Abstract

A general strategy for the accurate computation of conformational and spectroscopic properties of flexible molecules in the gas phase is applied to two representative proteinogenic amino acids with aromatic side chains, namely, phenylalanine and tyrosine. The main features of all the most stable conformers predicted by this computational strategy closely match those of the species detected in microwave and infrared experiments. Together with their intrinsic interest, the accuracy of the results obtained with reasonable computer times paves the route for accurate investigations of other flexible bricks of life.
Introduction
Amino acids represent central targets for accurate experimental and theoretical studies because their rich conformational landscape is tuned by the competition among different kinds of noncovalent interactions.1,2 While zwitterionic forms of amino acids are found in crystals3 and aqueous solutions,4 neutral forms are observed in the gas phase5−7 and in low-temperature inert matrices.8,9 In the latter case, van der Waals forces between the sample and inner gas molecules have usually a negligible effect on spectroscopic parameters,10 so high-resolution spectroscopic studies both in the gas phase (microwave, MW) and in inert matrices (infrared, IR) allow an unbiased disentanglement of intrinsic stereoelectronic effects without any strong perturbation from the environment.
Systematic investigations have revealed that the natural amino acids containing simple nonpolar side chains present two dominant conformers (usually referred to as type I and type II), stabilized by intrabackbone hydrogen bonds.7,11,12 On the other hand, the conformational landscape of natural amino acids with polar side chains is much richer due to the synergy or competition between intrabackbone and backbone–(side chain) hydrogen bonds.13−16 Here we will analyze in detail the situation for prototypical amino acids containing aromatic side chains (phenylalanine (Phe) and tyrosine (Tyr)), which can show an additional kind of noncovalent interaction between polar hydrogen atoms of the backbone and the π-system of the side chain.
Phenylalanine and tyrosine have been studied in the gas phase by high-resolution laser-induced fluorescence (LIF), hole-burning UV–UV, ion dip IR–UV, resonance-enhanced multiphoton ionization (REMPI) spectroscopy,6,17−22 and, more recently, by microwave (MW) spectroscopy in combination with supersonic jets.23,24 Phenylalanine has been studied also by matrix-isolation infrared spectroscopy.10 These experiments have been interpreted in terms of a variable number of low-energy conformers due to the presence of effects (e.g., fast relaxation of some structures to more stable counterparts in the presence of low energy barriers and photodissociation of some products) playing a different role under different experimental conditions.20,25,26
Accurate quantum chemical (QC) computations can help to solve this kind of problems,27−29 but the size of the systems and the number of different structures to be analyzed prevent a systematic use of very accurate but very expensive state-of-the-art QC methodologies.30−32 Furthermore, an effective exploration of flat potential energy surfaces (PES) cannot be performed by straightforward systematic searches and/or local optimizations.33,34 We have recently developed an integrated computational approach combining different QC methods driven by machine learning (ML) tools for the effective exploration of conformational PES and the successive refinement of the most significant stationary points.35−37 In this framework, once a suitable panel of low-energy minima has been defined, accurate structures are obtained by refining the optimized geometries by a linear regression approach38,39 and accurate relative energies are computed by reduced-scaling composite methods.40−45 The zero point energies (ZPEs) and thermal contributions leading from electronic to free energies are evaluated by parameter-free anharmonic approaches rooted in the second order vibrational perturbation theory (VPT2)46−53 for sufficiently stiff degrees of freedom and proper treatment of hindered rotations.54,55 Finally, accurate spectroscopic parameters of the energy minima with non negligible populations under the experimental conditions of interest are computed.56
This procedure has been recently validated for several amino acids containing polar side chains.57−59 Here we tackle the problem of aromatic side chains analyzing both the rotational and vibrational spectra of the prototypical phenylalanine and tyrosine amino acids. In fact, previous studies of these amino acids employed QC methods of limited accuracy, payed marginal attention to the geometrical parameters and neglected vibrational corrections beyond the standard rigid rotor harmonic oscillator (RRHO) model. However, these limitations hamper any a priori prediction of the spectroscopic outcome, allowing at most its a posteriori interpretation in terms of the agreement between experimental and computed spectroscopic parameters for a predefined number of conformers. Our approach allows, instead, an unbiased comparison with spectroscopic results thanks to the accuracy of the computational results, which provide mean unsigned errors (MUEs) within 10 MHz for rotational constants and 10 cm–1 for both relative energies and vibrational frequencies.41,60
Methods
A first characterization of low-energy conformers is performed at the B3LYP/jun-cc-pVDZ level,61,62 also including Grimme’s D3BJ dispersion corrections.63 In the following this computational model is referred to simply as B3 and is used also for the computation of anharmonic contributions. The geometries and harmonic force fields of conformers lying within a predefined energy range are then refined by the revDSD-PBEP86-D3BJ/jun-cc-pv(T+d)Z model64−67 (hereafter rDSD), which provides excellent conformational landscapes,30,68,69 geometries,39 dipole moments,70 and spectroscopic parameters.56,71 In the present context, we will analyze in detail only conformers not involved in fast relaxation processes and with a relative free energy below 620 cm–1 at room temperature (i.e., a relative population of about 5% when only the considered conformer and the absolute free energy minimum are taken into account, since kT/hc = 207 cm–1).
The typical MUEs of rDSD bond lengths and valence angles (0.3%)39 are sufficient to obtain accurate relative electronic energies and vibrational spectra of different conformers, but the situation is different for the leading terms of MW spectra, namely, rotational constants of the vibrational ground state (B0i, where i refers to the inertial axes a, b, c). These parameters include vibrational corrections (ΔBvib) in addition to equilibrium rotational constants (Bei).72 In the framework of the VPT2 approximation,73 the ground-state rotational constants can be expressed as
| 1 |
where the αr’s are the vibration–rotation interaction constants and the sum runs over all r vibrational modes (for details, see refs (31), (74), and (75)). With ΔBvibi contributions typically well below 1% of the corresponding Be rotational constants,76 they can be safely determined at affordable levels of theory (B3 in the present context), which deliver typical errors within 10% (i.e., less than 0.1% of typical rotational constants).31,77 On the other hand, inclusion of vibrational corrections is not warranted if the errors on the computed rotational constants are not much lower than 1% (10 MHz for a constant of 1000 MHz). Therefore, equilibrium rotational constants require very accurate geometrical parameters, which can be obtained only with state-of-the-art QC methods,78−81 although reasonable relative errors (typically 0.4–0.5%) are obtained at the rDSD level.39 We have recently shown that the systematic nature of the errors permits to improve significantly the rDSD geometrical parameters, and thus equilibrium rotational constants, by a linear regression approach (hereafter LRA).39,82,83 In this model, the computed geometrical parameters (rcomp) are corrected for systematic errors by means of scaling factors (a) and offset values (b) depending on the nature of the involved atoms and determined once for ever from a large database of accurate semiexperimental (SE) equilibrium geometries:39,59,83
| 2 |
Since at least one 14N quadrupolar nucleus is present in all amino acids, nuclear quadrupole coupling constants (χii, with i referring to the inertia axis a, b, or c) are important for accurate predictions of rotational spectra because they determine a splitting of the rotational transitions, which generates the so-called hyperfine structure.78,84 Furthermore, the components of dipole moments (μi) determine the intensities of rotational transitions.74,84 Several studies have shown that dipole moments and quadrupole coupling constants can be computed with sufficient accuracy at the rDSD level.59,70 On the other hand, accurate electronic energies can be obtained by single-point energy evaluations at rDSD geometries using composite wave function methods rooted in the coupled cluster (CC) ansatz.85 In particular, the CC model including single, double, and perturbative estimate of triple excitations (CCSD(T))86 is employed here taking also into account complete basis set (CBS) extrapolation and core valence (CV) correlation. The starting point of the family of “cheap” schemes (ChS) developed in the last years28,36,40,41 is a frozen core (fc) CCSD(T) computation in conjunction with a (partially augmented) triple-ζ basis set.62,66,87 Then, in analogy with the correlation consistent composite approach (ccCA),88,89 CBS and CV terms are computed with good accuracy and negligible additional cost by means of second order Møller–Plesset perturbation theory (MP2).90 In particular, the CBS extrapolation by the standard n–3 two-point formula91 employs MP2/jun-cc-pV(X+d)Z energies with X = T and Q, whereas the CV contribution is incorporated as the difference between all-electron (ae) and fc MP2 calculations, both with the cc-pCVW(T+d)Z basis set.92 Replacement of conventional methods by the explicitly correlated (F12) variants93,94 leads to our current standard version of the approach, which is referred to as junChS-F12.42,58,60 Comparison with the most accurate results available for a panel of representative noncovalent complexes provided an average absolute error smaller than 10 cm–1.41,42,60
The relative stability of different conformers is determined by the corresponding relative enthalpy at 0 K (ΔH0°) or free energy (ΔG°) at a temperature depending on the experimental conditions. The vibrational contributions to thermodynamic functions of stiff degrees of freedom are computed in the framework of second-order vibrational perturbation theory (VPT2),46,48,50,73 employing rDSD harmonic and B3 anharmonic contributions. The same computations provide also anharmonic IR spectra.53,95,96 Low-frequency contributions are, instead, taken into account by means of the quasi-harmonic (QH) approximation.54,97
All the computations have been performed with the Gaussian package98 except the junChS-F12 and QH ones, which have been performed with the help of the Molpro99 and GoodVibes97 software, respectively.
Results and Discussion
The conformational landscape of Phe and Tyr is ruled by two soft degrees of freedom in the backbone (ϕ = H–N–Cα–C′ and ψ = N–Cα–C′–O(H) dihedral angles) and two others in the side chain (χ1 = NCα–Cβ–Cγ and χ2 = Cα–Cβ–Cγ–Cδ dihedral angles) (see Figure 1). However, the nonplanarity of the NH2 moiety suggests to replace the ϕ dihedral angle with ϕ′ = LP–N–Cα–C′ = ϕ + 120°, where LP is the nitrogen lone-pair perpendicular to the plane defined by the two amine hydrogens and the Cα atom. The customary c, g, s, t labels are then used to indicate the cis, gauche, skew, and trans conformations determined by ϕ′, ψ, χ1, and χ2 dihedral angles.
Figure 1.

Structure and dihedral angles of phenylalanine (conformer of type I) and tyrosine (conformer of type II).
Only planar conformations are allowed for the carboxylic moiety of both amino acids (ω = Cα–C′–O–H = 0° or 180°) and the phenol group of tyrosine (χ3 = Cϵ–Cζ–O–H = 0° or 180°). In the former case ω = 0° is strongly preferred (due the formation of a weak hydrogen bridge with the carbonyl oxygen) except when the oxidryl hydrogen is involved in stronger hydrogen bonds with other acceptor groups. In the case of χ3 the two nonequivalent arrangements of the phenol OH moiety are labeled in the following as N and C with reference to the H(O) placement on the same side of the backbone N or C′ atom, respectively.
The exploration of the conformational landscape of phenylalanine provided several low-energy conformers stabilized by the formation of hydrogen bonds, classified as type I (bifurcated NH2···O=C, ϕ′ ≈ 180°, ψ ≈ 180°, ω ≈ 180°), I′ (HNH···O=C, ϕ′ ≈ 90°, ψ ≈ 180°, ω ≈ 180°), and II (N···HO, ϕ′ ≈ 0°, ψ ≈ 0°, ω ≈ 0°).100 Some conformers in which the O(H) oxygen of the carboxyl group forms a bifurcated hydrogen bond with the NH2 moiety (type III, ϕ′ ≈ 180°, ψ ≈ 0°, ω ≈ 180°) have been also found, but they can easily relax to more stable I conformers by rotation of 180° around ψ. In fact, III conformers have been observed in MW spectra only in very special situations.101 Also some conformers of type I relax to more stable structures (still of type I) through rotation around the χ1 dihedral angle. As a consequence, we are left with the six conformers shown in Figure 2, with the two I′ conformers corresponding to the nonequivalent g and g– orientations of the ϕ′ dihedral angle (see Table 1).
Figure 2.

Conformers of phenylalanine with the computed relative free energies at room temperature (in cm–1) given in parentheses. H-bonds are highlighted by dashed lines.
Table 1. rDSD Relative Electronic Energies (ΔErDSD), Harmonic Zero Point Energies (ΔZPEH), Thermal Contributions (ΔThH), and Quasi-Harmonic Corrections to Free Energies (TΔSQH–H), Together with Differences between junChS-F12 and rDSD Electronic Energies (ΔChS) and B3 Anharmonic Corrections to ZPEs (ΔZPEanh–H) for the Low-Lying Conformers of Phenylalaninea.
| label | ΔErDSD | ΔChS | ΔZPEH | ΔThH | ΔZPEanh–H | TΔSQH–H | ΔG°b | ϕ′ | ψ | ω | χ1 | χ2 |
|---|---|---|---|---|---|---|---|---|---|---|---|---|
| Igg | 317.3 | –4.6 | –152.1 | –142.2 | 49.4 | –2.2 | 65.6 | –178.4 | 174.3 | –178.9 | 62.6 | 89.6 |
| Itg | 596.9 | 36.6 | –180.8 | –236.2 | 41.3 | 54.2 | 312.0 | 173.5 | 129.7 | 177.4 | 179.9 | 81.4 |
| I′g–g | 660.4 | –11.0 | –190.7 | –206.1 | 43.1 | 26.9 | 322.6 | –74.4 | 151.9 | 176.2 | –60.4 | 99.1 |
| I′*g–g | 762.1 | –47.7 | –212.8 | –210.3 | 73.0 | 3.5 | 367.8 | 80.2 | –169.2 | –177.2 | –61.9 | 93.9 |
| IIgg | 0.0 | 0.0 | 0.0 | 0.0 | 0.0 | 0.0 | 0.0 | –26.1 | 12.2 | –2.6 | 52.7 | 81.0 |
| IIg–g | 176.9 | 7.5 | –31.6 | –92.8 | 31.6 | –0.9 | 90.7 | 33.8 | –18.0 | 4.0 | –63.2 | 103.3 |
Best estimates of relative free energies at room temperature (ΔG°) and dihedral angles optimized at the rDSD level (ϕ′, ψ, ω, χ1, and χ2) are also given. All the energetic quantities are in cm–1 and the angles in degrees.
Sum of columns 2–7.
The structural and energetic characteristics of those six low-energy conformers are reported in Table 1, whereas the different contributions to the final electronic energies are given in Table S1 and the main spectroscopic parameters are given in Tables S2 and S3. The first noteworthy feature is the remarkable accuracy of rDSD conformational energies (MUE and maximum unsigned error (MAX) of 21.5 and 47.7 cm–1 with respect to the junChS-F12 reference). Although rDSD electronic energies are not directly used to estimate relative stabilities, this finding gives further support to the use of rDSD geometries and harmonic force fields. Different staggered conformations are populated for χ1, which determines the relative position of the phenyl ring with respect to the amino acid backbone (see Figure 1). On the other hand, all low-energy conformers show a Cα–Cβ bond nearly perpendicular to the plane of the phenyl ring (χ2 ≈ 90°) and this feature has been confirmed experimentally by ENDOR spectroscopy for spin-labeled l-phenylalanine.102 Interestingly, the same region of χ2 is dominantly populated and characteristic of Phe residues in proteins.103
Contrary to the usual situation for aliphatic amino acids,59 conformers of type II are more stable than their types I and I′ counterparts in spite of a less favorable orientation of the OH group in the carboxyl moiety. The increased stability of these conformers is related to the existence of a “daisy chain” sequence of interactions involving aromatic π electrons, together with the NH2 and OH groups of the backbone.19 Zero-point and thermal effects stabilize all conformers with respect to the IIgg absolute energy minimum, with this effect being particularly significant for conformers of types I and I′. As a consequence, conformer Igg becomes the second most stable form, with a relative free energy (with respect to IIgg) marginally lower than that of the IIg–g conformer. Inclusion of anharmonic contributions in ZPEs is needed for obtaining quantitative results, but it does not alter the stability order of the different conformers. Finally, the main effect of the QH corrections is to reduce the overestimation of entropic contributions produced by the harmonic oscillator approximation, with the consequent excessive stabilization of all conformers with respect to the IIgg species. The final results suggest that Igg and IIg–g conformers have comparable populations (about 23% each at room temperature) only slightly lower than that of the IIgg conformer (about 33% at room temperature), whereas the other three conformers have equilibrium populations around 7% each at room temperature and their unequivocal characterization could be, therefore, quite difficult.
The rotational constants are very sensitive to the value of the χ1 dihedral angle, with g conformers showing values around 1650, 650, and 450 MHz, whereas g– conformers show values around 2450, 450, and 400 MHz (see Tables 2 and S2). On the other hand, much larger dipole moments are computed for conformers of type II with respect to those of type I, irrespective of the orientation of the side chain (see Table S3). These general trends were used to support the assignment of experimental MW spectra23 and are not very sensitive to the level of the QC computations. However, comparison of Table 2 (and Table S2) with the results of previous investigations23 shows that our computational approach strongly improves the quantitative agreement between the computed and experimental rotational constants of the detected conformers concerning both the parent species and the 15N isotopomers (MUE = 2.2 MHz and MAX = 7.7 MHz).
Table 2. Ground-State Rotational Constants (A0, B0, and C0 in MHz) and 14N-Nuclear Quadrupole Coupling Constants (χ in MHz) of the Two Most Stable Phenylalanine Conformersa.
| IIgg |
IIg–g |
|||||||
|---|---|---|---|---|---|---|---|---|
|
14N |
15N |
14N |
15N |
|||||
| param. | exp.b | calc.c | exp.b | calc.c | exp.b | calc.c | exp.b | calc.c |
| A0 | 1666.0436(14) | 1658.3 | 1646.7381(17) | 1639.2 | 2457.05490(48) | 2456.6 | 2425.69(30) | 2425.7 |
| B0 | 638.56314(12) | 641.3 | 636.56611(19) | 639.4 | 460.659722(79) | 461.1 | 459.64825(32) | 460.2 |
| C0 | 568.76843(15) | 571.1 | 569.83617(20) | 568.3 | 424.74604(13) | 424.5 | 423.71892(25) | 423.6 |
| χaa | –0.283(16) | –0.403 | –0.777(22) | –0.505 | ||||
| χbb | 1.275(55) | 1.479 | 0.0695(36) | –0.112 | ||||
| χcc | –0.992(39) | –1.076 | 0.7075(14) | 0.617 | ||||
Rotational constants of the 15N isotopomers are also reported.
From ref (23).
rDSD-LRA equilibrium geometries, rDSD properties and B3 vibrational corrections.
Despite the accessible relative energies of conformers of types I and I′′ (especially Igg), none of them has been detected in the MW study of ref (23), contrary to the conclusions issued from electronic and IR spectra.10,19 As a matter of fact, the laser energy needed to obtain MW spectra is much higher than that required by other spectroscopic techniques. This leads to an increased photofragmentation of Phe, with this explaining the weakness of the observed spectra.23 At the same time, both theory and experiments agree in indicating that conformers of type II have higher ionization energies than their I and I′ counterparts.20 Thus, if the non observed conformers are preferentially ionized in the laser ablation process, their populations in the supersonic jet would be much lower than those estimated from a Boltzmann distribution.
Based on these premises, the IR spectra in inert matrices should show the signatures of conformers of types I, I′, II, and, possibly, III. However, in the most relevant spectral regions (mainly OH, C=O and C–O stretching together with COH bending regions) conformers of type III are predicted to absorb at frequencies very close to those of the corresponding type II structures. Therefore, they cannot be characterized by this technique too. Computed and experimental spectra in the mid-IR region are compared in Figure 3. The general agreement between theory and experiment confirms that different conformers are present in the experimental sample and that their relative populations are correctly predicted by our computations. The most distinctive features of the spectra concern the regions of C–O (1060–1160 cm–1) and C=O (1750–1800 cm–1) stretchings. The presence of multiple bands in the former region proves that several conformers of type I and I′ are present in the matrix. In the latter region, the bands originating from conformers of type I and I′ fall at lower wave numbers than those of type II conformers due to the involvement of the O=C group in a hydrogen bond with the amine hydrogens. However, this kind of hydrogen bond is not very strong, so that the wavenumber shift is quite limited.
Figure 3.
Comparison between the FT-IR spectrum of matrix-isolated phenylalanine (Exp) and the computed anharmonic spectra for the six most stable conformers and their Boltzmann average (Weight.) in the 500–1250 cm–1 and 1250–1800 cm–1 regions. See main text for details.
The NH stretching vibrations are observed in the experimental spectrum as two broad bands of quite low intensity, centered at about 3340 and 3410 cm–1, and the computed values are in good agreement with this finding for all low-lying conformers (see Table 3). In the case of I and I′ conformers these vibrations correspond to symmetric and antisymmetric combinations of the amine NH stretching modes. However, for conformers of type II, the band above 3400 cm–1 is due to the antisymmetric NH2 stretching, but the band at about 3330 cm–1 can be assigned to the out-of-phase combination of NH and OH stretchings. The broad experimental band centered at 3178 cm–1 is assigned to the in phase combination of NH and OH stretchings of conformers of type II, which is computed slightly above 3200 cm–1. Finally, the OH stretching of conformers of type I and I′ is computed between 3564 and 3579 cm–1 and found experimentally between 3557 and 3567 cm–1. In more general terms, a remarkable quantitative agreement with experiment is obtained for the positions of all the computed IR bands without employing any adjustable parameter.
Table 3. Computed Wavenumbers for the NH and OH Stretching Vibrations of Phenylalanine and Tyrosyne Low-Energy Conformersa.
| phenylalanine | ||||
|---|---|---|---|---|
| assign. | Igg | Itg | I′g–g | I′*g–g |
| νNHsym | 3325.7 | 3318.1 | 3340.8 | 3364.7 |
| νNHasym | 3403.9 | 3406.6 | 3421.7 | 3444.3 |
| νOH | 3572.1 | 3563.6 | 3572.6 | 3579.3 |
| IIgg | IIg–g | |||
| νNHsym + νOH in phase | 3217.8 | 3268.8 | ||
| νNHsym + νOH out of phase | 3339.7 | 3325.9 | ||
| νNHasym | 3427.6 | 3412.0 | ||
| tyrosine | ||||
|---|---|---|---|---|
| assign. | IICgg | IINgg | IICg–g | IINg–g |
| νNHsym + νOH in phase | 3208.5 | 3213.6 | 3243.9 | 3255.9 |
| νNHsym + νOH out-of-phase | 3338.6 | 3337.5 | 3364.5 | 3364.5 |
| νNHasym | 3428.1 | 3426.6 | 3409.9 | 3408.9 |
| νOH phenol | 3652.3 | 3661.0 | 3659.3 | 3661.1 |
See main text for details.
The six Phe conformers discussed above give rise to 12 conformers of tyrosine (Tyr) due to the already mentioned presence of two nonequivalent orientations of the OH group in the para position of the phenyl ring. All these conformers are true energy minima, but, for the same reasons discussed above for Phe, we can expect that only species of type II are detected in MW experiments (see Figure 4). The structural and energetic characteristics of those four low-energy conformers are collected in Table 4, whereas the different contributions to the final electronic energies are given in Table S4. The situation envisaged by our computations was indeed confirmed by MW experiments, but, unfortunately, for conformers IINg–g and IICg–g too few lines could be assigned to allow an unequivocal characterization.24
Figure 4.
Conformers of tyrosine with the computed relative free energies at room temperature (in cm–1) given in parentheses. H-bonds are highlighted by dashed lines.
Table 4. rDSD Relative Electronic Energies (ΔErDSD), Harmonic Zero Point Energies (ΔZPEH), Thermal Contributions (ΔThH), and Quasi-Harmonic Corrections to Free Energies (TΔSQH–H), Together with Differences between junChS-F12 and rDSD Electronic Energies (ΔChS) and B3 Anharmonic Corrections to ZPEs (ΔZPEanh–H) for the Low-Lying Conformers of Tyrosinea.
| label | ΔErDSD | ΔChS | ΔZPEH | ΔThH | ΔZPEanh–H | TΔSQH–H | ΔG°b | ϕ′ | ψ | ω | χ1 | χ2 |
|---|---|---|---|---|---|---|---|---|---|---|---|---|
| IICggc | 0.0 | 0.0 | 0.0 | 0.0 | 0.0 | 0.0 | 0.0 | –26.32 | 12.33 | –2.55 | 53.08 | 80.97 |
| IINggd | 117.5 | 3.3 | –6.8 | –9.5 | 10.7 | 4.0 | 119.2 | –26.18 | 12.34 | –2.64 | 52.72 | 80.42 |
| IICg–gc | 278.0 | 11.5 | –32.9 | –101.9 | –10.4 | 35.6 | 179.9 | 33.92 | –18.25 | 4.09 | –62.64 | 104.25 |
| IINg–gd | 275.7 | 10.8 | –37.3 | –99.7 | –0.9 | 36.7 | 185.3 | 34.06 | –18.33 | 4.08 | –63.05 | 102.84 |
Best estimates of relative free energies at room temperature (ΔG°) and dihedral angles optimized at the rDSD level (ϕ′, ψ, ω, χ1, and χ2) are also given. All the energetic quantities are in cm–1 and the angles in degrees.
Sum of columns 2–7.
χ3 = 0°.
χ3 = 180°.
We are thus left with the IINgg and IICgg conformers, whose computed spectroscopic parameters are compared to their experimental counterparts in Table 5 (see also Table S5). For purposes of completeness, the computed spectroscopic parameters of conformers IINg–g and IICg–g are given in the same tables. The agreement between experimental and computed rotational constants is remarkable, and even more importantly, all the trends and differences between the two conformers are reproduced quantitatively. The percentage MUE for the rotational constants is 0.38% and the percentage MAX 0.56%. These values are comparable to those obtained for the IIgg conformer of Phe and are more than three times smaller than those issued from the MP2 computations reported in ref (24) (percentage MUE = 1.42%, percentage MAX = 1.88%). Larger errors affect the quadrupolar coupling constants, but the agreement between computed and experimental values is largely sufficient to allow an unbiased assignment of different conformers.
Table 5. Ground-State Rotational Constants (A0, B0, and C0 in MHz), 14N-Nuclear Quadrupole Coupling Constants (χ in MHz), and Electric Dipole Moment Components (μ in debye) of the Two Observed Tyrosine Conformersa.
| parameter | IINggexpb | IINggcalcc | IICggexpb | IICggcalcc | IINg–gcalcc | IICg–gcalcc |
|---|---|---|---|---|---|---|
| A0 | 1529.6791(40) | 1522.2 | 1525.2543(29) | 1517.6 | 2400.9 | 2404.3 |
| B0 | 463.94021(32) | 464.5 | 465.48173(25) | 467.7 | 341.0 | 340.7 |
| C0 | 425.76168(40) | 427.1 | 427.31023(27) | 429.4 | 321.1 | 321.6 |
| χaa | 0.709(14) | 0.5819 | 0.740(15) | 0.6111 | –0.5451 | –0.5301 |
| χbb | 0.236(88) | 0.3022 | 0.247(92) | 0.3134 | –0.1324 | –0.2353 |
| χcc | –0.945(88) | –0.8832 | –0.988(92) | –0.9245 | 0.6775 | 0.7654 |
| μa | 2.0713 | 1.4039 | 4.6739 | 4.7805 | ||
| μb | 5.7026 | 3.3171 | –0.7791 | –3.0797 | ||
| μc | 0.9601 | 0.4933 | 2.1722 | 1.2795 |
The spectroscopic parameters computed for the IINg–g and IINg–g conformers are also reported.
From ref (24).
rDSD-LRA equilibrium geometries, rDSD properties and B3 vibrational corrections.
The spectra of tyrosine in the mid-IR region computed taking into account all the type II conformers are shown in Figure 5. Although experimental spectra would probably show some characteristic signatures of conformers of types I and I′ (as discussed above for phenylalanine), in the present context we prefer to avoid any reference to species not well characterized experimentally.
Figure 5.
Computed anharmonic spectra for the four conformers and their Boltzmann average (Weight.) in the 500–1250 cm–1 and 1250–1800 cm–1 regions. See main text for details.
The spectra are very similar to those of the corresponding conformers of phenylalanine except for the presence of a quite intense band at about 1200 cm–1 (1197, 1193, 1198, and 1194 cm–1 for conformer IICgg, IINgg, IICgg–, and IINgg–, respectively) assigned to the bending of the phenyl OH moiety. The same trend is observed also in the NH/OH stretching region, where an additional band due to the phenol OH stretching is found at about 3650 cm–1 (see Table 3). A more detailed analysis is not pursued because it would be very similar to that given above for phenylalanine.
Conclusions
In this paper, a general strategy aimed to the unbiased disentanglement of the conformational bath of flexible biomolecule building blocks has been applied to prototypical α-amino acids containing aromatic side chains. Accurate structures and relative energies are obtained by the rDSD-LRA approach and the junChS-F12 composite method, respectively. Next, the spectroscopic parameters of sufficiently populated conformers can be safely computed at the rDSD level. The results obtained for phenylalanine and tyrosine are in full agreement with the available spectroscopic data, and permit their unbiased interpretation in terms of the cooperation or competition between intrabackbone and backbone-(side chain) hydrogen bonds.
Together with the intrinsic interest of the studied molecules, the results of the present investigation show that highly reliable analysis of structural and spectroscopic features is today possible for flexible building blocks of biomolecules in the gas phase.
Acknowledgments
Funding from the Italian Ministry of University and Research (MUR, Grant 2017A4XRCA) and Italian Space Agency (ASI, “Life in Space” project no. 2019-3-U.0) is gratefully acknowledged.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpca.3c01174.
Additional computed data for phenylalanine and tyrosine, cartesian coordinates of rDSD optimized geometries for all the low-energy conformers of phenylalanine and tyrosine (PDF)
The authors declare no competing financial interest.
Supplementary Material
References
- Barone V.; Biczysko M.; Bloino J.; Puzzarini C. Accurate Structure, Thermochemistry and Spectroscopic Parameters from CC and CC/DFT Schemes: the Challenge of the Conformational Equilibrium in Glycine. Phys. Chem. Chem. Phys. 2013, 15, 10094–10111. 10.1039/c3cp50439e. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alonso E. R.; León I.; Alonso J. L.. The Role of the Intramolecular Interactions in the Structural Behavior of Biomolecules: Insights from Rotational Spectroscopy. In Intra- and Intermolecular Interactions Between Non-Covalently Bonded Species; Elsevier, 2020; pp 93–141. [Google Scholar]
- Simpson H. J. J.; Marsh R. E. The Crystal Structure of L-alanine. Acta Crystallogr. 1966, 20, 550–555. 10.1107/S0365110X66001221. [DOI] [Google Scholar]
- Freedman T. B.; Diem M.; Polavarapu P. L.; Nafie L. A. Vibrational Circular Dichroism in Amino Acids and Peptides. 6. Localized Molecular Orbital Calculations of the Amino Acids Carbon-Hydrogen Stretching Vibrational Circular Dichroism in Deuterated Isotopomers of Alanine. J. Am. Chem. Soc. 1982, 104, 3343–3349. 10.1021/ja00376a016. [DOI] [Google Scholar]
- Godfrey P. D.; Brown R. D.; Rodgers F. M. The Missing Conformers of Glycine and Alanine: Relaxation in Seeded Supersonic Jets. J. Mol. Struct. 1996, 376, 65–81. 10.1016/0022-2860(95)09065-7. [DOI] [Google Scholar]
- Lee Y.; Jung J.; Kim B.; Butz P.; Snoek L. C.; Kroemer R. T.; Simons J. P. Alanyl Side Chain Folding in Phenylalanine: Conformational Assignment Through Ultraviolet Rotational Band Contour Analysis. J. Phys. Chem. A 2004, 108, 69–73. 10.1021/jp0368280. [DOI] [Google Scholar]
- Lesarri A.; Sanchez R.; Cocinero E. J.; López J. C.; Alonso J. L. Coded Amino Acids in Gas Phase: the Shape of Isoleucine. J. Am. Chem. Soc. 2005, 127, 12952–12956. 10.1021/ja0528073. [DOI] [PubMed] [Google Scholar]
- Lambie B.; Ramaekers R.; Maes G. Conformational Behavior of Serine: an Experimental Matrix-Isolation FT-IR and Theoretical DFT(B3LYP)/6-31++G**Study. J. Phys. Chem. A 2004, 108, 10426–10433. 10.1021/jp047192v. [DOI] [Google Scholar]
- Jarmelo S.; Lapinski L.; Nowak M.; Carey P. R.; Fausto R. Preferred Conformers and Photochemical (λ > 200 nm) Reactivity of Serine and 3,3-Dideutero-Serine in the Neutral Form. J. Phys. Chem. A 2005, 109, 5689–5707. 10.1021/jp0511202. [DOI] [PubMed] [Google Scholar]
- Kaczor A.; Reva I. D.; Proniewicz L. M.; Fausto R. Importance of Entropy in the Conformational Equilibrium of Phenylalanine: A Matrix-Isolation Infrared Spectroscopy and Density Functional Theory Study. J. Phys. Chem. A 2006, 110, 2360–2370. 10.1021/jp0550715. [DOI] [PubMed] [Google Scholar]
- Lesarri E. J.; Cocinero J. C.; López J. C.; Alonso J. L. The Shape of Neutral Valine. Angew. Chem., Int. Ed. Engl. 2004, 43, 605–610. 10.1002/anie.200352543. [DOI] [PubMed] [Google Scholar]
- Cocinero J. C.; Lesarri E. J.; Grabow J. U.; López J. C.; Alonso J. L. The Shape of Leucine in the Gas Phase. ChemPhysChem 2007, 8, 599–604. 10.1002/cphc.200600730. [DOI] [PubMed] [Google Scholar]
- Blanco S.; Sanz M. E.; López J. C.; Alonso J. L. Revealing the Multiple Structures of Serine. Proc. Natl. Acad. Sci. U.S.A. 2007, 104, 20183–20188. 10.1073/pnas.0705676104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alonso J. L.; Pérez M.; Eugenia Sanz M.; López J. C.; Blanco S. Seven Conformers of l-Threonine in the Gas Phase: a LA-MB-FTMW Study. Phys. Chem. Chem. Phys. 2009, 11, 617–627. 10.1039/B810940K. [DOI] [PubMed] [Google Scholar]
- Cabezas C.; Varela M.; Peña I.; Mata S.; López J. C.; Alonso J. L. The Conformational Locking of Asparagine. Chem. Commun. 2012, 48, 5934–5936. 10.1039/c2cc32109b. [DOI] [PubMed] [Google Scholar]
- Sanz M. E.; Cabezas C.; Mata S.; Alonso J. L. Rotational Spectrum of Tryptophan. J. Chem. Phys. 2014, 140, 204308. 10.1063/1.4876001. [DOI] [PubMed] [Google Scholar]
- Martinez S. j.; Alfano J. C.; Levy D. H. The Electronic Spectroscopy of the Amino Acids Tyrosine and Phenylalanine in a Supersonic Jet. J. Mol. Spectrosc. 1992, 156, 421–430. 10.1016/0022-2852(92)90243-H. [DOI] [Google Scholar]
- Martinez S. j.; Alfano J. C.; Levy D. H. The Electronic Spectroscopy of Tyrosine and Phenylalanine in a Supersonic Jet: Basic Analogs. J. Mol. Spectrosc. 1993, 158, 82–92. 10.1006/jmsp.1993.1056. [DOI] [Google Scholar]
- Snoek L. C.; Robertson E. G.; Kroemer R. T.; Simons J. P. Conformational Landscapes in Amino Acids: Infrared and Ultraviolet Ion-Dip Spectroscopy of Phenylalanine in the Gas Phase. Chem. Phys. Lett. 2000, 321, 49–56. 10.1016/S0009-2614(00)00320-1. [DOI] [Google Scholar]
- Lee K. T.; Sung J.; Lee K. J.; Park Y. D.; Kim S. K. Conformation-Dependent Ionization Energies of L-Phenylalanine. Angew. Chem., Int. Ed. Engl. 2002, 41, 4114–4117. . [DOI] [PubMed] [Google Scholar]
- Inokuchi Y.; Kobayashi Y.; Ito T.; Ebata T. Conformation of L-Tyrosine Studied by Fluorescence-Detected UV-UV and IR-UV Double-Resonance Spectroscopy. J. Phys. Chem. A 2007, 111, 3209–3215. 10.1021/jp070163a. [DOI] [PubMed] [Google Scholar]
- Shimozono Y.; Yamada K.; Ishiuchi S.; Tsukiyama K.; Fujii M. Revised Conformational Assignments and Conformational Evolution of Tyrosine by Laser Desorption Supersonic Jet Laser Spectroscopy. Phys. Chem. Chem. Phys. 2013, 15, 5163–5175. 10.1039/c3cp43573c. [DOI] [PubMed] [Google Scholar]
- Perez C.; Mata S.; Blanco S.; Lopez J. C.; Alonso J. L. Jet-Cooled Rotational Spectrum of Laser-Ablated Phenylalanine. J. Phys. Chem. A 2011, 115, 9653–9657. 10.1021/jp200800a. [DOI] [PubMed] [Google Scholar]
- Perez C.; Mata S.; Cabezas C.; Lopez J. C.; Alonso J. L. The Rotational Spectrum of Tyrosine. J. Phys. Chem. A 2015, 119, 3731–3735. 10.1021/acs.jpca.5b01792. [DOI] [PubMed] [Google Scholar]
- Godfrey P. D.; Brown R. D. Proportions of Species Observed in Jet Spectroscopy-Vibrational Energy Effects: Histamine Tautomers and Conformers. J. Am. Chem. Soc. 1998, 120, 10724–10732. 10.1021/ja980560m. [DOI] [Google Scholar]
- Florio G. M.; Christie R. A.; Jordan K. D.; Zwier T. S. Conformational Preferences of Jet-Cooled Melatonin: Probing trans- and cis-Amide Regions of the Potential Energy Surface. J. Am. Chem. Soc. 2002, 124, 10236–10247. 10.1021/ja0265916. [DOI] [PubMed] [Google Scholar]
- Helgaker T.; Klopper W.; Tew D. P. Quantitative Quantum Chemistry. Mol. Phys. 2008, 106, 2107–2143. 10.1080/00268970802258591. [DOI] [Google Scholar]
- Puzzarini C.; Barone V. Extending the Molecular Size in Accurate Quantum-Chemical Calculations: The Equilibrium Structure and Spectroscopic Properties of Uracil. Phys. Chem. Chem. Phys. 2011, 13, 7189–7197. 10.1039/c0cp02636k. [DOI] [PubMed] [Google Scholar]
- Karton A. A Computational Chemist’s Guide to Accurate Thermochemistry for Organic Molecules. WIRES, Comp. Mol. Sci. 2016, 6, 292–310. 10.1002/wcms.1249. [DOI] [Google Scholar]
- Kesharwani M. K.; Karton A.; Martin J. M. Benchmark Ab Initio Conformational Energies for the Proteinogenic Amino Acids Through Explicitly Correlated Methods. Assessment of Density Functional Methods. J. Chem. Theory Comput. 2016, 12, 444–454. 10.1021/acs.jctc.5b01066. [DOI] [PubMed] [Google Scholar]
- Puzzarini C.; Bloino J.; Tasinato N.; Barone V. Accuracy and Interpretability: The Devil and the Holy Grail. New Routes Across Old Boundaries in Computational Spectroscopy. Chem. Rev. 2019, 119, 8131–8191. 10.1021/acs.chemrev.9b00007. [DOI] [PubMed] [Google Scholar]
- Wang P.; Shu c.; Ye H.; Biczysko M. Structural and Energetic Properties of Amino Acids and Peptides Benchmarked by Accurate Theoretical and Experimental Data. J. Phys. Chem. A 2021, 125, 9826–9837. 10.1021/acs.jpca.1c06504. [DOI] [PubMed] [Google Scholar]
- Mancini G.; Fusè M.; Lazzari F.; Chandramouli B.; Barone V. Unsupervised Search of Low-Lying Conformers With Spectroscopic Accuracy: A Two-Step Algorithm Rooted Into the Island Model Evolutionary Algorithm. J. Chem. Phys. 2020, 153, 124110. 10.1063/5.0018314. [DOI] [PubMed] [Google Scholar]
- Ferro-Costas D.; Mosquera-Lois I.; Fernandez-Ramos A. Torsiflex: an Automatic Generator of Torsional Conformers. Application to the Twenty Proteinogenic Amino Acids. J. Cheminf. 2021, 13, 100. 10.1186/s13321-021-00578-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chandramouli B.; Del Galdo S.; Fusè M.; Barone V.; Mancini G. Two-Level Stochastic Search of Low-Energy Conformers For Molecular Spectroscopy: Implementation and Validation of MM and QM Models. Phys. Chem. Chem. Phys. 2019, 21, 19921–19934. 10.1039/C9CP03557E. [DOI] [PubMed] [Google Scholar]
- Barone V.; Lupi J.; Salta Z.; Tasinato N. Development and Validation of a Parameter-Free Model Chemistry for the Computation of Reliable Reaction Rates. J. Chem. Theory Comput. 2021, 17, 4913–4928. 10.1021/acs.jctc.1c00406. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mancini G.; Fusè M.; Lazzari F.; Barone V. Fast Exploration of Potential Energy Surfaces With a Joint Venture of Quantum Chemistry, Evolutionary Algorithms and Unsupervised Learning. Digit. Discovery 2022, 1, 790–805. 10.1039/D2DD00070A. [DOI] [Google Scholar]
- Penocchio E.; Piccardo M.; Barone V. Semiexperimental Equilibrium Structures for Building Blocks of Organic and Biological Molecules: The B2PLYP Route. J. Chem. Theory Comput. 2015, 11, 4689–4707. 10.1021/acs.jctc.5b00622. [DOI] [PubMed] [Google Scholar]
- Ceselin G.; Barone V.; Tasinato N. Accurate Biomolecular Structures by the Nano-LEGO Approach: Pick the Bricks and Build Your Geometry. J. Chem. Theory Comput. 2021, 17, 7290–7311. 10.1021/acs.jctc.1c00788. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Puzzarini C.; Biczysko M.; Barone V.; Largo L.; Pena I.; Cabezas C.; Alonso J. L. Accurate Characterization of the Peptide Linkage in the Gas Phase: a Joint Quantum-Chemistry and Rotational Spectroscopy Study of the Glycine Dipeptide Analogue. J. Phys. Chem. Lett. 2014, 5, 534–540. 10.1021/jz402744a. [DOI] [PubMed] [Google Scholar]
- Alessandrini S.; Barone V.; Puzzarini C. Extension of the “Cheap” Composite Approach to Noncovalent Interactions: The jun-ChS Scheme. J. Chem. Theory Comput. 2020, 16, 988–1006. 10.1021/acs.jctc.9b01037. [DOI] [PubMed] [Google Scholar]
- Lupi J.; Alessandrini S.; Barone V.; Puzzarini C. junChS and junChS-F12 Models: Parameter-free Efficient yet Accurate Composite Schemes for Energies and Structures of Noncovalent Complexes. J. Chem. Theory Comput. 2021, 17, 6974–6992. 10.1021/acs.jctc.1c00869. [DOI] [PubMed] [Google Scholar]
- Gyevi-Nagy L.; Kállay M.; Nagy P. R. Accurate Reduced-Cost CCSD(T) Energies: Parallel Implementation, Benchmarks and Large-Scale Applications. J. Chem. Theory Comput. 2021, 17, 860–878. 10.1021/acs.jctc.0c01077. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kállay M.; Horvath R. A.; Gyevi-Nagy L.; Nagy P. R. Size-Consisten Explicitly Correlated Triple Excitation Correction. J. Chem. Phys. 2021, 155, 034107. 10.1063/5.0057426. [DOI] [PubMed] [Google Scholar]
- Nagy P. R.; Gyevi-Nagy L.; Lorincz B. D.; Kallay M. Pursuing the Basis Set Limit of CCSD(T) Non-Covalent Interaction Energies for Medium-Sized Complexes: Case Study on the S66 Compilation. Mol. Phys. 2022, 120, e2109526. 10.1080/00268976.2022.2109526. [DOI] [Google Scholar]
- Papousek D.; Aliev M. R.. Molecular Vibrational-Rotational Spectra; Elsevier Scientific Publishing Company, 1982. [Google Scholar]
- Gaw F.; Willetts A.; Handy N.; Green W. In Advances in Molecular Vibrations and Collision Dynamics; Bowman J. M., Ed.; JAI Press, 1992; Vol. 1; pp 186–195. [Google Scholar]
- Clabo D. A. Jr.; Allen W. D.; Remington R. B.; Yamaguchi Y.; Schaefer H. F. III A Systematic Study of Molecular Vibrational Anharmonicity and Vibration-Rotation Interaction by Self-Consistent-Field Higher-Derivative Methods. Asymmetric Top Molecules. Chem. Phys. 1988, 123, 187–239. 10.1016/0301-0104(88)87271-9. [DOI] [Google Scholar]
- Burcl R.; Carter S.; Handy N. C. On the Representation of Potential Energy Surfaces of Polyatomic Molecules in Normal Coordinates: II. Parameterisation of the Force Field. Chem. Phys. Lett. 2003, 373, 357–365. 10.1016/S0009-2614(03)00552-9. [DOI] [Google Scholar]
- Barone V. Anharmonic Vibrational Properties by a Fully Automated Second Order Perturbative Approach. J. Chem. Phys. 2005, 122, 014108. 10.1063/1.1824881. [DOI] [PubMed] [Google Scholar]
- Rosnik A. M.; Polik W. F. VPT2+K Spectroscopic Constants and Matrix Elements of the Transformed Vibrational Hamiltonian of a Polyatomic Molecule With Resonances Using Van Vleck Perturbation Theory. Mol. Phys. 2014, 112, 261–300. 10.1080/00268976.2013.808386. [DOI] [Google Scholar]
- Franke P. R.; Stanton J. F.; Douberly G. E. How to VPT2: Accurate and Intuitive Simulations of CH Stretching Infrared Spectra Using VPT2+ K with Large Effective Hamiltonian Resonance Treatments. J. Phys. Chem. A 2021, 125, 1301–1324. 10.1021/acs.jpca.0c09526. [DOI] [PubMed] [Google Scholar]
- Mendolicchio M.; Bloino J.; Barone V. Perturb-Then-Diagonalize Vibrational Engine Exploiting Curvilinear Internal Coordinates. J. Chem. Theory Comput. 2022, 18, 7603–7619. 10.1021/acs.jctc.2c00773. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Grimme S. Supramolecular Binding Thermodynamics by Dispersion-Corrected Density Functional Theory. Chemistry, Europ. J. 2012, 18, 9955–9964. 10.1002/chem.201200497. [DOI] [PubMed] [Google Scholar]
- Li S.-C.; Lin Y.-C.; Li Y.-P. Comparative Analysis of Uncoupled Mode Approximations for Molecular Thermochemistry and Kinetics. J. Chem. Theory Comput. 2022, 18, 6866–6877. 10.1021/acs.jctc.2c00664. [DOI] [PubMed] [Google Scholar]
- Barone V.; Ceselin G.; Fusè M.; Tasinato N. Accuracy Meets Interpretability for Computational Spectroscopy by Means of Hybrid and Double-Hybrid Functionals. Front. Chem. 2020, 8, 584203. 10.3389/fchem.2020.584203. [DOI] [PMC free article] [PubMed] [Google Scholar]
- León I.; Fusè M.; Alonso E. R.; Mata S.; Mancini G.; Puzzarini C.; Alonso J. l.; Barone V. Unbiased Disentanglement of Conformational Baths With the Help of Microwave Spectroscopy, Quantum Chemistry, and Artificial Intelligence: The Puzzling Case of Homocysteine. J. Chem. Phys. 2022, 157, 074107. 10.1063/5.0102841. [DOI] [PubMed] [Google Scholar]
- Barone V.; Fusè M.; Aguado R.; Potenti S.; León I.; Alonso E. R.; Mata S.; Lazzari F.; Mancini G.; Spada L. Bringing Machine-Learning Enhanced Quantum Chemistry and Microwave Spectroscopy to Conformational Landscape Exploration: The Paradigmatic Case of 4-Fluoro-Threonine. Chem. - Eur. J. 2023, e202203990. 10.1002/chem.202203990. [DOI] [PubMed] [Google Scholar]
- Barone V.; Fusè M.; Lazzari F.; Mancini G. Benchmark Structures and Conformational Landscapes of Amino Acids in the Gas Phase: a Joint Venture of Machine Learning, Quantum Chemistry, and Rotational Spectroscopy. J. Chem. Theory Comput. 2023, 19, 1243–1260. 10.1021/acs.jctc.2c01143. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Barone V.; Di Grande S.; Puzzarini C. Toward Accurate Yet Effective Computations of Rotational Spectroscopy Parameters for Biomolecule Building Blocks. Molecules 2023, 28, 913. 10.3390/molecules28020913. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Becke A. D. Density-Functional Exchange-Energy Approximation With Correct Asymptotic Behavior. Phys. Rev. A 1988, 38, 3098–3100. 10.1103/PhysRevA.38.3098. [DOI] [PubMed] [Google Scholar]
- 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. 10.1063/1.456153. [DOI] [Google Scholar]
- Grimme S.; Antony J.; Ehrlich S.; Krieg H. A Consistent and Accurate Ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010, 132, 154104. 10.1063/1.3382344. [DOI] [PubMed] [Google Scholar]
- Santra G.; Sylvetsky N.; Martin J. M. L. Minimally Empirical Double-Hybrid Functionals Trained against the GMTKN55 Database: revDSD-PBEP86-D4, revDOD-PBE-D4, and DOD-SCAN-D4. J. Phys. Chem. A 2019, 123, 5129–5143. 10.1021/acs.jpca.9b03157. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dunning T. H.; Peterson K. A.; Wilson A. K. Gaussian Basis Sets for Use in Correlated Molecular Calculations. X. The Atoms Aluminum Through Argon Revisited. J. Chem. Phys. 2001, 114, 9244–9253. 10.1063/1.1367373. [DOI] [Google Scholar]
- Papajak E.; Zheng J.; Xu X.; Leverentz H. R.; Truhlar D. G. Perspectives on Basis Sets Beautiful: Seasonal Plantings of Diffuse Basis Functions. J. Chem. Theory Comput. 2011, 7, 3027–3034. 10.1021/ct200106a. [DOI] [PubMed] [Google Scholar]
- Biczysko M.; Panek P.; Scalmani G.; Bloino J.; Barone V. Harmonic and Anharmonic Vibrational Frequency Calculations with the Double-Hybrid B2PLYP Method: Analytic Second Derivatives and Benchmark Studies. J. Chem. Theory Comput. 2010, 6, 2115–2125. 10.1021/ct100212p. [DOI] [PubMed] [Google Scholar]
- Kang Y. K.; Park H. S. Assessment of CCSD(T), MP2, DFT- D, CBS-QB3, and G4(MP2) Methods for Conformational Study of Alanine and Proline Dipeptides. Chem. Phys. Lett. 2014, 600, 112–117. 10.1016/j.cplett.2014.03.067. [DOI] [Google Scholar]
- Kang Y. K.; Park H. S. Exploring Conformational Preferences of Alanine Tetrapeptide by CCSD(T), MP2, and Dispersion- Corrected DFT Methods. Chem. Phys. Lett. 2018, 702, 69–75. 10.1016/j.cplett.2018.05.006. [DOI] [Google Scholar]
- Hait D.; Head-Gordon M. How Accurate is Density Functional Theory at Predicting Dipole Moments? An Assessment Using a New Database of 200 Benchmark Values. J. Chem. Theory Comput. 2018, 14, 1969–1981. 10.1021/acs.jctc.7b01252. [DOI] [PubMed] [Google Scholar]
- Zapata Trujillo J. C.; McKemmish L. K. Model Chemistry Recommendations for Scaled Harmonic Frequency Calculations: A Benchmark Study. J. Phys. Chem. A 2023, 127 (7), 1715–1735. 10.1021/acs.jpca.2c06908. [DOI] [PubMed] [Google Scholar]
- Pulay P.; Meyer W.; Boggs J. E. Cubic Force Constants and Equilibrium Geometry of Methane from Hartree–Fock and Correlated Wavefunctions. J. Chem. Phys. 1978, 68, 5077–5085. 10.1063/1.435626. [DOI] [Google Scholar]
- Mills I. M. Vibration–Rotation Structure in Asymmetric- and Symmetric-Top Molecules. Mol. Spectrosc.: Mod. Res. 1972, 1, 115–140. 10.1016/B978-0-12-580640-4.50013-3. [DOI] [Google Scholar]
- Puzzarini C.; Stanton J. F.; Gauss J. Quantum-Chemical Calculation of Spectroscopic Parameters for Rotational Spectroscopy. Int. Rev. Phys. Chem. 2010, 29, 273–367. 10.1080/01442351003643401. [DOI] [Google Scholar]
- Liévin J.; Demaison J.; Herman M.; Fayt A.; Puzzarini C. Comparison of the Experimental, Semi-Experimental and Ab Initio Equilibrium Structures of Acetylene: Influence of Relativisitic Effects and of the Diagonal Born–Oppenheimer Corrections. J. Chem. Phys. 2011, 134, 064119. 10.1063/1.3553203. [DOI] [PubMed] [Google Scholar]
- Puzzarini C.; Stanton J. F. Connections Between the Accuracy of Rotational Constants and Equilibrium Molecular Structures. Phys. Chem. Chem. Phys. 2023, 25, 1421–1429. 10.1039/D2CP04706C. [DOI] [PubMed] [Google Scholar]
- Piccardo M.; Penocchio E.; Puzzarini C.; Biczysko M.; Barone V. Semi-Experimental Equilibrium Structure Determinations by Employing B3LYP/SNSD Anharmonic Force Fields: Validation and Application to Semirigid Organic Molecules. J. Phys. Chem. A 2015, 119, 2058–2082. 10.1021/jp511432m. [DOI] [PubMed] [Google Scholar]
- Puzzarini C.; Heckert M.; Gauss J. The Accuracy of Rotational Constants Predicted by High-Level Quantum-Chemical Calculations. I. Molecules Containing First-Row Atoms. J. Chem. Phys. 2008, 128, 194108. 10.1063/1.2912941. [DOI] [PubMed] [Google Scholar]
- Heim Z. N.; Amberger B. K.; Esselman B. J.; Stanton J. F.; Woods R. C.; McMahon R. J. Molecular Structure Determination: Equilibrium Structure of Pyrimidine (m–C4H4N2) From Rotational Spectroscopy (reSE) and High-Level Ab Initio Calculation (re) Agree Within the Uncertainty of Experimental Measurement. J. Chem. Phys. 2020, 152, 104303. 10.1063/1.5144914. [DOI] [PubMed] [Google Scholar]
- Gardner M. B.; Westbrook B. R.; Fortenberry R. C.; Lee T. J. Highly-Accurate Quartic Force Fields for the Prediction of Anharmonic Rotational Constants and Fundamental Vibrational Frequencies. Spectrochim. Acta, Part A 2021, 248, 119184. 10.1016/j.saa.2020.119184. [DOI] [PubMed] [Google Scholar]
- Watrous A. G.; Westbrook B. R.; Fortenberry R. F12-TZ-cCR: a Methodology for Faster and Still Highly Accurate Quartic Force Fields. J. Phys. Chem. A 2021, 125, 10532–10540. 10.1021/acs.jpca.1c08355. [DOI] [PubMed] [Google Scholar]
- Alonso E. R.; Fusè M.; Leòn I.; Puzzarini C.; Alonso J. L.; Barone V. Exploring the Maze of Cycloserine Conformers in the Gas Phase Guided by Microwave Spectroscopy and Quantum Chemistry. J. Phys. Chem. A 2021, 125, 2121–2129. 10.1021/acs.jpca.1c00455. [DOI] [PubMed] [Google Scholar]
- Melli A.; Tonolo F.; Barone V.; Puzzarini C. Extending the Applicability of the Semi-experimental Approach by Means of Template Molecule and Linear Regression Models on Top of DFT Computations. J. Phys. Chem. A 2021, 125, 9904–9916. 10.1021/acs.jpca.1c07828. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gordy W.; Cook R. L.; Weissberger A.. Microwave Molecular Spectra; Wiley: New York, 1984; Vol. 18. [Google Scholar]
- Shavitt I.; Bartlett R. J.. Many-Body Methods in Chemistry and Physics: MBPT and Coupled-Cluster Theory; Cambridge University Press, 2009. [Google Scholar]
- Raghavachari K.; Trucks G. W.; Pople J. A.; Head-Gordon M. A Fifth-Order Perturbation Comparison of Electron Correlation Theories. Chem. Phys. Lett. 1989, 157, 479–483. 10.1016/S0009-2614(89)87395-6. [DOI] [Google Scholar]
- Peterson K. A.; Dunning T. H. Accurate Correlation Consistent Basis Sets for Molecular Core-Valence Correlation effects: The Second Row Atoms Al-Ar, and the First Row Atoms B-Ne Revisited. J. Chem. Phys. 2002, 117, 10548–10560. 10.1063/1.1520138. [DOI] [Google Scholar]
- DeYonker N. J.; Cundari T. R.; Wilson A. K. The Correlation Consistent Composite Approach (ccCA): an Alternative to the Gaussian-n Methods. J. Chem. Phys. 2006, 124, 114104. 10.1063/1.2173988. [DOI] [PubMed] [Google Scholar]
- DeYonker N. J.; Wilson B. R.; Pierpont A. W.; Cundari T. R.; Wilson A. K. Towards the Intrinsic Error of the Correlation Consistent Composite Approach (ccCA). Mol. Phys. 2009, 107, 1107–1121. 10.1080/00268970902744359. [DOI] [Google Scholar]
- Møller C.; Plesset M. S. Note on an Approximation Treatment for Many-Electron Systems. Phys. Rev. 1934, 46, 618–622. 10.1103/PhysRev.46.618. [DOI] [Google Scholar]
- Helgaker T.; Klopper W.; Koch H.; Noga J. Basis-Set Convergence of Correlated Calculations on Water. J. Chem. Phys. 1997, 106, 9639–9646. 10.1063/1.473863. [DOI] [Google Scholar]
- Hill J. G.; Mazumder S.; Peterson K. A. Correlation Consistent Basis Sets for Molecular Core-Valence Effects With Explicitly Correlated Wave Functions: The Atoms B–Ne and Al–Ar. J. Chem. Phys. 2010, 132, 054108. 10.1063/1.3308483. [DOI] [PubMed] [Google Scholar]
- Werner H.-J.; Adler T. B.; Manby F. R. General Orbital Invariant MP2-F12 Theory. J. Chem. Phys. 2007, 126, 164102. 10.1063/1.2712434. [DOI] [PubMed] [Google Scholar]
- Knizia G.; Adler T. B.; Werner H.-J. Simplified CCSD(T)-F12 Methods: Theory and Benchmarks. J. Chem. Phys. 2009, 130, 054104. 10.1063/1.3054300. [DOI] [PubMed] [Google Scholar]
- Barone V. Accurate Vibrational Spectra of Large Molecules by Density Functional Computations Beyond the Harmonic Approximation: the Case of Pyrrole and Furan. Chem. Phys. Lett. 2004, 383, 528–532. 10.1016/j.cplett.2003.11.066. [DOI] [Google Scholar]
- Barone V.; Festa G.; Grandi A.; Rega N.; Sanna N. Accurate Vibrational Spectra of Large Molecules by Density Functional Computations Beyond the Harmonic Approximation: the Case of Uracil and 2-thiouracil. Chem. Phys. Lett. 2004, 388, 279–283. 10.1016/j.cplett.2004.03.024. [DOI] [Google Scholar]
- Luchini G.; Alegre-Requena J.; Funes-Ardoiz I.; Paton R. GoodVibes: Automated Thermochemistry for Heterogeneous Computational Chemistry Data. F1000Res. 2020, 9, 291. 10.12688/f1000research.22758.1. [DOI] [Google Scholar]
- 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.; Li X.; Caricato M.; Marenich A. V.; Bloino J.; Janesko B. G.; Gomperts R.; Mennucci B.; Hratchian H. P.; Ortiz J. V.; Izmaylov A. F.; Sonnenberg J. L.; Williams-Young D.; Ding F.; Lipparini F.; Egidi F.; Goings J.; Peng B.; Petrone A.; Henderson T.; Ranasinghe D.; Zakrzewski V. G.; Gao J.; Rega N.; Zheng G.; Liang W.; Hada M.; Ehara M.; Toyota K.; Fukuda R.; Hasegawa J.; Ishida M.; Nakajima T.; Honda Y.; Kitao O.; Nakai H.; Vreven T.; Throssell K.; Montgomery J. A. Jr.; Peralta J. E.; Ogliaro F.; Bearpark M.; Heyd J. J.; Brothers E. N.; Kudin K. N.; Staroverov V. N.; Kobayashi R.; Normand J.; Raghavachari K.; Rendell A.; Burant J. C.; Iyengar S. S.; Tomasi J.; Cossi M.; Millam J. M.; Klene M.; Adamo C.; Cammi R.; Ochterski J. W.; Martin R. L.; Morokuma K.; Farkas O.; Foresman J. B.; Fox D. J.. Gaussian 16, revision C.01; Gaussian, Inc.: Wallingford, CT, 2016.
- Werner H.-J.; Knowles P. J.; Manby F. R.; Black J. A.; Doll K.; Heßelmann A.; Kats D.; Köhn A.; Korona T.; Kreplin D. A.; et al. The Molpro Quantum Chemistry Package. J. Chem. Phys. 2020, 152, 144107. 10.1063/5.0005081. [DOI] [PubMed] [Google Scholar]
- Alonso J. L.; López J. C.. Microwave Spectroscopy of Biomolecular Building Blocks. In Gas-Phase IR Spectroscopy and Structure of Biological Molecules; Rijs A. M.; Oomens J., Eds.; Springer, 2015; pp 335–401. [DOI] [PubMed] [Google Scholar]
- Jimenéz A. I.; Vaquero V.; Cabezas C.; López J. C.; Cativiela C.; Alonso J. L. The Singular Gas-Phase Structure of 1-Aminocyclopropanecarboxylic Acid (Ac3c). J. Am. Chem. Soc. 2011, 133, 10621–10628. 10.1021/ja2033603. [DOI] [PubMed] [Google Scholar]
- Joela H.; Mustafi D.; Fair C. C.; Makinen M. W. Structure and Conformation of Spin-Labeled Methyl L-Phenylalaninate and L-Phenylalanine Determined by Electron Nuclear Double Resonance Spectroscopy. J. Phys. Chem. 1991, 95, 9135–9144. 10.1021/j100176a021. [DOI] [Google Scholar]
- Duan G. L.; Smith V. H. Jr.; Weaver D. F. Data Mining, Ab Initio, and Molecular Mechanics Study on Conformation of Phenylalanine and its Interaction with Neighboring Backbone Amide Groups in Proteins. Int. J. Quantum Chem. 2002, 90, 669–683. 10.1002/qua.956. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.



