Joanna
Sadlej
*ab,
Jan Cz.
Dobrowolski
bc,
Joanna E.
Rode
c and
Michał H.
Jamróz
c
aDepartment of Chemistry, Warsaw University, 1 Pasteura Street, 02-093, Warsaw, Poland
bNational Institute of Public Health, 30/34 Chełmska Street, 00-725, Warsaw, Poland
cIndustrial Chemistry Research Institute, 8 Rydygiera Street, 01-793, Warsaw, Poland
First published on 16th November 2005
This paper presents a discussion of the interaction energies, conformations, vibrational absorption (VA, harmonic and anharmonic) and vibrational circular dichroism (VCD) spectra for conformers of monomeric chiral D(−)-lactic acid and their complexes with water at the DFT(B3LYP)/aug-cc-pVDZ and DFT(B3LYP)/aug-cc-pVTZ levels. A detailed analysis has been performed principally for the two most stable complexes with water, differing by lactic acid conformation. The VCD spectra were found to be sensitive to conformational changes of both free and complexed molecules, and to be especially useful for discriminating between different chiral forms of intermolecular hydrogen bonding complexes. In particular, we show that the VCD modes of an achiral water molecule after complex formation acquire significant rotational strengths whose signs change in line with the geometry of the complex. Using the theoretical prediction, we demonstrate that the VCD technique can be used as a powerful tool for structural investigation of intermolecular interactions of chiral molecules and can yield information complementary to data obtained through other molecular spectroscopy methods.
VCD is an attractive technique, since the VCD spectrum provides a unique fingerprint information for a chiral molecule. It requires the molecule to have at least one chiral center (element), which allows the molecule to exhibit chirality and optical activity. The VCD spectra of chiral molecules are of interest from both a theoretical and biophysical standpoint. Each VCD band reports information regarding the molecular structure and the coupling of particular vibrational modes in the chiral molecules. Thus, the VCD spectrum is a subtle, stereospecific effect associated with the interaction of a chiral molecule with light. Although the intensity scale of the VCD spectra is approximately four orders of magnitude smaller than the scale for the parent VA (IR) spectra, nowadays advances in laser technology enables one to collect spectra of biologically interesting molecules and to reveal sensitive structural and stereochemical details of molecules in an aqueous environment.2
The ab initio calculations of VCD spectra prove a difficult and time-consuming computational task. In order to predict a VCD spectrum, it is necessary to evaluate the vibrational electric and magnetic dipole transition moments. Moreover, as so frequently occurs in the case of calculations of properties involving magnetic dipole operators, there arises a question of the origin-dependence of the results. As a consequence, most ab initio studies of the VCD effect focused on medium-sized molecules exhibiting optical activity.3 Theoretical prediction of the IR absorption and VCD spectra within the harmonic approximation requires calculations of harmonic frequencies and dipole and rotational strengths. These quantities, in turn, demand calculation of the molecular Hessian, harmonic force field (HFF), atomic polar tensors (APT), and atomic axial tensors (AAT).4–6
Several recent developments have enhanced the efficiency of calculations of VCD spectra involving the application of density functional theory (DFT) methods. Recently, DFT has been accepted by the ab initio quantum chemistry community as a cost-effective approach to computations of molecular structures and spectra (vibrational and NMR) of molecules of chemical interest. Many studies have shown that vibrational frequencies calculated by DFT methods are more reliable than those obtained at MP2 level.7,8
For VCD spectra calculations, the implementation of direct analytical derivative methods for calculating HFFs, APTs, and AATs9 is of most importance. Incorporation of gauge invariant atomic orbitals (GIAO) into the calculation of AATs makes it possible to predict the VCD spectra using the DFT method. It was shown that the spectra predicted using the DFT hybrid functional methods were of greater accuracy than those produced by the HF and MCSCF calculations.10,11
Numerous experimental and theoretical studies were devoted to proving the advantage of engaging the VCD spectra in determinations of conformational structure shaped by intramolecular hydrogen bonding.12 However, the subject of the influence of intermolecular H-bonds was undertaken in relatively few papers, where geometrical arrangement of the H-bond partners was shown to be determinable based on the VCD spectra.13–16 Our current interest in the VCD spectra is provoked by the potential use of the VCD technique in this way.
This paper aims to explore the role of VCD spectra in determining the intermolecular H-bond geometry of the complex of D-lactic acid with a water molecule. Lactic acid belongs to the group of α-hydroxyacids and is one of the simplest chiral molecules usually chosen as a model for studying biological systems exhibiting organic acid-type bonding. In eukaryotic cells, lactic acid is produced as a result of glucose anaerobic metabolism. It is also the main metabolic product of lactic acid bacteria. The molecule displays a quasi-planar HOC–COOH molecular skeleton with the oxygen atom of the aliphatic hydroxyl groups beside the CO group in the crystal structure.17 The acidic hydroxyl group is involved in intermolecular H-bonding as a proton donor to the aliphatic hydroxyl group of a neighboring molecule, while the aliphatic hydroxyl group also participates in a bifurcated intramolecular H-bond, in which the aliphatic hydroxyl and acid carbonyl oxygen atoms of a symmetry-related molecule participate as acceptors. The microwave spectrum analysis of lactic acid in the gaseous phase shows that a conformation with a H-bond from the α-hydroxyl group to the carbonyl oxygen is the most stable form of the molecule.18 The Raman spectra of lactic acid have been collected in aqueous solution,19,20 where the lactic acid is shown to be dissociated.
Lactic acid was studied theoretically by Pecul et al.21 at the SCF and MP2/aug-cc-pVDZ levels. However, only two low energy minima on the potential energy surface (PES) were discussed in their work. Recently, an analysis of argon and xenon matrix-isolation FT-IR spectra at 9 K has been published and supported with a systematic search for the possible minima on the PES, using the DFT(B3LYP)/6-311++G(d,p) and MP2/6-31G(d,p) calculations.22 That very experiment was the first to show the existence of three low energy conformers of lactic acid.
To the best of our knowledge, no experimental VCD spectra of the lactic acid optical isomers are available in the literature. A comparison of theoretical predictions and experimental spectra would be difficult even if more experimental data were available, because lactic acid may be present in most solvents in the form of dimers (and/or complexes with the solvent). However, according to the authors of ref. 22, “water doping of the matrix allowed the identification of spectral features ascribable to weakly bound complexes of lactic acid with water”. In the FT-IR spectra the isolated lactic acid monomers (three conformers) and weakly bonded lactic 1 : 1 acid⋯water complexes were detected.22 These findings yielded us an additional argument to study the complexes theoretically.
The lactic acid molecule (Scheme 1) has four sites potentially amenable to forming H-bonds with water, i.e., the proton donor carboxylic OH group, alcoholic COH group, proton acceptor carboxylic CO group and alcoholic oxygen atom. The H-bond of water with the whole –COOH group seems to be of particular importance, as both the water molecule and the –COOH group simultaneously play proton donor and proton acceptor roles. All these possibilities turn the hydration of a lactic acid molecule into an interesting subject to be investigated by theoretical means. The interaction with water is expected to affect the conformational stability of the lactic acid molecule; hence characterization of the complexes with water seems to be important for understanding the dissociation process as well as for acquiring other valuable information on the properties of the H-bonding of this molecule.
![]() | ||
Scheme 1 |
As far as we are aware, with the exception of the study of Borba et al.,22 the lactic acid–water complex has not been observed in spectroscopic experiments or considered in theoretical studies. In this study we analyze the conformational stability of 1 : 1 H-bonded complexes formed between lactic acid and water. Thus, the first aim of this paper is to study possible structures and the infrared spectra of stable hydrated complexes. The second priority of this paper is to obtain some insight into the influence of the molecular conformation and presence of intermolecular H-bonds on the VCD spectra using lactic acid as a model system. We hope that the analysis of the influence of molecular conformation on the VCD spectra will contribute to the development of VCD spectroscopy as a tool for structural investigations of biologically active compounds. We wish to utilize this feature in our computational research.
![]() | ||
Fig. 1 Considered conformers of D-lactic acid. |
aug-cc-pVDZ | aug-cc-pVTZ | |||||||
---|---|---|---|---|---|---|---|---|
Conformer | D e(X–I)/kJ mol−1a | D 0(X–I)/kJ mol−1a | ΔG298/kJ mol−1 | μ/D | D e(X–I)kJ mol−1a | D 0(X–I)kJ mol−1a | ΔG298/kJ mol−1 | μ/D |
a D(X–I) = D(X)–D(I) | ||||||||
I | 0 | 0 | 0 | 2.383 | 0 | 0 | 0 | 2.370 |
II | 8.79 | 8.70 | 8.12 | 1.616 | 9.29 | 9.25 | 8.28 | 1.519 |
III | 10.63 | 10.46 | 10.04 | 4.928 | 10.75 | 10.33 | 9.46 | 4.888 |
IV | 19.37 | 18.74 | 18.07 | 3.278 | 20.08 | 18.95 | 17.74 | 3.258 |
According to the methods used in this work, conformer I is predicted to be the most stable, with the energy differences with conformers II, III, and IV being 9.25, 10.33, and 18.95 kJ mol−1, respectively. The DFT(B3LYP)/aug-cc-pVDZ method used throughout this study is fairly sophisticated for monomers, and it yields relative stabilization and Gibbs free energies concordant with those obtained using the DFT(B3LYP)/aug-cc-pVTZ level (Table 1). Moreover, these results are in good agreement with the recently published DFT and MP2 results.22 For example, at the MP2/ 6-31++G(d,p) level the energies of the conformers GskC, AaT, and AsC (here denoted as II, III, and IV), in relation to the energy of the ScC conformer (I in our notation), equal 6.5, 11.4, and 18.5 kJ mol−1, respectively.22
Conformer I (Fig. 1) is stabilized by the intramolecular H-bond between the H12 proton of the alcohol O6–H12 group and the O4 oxygen atom of the carbonyl C3O4 group. It is defined by the following dihedral angles: O4
C3–O5–H11, C3–C2–O6–H12, and C2–C3–O5–H11, which are equal to ca. 0.0, 0.2 and 179.5°, respectively. This conformer, with an almost planar structure of the five-membered intramolecular H-bond ring and a relatively short intramolecular H-bond H12⋯O4 distance (2.093 Å), is the global PES minimum. The second most stable conformer, II, (Fig. 1) is fixed by an O6–H12(alcohol)⋯O5–H11(carboxyl) intramolecular H-bond, where the O5 oxygen atom of the carboxylic OH group acts as a proton acceptor. For conformer II, the above-mentioned dihedrals equal −0.2, 46.0, and −177.3°, respectively. Thus, these two conformers differ in their rotation around the C2–C3 axis. In the third conformer, III, (Fig. 1), an intramolecular H-bond is formed between O5–H11(carboxylic)⋯O6–H12(alcoholic) groups, where (as opposed to conformation I) the carboxylic C
O and alcoholic O–H groups are in the energetically less favorable trans configuration and the alcoholic oxygen atom acts as a proton acceptor. In this case, the dihedrals equal 179.0, −164.4, and −1.7°, respectively. The fourth conformer considered here, IV, is quite high in energy (Fig. 1, Table 1), and is not stabilized by an intramolecular H-bond. As may be seen in ref. 22, the remaining conformers of lactic acid are even higher in energy and are not studied in this paper. As one may expect from energy differences between the first three conformers, all of them were identified in a most impressive experimental study of the lactic acid Ar and Xe low temperature matrix IR spectra.22
Each of the three conformers has an intramolecular H-bond. In two of them, I and II, the alcoholic proton is involved in the intramolecular H-bond as a proton donor, while in the conformer III, the oxygen atom of the alcoholic group acts as a proton acceptor. That is, the O–H bond in the conformer III is much shorter (0.964 Å) than its counterpart in conformers I and II (0.971 and 0.966 Å, respectively). Unlike in the above-described cases, the carboxylic OH group is a proton donor only in the conformer III, and therefore the group exhibits the greatest O5–H11 distance of 0.975 Å, whereas in conformers I and II the analogous distances equal 0.971 Å. Enlargement of the basis set from aug-cc-pVDZ to aug-cc-pVTZ did not change the results significantly.
The vibrational frequencies are discussed solely for conformers I and II. All 30 IR active fundamental vibrations are considered and interpreted. The frequencies of the most important modes: stretching ν(OHC), ν(OHA), ν(CO), ν(C–H), ν(C–O), bending and out-of plane were found to be close to the other calculated values, and to the data found in experimental literature.22 This may serve as evidence that the calculated frequencies of monomeric lactic acid described in this paper constitute reliable reference data for calculations of frequency shifts in H-bonding complexes with water molecules.
The coordinate definitions for the analysis of the vibrational spectra of the lactic acid are presented in Table 2. Tables 3 and 4 contain the interpretation of the calculated IR frequencies, VA intensities, and VCD spectra for conformers I and II.
Number | Mode description | Name | Monomer |
---|---|---|---|
Modes: ν-stretching, β-bending, τ-torsion, δ-out-of-plane bending, σ-hb stretching, λ-libration Indices: W-water, A-alcoholic, C-carboxylic, hb-H-bonded, free-not H-bonded, s-symmetric, a-asymmetric, S-skeleton, butt-butterfly, MI, MII-monomers I, II. | |||
1 | R(O13–H15) | ν(OHW,free) | |
2 | R(O6–H12) | ν(OHA) | |
3 | R(O13–H14) | ν(OHW,hb) | |
4 | R(O5–H11) | ν(OHC) | |
5 | R(C1–H7) − R(C1–H9) | ν a(CH3)1 | |
6 | R(C1–H8) − R(C1–H9) | ν a(CH3)2 | |
7 | R(C1–H7) + R(C1–H8) + R(C1–H9) | ν s(CH3) | |
8 | R(C2–H10) | ν(C*H) | |
9 | R(C3–O4) |
ν(C![]() |
|
10 | A(H14–O13–H15) | β(H2O) | |
11 | A(H7–C1–H9) − A(H8–C1–H9) | β a(CH3) | |
12 | A(H7–C1–H8) − A(H7–C1–H9) − A(C3–O5–H11) | β(CH + OH)1 | |
12MI, MII | A(H7–C1–H8) − A(H8–C1–H9) | β(CH)1 | |
13 | A(H7–C1–H8) + A(H10–C2–C3) + A(C3–O5–H11) − A(C2–O6–H12) | β(CH + OH)2 | |
13MI, MII | A(H10–C2–C6) + A(C3–O5–H11) + A(C2–O6–H12) | β(CH + OH)2 | |
14 | A(H7–C1–H8) − A(H10–C2–C3) + A(C3–O5–H11) + A(C2–O6–H12) | β(CH + OH)3 | |
14MI | A(H7–C1–H8) − A(H10–C2–C3) − A(C3–O5–H11) + A(C2–O6–H12) | β(CH + OH)3 | |
14MII | A(H7–C1–H9) + A(H10–C2–C3) − A(C2–O6–H12) | β(CH + OH)3 | |
15 | A(H7–C1–H8) + A(H7–C1–H9) + A(H8–C1–H9) | β s(CH3) | |
16 | D(H10–C1–C3–C2) | τ(C*H) | |
17 | A(H10–C2–C3) + A(H12–O6–C2) | β(C*H + OHA) | |
18; 18MII | R(C3–O5) | ν(C–O) | |
18MI | R(C3–O5) + R(C2–O6) | ν s(C–O) | |
19; 19MII | R(C1–C2) − R(C2–O6) | ν(C*O) | |
19MI | R(C3–O5) − R(C2–O6) | ν a(C–O) | |
20 | D(H8–C2–C3–C1) − D(H9–C2–O6–C1) | τ(CH3) | |
21 | A(C2–C1–H9) − A(C3–C2–H10) | β(C*H) | |
22 | R(C1–C2) − R(C2–C3) − R(C3–O5) | ν(S)1 | |
23 | D(H11–O5–C3–C2) + D(H14–O13–H11–O5) | δ(OHC) | |
23MI, MII | D(H11–O5–C3–C2) | δ(OHC) | |
24 | R(C1–C2) + R(C2–C3) + R(C3–O5) | ν(S)2 | |
25 | D(C2–O4–O5–C3) |
τ(OC![]() |
|
26 | A(H11–O13–H14) + A(H14–O4–C3) | β(OHW⋯O) | |
27 | A(O4–C3–O5) |
β(OC![]() |
|
28 | Ia A(C2–C3–O5) | ||
IIa A(C2–C3–O5) − A(C3–C2–O6) | β(CC–OC) | ||
29 | A(C1–C2–O6) | β(CC–OA) | |
30 | D(H11–O13–H14–O4) + D(H12–O6–C2–C3) | λ(H2O + OHA) | |
30MI | D(H12–O6–C2–C3) | δ(OHA) | |
31 | Ia A(C2–C3–O4) + A(C3–C2–O6) | ||
IIa A(C1–C2–O3) + A(C2–C3–O5) |
β(O![]() |
||
32 | Ia D(H11–O13–H14–O4) − D(H12–O6–C2–C3) | λ(H2O–OHA) | |
IIa D(H12–O6–C2–C3) | δ(OHA) | ||
33 | A(C1–C2–C3) | β(CCC) | |
34 | D(H15–O13–H14–O4) | δ(OHW,free) | |
35 | D(H7–C1–C2–C3) + D(H8–C1–C2–C3) + D(H9–C1–C2–C3) | λ(CH3) | |
36 | R(H11–O13) | σ(OHC⋯O) | |
37 | R(O4–H14) − R(H11–O13) | σ(OHW⋯O) | |
38 | Ia D(H14–O4–C3–C2) + D(C1–C2–C3–O4) | ||
IIa D(H14–O4–C3–C2) | τ(butt)1 | ||
39 | Ia D(H14–O4–C3–C2) − D(C1–C2–C3–O4) | ||
IIa D(C1–C2–C3–O4) | τ(butt)2 | ||
39MI, MII | D(C1–C2–C3–O4) | τ(butt) |
Complex Ia | Monomer I | ||||||||||
---|---|---|---|---|---|---|---|---|---|---|---|
ν
harm
cm−1 |
ν
scal
a
cm−1 |
ν
anharm
cm−1 |
I
IR
km mol−1 |
R
i
10−44 esu2 cm2 |
PED
% |
ν
harm
cm−1 |
ν
scal
a
cm−1 |
ν
anharm
cm−1 |
I
IR
km mol−1 |
R
i
10−44 esu2 cm2 |
PED
% |
a ν scal = 0.955 νcalc + 29.6 | |||||||||||
3863 | 3719 | 3679 | 101.5 | −28.90 | 95 ν(OHW,free) | ||||||
3719 | 3581 | 3526 | 83.6 | 13.56 | 100 ν(OHA) | 3719 | 3581 | 3533 | 86.0 | 16.3 | 100 ν(OHA) |
3635 | 3501 | 3444 | 295.1 | 10.17 | 92 ν(OHW,hb) | ||||||
3312 | 3192 | 3108 | 805.3 | 1.89 | 95 ν(OHC) | 3738 | 3599 | 3540 | 73.0 | −2.2 | 100 ν(OHC) |
3137 | 3025 | 2980 | 14.3 | −1.90 | 94 νa(CH3)1 | 3136 | 3024 | 2982 | 12.6 | −3.7 | 96 νa(CH3)1 |
3126 | 3015 | 2967 | 18.9 | 4.11 | 92 νa(CH3)2 | 3127 | 3016 | 2970 | 18.4 | 3.8 | 95 νa(CH3)2 |
3046 | 2938 | 2863 | 14.7 | −0.98 | 99 νs(CH3) | 3047 | 2939 | 2923 | 13.3 | −0.9 | 99 νs(CH3) |
2996 | 2891 | 2846 | 25.2 | −23.22 | 99 ν(C*H) | 3002 | 2896 | 2844 | 24.3 | −22.3 | 99 ν(C*H) |
1744 | 1695 | 1716 | 260.9 | 49.19 | 74 ν(C![]() |
1790 | 1739 | 1758 | 300.4 | 0.8 | 82 ν(C![]() |
1607 | 1564 | 1558 | 131.9 | −134.74 | 89 β(H2O) | ||||||
1489 | 1452 | 1440 | 13.1 | 13.22 | 65 β(CH + OH)2 + 17 ν(C–O) | 1330 | 1300 | 1312 | 74.2 | 44.2 | 57 β(CH + OH)2 |
1470 | 1433 | 1441 | 7.44 | 6.61 | 77 βa(CH3) – 13 τ(CH3) | 1469 | 1432 | 1432 | 8.3 | 6.3 | 83 βa(CH3) −10 τ(CH3) |
1454 | 1418 | 1414 | 1.35 | −16.21 | 74 β(CH + OH)1 | 1476 | 1439 | 1437 | 2.7 | 2.4 | 71 β(CH)1 |
1397 | 1364 | 1344 | 24.4 | −13.36 | 70 β(CH + OH)3 | 1424 | 1389 | 1371 | 5.1 | −25.6 | 56 β(CH + OH)3 |
1384 | 1351 | 1353 | 5.7 | 22.44 | 83 βs(CH3) | 1386 | 1353 | 1515 | 9.4 | 11.3 | 87 βs(CH3) |
1336 | 1305 | 1296 | 6.2 | −19.96 | 71 τ(C*H) | 1335 | 1304 | 1301 | 15.6 | −37.3 | 60 τ(C*H) |
1287 | 1259 | 1246 | 259.2 | 133.21 | 30 β(C*H + OHA) + 31 ν(C–O) | 1187 | 1163 | 1148 | 41.2 | 17.7 | 36 νs(C–O) −10 β(CH + OH)2 |
1249 | 1222 | 1221 | 17.1 | −35.57 | 54 β(C*H + OHA)–16 ν(C–O) | 1261 | 1234 | 1235 | 42.2 | 70.7 | 75 β(C*H + OHA) |
1155 | 1133 | 1130 | 128.5 | 13.70 | 42 ν(C*O) − 14 τ(CH3) + 11 β(C*H) | 1139 | 1117 | 1106 | 262.9 | 18.3 | 50 νa(C–O) − 11 β(CH + OH)3 − 12 τ(CH3) |
1104 | 1084 | 1073 | 26.6 | −29.02 | 16 ν(C*O) + 34 τ(CH3) − 13 ν(S)1 | 1103 | 1083 | 1073 | 39.4 | −41.6 | 35 τ(CH3) −20 ν(S)1 − 16 ν(S)2 |
1044 | 1027 | 1023 | 36.1 | 35.66 | 21 ν(C*O) − 45 β(C*H) | 1042 | 1025 | 1020 | 38.5 | 34.3 | 49 β(C*H) + 20 ν(S)1 |
935 | 922 | 910 | 2.3 | 4.29 | 17 τ(CH3) − 13 β(C*H) + 47 ν(S)1 | 931 | 919 | 908 | 1.3 | 10.5 | 28 ν(S)1 − 10 νs(C–O) |
917 | 905 | 839 | 106.8 | −7.19 | 85 δ(OHC) | 579 | 582 | 551 | 54.3 | 83.1 | 69 δ(OHC) |
823 | 815 | 806 | 20.1 | −17.16 | 45 ν(S)2 | 805 | 798 | 792 | 23.2 | −12.4 | 33 ν(S)2 + 13 τ(OC![]() |
744 | 740 | 732 | 6.3 | −3.43 | 36 τ(OC![]() ![]() |
735 | 731 | 721 | 39.4 | −12.5 | 38 τ(OC![]() ![]() |
628 | 629 | 612 | 25.1 | 17.67 | 12 ν(C*O) + 14 τ(OC![]() ![]() |
632 | 633 | 615 | 42.6 | −74.9 | 43 β(OC![]() |
614 | 616 | 491 | 179.8 | −16.06 | 74 β(OHW⋯O) | ||||||
515 | 521 | 504 | 10.0 | 0.09 | 12 ν(S)2 + 44 β(CC–OC) + 13 β(CC–OA) | 494 | 501 | 486 | 13.6 | −4.7 | 40 β(CC–OC) + 18 β(CC–OA) |
422 | 433 | 410 | 8.5 | 6.50 | 60 β(CC–OA) | 408 | 419 | 393 | 15.6 | 1.7 | 51 β(CC–OA) + 10δ(OHA) |
376 | 389 | 337 | 81.7 | −184.31 | 49 λ(H2O + OHA) − 10 β(O![]() |
||||||
360 | 373 | 330 | 32.8 | 114.25 | 62 λ(H2O–OHA) | 353 | 367 | 268 | 48.9 | −20.6 | 63 δ(OHA) − 11 β(CCC) |
309 | 325 | 297 | 61.8 | 55.91 | 14 λ(H2O + OHA) + 40 β(O![]() |
298 | 314 | 254 | 28.4 | 13.2 | 41 β(O![]() |
276 | 293 | 261 | 69.9 | 18.00 | 36 β(CCC) − 15 σ(OHC⋯O) | 243 | 261 | 237 | 0.7 | 8.6 | 43 β(CCC) + 10 τ(OC![]() |
245 | 264 | 288 | 73.7 | 38.02 | 64 δ(OHW,free) | ||||||
215 | 235 | 259 | 3.6 | 31.83 | 88 λ(CH3) | 218 | 238 | 202 | 0.5 | −11.4 | 82 λ(CH3) |
185 | 206 | 168 | 5.0 | 2.31 | 68 σ(OHC⋯O) | ||||||
133 | 157 | 106 | 17.7 | 12.01 | 75 σ(OHW⋯O) | ||||||
67 | 93 | 79 | 1.7 | −5.47 | 74 τ(butt)1 − 10 λ(H2O + OHA) | ||||||
50 | 77 | 65 | 6.3 | −9.68 | 74 τ(butt)2 | 51 | 78 | 39 | 2.6 | 0.1 | 84 τ(butt) |
Complex 2a | Monomer 2 | ||||||||||
---|---|---|---|---|---|---|---|---|---|---|---|
ν
harm
cm−1 |
ν
scal
a
cm−1 |
ν
anharm
cm−1 |
I
IR
km mol−1 |
R
i
10−44 esu2 cm2 |
PED
% |
ν
harm
cm−1 |
ν
scal
a
cm−1 |
ν
anharm
cm−1 |
I
IR
km mol−1 |
R
i
10−44 esu2 cm2 |
PED
% |
a ν scal = 0.955 νcalc + 29.6 | |||||||||||
3862 | 3718 | 3675 | 97.8 | −29.66 | 96 ν(OHW,free) | ||||||
3803 | 3661 | 3617 | 52.0 | 11.89 | 100 ν(OHA) | 3813 | 3671 | 3618 | 51.1 | 10.3 | 100 ν(OHA) |
3603 | 3470 | 3403 | 364.5 | −25.05 | 92 ν(OHW,hb) | ||||||
3293 | 3174 | 3080 | 810.9 | 60.20 | 95 ν(OHC) | 3744 | 3605 | 3547 | 73.8 | 0.6 | 100 ν(OHC) |
3148 | 3036 | 2992 | 10.3 | −2.28 | 93 νa(CH3)1 | 3149 | 3037 | 2991 | 10.2 | −2.9 | 93 νa(CH3)1 |
3132 | 3021 | 2970 | 17.6 | 3.02 | 93 νa(CH3)2 | 3133 | 3022 | 2975 | 16.2 | 1.2 | 93 νa(CH3)2 |
3055 | 2947 | 2928 | 13.9 | 0.87 | 98 νs(CH3) | 3057 | 2949 | 2931 | 10.2 | 0.9 | 99 νs(CH3) |
2955 | 2852 | 2816 | 34.3 | −9.01 | 99 ν(C*H) | 2961 | 2857 | 2821 | 34.0 | −8.0 | 99 ν(C*H) |
1764 | 1714 | 1734 | 226.9 | 53.02 | 75 ν(C![]() |
1816 | 1764 | 1791 | 269.7 | 1.9 | 83 ν(C![]() |
1610 | 1567 | 1560 | 115.3 | −109.75 | 89 β(H2O) | ||||||
1474 | 1437 | 1476 | 2.2 | 2.30 | 71 βa(CH3) − 11 τ(CH3) | 1462 | 1426 | 1432 | 9.6 | 14.0 | 67 βa(CH3) + 18 β(CH)1 − 10 β(CH + OH)3 |
1464 | 1428 | 1421 | 13.5 | 10.07 | 79 β(CH + OH)1 + 10 βa(CH3) | 1473 | 1436 | 1437 | 0.7 | −1.02 | 65 β(CH)1 − 10βa(CH3) + 12τ(CH3) |
1446 | 1411 | 1391 | 14.7 | 21.49 | 59 β(CH + OH)2 + 12 ν(C–O) | 1314 | 1284 | 1262 | 19.5 | 3.2 | 56 β(CH + OH)2 − 16 β(C*H + OHA) |
1409 | 1375 | 1375 | 29.5 | −26.49 | 44 β(CH + OH)3 − 22 βs(CH3) | 1411 | 1377 | 1364 | 17.1 | −21.7 | 47 β(CH + OH)3 + 18βs(CH3) |
1382 | 1349 | 1346 | 48.2 | −38.79 | 69 βs(CH3) + 18 β(CH + OH)3 | 1382 | 1349 | 1348 | 41.5 | −37.8 | 74 βs(CH3) |
1329 | 1300 | 1297 | 27.0 | −16.19 | 58 τ(C*H) + 11 β(C*H + OHA) | 1352 | 1321 | 1325 | 17.6 | −25.0 | 44 τ(C*H) + 10 βs(CH3) |
1274 | 1246 | 1242 | 53.4 | 108.73 | 63 β(C*H + OHA) − 14 τ(C*H) + 13 τ(CH3) | 1274 | 1246 | 1241 | 64.6 | 94.3 | 50 β(C*H + OHA) + 14 τ(CH3) − 23β(CH)1 |
1261 | 1234 | 1224 | 231.0 | −69.81 | 53 ν(C–O) + 12 β(CH + OH)3 − 11 β(CH + OH)2 | 1157 | 1135 | 1117 | 229.7 | −75.2 | 43 ν(C–O) − 17β(CH + OH)2 − 11β(CH + OH)3 |
1156 | 1134 | 1124 | 60.8 | −25.63 | 49 ν(C*O) − 16 β(C*H) | 1149 | 1127 | 1118 | 57.9 | −35.9 | 49 ν(C*O) + 17 β(C*H) |
1092 | 1072 | 1067 | 19.2 | −16.22 | 41 τ(CH3) − 12 β(C*H + OHA) − 13 ν(S)1 | 1084 | 1065 | 1057 | 44.6 | −7.2 | 40 τ(CH3) − 11 ν(S)1 − 12β(CH + OH)3 |
1042 | 1025 | 1021 | 49.0 | 55.24 | 44 β(C*H) + 21 ν(C*O) + 16 ν(S)1 | 1042 | 1025 | 1018 | 54.6 | 72.4 | 42 β(C*H) + 16 ν(S)1 − 20 τ(C*H) |
931 | 919 | 895 | 40.9 | −0.81 | 31 ν(S)1 + 25 δ(OHC) + 13 τ(CH3) | 921 | 909 | 902 | 7.7 | −9.5 | 43 ν(S)1 + 16β(C*H + OHA) |
926 | 914 | 858 | 70.7 | −9.73 | 60 δ(OHC) − 15 ν(S)1 | 614 | 616 | 588 | 91.5 | 59.6 | 75 δ(OHC) − 13β(OC![]() |
824 | 816 | 810 | 9.8 | −14.07 | 37 ν(S)2
− 13 τ(OC![]() ![]() |
810 | 803 | 795 | 7.5 | 27.2 | 26 τ(OC![]() |
743 | 739 | 733 | 11.1 | 26.07 | 33 τ(OC![]() ![]() |
726 | 723 | 717 | 31.9 | −56.1 | 26 τ(OC![]() ![]() |
647 | 647 | 526 | 197.0 | −14.94 | 73 β(OHW⋯O) | ||||||
573 | 577 | 568 | 22.7 | −20.36 | 36 β(OC![]() ![]() |
535 | 541 | 525 | 32.5 | −21.6 | 33 β(OC![]() ![]() |
517 | 523 | 514 | 16.6 | −15.78 | 45 β(CC–OC) + 11 ν(C*O) + 13 β(O![]() |
503 | 510 | 499 | 13.0 | 4.2 | 39 β(O![]() |
429 | 439 | 423 | 6.3 | −13.28 | 73 β(CC–OA) | 412 | 423 | 407 | 8.9 | −15.6 | 69 β(CC–OA) |
384 | 396 | 309 | 68.0 | −109.13 | 61 λ(H2O + OHA) + 14 δ(OHW,free) | ||||||
357 | 371 | 337 | 31.4 | 94.34 | 11 τ(OC![]() ![]() |
338 | 352 | 312 | 50.0 | −6.7 | 38δ(OHA) + 16τ(OC![]() ![]() |
304 | 320 | 252 | 94.2 | 75.60 | 79 λ(H2O–OHA) | ||||||
288 | 305 | 249 | 95.1 | 16.33 | 39 β(CCC) + 12 β(O![]() |
245 | 264 | 235 | 4.7 | −22.4 | 40 β(CC–OC) − 37 β(CCC) |
262 | 280 | 163 | 65.7 | 40.62 | 12 λ(H2O + OHA) − 11 β(CCC) + 61 δ(OHW,free) | 313 | 329 | 276 | 71.3 | 54.2 | 49 δ(OHA) + 22 β(CCC) |
213 | 233 | 138 | 0.5 | −16.68 | 89 λ(CH3) | 205 | 225 | 157 | 3.2 | −11.5 | 94 λ(CH3) |
185 | 206 | 160 | 3.8 | −2.92 | 74 σ(OHC⋯O) | ||||||
136 | 159 | 114 | 19.6 | 22.73 | 69 σ(OHW⋯O) − 15 β(CCC) | ||||||
68 | 95 | 60 | 1.9 | 0.96 | 86 τ(butt)1 | ||||||
44 | 72 | 29 | 2.2 | 0.89 | 93 τ(butt)2 | 42 | 67 | 36 | 2.8 | −4.4 | 92 τ(butt) |
The high frequency stretching modes (Tables 3 and 4) constitute the most localized vibrations. As far as the other vibrations are concerned, except in a few cases, they are the combinations of different bending and stretching modes (Tables 3 and 4). There are several differences between the IR spectra of monomers I and II. Firstly, the ν(OHA) band in conformer I is predictably located at wavenumbers ca. 90 cm−1 lower than that of conformer II. This is due to the higher strength of the OHA⋯OC intramolecular H-bond in I than the OHA⋯OHC intramolecular H-bond in conformer II. The band intensity is much higher in conformer I than in II as well. Secondly, the ν(C*H) band position differs in the two spectra by ca. 40 cm−1, at 2986 cm−1 for Ivs. 2857 cm−1 for II. Such a difference in this region of the spectra is quite meaningful. Thirdly, the location of the ν(C
O) band is also significant: in I, where the C
O group is H-bonded, it is shifted 25 cm−1 towards lower wavenumbers than in the conformer II, where such a H-bond is absent.
The main differences between the VCD spectra of the two conformers are illustrated in Table 5 and Fig. S1 of the ESI.† The VCD effect is greater for the medium frequency bending vibrations than for the high frequency stretching ones. The bending β(C*H + OHA) mode of the H10–C2–C3 moiety, including the chirality center at the C*2 atom, displays a relatively high optical activity Ri (70.7 and 94.3 × 10−44 esu2 cm2, for conformer I and II, respectively). A similar characteristic is observed for the second C*H bending mode, denoted as β(C*H), which is not as localized as the previous one (Tables 3 and 4). For this mode, Ri is positive and equals 34.3 and 72.4 × 10−44 esu2 cm2 for conformers I and II, respectively. The medium frequency region also contains the νa(C–O) stretching vibrations mode of the C2–O6 moiety, at 1117 (conformer I) and 1135 cm−1 (conformer II), which is one of the most intense bands in the IR spectra and has opposite signs in the VCD spectra, 18.3 and −75.2 × 10−44 esu2 cm2 for conformers I and II, respectively. It is remarkable that in the former case the O6-H group is engaged in the H-bond with the carbonyl group, namely O6-H12⋯O4C3, and the ν(C2–O6) mode couples to ν(C3–O5) one, whereas in the latter the O5–H11 moiety acts as H-bond acceptor and the ν(C2–O6) local mode is not coupled (Tables 3 and 4). For the two conformers, the out-of-plane bending δ(OHC) mode absorbs near 600 cm−1. The band has a relatively high rotational strength Ri: 83.1 and 59.6 × 10−44 esu2 cm2 for conformers I and II, respectively. In the low frequency region, a bending vibration with an important β(CCC) mode contribution has a positive Ri value for conformer I (8.6 × 10−44 esu2 cm2), while this is negative and quite significant for conformer II (−22.4 × 10−44 esu2 cm2).
ν harm/cm−1 | ν scal/cm−1 | ν anharm/cm−1 | I IR/km mol−1 | R i/× 10−44esu2cm2 | PED(%) |
---|---|---|---|---|---|
1261 | 1234 | 1235 | 42.2 | 70.7 | 75 β(C*H + OHA) |
1274 | 1246 | 1241 | 64.6 | 94.3 | 50 β(C*H + OHA) + 14 τ(CH3) − 23β(CH)1 |
1139 | 1117 | 1106 | 262.9 | 18.3 | 50 νa(C–O) − 11 β(CH + OH)3 − 12 τ(CH3) |
1157 | 1135 | 1117 | 229.7 | −75.2 | 43 ν(C–O) − 17β(CH + OH)2 − 11β(CH + OH)3 |
1042 | 1025 | 1020 | 38.5 | 34.3 | 49 β(C*H) + 20 ν(S)1 |
1042 | 1025 | 1018 | 54.6 | 72.4 | 42 β(C*H) + 16 ν(S)1 − 20 τ(C*H) |
579 | 582 | 551 | 54.3 | 83.1 | 69 δ(OHC) |
614 | 616 | 588 | 91.5 | 59.6 | 75 δ(OHC) − 13β(OC![]() |
242 | 261 | 237 | 0.7 | 8.6 | 43 β(CCC) + 10 τ(OC![]() |
245 | 264 | 235 | 4.7 | −22.4 | 40 β(CC–OC) − 37 β(CCC) |
218 | 238 | 202 | 0.5 | −11.4 | 82 λ(CH3) |
205 | 225 | 157 | 3.2 | −11.5 | 94 λ(CH3) |
Summarizing, the VCD spectra change quite remarkably when the D-lactic acid molecule assumes different conformations.
![]() | ||
Fig. 2 Considered forms of D-lactic acid complexes with a water molecule. |
Complex | ΔEint/kJ mol−1 | ΔEint+BSSE/kJ mol−1 | D e/kJ mol−1 | ΔEdef/kJ mol−1 | D 0/kJ mol−1 | ΔG298/kJ mol−1 |
---|---|---|---|---|---|---|
ΔE = int uncorrected interaction energy, ΔE = int+BSSE interaction energy counterpoise corrected, ΔE = def deformation energy. | ||||||
aug-cc pVDZ | ||||||
Ia | −40.88 | −41.84 | −38.91 | 2.93 | −31.34 | 0 |
Ib | −21.38 | −26.28 | −19.41 | 6.96 | −13.60 | 12.35 |
Ic | −21.25 | −19.75 | −19.08 | 0.67 | −13.72 | 11.44 |
Id | −10.29 | −8.87 | −8.20 | 0.67 | −5.10 | 17.01 |
IIa | −41.92 | −43.56 | −40.00 | 3.56 | −32.22 | 7.65 |
IIb | −21.00 | −22.55 | −18.70 | 3.81 | −11.05 | 27.89 |
IIc | −19.83 | −18.58 | −18.03 | 0.54 | −12.72 | 18.92 |
IId | −19.04 | −8.87 | −16.99 | −8.12 | −12.09 | 18.89 |
aug-cc-pVTZ | ||||||
Ia | −40.63 | −41.92 | −41.17 | 1.76 | −30.42 | 0 |
IIa | −41.17 | −44.27 | −39.71 | 4.56 | −29.92 | 7.44 |
Any of the lactic acid conformers could possibly form four complexes with a water molecule. Indeed, water could act as an electron donor to both carboxylic and alcoholic OH groups in structures Ia and IIa, and Ib and IIb, respectively. Moreover, the water molecule could act as a proton donor to the oxygen atom of both alcoholic oxygen and carboxylic oxygen OH atoms as in structures Ic and Id, respectively. Formally, a linear water OH⋯OC is also possible, however, this supposed form converges to a cyclic H-bond as, for example, in Ia.
Structures Ia and IIa are the most stable, yet Ia remains the most populated (Table 6). Not including the intramolecular H-bond, the two complex forms exhibit two intermolecular H-bonds between the carboxylic O5–H11 group and O13 atom of water, and between the carbonylic O4 atom and water H14–O13 moiety. In this way a 6-membered H-bonded cycle is formed. Note that an analogous water complex with conformer III cannot exist because the carboxylic OH group is in a trans position and a cyclic H-bond with the COOH group cannot form.
The calculated energies at the B3LYP level, carried out with the two basis sets aug-cc-pVDZ and aug-cc-pVTZ for Ia and IIa, are given in Table 6. The binding energies of the two most stable complexes do not differ substantially, and equal −38.91 and −40.00 kJ mol−1 for Ia and IIa, respectively. Moreover, an extension of the basis set to aug-cc-pVTZ changes the binding energies to a fairly insignificant extent, i.e. to −41.17 and −39.71 kJ mol−1 (Table 6), respectively.
The intermolecular H-bond between the alcoholic OHA group, acting as a proton donor, and the water oxygen atom, acting as an acceptor, together with the water OH group donating the proton to either the carboxylic O4 (Ib) or O5 (IIb) atom, give rise to formation of a cyclic 7-membered H-bond system stabilizing the structures Ib and IIb with energies much lower than those of Ia and IIa (−19.41 and −18.70 kJ mol−1 for Ib and IIb, respectively). This is probably due to weaker H-bonds between the alcoholic hydroxyl group with water in Ib and IIb compared to those of the carboxyl OH group in Ia and IIa, and a weaker stabilization of the 7-membered H-bond ring in Ib and IIb than the 6-membered H-bonded ring in Ia and IIa.
Complexes of type c, Ic and IIc, in which a water molecule is the proton donor to the oxygen atom of alcoholic group, also form structures less stable than those of type a. They display binding energies of −19.08 and −18.03 kJ mol−1 for Ic and IIc, respectively. In the two type d structures, the water molecule plays the role of proton donor, but in Id it is donated to the oxygen atom of the carboxylic O–H group, while in IId it is donated to the oxygen atom of the carboxylic CO group. The binding energies of Id and IId are the lowest in the series studied, i.e., −8.20 and −16.99 kJ mol−1, for Id and IId, respectively, while the ΔG298 values are much closer (17.01 and 18.89 kJ mol−1, respectively), but this indicates that the d-type complexes would hardly be observable in the mixture of complexes (Table 6).
To investigate the role of basis set effects on binding energy between lactic acid and a single water molecule in more detail, let us turn to Table 6, in which the results of the binding energy obtained with the two basis sets are listed. The numbers shown are corrected for the BSSE effect. As expected, basis set sensitivity exists, but is small, and as the basis set is enlarged, the role of the BSSE decreases. Also, the binding energy values for the complexes calculated at the B3LYP/aug-cc-pVDZ and B3LYP/aug-cc-pVTZ levels are convincing proof that the former level is quite suitable for the purpose of studying H-bonding.
To analyze the influence of intermolecular interactions on the lactic acid molecule, let us comment in more detail on the changes to selected geometrical parameters (Table 7). The alcoholic O6–H12 distance is almost insensitive to the side on which water molecule approaches the acid. For three complexes, Ia, Ic and Id, this is connected with the intramolecular H-bond O6–H12⋯O4, which stabilizes the geometry of this moiety, whereas for Ib the distance is slightly greater, which means that the intermolecular interaction with water exerts a slightly stronger influence than the intramolecular H-bond does. Similarly, the carboxylic O5–H11 distance is constant, except in structure Ia where it is elongated by the intermolecular H-bond (Fig. 2). The changes in the C3O4 distance are in line with the changes in binding energy. The intramolecular H12⋯O4 distance reveals that in the case of Ib, water constrains any significant change in the intramolecular H-bond O6–H12⋯O4 geometry. In the complexes of conformer II, it is interesting to note that the O6–H12 distance is far shorter than for conformer I. This is due to a change in the type of intramolecular H-bond from O6–H12⋯O4
C3 to O6–H12⋯O5(H). As shown below, all of the above-mentioned changes are reflected even more in the vibrational spectra.
Complex | r(O6–H12)/Å | r(O5–H11)/Å |
r(C3![]() |
r(H12⋯O4)/Å | r(H11,12⋯O13)/Å | r(On⋯H14)/Å |
---|---|---|---|---|---|---|
a H11⋯O13. b H12⋯O13. c H14⋯O4. d H14⋯O6. e H14⋯O5. | ||||||
Ia | 0.971 | 0.993 | 1.227 | 2.070 | 1.757a | 2.008c |
Ib | 0.974 | 0.971 | 1.216 | 2.496 | 1.924b | 1.947c |
Ic | 0.972 | 0.972 | 1.214 | 2.074 | 1.931d | |
Id | 0.970 | 0.972 | 1.211 | 2.126 | 2.103e | |
r(H12⋯O5)/Å | ||||||
IIa | 0.966 | 0.994 | 1.222 | 2.170 | 1.748a | 1.971c |
IIb | 0.977 | 0.971 | 1.207 | 2.670 | 1.908b | 1.987e |
IIc | 0.967 | 0.971 | 1.208 | 2.188 | 1.937d | |
IId | 0.965 | 0.972 | 1.215 | 2.274 | 1.942c |
As expected,31 the most conspicuous change in the IR spectra of I after formation of complex Ia is a shift of the ν(OHC) band towards lower frequency by ca. 400 cm−1 and a simultaneous, ca. tenfold, increase of the band intensity. The ν(CO) band undergoes a relatively small, yet significant, shift towards lower frequency, ca. 40 cm−1, and a small intensity decrease of no essential significance. There are three bands in the bending vibrations region that are shifted considerably: firstly, the β(CH + OH)2 band is shifted from 1300 to ca. 1450 cm−1; secondly, the νS(C–O) band which changes its form from symmetric ν(C2–O6) + ν(C3–O5) to uncoupled ν(C3–O5), and shifts from 1165 to 1260 cm−1; and finally, the δ(OHC), which shifts as much as 325 cm−1 towards higher frequency from 580 to 905 cm−1.
In our previous paper14a we have shown that after complex formation with a chiral molecule, an achiral molecule becomes active in the VCD spectra. This is also the case for the system studied in this paper. An achiral water molecule exhibits the VCD absorptions of the ν(OHW,free), ν(OHW,hb), and β(H2O) modes at 3719, 3501, and 1564 cm−1, respectively. Moreover, although the rotational strengths of the two stretching modes of water are of average value, −29 and 10 × 10−44 esu2 cm2, the strength of the bending vibration mode is one of the most intense in the VCD spectrum of the Ia complex, i.e., it equals −135 × 10−44 esu2 cm2 (Table 3). Furthermore, the intramolecular H-bond bending vibration band β(OHW⋯O), water librations λ(H2O + OHA), and out-of-plane bending δ(OHW,free) vibration of water come into view and absorb at 616, 389, and 264 cm−1 with rotational strengths in the VCD spectrum of ca. −16, −184 and 38 × 10−44 esu2 cm2, respectively. It is worth noting that the λ(H2O + OHA) librations are of greater magnitude, yet they are placed in the beginning of the far-IR region, and therefore have lesser value as an indicator of the complex. Thus, the new VCD bands appear in quite specific regions, they are fairly appreciable, and none of the bands of water that are not H-bonded to the chiral molecule manifest themselves in the VCD spectrum.
After complexation the Ri of the ν(CO) band increases from 1.0 to 50 × 10−44 esu2 cm2, but the ν(OHC) VCD band changes negligibly (Table 3). The out-of-plane bending δ(OHC) vibration VCD band changes remarkably in its magnitude as well as in sign alteration, from 83 to −7 × 10−44 esu2 cm2, for I and Ia, respectively. Surprisingly, this is also the case for the out-of-plane bending δ(OHA) vibration of the OHA engaged in the intramolecular H-bond.
The Ia complex formation induces several changes in the bending vibration region between 1470 and 1170 cm−1, however the modes are difficult to interpret since the PED contributions of the particular local modes are small and we will not go into the interpretational details.
The differences between the IR spectra of II and IIa are fairly analogous to those for conformer I (Table 4). In comparison with II, the new bands in the VCD spectrum of IIa originating from water exhibit important VCD intensities: −30, −25, −110, −15, and 76 × 10−44 esu2 cm2, for the ν(OHW,free), ν(OHW,hb), β(H2O), β(OHW⋯O), and λ(H2O–OHA) bands, respectively. Also, as before, the ν(CO) of negligible intensity in the spectrum of conformer II becomes quite considerable in the VCD spectrum of IIa. In contrast to the VCD spectrum of Ia, the ν(OHC) in IIa is increased 100-times when compared to that of II. A change can be seen for the δ(OHC) VCD band rotational strength, too; it equals ca. 60 × 10−44 esu2 cm2 for II while it is ca. −10 × 10−44 esu2 cm2 for IIa.
Below, we shall comment on the differences in the VCD spectra of Ia and IIa (Table 8, Fig. S3 in the ESI).† Firstly, and most importantly, dissimilarity between the VCD spectra gives the rotational strength of the ν(OHW,hb) band a different sign: 10 vs. −25 × 10−44 esu2 cm2 for Ia and IIa, respectively, which also differs with respect to the position of maxima by ca. 30 cm−1, yet their IR intensities are fairly similar. Next, the ν(OHC) band is almost absent in the VCD spectra of Ia, whereas it is quite prominent for IIa (2 vs. 60 × 10−44 esu2 cm2). It is worth recording the fact that the large IR intensity of this mode is a paramount feature of H-bond formation in the mid-IR region, which does not provide a basis for differentiation between the Ia and IIa complexes. In this context, the ν(OHW,hb) and ν(OHC) VCD bands yield a very important distinction between the two similar complex structures.
ν harm/cm−1 | ν scal/cm−1 | ν anharm/cm−1 | I IR/km mol−1 | R i/× 10−44 esu2 cm2 | PED(%) |
---|---|---|---|---|---|
3635 | 3501 | 3444 | 295.1 | 10.2 | 92 ν(OHW,hb) |
3603 | 3470 | 3403 | 364.5 | −25.1 | 92 ν(OHW,hb) |
3312 | 3192 | 3108 | 805.3 | 1.9 | 95 ν(OHC) |
3293 | 3174 | 3080 | 810.9 | 60.2 | 95 ν(OHC) |
2996 | 2891 | 2846 | 25.2 | −23.22 | 99 ν(C*H) |
2955 | 2852 | 2816 | 34.3 | −9.01 | 99 ν(C*H) |
1607 | 1564 | 1558 | 131.9 | −134.7 | 89 β(H2O) |
1610 | 1567 | 1560 | 115.3 | −109.8 | 89 β(H2O) |
1454 | 1418 | 1414 | 1.4 | −16.2 | 74 β(CH + OH)1 |
1464 | 1428 | 1421 | 13.5 | 10.1 | 79 β(CH + OH)1 + 10 βa(CH3) |
1384 | 1351 | 1353 | 5.7 | 22.4 | 83 βs(CH3) |
1382 | 1349 | 1346 | 48.2 | −38.8 | 69 βs(CH3) + 18 β(CH + OH)3 |
1287 | 1259 | 1246 | 259.2 | 133.21 | 30 β(C*H + OHA) + 31 ν(C–O) |
1274 | 1246 | 1242 | 53.4 | 108.73 | 63 β(C*H + OHA) − 14 τ(C*H) + 13 τ(CH3) |
1155 | 1133 | 1130 | 128.5 | 13.7 | 42 ν(C*O) − 14 τ(CH3) + 11 β(C*H) |
1156 | 1134 | 1124 | 60.8 | −25.6 | 49 ν(C*O) − 16 β(C*H) |
628 | 629 | 612 | 25.1 | 17.7 | 43 β(OC![]() ![]() |
573 | 577 | 568 | 22.7 | −20.4 | 36 β(OC![]() ![]() |
422 | 433 | 410 | 8.5 | 6.5 | 60 β(CC–OA) |
429 | 439 | 423 | 6.3 | −13.3 | 73 β(CC–OA) |
376 | 389 | 337 | 81.7 | −184.3 | 49 λ(H2O + OHA) − 10 β(O![]() |
384 | 396 | 309 | 68.0 | -109.1 | 61 λ(H2O + OHA) + 14 δ(OHW,free) |
360 | 373 | 330 | 32.8 | 114.3 | 62 λ(H2O–OHA) |
304 | 320 | 252 | 94.2 | 75.6 | 79 λ(H2O–OHA) |
215 | 235 | 259 | 3.6 | 31.8 | 88 λ(CH3) |
213 | 233 | 138 | 0.5 | −16.7 | 89 λ(CH3) |
The separation of the ν(C*H) bands in the spectra of the two complexes, Ia and IIa, seems to be important for complex identification when they form a mixture. Although they display similar characteristics in terms of IR intensity, and magnitude and sign of VCD rotational strength, they are separated by as much as 40 cm−1 at a very specific lower limit of the ν(CH) vibrations’ region. The VCD intense β(H2O) band is comparable for the two complexes. In fact, the same holds true for the water libration VCD λ(H2O + OHA) band, whereas the other libration, λ(H2O–OHA), has similar VCD intensities for Ia and IIa. Indeed, the position of λ(H2O–OHA) in the two spectra differs by as much as 50 cm−1. This difference would be important for discriminating between the complexes, providing that the bands would not be positioned in the far-IR range (Table 8). An inspection of two kinds of vibrational spectra, IR and VCD, in the region near 1250 cm−1 also possibly allows one to determine whether the complex mixture is composed of two complexes or not. Indeed, in the IR spectrum at 1260 cm−1, the band corresponding to the β(C*H + OHA) mode of Ia has an intensity five times higher in magnitude than that of IIa, positioned at 1245 cm−1. On the other hand, in the VCD spectrum these two bands have similar rotational strengths, i.e. 133 and 109 × 10−44 esu2 cm2, for Ia and IIa, respectively.
The successful discrimination based on the VCD spectra between the two structures is most likely when bands in the two spectra have opposite signs, are intense, and are positioned at (at least slightly) different wavenumbers. If they satisfy the first two conditions, but they are not positioned at different places, they may compensate each other partially or even annihilate each other. Presumably, this would be the case for the βs(CH3), ν(C*O), and λ(CH3) VCD bands, whose maxima differ by ca. 2 cm−1. Also, for the β(CC–OA) VCD bands of Ia and IIa, which exhibit weak VCD absorptions but different signs (7 (Ia) vs. −13 × 10−44 esu2 cm2 (IIa)), the band maxima are separated by only 6 cm−1 (Table 8) and it depends on the band halfwidths as to whether they may mutually compensate each other or not. This is why only two additional bands are fairly good markers for complex discrimination, namely β(CH + OH)1 and β(OCO). Although they are not intense, they differ in sign; and for the Ia and IIa complexes the difference in band location of the former equals ca. 10 cm−1 (the β(CH) bands are fairly narrow) whereas for the latter it equals 50 cm−1.
This paper aimed to address the important question of how to relate the characteristic pattern of the VCD spectra to the intermolecular H-bonding present in the studied system. One may conclude, based on investigations of the formation of H-bonded systems, that the O–H⋯OC H-bond formation results in a significant increase of the ν(C
O) rotational strength (Ia and IIa, Tables 3, 4, and 8). In case of the Ib system, where the alcoholic OHA moiety of the D-lactic acid is a proton donor in the intermolecular H-bond with water, there is a tenfold increase of VCD intensity. Moreover, when the water molecules form a cyclic H-bonded ring, as in Ia and IIa as well as Ib, the bending β(H2O) mode is one of the most intense bands in the VCD spectrum (and in our particular systems bears a negative sign). On the other hand, when the H-bond is not cyclic, as in Ic and Id, the rotational strength of the β(H2O) band is rather small and either positive (Id) or negative (Ic) (Table 9).
System | Mode | ν harm/cm−1 | I IR/km mol−1 | R i/ × 10−44 esu2 cm2 |
---|---|---|---|---|
I | ν(OHA) | 3719 | 86.0 | 16.3 |
ν(OHC) | 3738 | 73.0 | −2.2 | |
ν(C*H) | 3002 | 24.3 | −22.3 | |
ν(C![]() |
1790 | 300.4 | 0.8 | |
Ia | ν(OHA) | 3719 | 83.6 | 13.57 |
ν(OHC) | 3312 | 805.3 | 1.89 | |
ν(C*H) | 2996 | 25.2 | −23.22 | |
ν(C![]() |
1744 | 261 | 49.19 | |
Ib | ν(OHA) | 3628 | 256.9 | 116.4 |
ν(OHC) | 3738 | 68.2 | −6.4 | |
ν(C*H) | 2959 | 38.1 | −30.0 | |
ν(C![]() |
1778 | 361.4 | 18.5 | |
Ic | ν(OHA) | 3702 | 58.9 | 7.3 |
ν(OHC) | 3736 | 77.4 | −1.7 | |
ν(C*H) | 3043 | 10.6 | −23.3 | |
ν(C![]() |
1789 | 292.8 | 3.5 | |
Id | ν(OHA) | 3734 | 54.6 | 20.4 |
ν(OHC) | 3733 | 97.7 | −1.0 | |
ν(C*H) | 3047 | 17.7 | 5.1 | |
ν(C![]() |
1799 | 306.8 | −5.4 | |
Ia | ν(OHfree) | 3863 | 101.5 | −28.9 |
ν(OHHB) | 3635 | 295 | 10.2 | |
β(HOH) | 1607 | 132 | −134.7 | |
Ib | ν(OHfree) | 3878 | 104.3 | −39.2 |
ν(OHHB) | 3689 | 431.8 | −87.9 | |
β(HOH) | 1617 | 87.1 | −47.4 | |
Ic | ν(OHfree) | 3878 | 95.8 | −10.9 |
ν(OHHB) | 3668 | 408.7 | −5.0 | |
β(HOH) | 1643 | 58.9 | −14.7 | |
Id | ν a(OH) | 3888 | 127.2 | 18.9 |
ν s(OH) | 3760 | 88.2 | −4.0 | |
β(HOH) | 1627 | 73.9 | 21.3 |
Finally, let us comment on the anharmonic frequencies calculated for the two most stable conformers of the monomers and complexes (Tables 3 and 4). The high frequency anharmonic stretchings are red-shifted ca. 150–200 cm−1, in comparison to harmonic wavenumbers, and the extent of their shift is greater than that in the case of the scaled values. Below 1800 cm−1, the anharmonic frequencies fluctuate around the appropriate scaled values. For example, in comparison to scaled values, the anharmonic ν(CO) frequency is ca. 20 cm−1 blue-shifted in both monomers and complexes, whereas for instance the δ(OHC) is significantly red-shifted (especially in complexes). For numerous modes the difference between the scaled and anharmonic data is negligible (Tables 3 and 4).
1. The calculations predicted the conformer (I) stabilized by a C–OHA⋯OC intramolecular H-bond to be the most stable form. The second stable form (II), higher in energy by ca. 9.2 kJ mol−1, is also stabilized by the intramolecular C–OHA⋯OHC H-bond. Subsequently, more conformers were found, but they were ruled out of further considerations.
2. For each of the conformers I and II, interacting with a water molecule, four stable minima have been found and studied. The complexes with a cyclic intermolecular H-bond between the carboxylic group of the acid and water (Ia and IIa) proved most stable, playing both a proton donor and proton acceptor role. The binding energies of the two most stable complexes do not differ substantially, with values of −38.9 and −40.0 kJ mol−1, for Ia and IIa, respectively. The Gibbs free energy of the Ia complex is ca. 7.5 kJ mol−1 lower than that of the IIa complex, and thus the former is predicted to be much more populated in a mixture of these two complexes.
VA spectroscopy discriminates between lactic acid conformers Ia and IIa, on the basis of frequency shifts for ν(OHA), ν(C*H), and ν(CO) where the OHA hydroxyl group is a proton donor to an intramolecular H-bond.
The calculated VCD spectra change when the interacting molecules of D-lactic acid⋯water assume different configurations. First, after forming a complex with the D-lactic acid molecule, the VCD modes of the achiral water molecule acquire significant rotational strengths, whose signs change in parallel with the geometry of the complex. Second, several VCD bands enable unequivocal differentiation between conformer structures, which may give new insight into the interpretation of vibrational spectra of complex mixtures. As well as for VCD bands such as ν(OHW,hb) and ν(OHC), which highlight the differences between patterns in the VA and VCD spectra and in this way help interpretation, there are modes like, for example, β(CH + OH)1 and β(OCO) which change their signs upon a change in the complex structure. Third, the stretching vibration ν(C
O) VCD band has a small rotational strength for a monomeric form of D-lactic acid, whereas it turns quite intense when the carbonyl group is engaged in an intramolecular H-bond. In conclusion, the VCD spectra of intramolecularly interacting molecules can be used as a powerful tool for the structural investigation of intermolecular interactions of chiral molecules, and can yield information complementary to data obtained from other molecular spectroscopy techniques.
Footnote |
† Electronic supplementary information (ESI) available: Optimised geometries and a graphical comparison of the calculated VCD spectra. See DOI: 10.1039/b509351a |
This journal is © the Owner Societies 2006 |