 2967 2967 2967 2954
νCN A1 1656 2308 2152 2178 2174 2174 2173 2173 2173 2173 2173 2267
δCH3,as E 1099 1547 1438 1456 1453 1453 1453 1453 1453 1453 1453 1454
δCH3,as E 1099 1550 1437 1456 1453 1453 1453 1453 1453 1453 1453 1454
δCH3 ,s A1 1047 1487 1378 1397 1394 1394 1394 1393 1393 1393 1393 1389
ρCH3 E 788 1121 1037 1052 1049 1049 1049 1049 1049 1049 1049 1041
ρCH3 E 789 1122 1038 1053 1050 1050 1050 1050 1050 1050 1050 1041
νCC A1 692 983 923 932 930 930 930 930 930 930 930 920
δCCN E 270 395 361 367 366 366 366 366 366 366 366 361
δCCN E 268 395 359 365 364 364 364 364 364 364 364 361
µ 24.4 6.4 1.1 1.1 1.0 1.0 1.0 1.0 1.0 1.0 1.0 n.a.
C2H4(D2h) Sym. 6 7 8 9 10 11 12 13 14 15 16 Expt. [59,60,101,102]
νCH2,as B2u 2399 3358 3122 3161 3155 3154 3154 3154 3154 3154 3154 3106
νCH2,as B1g 2379 3328 3095 3133 3127 3126 3126 3126 3126 3126 3126 3103
νCH2,s Ag 2297 3152 3047 3055 3052 3051 3051 3051 3051 3051 3051 3026
νCH2,s B3u 2318 3253 3023 3060 3055 3054 3054 3054 3054 3054 3054 2989
νCC Ag 1273 1750 1628 1648 1644 1644 1644 1644 1644 1644 1644 1623
δCH2 B3u 1095 1554 1439 1458 1455 1455 1455 1455 1455 1455 1455 1444
δCH2 Ag 1051 1432 1345 1361 1359 1358 1358 1358 1358 1358 1358 1342
ρCH2 B1g 922 1318 1217 1235 1232 1232 1232 1232 1232 1232 1232 1236
τCH2 Au 795 1129 1046 1060 1058 1058 1058 1058 1058 1058 1058 1023
ωCH2 B1u 730 1036 960 973 971 971 971 971 971 971 971 949
ωCH2 B2g 712 1013 939 951 949 949 949 949 949 949 949 943
ρCH2 B2u 619 893 824 836 834 834 834 834 834 834 834 826
µ 23.4 7.7 0.8 1.5 1.4 1.4 1.3 1.3 1.3 1.3 1.3 n.a.
C2H4O (C2v) Sym. 6 7 8 9 10 11 12 13 14 15 16 Expt. [59,60,103,104]
νCH2,as B2 2372 3313 3083 3120 3114 3114 3113 3113 3113 3113 3113 3065
νCH2,as A2 2361 3296 3067 3104 3098 3098 3097 3097 3097 3097 3097 n.o.
νCH2,s A1 2378 3097 2939 3047 3064 3058 3057 3056 3056 3056 3056 3018
νCH2,s B1 2295 3214 2988 3024 3019 3018 3018 3018 3018 3018 3018 3006
δCH2,s A1 1153 1634 1500 1518 1515 1515 1515 1515 1515 1515 1515 1498
δCH2,as B1 1121 1590 1472 1492 1489 1489 1489 1489 1489 1489 1489 1472
νCC A1 1070 1328 1277 1288 1286 1285 1285 1285 1285 1285 1285 1270
ρCH2 A2 876 1253 1157 1174 1171 1171 1171 1171 1171 1171 1171 n.o.
τCH2 B2 871 1244 1149 1166 1163 1163 1163 1163 1163 1163 1163 1142
ωCH2 B1 859 1229 1135 1152 1149 1149 1149 1149 1149 1149 1149 1151
ωCH2 A1 882 1203 1128 1141 1139 1138 1138 1138 1138 1138 1138 1148
τCH2 A2 783 1117 1032 1047 1044 1044 1044 1044 1044 1044 1044 n.o.
νOC2,s A1 669 946 877 889 887 887 887 887 887 887 887 877
νOC2,as B1 624 891 822 833 832 831 831 831 831 831 831 872
ρCH2 B2 615 885 816 829 827 827 827 827 827 827 827 821
µ 23.3 6.5 1.2 1.4 1.4 1.3 1.3 1.3 1.3 1.3 1.3 n.a.
CH3NH2(Cs) Sym. 6 7 8 9 10 11 12 13 14 15 16 Expt. [59,105,106,107]
νNH2,as A” 2629 3517 3318 3348 3343 3343 3343 3343 3343 3343 3343 3411
νNH2,s A’ 2539 3431 3362 3352 3347 3347 3346 3346 3346 3346 3346 3349
νCH3,as A” 2283 3110 2919 2947 2943 2942 2942 2942 2942 2942 2942 2985
νCH3,as A’ 2283 3079 2907 2960 2957 2956 2956 2956 2955 2955 2955 2963
νCH3,s A’ 2201 3010 2833 2889 2911 2906 2904 2904 2904 2904 2904 2816
δNH2 A’ 1223 1733 1596 1610 1608 1608 1608 1608 1608 1608 1608 1642
δCH3,as A” 1132 1601 1483 1503 1500 1494 1500 1500 1500 1500 1500 1481
δCH3,as A’ 1119 1581 1467 1487 1484 1483 1483 1483 1483 1483 1483 1463
δCH3,s A’ 1083 1534 1421 1440 1437 1437 1436 1436 1436 1436 1436 1450
ρCH3 A” 998 1420 1314 1331 1328 1328 1328 1328 1328 1328 1328 n.o.
ρCH3 A’ 891 1237 1152 1169 1167 1167 1167 1167 1167 1167 1167 1144
νCN A’ 845 1109 1045 1059 1057 1057 1057 1057 1057 1057 1057 1050
τNH2 A” 725 1052 967 981 979 979 979 979 979 979 979 n.o.
ωNH2 A’ 637 909 840 852 851 850 850 850 850 850 850 816
CN axis torsion A” 245 316 345 337 307 313 316 313 313 314 314 304
µ 22.8 5.9 2.4 2.3 1.6 1.7 1.8 1.7 1.7 1.7 1.7 n.a.
C2H6(D3d) Sym. 6 7 8 9 10 11 12 13 14 15 16 Expt. [59,60,61,62,108,109]
νCH3,as Eu 2309 3221 2999 3037 3031 3030 3030 3030 3030 3030 3030 2985
νCH3,as Eg 2290 3185 2967 3011 3007 3006 3006 3006 3006 3006 3006 2969
νCH3,s A1g 2470 3006 2950 2981 3004 2996 2995 2995 2994 2994 2994 2954
νCH3,s A2u 2250 3172 2943 2982 2976 2975 2975 2975 2975 2975 2975 2896
δCH3,as Eg 1123 1596 1474 1494 1491 1491 1491 1491 1491 1491 1491 1468
δCH3,as Eu 1123 1593 1475 1496 1493 1493 1492 1492 1492 1492 1492 1469
δCH3,s A1g 1059 1501 1393 1411 1408 1408 1408 1408 1408 1408 1408 1388
δCH3,s A2u 1042 1482 1372 1392 1389 1389 1388 1388 1388 1388 1388 1379
ρCH3 Eg 909 1294 1197 1215 1212 1212 1212 1212 1212 1212 1212 1190
νCC A1g 818 1051 1003 1010 1008 1008 1008 1008 1008 1008 1008 995
ρCH3 Eu 620 899 828 842 839 839 839 839 839 839 839 822
µ 22.3 7.5 0.6 1.7 1.6 1.6 1.6 1.6 1.6 1.6 1.6 n.a.
CH3OCH3(C2v) Sym. 6 7 8 9 10 11 12 13 14 15 16 Expt. [59,60,110,111]
νCH3,as A1 2339 3142 2945 3016 3013 3012 3012 3011 3011 3011 3011 2996
νCH3,as B1 2313 3145 2949 2976 2976 2976 2975 2975 2975 2975 2975 2996
νCH3,as A2 2242 3126 2906 2942 2936 2936 2936 2936 2936 2936 2935 2952
νCH3,as B2 2236 3117 2899 2934 2928 2928 2928 2928 2928 2928 2928 2925
νCH3,s A1 2206 2949 2825 2869 2852 2856 2855 2855 2855 2855 2854 2817
νCH3,s B1 2200 3080 2861 2897 2891 2891 2891 2891 2891 2891 2891 2817
δCH3,as A1 1142 1597 1488 1506 1503 1503 1503 1503 1503 1503 1503 1464
δCH3,as B2 1122 1587 1475 1494 1492 1491 1491 1491 1491 1491 1491 1464
δCH3,as B1 1118 1589 1475 1494 1491 1491 1491 1491 1491 1491 1491 1464
δCH3,as A2 1111 1579 1466 1485 1482 1482 1482 1482 1482 1482 1482 1464
δCH3,s A1 1114 1574 1460 1479 1476 1475 1475 1475 1475 1475 1475 1452
δCH3,s B1 1082 1541 1426 1445 1442 1442 1442 1442 1442 1442 1442 1452
ρCH3 A1 944 1349 1248 1265 1263 1262 1262 1262 1262 1262 1262 1244
ρCH3 B1 897 1275 1181 1197 1194 1194 1194 1194 1194 1194 1194 1227
ρCH3 B2 893 1273 1180 1194 1195 1190 1189 1188 1188 1188 1188 1179
ρCH3 A2 871 1245 1153 1168 1166 1166 1166 1166 1166 1166 1166 1150
νC2O,as B1 839 1207 1113 1129 1127 1127 1126 1126 1126 1126 1126 1102
νC2O,s A1 780 988 941 950 949 949 949 948 948 948 948 928
δC2O A1 333 452 423 429 428 428 428 428 428 428 428 418
µ 23.0 7.2 1.1 1.7 1.6 1.5 1.5 1.5 1.5 1.5 1.5 n.a.
C2H5NO2(Cs) Sym. 6 7 8 9 10 11 12 13 14 15 16 Expt. [112,113]
νOH A’ 2773 3737 3524 3557 3552 3551 3551 3550 3550 3550 3550 3560
νNH2,as A” 2609 3556 3337 3370 3365 3364 3364 3364 3364 3364 3364 3410
νNH2,s A’ 2629 3480 3346 3357 3352 3351 3351 3351 3351 3350 3350 n.o.
νCH2,as A” 2268 3101 2907 2937 2932 2932 2932 2931 2931 2931 2931 n.o.
νCH2,s A’ 2265 3059 2982 2957 2953 2952 2952 2952 2952 2952 2952 2958
νC=O A’ 1346 1921 1777 1802 1798 1798 1798 1798 1798 1798 1798 1779
δNH2 A’ 1228 1750 1602 1620 1617 1617 1617 1617 1617 1617 1617 1630
δCH2 A’ 1085 1532 1421 1440 1437 1437 1437 1437 1437 1437 1437 1429
νC−O A’ 1043 1482 1372 1391 1388 1388 1388 1387 1387 1387 1387 1373
τCH2,NH2 A” 1030 1466 1355 1374 1371 1371 1371 1371 1370 1370 1370 n.o.
ωCH2 A’ 968 1382 1275 1293 1290 1290 1290 1290 1290 1290 1290 n.o.
δCCN,oop A” 881 1262 1164 1181 1178 1178 1178 1178 1178 1178 1178 n.o.
νCN A’ 905 1245 1156 1172 1169 1169 1169 1169 1169 1169 1169 1136
δCO2 A’ 832 1186 1097 1113 1110 1110 1110 1110 1110 1110 1110 1101
δCNH2 A’ 716 1002 922 938 935 935 934 934 934 934 934 907
δC=O,oop A” 685 990 910 924 922 922 922 922 922 922 922 883
νC −C A’ 658 863 815 823 822 822 822 822 822 822 822 801
δC=O,ip A’ 486 678 629 637 636 636 636 636 636 636 636 619
C-O axis torsion A” 369 529 475 485 483 483 483 483 483 483 483 500
O-C2N axis shear A’ 365 495 461 467 466 466 466 466 466 466 466 463
O=C2N axis shear A’ 194 286 262 266 266 266 266 266 266 266 266 n.o.
µ 22.8 7.5 1.5 1.8 1.7 1.7 1.7 1.7 1.7 1.7 1.7 n.a.

The obtained values for µ are generally smaller than 2%, which is in good agreement with the recent work by Roy et al. [19] who concluded that MP2/cc-pVTZ based PT2-VSCF evaluations exhibit a mean unsigned error of under 2%. MP2 is a topical ab initio method that accounts for electron correlation effects to a certain extent and it seems as if for most molecules, this method indeed delivers satisfactory results. Acetonitrile is an exception due to its CN triple bond. However, the particularities of this molecule are discussed elsewhere [63] and it was found that state-of-the-art quantum chemical methods [64,65,66,67,68,69] are required for a proper description of νCN. Importantly, the more or less ubiquitous stretching motions arising from a methyl group are not described in a reliable manner and this has been recently ascribed to the nature of the MP2 method [19]. Conversely, it was shown that when higher order coupling terms (i.e., i<j<kNVi,j,ktriples(Qi,Qj,Qk)) are included in a DPT2-VSCF [44] treatment, MP2 indeed seems to be a viable ab initio method of choice [63]. VSCF data relies on the HOA and hence makes use of the rigid rotor approximation [70] which inherently is co-responsible for discrepancies between experiment and theory. Considering that still a large number of approximations are involved even in state-of-the-art VSCF calculations and their underlying ab initio energy evaluations, it must be concluded that computationally derived spectroscopic data are prone to a certain degree of fortuitous error compensation. However, both the data presented by Roy et al. [19] and our results indicate that triple-ζ based MP2 calculations in a VSCF treatment rather reliably deliver data with no more than 2% of unsigned error.

3.2. Performance at Reduced Grid Densities

As a VSCF evaluation with 16 grid points per mode is computationally highly demanding, a reduction of the grid density is desirable. From the data in Table 1 it becomes evident that for most cases, a reduction to 10 grid points yields identical results as the corresponding calculation with 16 grid points, indicating that the computational demand in a VSCF treatment of i=1NVidiag(Qi) and i<jNVi,jpairs(Qi,Qj) contributions can be reduced safely by more than 50%. It has to be kept in mind, however, that for every set of reduced densities, the grid is interpolated to the original resolution and the displacement boundaries are not affected by the reduction. Hence, as long as the points on the potential energy surface are chosen properly, the accuracy of the VSCF results is not affected. Noteworthy, the default routine in GAMESS chooses the displacement boundaries in dependence to the underlying normal mode’s wavenumber but the user is nonetheless able to alter these settings if desired.

Reviewing the values for µ in Table 2 permits the conclusion that major errors are starting to appear at grid densities <8 points. The GAMESS manual states that the VSCF code is “thought to give accuracy to 50 cm−1 for the larger fundamentals” when MP2 with a triple-ζ basis set is employed and hence, the boundaries of accuracy are wider than the error inflicted by a reduction of the grid density to 8 points. In order to provide statistical evidence as to what extent a reduction of the grid density is viable, paired t-tests [114] of the obtained values for µ have been conducted with the following equation:

t=|i=1χμrefμcalcχ|sd×χ (12)

Table 2.

Obtained values for µ in %, their arithmetic means and the corresponding t-values according to Equation (12). n.a. means “not applicable”.

Number of Grid Points
Molecule 6 7 8 9 10 11 12 13 14 15 16
H2O 21.3 5.3 0.8 0.6 0.6 0.6 0.5 0.5 0.5 0.5 0.5
CO2 25.1 6.9 2.3 2.0 2.0 2.0 2.0 2.0 2.0 2.0 2.0
CH2O 23.0 6.4 0.8 0.8 0.7 0.7 0.7 0.7 0.7 0.7 0.7
C2H2 24.5 5.2 2.0 1.7 1.8 1.8 1.8 1.8 1.8 1.8 1.8
HCOOH 22.8 6.0 2.1 1.4 1.7 1.7 1.7 1.7 1.7 1.7 1.7
CH4 22.3 5.9 0.9 0.4 0.3 0.3 0.3 0.3 0.3 0.3 0.3
CH3Cl 23.0 7.0 0.9 1.4 1.2 1.2 1.2 1.2 1.2 1.2 1.2
CH3OH 22.9 6.4 0.9 1.3 1.2 1.2 1.2 1.2 1.1 1.1 1.1
CH3CN 24.4 6.4 1.1 1.1 1.0 1.0 1.0 1.0 1.0 1.0 1.0
C2H4 23.4 7.7 0.8 1.5 1.4 1.4 1.3 1.3 1.3 1.3 1.3
C2H4O 23.3 6.5 1.2 1.4 1.4 1.3 1.3 1.3 1.3 1.3 1.3
CH3NH2 22.8 5.9 2.4 2.3 1.6 1.7 1.8 1.8 1.7 1.7 1.7
C2H6 22.3 7.5 0.6 1.7 1.6 1.6 1.6 1.6 1.6 1.6 1.6
CH3OCH3 23.0 7.2 1.1 1.7 1.6 1.5 1.5 1.5 1.5 1.5 1.5
C2H5NO2 22.8 7.5 1.5 1.8 1.7 1.7 1.7 1.7 1.7 1.7 1.7
Mean µ 23.1 6.5 1.3 1.4 1.3 1.3 1.3 1.3 1.3 1.3 1.3
t-Value 102.96 25.59 0.09 2.43 0.81 0.78 1.71 1.34 0.66 2.39 n.a.

In Equation (12), µref is the reference value at a grid density of 16 points. µcalc is the mean absolute percentage deviation at the examined grid density. χ is the number of investigated molecules and the t-value is calculated with sd as the corrected sample standard deviation of the respective grid density. For a two-tailed problem, critical values for t can be specified according to Table 3.

Table 3.

Critical boundaries according to Student’s t-distribution for a two-tailed problem.

t-Value Difference
t ≤ 2.145 Insignificant
2.145 < t ≤ 2.977 Probable
2.977 < t ≤ 4.140 Significant
4.140 < t Highly significant

Except for the data obtained at grid densities of 6 and 7 points, no significant deviations from the reference data are observed, indicating that even a reduction to 8 grid points is a very viable option. The computational effort involved in a pair-approximated VSCF evaluation (Equation (7)) may thus be reduced by a factor of nearly 4. It has to be stressed, however, that the obtained t-value for r = 8 (i.e., 0.09) may not be confused as to indicate the objectively best agreement with the reference values since the statistical analysis is based on unsigned error values. Moreover, the anharmonic correction seems to be slightly over-estimated at r = 8, yielding more red-shifted absorptions, as can be learned from Table 1. This over-correction is responsible for a slightly improved agreement with experiment for most of the molecules, especially since it is known that MP2 tendentially yields blue-shifted IR absorptions [19]. Overall, it may be concluded that VSCF treatments involving at least 8 grid points reproduce the experimental values within the same boundaries of accuracy as the reference values with a mean absolute percentage deviation from experiment of ~1.3% (Table 2).

In their recent paper, Roy et al., reported that soft and semi-rigid molecules have to be treated at higher grid densities than rigid molecules and they mention ethylene oxide as a particular example where the VSCF equations converge at no less than 12 grid points. Our results contradict this statement since our VSCF calculations of ethylene oxide did indeed converge for grid densities ≥6 points and a density of 8 points yielded the same quality of results as observed for the other molecules. Given the fact that the identical code, the same ab initio method and the same basis set were employed, this outcome is rather puzzling. This disagreement may be ascribed to the use of symmetry during our VSCF calculations and the underlying reference geometries having been minimized to the most stringent criterion available. Overall, none of our VSCF evaluations exhibited convergence problems apart from the few excluded torsional degrees of freedom that have been mentioned earlier.

3.3. Application to Deprotonated Phosphoserine

Phosphorylation of proteins is considered a major signal transduction mechanism, mainly occurring at the OH terminus of the amino acids serine, threonine and tyrosine. As a major constituent of biofluids, phosphoserine was subject of a gas-chromatographic investigation of human urine samples [115]. In autopsied Alzheimer’s disease brain tissue, l-phosphoserine was found in elevated concentrations [116] and only very recently, plasma phosphoserine levels were found to be upregulated in sepsis patients [117]. Deprotonated phosphoserine, abbreviated as [pSer-H], was investigated via infrared multiple photon dissociation (IRMPD) spectroscopy and hence, experimental data of a few major gas phase IR absorptions of the molecule is available [118]. The authors identified the lowest energy conformer of [pSer-H] as the one exhibiting hydrogen bonds between the carboxylic OH and the phosphate O as well as between the phosphate OH and the amino N. This conformer was chosen as a benchmark for the 8 point VSCF analysis due to the fact that it represents rigid and weak structural motifs likewise. Figure 1 shows the minimum geometry obtained at the MP2/cc-pVTZ level of theory which served as a reference state for the VSCF evaluation. Table 4 lists the PT2-VSCF computed absorptions of [pSer-H], the respective intensities and the available experimental data [118]. Due to the cage-like structure, many normal modes couple among each other and hence, especially lower vibrations cannot be ascribed to distinct normal modes. In such cases, the main vibrational contributions are given in Table 4. For illustrative purposes, the calculated IR spectrum of [pSer-H] is shown in Figure 2. The intensities have been computed using harmonic dipole derivatives at the MP2/cc-pVTZ level of theory and band broadening was introduced via a Lorentzian function using a band width at half height of 20 cm−1.

Figure 1.

Figure 1

The lowest energy conformer of [pSer-H] exhibits a hydrogen bond between the carboxylic OH and the phosphate O and between the phosphate OH and the amino N.

Table 4.

The main absorptions of [pSer-H] and the experimentally obtained [118] values.

Band Number PT2-VSCF (cm−1) Intensity (km mol−1) IRMPD (cm−1) [118] Main Contributions
1 3430 225 n.o. νOH
2 3346 11 n.o. νNH2,as
3 3139 125 n.o. νNH2,s
4 2945 21 n.o. νCαH
5 2915 47 n.o. νCβH2,s
6 2888 35 n.o. νCβH2,as
7 1734 359 1728 νCcarb =O
8 1565 39 1610 δNH2
9 1512 244 1461 δCcarbOH,ip
1449 1 1419 δCβH2
1380 4 n.o. τNH2 and δNCαH
1371 9 n.o. ωCβH2
1337 68 n.o. νCcarb−O and τCβH2
10 1315 484 1291 νP =O
1292 21 n.o. ωCβH2 and δCcarbCαH and νCcarb−O
11 1251 24 n.o. νCαCβ and δCβCαH and νCα−N and τNH2
1230 6 n.o. νCcarbCα and δCcarbCαH and τNH2
12 1176 60 n.o. δCcarbOH,oop
13 1157 51 n.o. νCβO
14 1113 99 1108 δPOH
15 1084 74 n.o. νNCαCcarb,as and νCβO
16 1060 141 1052 ωNH2
17 1017 121 1028 νP − OH −bonded
18 957 113 n.o. νNCαCβ,as
19 915 17 n.o. νNCαCβ,s and νNCαCcarb,as and νCcarbCαCβ,as
20 830 98 836 νP −OH
823 24 812 δCcarbCαCβ and δCcarbO2,oop
787 7 n.o. δCcarbO2 and δCcarbCαCβ
21 736 120 738 νP −OCβ and δCcarbO2
22 692 39 n.o. δCαNCβCcarb,umbrella and νP − OCbeta
513 41 n.o. δPOH,oop

Figure 2.

Figure 2

The computed IR spectrum of [pSer-H]. For band assignments, refer to Table 4.

Considering the 11 available experimental values, µ was computed for the corresponding VSCF data as 1.38%. Since we did calculate the IR absorptions of [pSer-H] exclusively using 8 grid points per mode, no definitive conclusion can be drawn whether higher grid densities would enable an improved accuracy for this larger molecule. A re-evaluation of the potential energy surface at grid densities up to 16 points would be unmanageably expensive but nonetheless, the data presented fits well into the error boundaries as observed for the 15 molecules discussed earlier.

The band assignment in Table 4 clearly shows that empirical considerations are only applicable to the large fundamentals. For a proper description of low lying vibrations that are determined by a number of torsions, computational approaches are an important option aiding the spectroscopist. In their experimental work, Scuderi et al., have employed computational techniques during their band assignment of [pSer-H] as well, but they resorted to scaled HOA absorption data at the B3LYP/6-311+G(d,p) level of theory.

4. Conclusions

VSCF theory has grown an important field in computational spectroscopy and this is owed to the increasingly available computational resources and to efficient and easily applicable algorithms. With this systematic study it has been demonstrated that a largely reduced grid density in a VSCF evaluation does not only spare considerable resources but also does not significantly affect the resulting absorption data. It was found that the convergence of VSCF equations is not impaired by such a reduced-effort technique even when highly problematic densities of <8 grid points are employed. While further investigations are required with regard to the role of symmetry in a VSCF treatment, it was found that reduced grid densities may be safely applied to a wide set of molecular entities. Application to [pSer-H] showed that a 8 point VSCF calculation yields accurate data for a molecule exhibiting weak interactions. The routine presented herein should, in conjunction with other recent developments [35], prove valuable for larger molecules relevant to life sciences where a conventional VSCF treatment with 16 grid points per mode is not feasible.

Author Contributions

O.M.D.L., B.M.R., G.K.B. and C.W.H. designed research; O.M.D.L. performed research; O.M.D.L. analyzed data; O.M.D.L. wrote the paper. All authors read and approved the final manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

Footnotes

Sample Availability: Not available.

References

  • 1.Scott A.P., Radom L. Harmonic Vibrational Frequencies: An Evaluation of Hartree-Fock, Møller-Plesset, Quadratic Configuration Interaction, Density Functional Theory, and Semiempirical Scale Factors. J. Phys. Chem. 1996;100:16502–16513. [Google Scholar]
  • 2.Sinha P., Boesch S.E., Gu C., Wheeler R.A., Wilson A.K. Harmonic Vibrational Frequencies: Scaling Factors for HF, B3LYP, and MP2 Methods in Combination with Correlation Consistent Basis Sets. J. Phys. Chem. A. 2004;108:9213–9217. doi: 10.1021/jp048233q. [DOI] [Google Scholar]
  • 3.Bowman J.M. Self-Consistent Field Energies and Wavefunctions for Coupled Oscillators. J. Chem. Phys. 1978;68:608–610. doi: 10.1063/1.435782. [DOI] [Google Scholar]
  • 4.Cohen M., Greita S., McEarchran R. Approximate and Exact Quantum Mechanical Energies and Eigenfunctions for a System of Coupled Oscillators. Chem. Phys. Lett. 1979;60:445–450. doi: 10.1016/0009-2614(79)80609-0. [DOI] [Google Scholar]
  • 5.Gerber R., Ratner M. A Semiclassical Self-Consistent Field (SC SCF) Approximation for Eigenvalues of Coupled-Vibration Systems. Chem. Phys. Lett. 1979;68:195–198. [Google Scholar]
  • 6.Bowman J.M. The Self-Consistent-Field Approach to Polyatomic Vibrations. Acc. Chem. Res. 1986;19:202–208. doi: 10.1021/ar00127a002. [DOI] [Google Scholar]
  • 7.Carter S., Culik S.J., Bowman J.M. Vibrational Self-Consistent Field Method for Many-Mode Systems: A New Approach and Application to the Vibrations of CO Adsorbed on Cu(100) J. Chem. Phys. 1997;107:10458–10469. doi: 10.1063/1.474210. [DOI] [Google Scholar]
  • 8.Carney G.D., Sprandel L.L., Kern C.W. Variational Approaches to Vibration-Rotation Spectroscopy for Polyatomic Molecules. In: Prigogine I., Rice S.A., editors. Advances in Chemical Physics. Volume 37. John Wiley & Sons,Inc.; Hoboken, NJ, USA: 2007. pp. 305–379. [Google Scholar]
  • 9.Barone V. Vibrational Zero-Point Energies and Thermodynamic Functions Beyond the Harmonic Approximation. J. Chem. Phys. 2004;120:3059–3065. doi: 10.1063/1.1637580. [DOI] [PubMed] [Google Scholar]
  • 10.Barone V. Anharmonic Vibrational Properties by a Fully Automated Second-order Perturbative Approach. J. Chem. Phys. 2005;122 doi: 10.1063/1.18248810. [DOI] [PubMed] [Google Scholar]
  • 11.Frisch M.J., Trucks G.W., Schlegel H.B., Scuseria G.E., Robb M.A., Cheeseman J.R., Scalmani G., Barone V., Mennucci B., Petersson G.A., et al. Gaussian Inc.; Wallingford, CT, USA: 2013. Gaussian 09, Revision D.01. [Google Scholar]
  • 12.Barone V., Biczysko M., Bloino J. Fully Anharmonic IR and Raman Spectra of Medium-Size Molecular Systems: Accuracy and Interpretation. Phys. Chem. Chem. Phys. 2014;16:1759–1787. doi: 10.1039/c3cp53413h. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Biczysko M., Bloino J., Carnimeo I., Panek P., Barone V. Fully Ab Initio IR Spectra for Complex Molecular Systems from Perturbative Vibrational Approaches: Glycine as a Test Case. J. Mol. Struct. 2012;1009:74–82. doi: 10.1016/j.molstruc.2011.10.012. [DOI] [Google Scholar]
  • 14.Chaban G.M., Jung J.O., Gerber R.B. Ab Initio Calculation of Anharmonic Vibrational States of Polyatomic Systems: Electronic Structure Combined with Vibrational Self-Consistent Field. J. Chem. Phys. 1999;111:1823–1829. doi: 10.1063/1.479452. [DOI] [Google Scholar]
  • 15.Jung J.O., Gerber R.B. Vibrational Wave Functions and Spectroscopy of (H2O)n, n = 2, 3, 4, 5: Vibrational Self-Consistent Field with Correlation Corrections. J. Chem. Phys. 1996;105:10332–10348. doi: 10.1063/1.472960. [DOI] [Google Scholar]
  • 16.Chaban G.M., Jung J.O., Gerber R.B. Anharmonic Vibrational Spectroscopy of Hydrogen-Bonded Systems Directly Computed from Ab Initio Potential Surfaces: (H2O)n, n = 2, 3; Cl−(H2O)n, n = 1, 2; H+(H2O)n, n = 1, 2; H2O-CH3OH. J. Phys. Chem. A. 2000;104:2772–2779. doi: 10.1021/jp993391g. [DOI] [Google Scholar]
  • 17.Chaban G.M., Jung J.O., Gerber R.B. Anharmonic Vibrational Spectroscopy of Glycine: Testing of Ab Initio and Empirical Potentials. J. Phys. Chem. A. 2000;104:10035–10044. doi: 10.1021/jp002297t. [DOI] [Google Scholar]
  • 18.Miller Y., Chaban G., Gerber R. Theoretical Study of Anharmonic Vibrational Spectra of HNO3, HNO3-H2O, HNO4: Fundamental, Overtone and Combination Excitations. Chem. Phys. 2005;313:213–224. doi: 10.1016/j.chemphys.2005.01.012. [DOI] [Google Scholar]
  • 19.Roy T.K., Carrington T., Gerber R.B. Approximate First-Principles Anharmonic Calculations of Polyatomic Spectra Using MP2 and B3LYP Potentials: Comparisons with Experiment. J. Phys. Chem. A. 2014;118:6730–6739. doi: 10.1021/jp5060155. [DOI] [PubMed] [Google Scholar]
  • 20.Schmidt M.W., Baldridge K.K., Boatz J.A., Elbert S.T., Gordon M.S., Jensen J.H., Koseki S., Matsunaga N., Nguyen K.A., Su S., et al. General Atomic and Molecular Electronic Structure System. J. Comput. Chem. 1993;14:1347–1363. doi: 10.1002/jcc.540141112. [DOI] [Google Scholar]
  • 21.Valiev M., Bylaska E., Govind N., Kowalski K., Straatsma T., Dam H.V., Wang D., Nieplocha J., Apra E., Windus T., et al. NWChem: A Comprehensive and Scalable Open-Source Solution for Large Scale Molecular Simulations. Comput. Phys. Commun. 2010;181:1477–1489. doi: 10.1016/j.cpc.2010.04.018. [DOI] [Google Scholar]
  • 22.Werner H., Knowles P.J., Knizia G., Manby F.R., Schütz M., Celani P., Korona T., Lindh R., Mitrushenkov A., Rauhut G., et al. MOLPRO, Version 2012.1, A Package of Ab Initio Programs. 2012. [(accessed on 15 December 2014)]. Available online: http://www.molpro.net/
  • 23.Møller C., Plesset M.S. Note on an Approximation Treatment for Many-Electron Systems. Phys. Rev. 1934;46:618–622. doi: 10.1103/PhysRev.46.618. [DOI] [Google Scholar]
  • 24.Fletcher G.D., Schmidt M.W., Gordon M.S. Developments in Parallel Electronic Structure Theory. In: Prigogine I., Rice S.A., editors. Advances in Chemical Physics. Volume 110. John Wiley & Sons, Inc.; Hoboken, NJ, USA: 2007. pp. 267–294. [Google Scholar]
  • 25.Aikens C.M., Gordon M.S. Parallel Unrestricted MP2 Analytic Gradients Using the Distributed Data Interface. J. Phys. Chem. A. 2004;108:3103–3110. doi: 10.1021/jp031142t. [DOI] [Google Scholar]
  • 26.Sachs E.S., Hinze J., Sabelli N.H. Frozen Core Approximation, a Pseudopotential Method Tested on Six States of NaH. J. Chem. Phys. 1975;62:3393–3398. doi: 10.1063/1.430993. [DOI] [Google Scholar]
  • 27.Dunning T.H. Gaussian Basis Sets for Use in Correlated Molecular Calculations. I. The Atoms Boron Through Neon and Hydrogen. J. Chem. Phys. 1989;90:1007–1023. doi: 10.1063/1.456153. [DOI] [Google Scholar]
  • 28.Bowman J.M., Carrington T., Meyer H.D. Variational Quantum Approaches for Computing Vibrational Energies of Polyatomic Molecules. Mol. Phys. 2008;106:2145–2182. doi: 10.1080/00268970802258609. [DOI] [Google Scholar]
  • 29.Chan M., Carrington T., Manzhos S. Anharmonic Vibrations of the Carboxyl Group in Acetic Acid on TiO2: Implications For Adsorption Mode Assignment in Dye-Sensitized Solar Cells. Phys. Chem. Chem. Phys. 2013;15:10028–10034. doi: 10.1039/c3cp00065f. [DOI] [PubMed] [Google Scholar]
  • 30.Sakurai J.J., Tuan S.F. Modern Quantum Mechanics. Volume 1 Addison-Wesley; Boston, MA, USA: 1985. [Google Scholar]
  • 31.Griffiths D.J., Harris E.G. Introduction to Quantum Mechanics. Volume 2 Prentice Hall; Upper Saddle River, NJ, USA: 1995. [Google Scholar]
  • 32.Ratner M.A., Gerber R.B. Excited Vibrational States of Polyatomic Molecules: The Semiclassical Self-Consistent Field Approach. J. Phys. Chem. 1986;90:20–30. doi: 10.1021/j100273a008. [DOI] [Google Scholar]
  • 33.Watson J.K. Simplification of the Molecular Vibration-Rotation Hamiltonian. Mol. Phys. 1968;15:479–490. doi: 10.1080/00268976800101381. [DOI] [Google Scholar]
  • 34.Bowman J.M., Carter S., Huang X. MULTIMODE: A Code to Calculate Rovibrational Energies of Polyatomic Molecules. Int. Rev. Phys. Chem. 2003;22:533–549. doi: 10.1080/0144235031000124163. [DOI] [Google Scholar]
  • 35.Roy T.K., Gerber R.B. Vibrational Self-Consistent Field Calculations for Spectroscopy of Biological Molecules: New Algorithmic Developments and Applications. Phys. Chem. Chem. Phys. 2013;15:9468–9492. doi: 10.1039/c3cp50739d. [DOI] [PubMed] [Google Scholar]
  • 36.Gregurick S.K., Fredj E., Elber R., Gerber R.B. Vibrational Spectroscopy of Peptides and Peptide-Water Complexes: Anharmonic Coupled-Mode Calculations. J. Phys. Chem. B. 1997;101:8595–8606. doi: 10.1021/jp971587f. [DOI] [Google Scholar]
  • 37.Gregurick S.K., Liu J.H.Y., Brant D.A., Gerber R.B. Anharmonic Vibrational Self-Consistent Field Calculations as an Approach to Improving Force Fields for Monosaccharides. J. Phys. Chem. B. 1999;103:3476–3488. doi: 10.1021/jp9826221. [DOI] [Google Scholar]
  • 38.Kowal A.T. First-principles Calculation of Geometry and Anharmonic Vibrational Spectra of Thioformamide and Thioformamide-d2. J. Chem. Phys. 2006;124 doi: 10.1063/1.2139995. [DOI] [PubMed] [Google Scholar]
  • 39.Carter S., Bowman J.M., Handy N.C. Extensions and Tests of MULTIMODE: A Code to Obtain Accurate Vibration/Rotation Energies of Many-Mode Molecules. Theor. Chem. Acc. 1998;100:191–198. [Google Scholar]
  • 40.Rheinecker J., Bowman J.M. The Calculated Infrared Spectrum of Cl H2O Using a New Full Dimensional Ab Initio Potential Surface and Dipole Moment Surface. J. Chem. Phys. 2006;125 doi: 10.1063/1.2209675. [DOI] [PubMed] [Google Scholar]
  • 41.Norris L.S., Ratner M.A., Roitberg A.E., Gerber R.B. Møller-Plesset Perturbation Theory Applied to Vibrational Problems. J. Chem. Phys. 1996;105:11261–11267. doi: 10.1063/1.472922. [DOI] [Google Scholar]
  • 42.Pele L., Gerber R.B. On the Number of Significant Mode-Mode Anharmonic Couplings in Vibrational Calculations: Correlation-Corrected Vibrational Self-Consistent Field Treatment of Di-, Tri-, and Tetrapeptides. J. Chem. Phys. 2008;128 doi: 10.1063/1.2909558. [DOI] [PubMed] [Google Scholar]
  • 43.Pele L., Brauer B., Gerber R. Acceleration of Correlation-Corrected Vibrational Self-Consistent Field Calculation Times for Large Polyatomic Molecules. Theor. Chem. Acc. 2007;117:69–72. doi: 10.1007/s00214-006-0132-2. [DOI] [Google Scholar]
  • 44.Matsunaga N., Chaban G.M., Gerber R.B. Degenerate Perturbation Theory Corrections for the Vibrational Self-Consistent Field Approximation: Method and Applications. J. Chem. Phys. 2002;117:3541–3547. [Google Scholar]
  • 45.Bowman J.M., Christoffel K., Tobin F. Application of SCF-SI Theory to Vibrational Motion in Polyatomic Molecules. J. Phys. Chem. 1979;83:905–912. doi: 10.1021/j100471a005. [DOI] [Google Scholar]
  • 46.Christoffel K.M., Bowman J.M. Investigations of Self-Consistent Field, SCF CI and Virtual Stateconfiguration Interaction Vibrational Energies for a Model Three-Mode System. Chem. Phys. Lett. 1982;85:220–224. doi: 10.1016/0009-2614(82)80335-7. [DOI] [Google Scholar]
  • 47.Christiansen O., Luis J.M. Beyond Vibrational Self-Consistent-Field Methods: Benchmark Calculations for the Fundamental Vibrations of Ethylene. Int. J. Quantum Chem. 2005;104:667–680. doi: 10.1002/qua.20615. [DOI] [Google Scholar]
  • 48.Neff M., Rauhut G. Toward Large Scale Vibrational Configuration Interaction Calculations. J. Chem. Phys. 2009;131 doi: 10.1063/1.3243862. [DOI] [PubMed] [Google Scholar]
  • 49.Kaulfuss U.B., Altenbokum M. Anharmonic Oscillator as a Test of the Coupled-Cluster Method. Phys. Rev. D. 1986;33:3658–3664. doi: 10.1103/PhysRevD.33.3658. [DOI] [PubMed] [Google Scholar]
  • 50.Nagalakshmi V., Lakshminarayana V., Sumithra G., Prasad M.D. Coupled Cluster Description of Anharmonic Molecular Vibrations. Application to O3 and SO2. Chem. Phys. Lett. 1994;217:279–282. doi: 10.1016/0009-2614(93)E1380-Y. [DOI] [Google Scholar]
  • 51.Thomsen B., Yagi K., Christiansen O. Optimized Coordinates in Vibrational Coupled Cluster Calculations. J. Chem. Phys. 2014;140 doi: 10.1063/1.4870775. [DOI] [Google Scholar]
  • 52.Culot F., Liévin J. A Multiconfigurational SCF Computational Method for the Resolution of the Vibrational Schrödinger Equation in Polyatomic Molecules. Theor. Chim. Acta. 1994;89:227–250. doi: 10.1007/BF01225116. [DOI] [Google Scholar]
  • 53.Meyer H.D. Studying Molecular Quantum Dynamics With the Multiconfiguration Time-Dependent Hartree Method. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2012;2:351–374. doi: 10.1002/wcms.87. [DOI] [Google Scholar]
  • 54.Njegic B., Gordon M.S. Exploring the Effect of Anharmonicity of Molecular Vibrations on Thermodynamic Properties. J. Chem. Phys. 2006;125 doi: 10.1063/1.2395940. [DOI] [PubMed] [Google Scholar]
  • 55.Bounouar M., Scheurer C. The Impact of Approximate VSCF Schemes and Curvilinear Coordinates on the Anharmonic Vibrational Frequencies of Formamide and Thioformamide. Chem. Phys. 2008;347:194–207. [Google Scholar]
  • 56.Strobusch D., Scheurer C. Hierarchical Expansion of the Kinetic Energy Operator in Curvilinear Coordinates for the Vibrational Self-Consistent Field Method. J. Chem. Phys. 2011;135 doi: 10.1063/1.3637629. [DOI] [PubMed] [Google Scholar]
  • 57.Boatz J.A., Gordon M.S. Decomposition of Normal-Coordinate Vibrational Frequencies. J. Phys. Chem. 1989;93:1819–1826. doi: 10.1021/j100342a026. [DOI] [Google Scholar]
  • 58.Yagi K., Hirao K., Taketsugu T., Schmidt M.W., Gordon M.S. Ab Initio Vibrational State Calculations with a Quartic Force Field: Applications to H2CO, C2H4, CH3OH, CH3CCH and C6H6. J. Chem. Phys. 2004;121:1383–1389. doi: 10.1063/1.1764501. [DOI] [PubMed] [Google Scholar]
  • 59.Shimanouchi T. Tables of Molecular Vibrational Frequencies Consolidated Volume I. Natl. Bur. Stand. 1972;1:1–164. [Google Scholar]
  • 60.Herzberg G. Infrared and Raman Spectra of Polyatomic Molecules. Van Nostrand; New York, NY, USA: 1949. [Google Scholar]
  • 61.Smith L.G. The Infra-Red Spectrum of C2H6. J. Chem. Phys. 1949;17:139–167. doi: 10.1063/1.1747206. [DOI] [Google Scholar]
  • 62.Weiss S., Leroi G.E. Direct Observation of the Infrared Torsional Spectrum of C2H6, CH3CD3, and C2D6. J. Chem. Phys. 1968;48:962–967. doi: 10.1063/1.1668849. [DOI] [Google Scholar]
  • 63.Lutz O.M., Rode B.M., Bonn G.K., Huck C.W. The Impact of Highly Correlated Potential Energy Surfaces on the Anharmonically Corrected IR Spectrum of Acetonitrile. Spectrochim. Acta Part A. 2014;131:545–555. doi: 10.1016/j.saa.2014.04.067. [DOI] [PubMed] [Google Scholar]
  • 64.Bartlett R.J., Purvis G.D. Many-Body Perturbation Theory, Coupled-Pair Many-Electron Theory, and the Importance of Quadruple Excitations for the Correlation Problem. Int. J. Quantum Chem. 1978;14:561–581. doi: 10.1002/qua.560140504. [DOI] [Google Scholar]
  • 65.Pople J.A., Krishnan R., Schlegel H.B., Binkley J.S. Electron Correlation Theories and Their Application to the Study of Simple Reaction Potential Surfaces. Int. J. Quantum Chem. 1978;14:545–560. doi: 10.1002/qua.560140503. [DOI] [Google Scholar]
  • 66.Gauss J., Stanton J.F. Perturbative Treatment of Triple Excitations in Coupled-cluster Calculations of Nuclear Magnetic Shielding Constants. J. Chem. Phys. 1996;104:2574–2583. doi: 10.1063/1.471005. [DOI] [Google Scholar]
  • 67.Pople J.A., Head-Gordon M., Raghavachari K. Quadratic Configuration Interaction. A General Technique for Determining Electron Correlation Energies. J. Chem. Phys. 1987;87:5968–5975. doi: 10.1063/1.453520. [DOI] [Google Scholar]
  • 68.Piecuch P., Włoch M. Renormalized Coupled-Cluster Methods Exploiting Left Eigenstates of the Similarity-Transformed Hamiltonian. J. Chem. Phys. 2005;123 doi: 10.1063/1.2137318. [DOI] [PubMed] [Google Scholar]
  • 69.Piecuch P., Włoch M., Gour J.R., Kinal A. Single-Reference, Size-Extensive, Non-Iterative Coupled-Cluster Approaches to Bond Breaking and Biradicals. Chem. Phys. Lett. 2006;418:467–474. doi: 10.1016/j.cplett.2005.10.116. [DOI] [Google Scholar]
  • 70.Dennison D.M. The Infrared Spectra of Polyatomic Molecules Part I. Rev. Mod. Phys. 1931;3:280–345. doi: 10.1103/RevModPhys.3.280. [DOI] [Google Scholar]
  • 71.Barker E.F., Sleator W.W. The Infrared Spectrum of Heavy Water. J. Chem. Phys. 1935;3:660–663. doi: 10.1063/1.1749572. [DOI] [Google Scholar]
  • 72.Benedict W.S., Gailar N., Plyler E.K. Rotation-Vibration Spectra of Deuterated Water Vapor. J. Chem. Phys. 1956;24:1139–1165. doi: 10.1063/1.1742731. [DOI] [Google Scholar]
  • 73.Amat G., Pimbert M. On Fermi Resonance in Carbon Dioxide. J. Mol. Spectrosc. 1965;16:278–290. doi: 10.1016/0022-2852(65)90123-2. [DOI] [Google Scholar]
  • 74.Suzuki I. General Anharmonic Force Constants of Carbon Dioxide. J. Mol. Spectrosc. 1968;25:479–500. doi: 10.1016/S0022-2852(68)80018-9. [DOI] [Google Scholar]
  • 75.Davidson D.W., Stoicheff B.P., Bernstein H.J. The Infrared and Raman Spectra of Formaldehyde-d1 Vapor. J. Chem. Phys. 1954;22:289–294. doi: 10.1063/1.1740054. [DOI] [Google Scholar]
  • 76.Blau H.H., Nielsen H.H. The Infrared Absorption Spectrum of Formaldehyde Vapor. J. Mol. Spectrosc. 1957;1:124–132. doi: 10.1016/0022-2852(57)90015-2. [DOI] [Google Scholar]
  • 77.Nakagawa T., Kashiwagi H., Kurihara H., Morino Y. Vibration-Rotation Spectra of Formaldehyde: Band Contour Analysis of the ν2 and ν3 Fundamentals. J. Mol. Spectrosc. 1969;31:436–450. doi: 10.1016/0022-2852(69)90373-7. [DOI] [Google Scholar]
  • 78.Lafferty W.J., Thibault R.J. High Resolution Infrared Spectra of C212H2, C12C13H2, and C213H2. J. Mol. Spectrosc. 1964;14:79–96. doi: 10.1016/0022-2852(64)90101-8. [DOI] [Google Scholar]
  • 79.Scott J., Rao K. Infrared Absorption Bands of Acetylene: Part I. Analysis of the Bands at 13.7µ. J. Mol. Spectrosc. 1965;16:15–23. doi: 10.1016/0022-2852(65)90081-0. [DOI] [Google Scholar]
  • 80.Scott J., Rao K.N. Infrared Absorption Bands of Acetylene: Part III. Analysis of the Bands at 3µ. J. Mol. Spectrosc. 1966;20:438–460. doi: 10.1016/0022-2852(66)90015-4. [DOI] [Google Scholar]
  • 81.Williams V.Z. Infra-Red Spectra of Monomeric Formic Acid and Its Deuterated Forms. I. High Frequency Region. J. Chem. Phys. 1947;15:232–242. doi: 10.1063/1.1746487. [DOI] [Google Scholar]
  • 82.Wilmshurst J.K. Vibrational Assignment in Monomeric Formic Acid. J. Chem. Phys. 1956;25:478–480. doi: 10.1063/1.1742947. [DOI] [Google Scholar]
  • 83.Millikan R.C., Pitzer K.S. Infrared Spectra and Vibrational Assignment of Monomeric Formic Acid. J. Chem. Phys. 1957;27:1305–1308. doi: 10.1063/1.1743996. [DOI] [Google Scholar]
  • 84.Miyazawa T., Pitzer K.S. Internal Rotation and Infrared Spectra of Formic Acid Monomer and Normal Coordinate Treatment of Out-Of-Plane Vibrations of Monomer, Dimer, and Polymer. J. Chem. Phys. 1959;30:1076–1086. doi: 10.1063/1.1730085. [DOI] [Google Scholar]
  • 85.Allen H.C., Plyler E.K. ν3 Band of Methane. J. Chem. Phys. 1957;26:972–973. doi: 10.1063/1.1743458. [DOI] [Google Scholar]
  • 86.Herranz J., Morcillo J., Gómez A. The ν2 Infrared Band of CH4 and CD4. J. Mol. Spectrosc. 1966;19:266–282. doi: 10.1016/0022-2852(66)90250-5. [DOI] [Google Scholar]
  • 87.King W.T., Mills I.M., Crawford B. Normal Coordinates in the Methyl Halides. J. Chem. Phys. 1957;27:455–457. doi: 10.1063/1.1743745. [DOI] [Google Scholar]
  • 88.Jones E., Popplewell R., Thompson H. Vibration—Rotation Bands of Methyl Chloride. Spectrochim. Acta. 1966;22:669–680. [Google Scholar]
  • 89.Maki A.G., Thibault R. Analysis of Some Perturbations in the ν4 Band of Methyl Chloride. J. Chem. Phys. 1968;48:2163–2167. doi: 10.1063/1.1669407. [DOI] [Google Scholar]
  • 90.Morillon-Chapey M., Graner G. Fine Structure in the ν1 Band of CH3Cl near 2970 cm−1. J. Mol. Spectrosc. 1969;31:155–169. doi: 10.1016/0022-2852(69)90348-8. [DOI] [Google Scholar]
  • 91.Tanaka C., Kuratani K., Mizushima S.I. In-Plane Normal Vibrations of Methanol. Spectrochim. Acta. 1957;9:265–269. doi: 10.1016/0371-1951(57)80141-6. [DOI] [Google Scholar]
  • 92.Van Thiel M., Becker E.D., Pimentel G.C. Infrared Studies of Hydrogen Bonding of Methanol by the Matrix Isolation Technique. J. Chem. Phys. 1957;27:95–99. doi: 10.1063/1.1743725. [DOI] [Google Scholar]
  • 93.Falk M., Whalley E. Infrared Spectra of Methanol and Deuterated Methanols in Gas, Liquid, and Solid Phases. J. Chem. Phys. 1961;34:1554–1568. doi: 10.1063/1.1701044. [DOI] [Google Scholar]
  • 94.Nakagawa I., Shimanouchi T. Rotation-Vibration Spectra and Rotational, Coriolis Coupling, Centrifugal Distortion and Potential Constants of Methyl Cyanide. Spectrochim. Acta. 1962;18:513–539. doi: 10.1016/S0371-1951(62)80163-5. [DOI] [Google Scholar]
  • 95.Tolonen A., Koivusaari M., Paso R., Schroderus J., Alanko S., Anttila R. The Infrared Spectrum of Methyl Cyanide Between 850 and 1150 cm−1: Analysis of the ν4, ν7, and 3ν18 Bands with Resonances. J. Mol. Spectrosc. 1993;160:554–565. doi: 10.1006/jmsp.1993.1201. [DOI] [Google Scholar]
  • 96.Venkateswarlu P. The Rotation-Vibration Spectrum of Methyl Cyanide in the Region 1.6 µ–20 µ. J. Chem. Phys. 1951;19:293–298. doi: 10.1063/1.1748197. [DOI] [Google Scholar]
  • 97.Venkateswarlu P. The Infrared Spectrum of Methyl Cyanide. J. Chem. Phys. 1952;20:923–923. doi: 10.1063/1.1700606. [DOI] [Google Scholar]
  • 98.Thompson H.W., Williams R.L. The Infra-red Spectra of Methyl Cyanide and Methyl Isocyanide. Trans. Faraday Soc. 1952;48:502–513. [Google Scholar]
  • 99.Parker F., Nielsen A., Fletcher W. The Infrared Absorption Spectrum of Methyl Cyanide Vapor. J. Mol. Spectrosc. 1957;1:107–123. doi: 10.1016/0022-2852(57)90014-0. [DOI] [Google Scholar]
  • 100.Huet T. The ν1 and ν5 Fundamental Bands of Methyl Cyanide. J. Mol. Struct. 2000;517:127–131. doi: 10.1016/S0022-2860(99)00243-4. [DOI] [Google Scholar]
  • 101.Arnett R.L., Crawford B.L. The Vibrational Frequencies of Ethylene. J. Chem. Phys. 1950;18:118–126. doi: 10.1063/1.1747428. [DOI] [Google Scholar]
  • 102.Allen H.C., Plyler E.K. The Structure of Ethylene from Infrared Spectra. J. Am. Chem. Soc. 1958;80:2673–2676. [Google Scholar]
  • 103.Lord R.C., Nolin B. Vibrational Spectra of Ethylene Oxide and Ethylene Oxide-d4. J. Chem. Phys. 1956;24:656–658. [Google Scholar]
  • 104.Potts W. The Fundamental Vibration Frequencies of Ethylene Oxide and Ethylene Imine. Spectrochim. Acta. 1965;21:511–527. doi: 10.1016/0371-1951(65)80143-6. [DOI] [Google Scholar]
  • 105.Kirby-Smith J.S., Bonner L.G. The Raman Spectra of Gaseous Substances 1. Apparatus and the Spectrum of Methylamine. J. Chem. Phys. 1939;7:880–883. doi: 10.1063/1.1750337. [DOI] [Google Scholar]
  • 106.Gray A.P., Lord R.C. Rotation-Vibration Spectra of Methyl Amine and Its Deuterium Derivatives. J. Chem. Phys. 1957;26:690–705. doi: 10.1063/1.1743369. [DOI] [Google Scholar]
  • 107.Tsuboi M., Hirakawa A.Y., Ino T., Sasaki T., Tamagake K. Amino Wagging and Inversion in Methylamines. J. Chem. Phys. 1964;41:2721–2734. doi: 10.1063/1.1726344. [DOI] [Google Scholar]
  • 108.Cole A., Lafferty W., Thibault R. Rotational Fine Structure of the Perpendicular Band, ν7, of Ethane. J. Mol. Spectrosc. 1969;29:365–374. doi: 10.1016/0022-2852(69)90113-1. [DOI] [Google Scholar]
  • 109.Nakagawa I., Shimanouchi T. Rotation-Vibration Spectra and Rotational, Coriolis Coupling and Potential Constants of Ethane, Ethane-d6 and Ethane-1, 1, 1-d3. J. Mol. Spectrosc. 1971;39:255–274. doi: 10.1016/0022-2852(71)90058-0. [DOI] [Google Scholar]
  • 110.Taylor R.C., Vidale G.L. Raman Spectrum and Vibrational Assignments of Gaseous Dimethyl Ether. J. Chem. Phys. 1957;26:122–123. doi: 10.1063/1.1743235. [DOI] [Google Scholar]
  • 111.Fateley W., Miller F.A. Torsional Frequencies in the Far Infrared-II: Molecules With Two or Three Methyl Rotors. Spectrochim. Acta. 1962;18:977–993. [Google Scholar]
  • 112.Stepanian S.G., Reva I.D., Radchenko E.D., Rosado M.T.S., Duarte M.L.T.S., Fausto R., Adamowicz L. Matrix-Isolation Infrared and Theoretical Studies of the Glycine Conformers. J. Phys. Chem. A. 1998;102:1041–1054. doi: 10.1021/jp973397a. [DOI] [Google Scholar]
  • 113.Brauer B., Chaban G.M., Gerber R.B. Spectroscopically-Tested, Improved, Semi-Empirical Potentials for Biological Molecules: Calculations for Glycine, Alanine and Proline. Phys. Chem. Chem. Phys. 2004;6:2543–2556. doi: 10.1039/b315326f. [DOI] [Google Scholar]
  • 114.Goulden C. Methods of Statistical Analysis. Wiley; New York, NY, USA: 2007. pp. 50–55. [Google Scholar]
  • 115.Kataoka H., Nakai K., Katagiri Y., Makita M. Analysis of Free and Bound O-Phosphoamino Acids in Urine by Gas Chromatography with Flame Photometric Detection. Biomed. Chromatogr. 1993;7:184–188. doi: 10.1002/bmc.1130070403. [DOI] [PubMed] [Google Scholar]
  • 116.Mason R., Trumbore M.W., Pettegrew J.W. Membrane Interactions of a Phosphomonoester Elevated Early in Alzheimer’s Disease. Neurobiol. Aging. 1995;16:531–539. doi: 10.1016/0197-4580(95)00057-L. [DOI] [PubMed] [Google Scholar]
  • 117.Chiarla C., Giovannini I., Siegel J.H. High Phosphoserine in Sepsis: Panel of Clinical and Plasma Amino Acid Correlations. SpringerPlus. 2014;3 doi: 10.1186/2193-1801-3-279. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Scuderi D., Correia C.F., Balaj O.P., Ohanessian G., Lemaire J., Maitre P. Structural Characterization by IRMPD Spectroscopy and DFT Calculations of Deprotonated Phosphorylated Amino Acids in the Gas Phase. ChemPhysChem. 2009;10:1630–1641. doi: 10.1002/cphc.200800856. [DOI] [PubMed] [Google Scholar]

Articles from Molecules are provided here courtesy of Multidisciplinary Digital Publishing Institute (MDPI)

RESOURCES