IR Spectra of Hydrogen-Bonded Complexes of Trifluoroacetic Acid

Mar 27, 2019 - The structural, spectroscopic, and electronic properties indicate an essential role of weak bonds in the total complexation energy of t...
0 downloads 0 Views 2MB Size
Subscriber access provided by OCCIDENTAL COLL

A: Spectroscopy, Molecular Structure, and Quantum Chemistry

IR Spectra of Hydrogen Bonded Complexes of Trifluoroacetic Acid with Acetone and Diethyl Ether in the Gas Phase. Interaction Between CH and OH Stretching Vibrations Ruslan E. Asfin J. Phys. Chem. A, Just Accepted Manuscript • DOI: 10.1021/acs.jpca.8b10215 • Publication Date (Web): 27 Mar 2019 Downloaded from http://pubs.acs.org on March 27, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

IR Spectra of Hydrogen Bonded Complexes of Trifluoroacetic Acid with Acetone and Diethyl Ether in the Gas Phase. Interaction between CH and OH Stretching Vibrations Ruslan E. Asfin* Department of Physics, Saint Petersburg State University, 7/9 Universitetskaya Nab., 199034 Saint Petersburg, Russian Federation *Corresponding

author.

Tel: +7 812 428 74 19 Fax: +7 812 428 72 40. Email address: [email protected]

1 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 36

Abstract The spectra of complexes of trifluoroacetic acid with acetone and diethyl ether and their perdeuterated isotopologues were extracted from the spectra of the mixture of the compounds recorded at room temperature. The ν(OH) bands of the complexes with protiated and deuterated acetone notable differ from each other, whereas these ν(OH) bands are practically not affected by the deuteration of the diethyl ether. An assumption about the interaction of CH and OH groups in the (CH3)-C=O···HO fragment is made. According DFT calculations, complexes of trifluoroacetic acid with both acetone and diethyl ether have a cyclic structure with one strong =O···HO hydrogen bond and one weak CH···O= bond. The structural, spectroscopic, and electronic properties indicate an essential role of weak bonds in the total complexation energy of the systems studied.

2 ACS Paragon Plus Environment

Page 3 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Introduction The nature of the ν(XH) band structure in the spectra of complexes with moderate and strong XH···B hydrogen bonds is still an actively discussed topic in spectroscopy.1–7 The complexes involving carboxylic acids play a significant role in the study of systems with hydrogen bonds, because of their physical properties and as models of biologically important interactions. There is a number attempts to describe the profile of a ν(OH) band in the spectra of these complexes using different physical models.8,9,18,10–17 Most of them describe the bands profile as a result of interaction of νOH stretching vibration with low-frequency intermolecular modes and Fermi resonances with combination levels of fingerprint vibrations. Consideration of isotope effects comes down to the COOH/COOD substitution in most cases. The influence of deuterium substitution in groups connected to COOH on the profile of the ν(OH) band in the spectra of formic and acetic acids cyclic dimers was studied by Flakus recently19 and in some earlier studies.20,21 For these effects Flakus has proposed the term 'longrange H/D isotopic effects and explained it by the "relatively strong vibronic coupling of the Herzberg–Teller type".19 In this study we looked for similar effects in the spectra of heterogeneous complexes. Here the spectra of trifluoroacetic acid (TFA, CF3COOH) in the mixture with acetone (dimethyl ketone, DMK, (CH3)2CO) and diethyl ether (DEE, (C2H5)2O) and their perdeuterated analogues DMKD6 ((CD3)2CO) and DEED10 ((C2D5)2O) are presented. From these spectra the spectra of individual TFA···B complexes (B = acetones or diethyl ethers) are extracted and discussed. Our attempt to record the spectra of analogous complexes with acetic acid has failed. Namely, in the spectra of mixtures of acetic acid with acetone or diethyl ether the bands of heterogeneous complexes were absent. The TFA···DMK complexation energy in the gas phase was measured by analyzing the temperature dependence of the vapor density22 and the intensity of the ν(OH) bands of monomers and complexes.23,24 The energy of TFA···DEE complexation was also reported 3 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 36

previously.23,24 The spectra of TFA···DMK and TFA···DEE and a number of their H/D isotopologues were recorded in the gas phase and in low-temperature inert matrices,25 however the spectra of complexes were not separated from the spectra of monomers and TFA dimer. Below we discuss the experimental data supported by DFT calculations. The main focus is on the spectral features in ν(C=O) and ν(OH) band regions, especially on the influence of H/D substitution in proton acceptors on the profiles of ν(OH) bands.

Experimental and Calculation Details Experimental. The samples were placed into a stainless cell with BaF2 windows and the gasphase spectra in the 4700 – 800 cm−1 range were recorded with a Bruker IFS125 HR spectrometer at a resolution of 0.5 cm−1. The optical path length was 8.5 cm. A liquid nitrogencooled MCT detector, a KBr/Ge beam splitter, and a globar as a light source were used. All spectra were obtained at a temperature of 20 ± 1 °C. The studied compounds used were obtained from commercial sources. The ampoules with compounds and the cell were attached to a glass vacuum system. After evacuating the cell, the required pressures of compounds were admitted. The acid was admitted firstly, and then acetone or ether was added. A number of spectra of trifluoroacetic acid in the range 2.4 – 10.6 Torr as well as spectra of both isotopologues of acetone (14 – 18 Torr) and diethyl ether (2.4 – 14.5 Torr) were recorded. These data were used to extract the spectra of TFA···DMK(D6) and TFA···DEE(D10) complexes from the spectra of mixtures. The spectra of mixtures of the acid with acetone or ether at total pressure of 4 – 23 Torr were recorded with the ratio of TFA:B in the range 1:1 – 1:4. The all pressures were measured during admission with an U-shaped oil gauge, and after that they could slightly change due to adsorption in the wall of the cell. This did not affect the extraction of spectra of monomers, dimers, and complexes, however did not allow us to estimate the absolute intensity of the bands (see Supporting Information for details).

4 ACS Paragon Plus Environment

Page 5 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

The spectra of complexes were obtained by sequential subtraction of the spectra of monomer and dimer of trifluoroacetic acid, and the spectrum of acceptor from the spectra of mixtures. The procedure is discussed in detail below. Calculations. The quantum-mechanical calculations were performed with a Gaussian 16, Revision A.03 package.26 The density functional theory (DFT) method with the B3LYP functional was used to optimize the geometry. Several starting configurations were chosen for the search of local minima. Additionally, some calculations were made with the wB97XD and M062X functionals. The final geometry optimization, the energy and frequency calculations were performed using the 6-311++G(2df,2pd) basis set. The basis set superposition error (BSSE) correction was done by the counterpoise method during the calculation of the geometry and frequencies of the complexes. The anharmonic calculations were performed with a 'freq=anharmonic' keyword. The analysis of the electron densities in the vicinity of bond critical points was performed with the AIM2000 program.27,28

Results and Discussions Results of Calculations. In Figure 1 the optimized structures of studied complexes are presented. The selected geometrical and energetic parameters are collected in Table 1. The structures of monomers and Cartesian coordinates of atoms in optimized structures of monomers and complexes are presented in Supplementary Information in Table S1. Table 1. Selected Calculated Geometric Parameters of TFA, DMK and DEE Monomers, as well as Geometric and Energetic Parameters of TFA···TFA Dimer, TFA···DMK and TFA···DEE Complexes and the Changes of Parameters Upon Complexation (c – m). Parameter Monomer Complex c−m TFA···TFA r(C=O), Å 1.193 1.212 +0.019 r(O−H), Å 0.969 0.999 +0.030 r(O···O), Å 2.671 ΔE, BSSE, kcal/mol 16.79 ΔE, BSSE, ZPE, kcal/mol 15.66 5 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 36

TFA···DMK r(C=O), TFA, Å 1.193 1.201 +0.008 r(O−H), TFA, Å 0.969 0.997 +0.028 r(C=O), DMK, Å 1.210 1.220 +0.010 r(C−H), DMK, Å 1.086 1.086 −0.00 r(O···O), Å 2.669 r(O···C), Å 3.480 ΔE, BSSE, kcal/mol 12.02 ΔE, BSSE, ZPE, kcal/mol 13.00 TFA···DEE r(C=O), TFA, Å 1.193 1.200 +0.007 r(O−H), TFA, Å 0.969 1.000 +0.031 r(C−O), DEE, Å 1.416 1.433 +0.017 r(O···O), Å 2.653 r(O···C), Å 3.658 ΔE, BSSE, kcal/mol 10.97 ΔE, BSSE, ZPE, kcal/mol 11.98 The dimer of trifluoroacetic acid has the structure of a planar cycle with two strong hydrogen bridges. This structure corresponds to that obtained previously in crystal29 and in the gas phase.30 The energy of dimerization, taking into account the zero-point vibrational energy (ZPE), is ΔE = 15.7 kcal/mol which is comparable with the experimental value of enthalpy of dimerization of ca. 14 kcal/mole obtained independently by various methods.31–36 The complex of trifluoroacetic acid with acetone has the structure similar to a planar ring. Beside the strong C=O···H-O bridge there is the additional weak interaction between the C=O group of the acid and the C-H group of the ketone’s methyl. The calculated energy of complexation ΔE = 13.0 kcal/mol is in a satisfactory agreement with the value of enthalpy of complexation of 14.4 (1.0) kcal/mol obtained by a vapor density technique22 and essentially higher than the value of 7 (1) kcal/mol derived from spectroscopic measurements at a higher temperature.23 In the complex of trifluoroacetic acid with diethyl ether there are two interactions between the molecules: the hydrogen bond between the oxygen atom of the ether and the hydroxyl group of the acid and another weak hydrogen bond between the C=O group of the acid and one of the hydrogen atoms of the methyl group of the ether. The calculations with functionals taking into account dispersion predict the weak bond with hydrogen atom of CH2 group. The energy of 6 ACS Paragon Plus Environment

Page 7 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

dimerization is again essentially higher than the experimental one23 (12 vs 8.4 (1) kcal/mol). Despite the fact that the hydrogen bond in this complex is slightly stronger than that in the TFA···DMK complex (as evidenced by the shorter r(O···O) distance) the calculated total complexation energy is less by 1 kcal/mol. At the same time the r(O···HC) distance in the weak hydrogen bond is essentially longer that in the complex with acetone. It means that the weak interactions play an important role in the complexation energy at least for the TFA···DMK complex. The close values of complexation energies of TFA···DMK and TFA···DEE indicate similar proton acceptor abilities of acetone and ether. The same conclusion was derived previously from the gas phase spectral properties and calculations of complexes of methanol with acetone and dimethyl ether.37 Spectra Extraction. In the spectra of mixtures of TFA with a proton acceptor B one expects to record a superposition of four types of spectra: the spectra of monomers of TFA and B, the spectrum of TFA dimers and most interesting for us spectrum of the TFA···B complex. To identify and isolate the latter the consequent subtraction of spectra should be performed. The spectra of monomeric and dimeric forms of TFA were obtained from the analysis of the acid’s spectra recorded at different pressures, presented in Figure 2(A). It is well known that TFA in the gas phase at temperatures near 300 K exists predominantly in monomeric and dimeric forms. The separation of monomer’s and dimer’s spectra in this case is a rather simple task because some bands are located in non-overlapping regions. The examples of such monomer bands are marked with the symbol "M" in Figure 2(A): the ν(OH) band near 3580 cm−1, the ν(C=O) band near 1830 cm−1, and the δ(OH) band near 1125 cm−1. The spectrum of individual species can be obtained by subtracting the spectrum recorded at a certain pressure from the spectrum obtained at a different pressure with a coefficient chosen to eliminate the characteristic bands of other species. Here, the series of four spectra recorded at different pressures of TFA between 2.5 and 10 Torr was treated in this way. The spectra of monomers and 7 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 36

dimers obtained from different pairs of experimental spectra were averaged and the results are presented in Figure 2(B). After the spectra of monomers and TFA dimer are established, the spectra of TFA···B can be extracted from the spectra of mixtures. The spectra of monomers could be easily subtracted because the monomers have separately located bands. However, the bands of TFA dimer and the bands of the complexes strongly overlap with each other and no isolated bands of TFA dimer were found. To overcome this problem, the information on the dimerization constant obtained by processing the spectra of pure TFA and the intensity of the TFA monomer bands were used to find out the appropriate coefficient with which to subtract the spectrum of TFA dimer. As an example of the result, the spectra of mixture of trifluoroacetic acid and acetone-D6, as well as the isolated spectrum of their complex are presented in Figure 3. Some additional aspects of spectra extraction are discussed in Supplementary Information. The Region of C=O Vibrations. In Figure 4 the spectra of the studied monomers and complexes are presented in the spectral region of the ν(C=O) band. The used resolution and probably the pressure and the temperature in the cell do not allow the rotational-vibrational structure of these bands even in the spectra of monomers. The bands of complexes have more or less symmetric shape and the wavenumbers of maxima were used as wavenumbers of vibrational transitions in the complexes. The bands in the spectra of monomers of trifluoroacetic acid and acetone both in protiated and deuterated forms are asymmetric. The central peak of the TFA monomer band (maximum of a Q-branch) was used to determine the wavenumber of the corresponding transition. The value of 1830.0(5) cm−1 is in satisfactory agreement with the previously reported ones: 1826 cm−1,38 1830 cm−1,35 1829 cm−1,39 1827 cm−1,23,24 and slightly differs from 1843 cm−1.40 It should be noted, that all these values were obtained with prism spectrometers. The recently reported value41,42 of 1830 cm−1 is also in a good agreement with the presented wavenumber. The spectrum of protiated acetone is in excellent agreement in the range of intersection with the 8 ACS Paragon Plus Environment

Page 9 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

spectrum reported in the GEISA spectroscopic database,43 recorded with a resolution of 0.015 cm−1 at 297.5 K. The absorption in the ν(C=O) band range has a complicated structure and there are two local maxima at 1739.8 and 1728.5 cm−1 in the spectra of DMK and 1743.7 and 1732.0 cm−1 in the spectra of DMKD6. The local minima are observed near 1731.6 and 1739.7 cm−1 for protiated and deuterated forms respectively. The centers of gravity of the bands are located near 1741 and 1736 cm−1. The previously reported wavenumbers of the ν(C=O) band for the protiated form of DMK are 1731,44 1738,45 and 1740 cm−1 (the latter is an implicit value, calculated from tabulated shifts).25 For DMKD6 the reported values are 1732,44 1741,45 and 1740 cm−1.25 It is also reported the value of the center of gravity of the DMK band at 1742 cm−1.3 The most probable origin of such a structure of the ν(C=O) band in the spectra of DMK is the Fermi resonance between the ν(C=O) stretching and combination levels. According to anharmonic calculations the most probably combination levels in protiated acetone are the overtones of γ(C-H)sym and δ(C-C)as as well as the combination γ(C=O)+ν(C-C)as. Fermi resonance with these bands leads to the appearance of four bands: the most intense one is the high-frequency component at 1771 cm−1; the lowest frequency component is located at 1722 cm−1. The center of gravity of these four components is near 1753.5 cm−1. In the calculated anharmonic spectra of deuterated acetone there are no resonances in the region of the ν(C=O) band. However in addition to the carbonyl band near 1747 cm−1, the combinational band of the ν(C-C)sym and the symmetric umbrella vibration of CH3 with a noticeable intensity is located at 1764 cm−1. The center of gravity of these two bands is located near 1758 cm−1. In calculated spectra of complexes of TFA with DMK and DMKD6 there are no the strong Fermi resonances in the region of the ν(C=O) band. Upon formations of the complexes the ν(C=O) bands of both TFA and acetone are red-shifted. The experimental and calculated wavenumbers of these bands and the shifts are collected in Table 2. 9 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 36

Table 2. The Experimental (ν(C=O) exp) And Calculated (ν(C=O) calc, Harmonic and Anharmonic) Wavenumbers of the ν(C=O) Bands in the Complexes and the Shifts (Δc-m) Relative To the Wavenumbers of These Bands in Monomers of Trifluoroacetic Acid and Acetone. ν(C=O) exp, Δc-m exp, ν(C=O) calc, Δc-m calc, cm−1 cm−1 harm/anharm, harm/anharm, cm−1 cm−1 ν(C=O) TFA TFA···TFA 1788.6 −41 1808 / 1772 −51 / −54 a b b −72 1779 / 1731 −80 / −95 TFA···DMK 1802.4 −28 1823 / 1788 −36 / −38 TFA···DMKD6 1802.2 −28 1823 / 1787 −36 / −39 TFA···DEE 1795.9 −34 1825/1791 −34 / −35 TFA···DEED10 1796.1 −34 1824/1789 −35/ −37 ν(C=O) DMK TFA··· DMK 1715.2 −26c/ −23d 1747 / 1699 −39 / −54 c d TFA···DMKD6 1706.2 −30 / −38 1737 / 1688 −40 / −70 a Including half of splitting between antisymmetric and symmetric vibrations. See text for details b The mean of symmetric and asymmetric wavenumbers of ν(C=O) c From the center of gravity of the monomer ν(C=O) band d From the maximum of the monomer ν(C=O) band

It is remarkable that the carbonyl band in the spectrum of TFA···DMK complex is located at higher frequency than that in the spectrum of TFA···DMKD6 complex, whereas the maxima of absorption (but not the centers of gravity!) of acetone monomers are located in the reverse order. The harmonic calculations overestimate the wavenumbers of the bands, while anharmonic calculations underestimate them. The shifts of the bands upon complexation are also slightly overestimated both in harmonic and anharmonic case, although for the spectra of TFA···DEE(D10) complexes the calculations perfectly reproduce the experimental data. The slightly larger shift of carboxyl ν(C=O) band upon complex formation in the case of the ethers comparing with the acetones indirectly indicates a stronger hydrogen bond in the TFA···DEE complex. The comparison of the shifts of carboxyl ν(C=O) band in the spectra of TFA···TFA and TFA···acceptor is pointless because both C=O bands in the acid dimer participate in strong 10 ACS Paragon Plus Environment

Page 11 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

hydrogen bonds. Another problem in estimating the hydrogen bond strength from the experimental shift of the ν(C=O) band is its splitting for TFA dimers into symmetric (Raman active) and antisymmetric (IR active) vibrations, because of the symmetry of the complex. Sometimes it is called Davydov splitting.10,13 To the best of our knowledge there are no data on the Raman shift of the symmetric ν(C=O) band of trifluoroacetic acid in the gas phase. The difference between the wavenumbers of antisymmetric (IR active) and Raman shift of symmetric (Raman active) vibrations of C=O groups in jet-cooled spectra is νas – νs = 56 cm−1;46 in solution in CCl4 this value is νas – νs ≈ 62 cm−1.35,38,40,47 Our calculation gave the value of 59 cm−1 in harmonic calculation and 82 cm−1 in anharmonic one. Thus, approximately 30 cm−1 should be added to the shift of ν(C=O) band upon dimerization of trifluoroacetic acid to connect these shift with the energy of hydrogen bond. This value is also listed in Table 2. The Region of OH Vibration. Interaction of ν(OH) with ν(CH) Modes. The maximum of the ν(OH) band of monomers of trifluoroacetic acid is located near 3580.0 cm−1. The previously reported value of 3587 cm−1,35,38 is seems to be the maximum of the band in low resolution spectra. In the spectra recorded with relatively high resolution this band has the maximum matching with the wavenumber reported here.48 In the spectra of TFA dimers the ν(OH) band is broad with several submaxima (Figure 2). The features of the band coincide generally with those described earlier.35,38 In the Figure 5 the extracted spectra of TFA···DMK(D6) and TFA···DEE(D10) complexes as well as spectra of DMK and DEE monomers are presented in the region of ν(OH) bands of complexes. The ν(OH) band of all complexes is broad and intense, which is characteristic for strong H-bonded complexes. The spectral features of the TFA···DMKD6 and TFA···DEED10 bands (the wavenumbers of submaxima) coincide with the previously reported values25 as a whole. These values are listed in Tables S13 and S14 in the Supporting Information. The most probable reason of some disagreements is that in the literature the unextracted spectra are described25, which included the absorption of monomers of TFA, deuterated acceptor monomers 11 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 36

and TFA dimers, whereas here the separated bands of TFA···acceptor complexes are considered. The comparison of the spectra in Figure 5 shows that the ν(CH) bands of protiated diethyl ether in the complex TFA···DEE are sufficiently narrow and well discernible against the background of the ν(OH) band. The ν(CH) bands are located close to the same bands of DEE monomer. On the other hand, in the spectra of the TFA···DMK complex there is no evidence of clearly visible narrow bands in the region of ν(CH) of the DMK monomer. The similar picture is observed in the ν(CD) region in the spectra of the TFA···B complexes: there are well distinguishable bands slightly shifted from the bands of DEED10 monomer in the spectra of TFA···DEED10 complex and there are no sharp bands in this region in the spectra of TFA···DMKD6 complex. In the harmonic approximation the calculations predict a decrease in the total intensity of the ν(CH) bands by 20% and ν(CD) bands by 5% upon formation of complexes of trifluoroacetic acid with corresponding acetones. This of course cannot explain the disappearance of these bands in the spectra. Again, one can see that the ν(OH) bands of TFA···DMK and TFA···DMKD6 noticeably differ from each other, especially near the hole at 3095 cm−1, where the redistribution of intensity is notable. Another distinction is in the low-frequency wing of the band near 2275 cm−1. In contrast, the ν(OH) bands in the spectra of TFA···DEE and TFA···DEED10 complexes are virtually the same except for the contributions of ν(CH) and ν(CD) bands of diethyl ether in the spectra of complexes. To eliminate the ν(CH) bands the part of the spectrum of the TFA···DEE complex in the region 3040 – 2800 cm−1 was replaced by the ν(OH) band of the TFA···DEED10 complex in this region. The same was done for spectra of TFA···DEED10 in the range of ν(CD) band region. The result is shown in Figure 5B by dashed lines. To describe the broad bands it is convenient to use a set of spectral moments of a spectral function S ( )  D( ) / , where D( )  lg( I 0 / I ) . The zero spectral moment M 0 



S ( )d is

band

the integrated intensity of the band. The first normalized spectral moment

12 ACS Paragon Plus Environment

Page 13 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

M 1  M 01



S ( )  d is the center of gravity of the band. The second normalized central

band

spectral moment M 2  M 01



S ( )  (  M 1 ) 2 d characterizes the effective half-width of the

band

band 1/ 2  2 M 2 that is close to the full width of the band at half maximum. In Table 3 these moments of the ν(OH) bands are collected for all studied complexes.

Table 3. The First M1 (cm−1) Normalized Moment and Effective Half-Width ν1/2 (cm−1) of the ν(OH) Band in the Spectra of the Complexes and the Shift of M1 from the Wavenumber of TFA Monomer Band νm. M1, cm−1 ν1/2, cm−1 M1 − νm, cm−1 TFA···TFA 3022 430 −558 TFA···DMK 3032 475 −548 TFA···DMKD6 3042 530 −538 TFA···DEE 2867 435 −713 TFA···DEED10 2868 415 −712 The values of the moments support the previously mentioned observation: the centers of gravity and effective half-widths of ν(OH) bands of TFA···DMK and TFA···DMKD6 are noticeable different from each other, whereas these values are practically the same for TFA···DEE and TFA···DEED10 complexes. Differences in the ν(OH) bands and the absence of pronounced ν(CH)/ν(CD) bands in the spectra TFA···DMK and TFA···DMKD6 evidence indicate a substantial interaction of CH/CD bonds of the methyl group of acetone and the OH bond of the acid, which is absent in case of complexes formed between trifluoroacetic acid and diethyl ether. Similar effects were observed previously for the dimers of acetic acid. In ragout-jet spectra the influence of CD3 ← CH3 substitution on the "fine structure" of the ν(OH) band was noted.10,49 Flakus et al. have studied in particular the influence of deuterium substitution in methyl group on the ν(OH) band.19 A mechanism of this influence was proposed based on the vibronic coupling of the Herzberg–Teller type. Dreyer had to include two γCH3 wagging modes to describe the ν(OH) band by multidimensional DFT calculation and consideration of Fermi-resonances and anharmonic couplings with low frequency modes.11 On the other hand, it seems that the 13 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 36

interaction between νOH and νCH modes was not considered in this study. In this study it was shown that vibrations of methyl group of one molecule influence the shape of the ν(OH) band of the other molecule when these molecules form the hydrogen bonded complex. Probably the electronic structure of the (CH3)-C=O···HO fragment plays an important role in this process. In any case one should take into account the CH vibrations when describing the ν(OH) band of complexes with the C=O···HO hydrogen bridge. The first moment M1 of the complicated ν(XH) band can be used to estimate the shift of the band upon complexation.50 This shift is often associated with the strength of the hydrogen bond. There are several different correlations proposed between these two values.51–54 Despite the different types of functions describing these correlations all of them have one common property: a larger shift corresponds to a stronger bond. Thus, one can conclude (see Table 3) that the hydrogen bond in the TFA···DMK complex is slightly weaker than that in TFA··· DEE and the strength of this bond is approximately the same as in the trifluoroacetic dimer (per one bond). It should be reminded that the same conclusion about the relative strength of H-bonds in complexes with acetone and ether was derived above based on the calculated r(O···O) distances and observed shifts of ν(C=O) bands. The results of the AIM analysis28 of the complexes are presented in Supporting Information in Figures S1-S3; the properties of the electron density in the vicinity of intermolecular bond critical points obtained from B3LYP calculations are presented in Table 4. The comparison of these properties calculated with different functionals a collected in Supplementary Information. The correlations between the potential energy density or Lagrangian form of kinetic energy density and the intermolecular hydrogen bond strength in crystals were proposed55,56 and checked for moderately strong H-bonds.57 We have speculatively applied these correlations to our system, to estimate the relative energy of different intermolecular bonds. It should be note, that absolute values of estimated energies are rather meaningless, because the relations predict the energy if the distances between the atoms in the bonds are obtained in the crystals. 14 ACS Paragon Plus Environment

Page 15 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Nevertheless these estimations show that the O···HO bond energy in the TFA···DMK complex is lower than that in the TFA···TFA dimer, and noticeably lower that in the TFA···DEE complex. On the other hand, the energy of the weak C-H···O= bond in the complex between the acid and the acetone is much higher than that in the TFA···DEE complex. Thus, the difference in the strength of the weak interactions makes the TFA···DMK complex more stable that the TFA···DEE complex. This also can be used to explain the difference in the complexation energies of TFA···DMK obtained by a vapor density technique22 in the temperature range 21 – 40 °C and by spectroscopic measurements23 in the range 20 – 200 °C. Indeed, at high temperatures, the weak intermolecular bond could be broken, and only the O-H···O hydrogen bond would contribute to the measured energy. Table 4. The AIM Properties of the Intermolecular Bond Critical Points of Complexes Studied: Electron Densitya (ρ), Laplacian of ρ (2ρ), Potential Energy Density (V), Lagrangian Form of Kinetic Energy Density (G), the Estimated Bond Energies from V(ΔE (V)) and from G (ΔE (G)) CP ρ, a.u. V, a.u. ΔE (G)b, ΔE (V)c, 2ρ, a.u. G, a.u. kcal/mol kcal/mol TFA···TFA BCP (3,-1) 0.04844 0.09443 0.03482 −0.0460 9.0 8.9 O···HO TFA···DMK BCP (3,-1) 0.04663 0.09728 0.03417 −0.0440 8.9 8.6 OH···O BCP (3,-1) 0.00976 0.03181 0.00689 −0.0058 1.2 1.1 O···HC TFA···DEE BCP (3,-1) 0.05019 0.09559 0.03599 −0.0481 9.4 9.4 OH···O BCP (3,-1) 0.00590 0.01890 0.00399 −0.0033 0.4 0.6 O···HC a ρ was calculated by B3LYP/6-311++G(2df,2pd) method b ΔE (G) = (0.448∙G − 0.0012)∙ 627.509 (Ref. 56) c ΔE (V) = −0.31∙V∙ 627.509(Ref. 56)

Conclusions The IR spectra of mixtures of trifluoroacetic acid with acetone and diethyl ether and their perdeuterated analogues have been recorded in the gas phase at the room temperature. The spectra of complexes formed between the acid and the proton acceptors have been extracted. 15 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 36

Additionally the equilibrium structure, harmonic and anharmonic vibrational frequencies of these complexes have been obtained by DFT calculations. It was found that the profile of the ν(OH) band in the spectra of complexes of trifluoroacetic acid with diethyl ether virtually does not depend on deuteration of the ether whereas these bands differ noticeably in the spectra of the complex of the acid with acetone and acetone-D6. Besides, there are no pronounced ν(CH) or ν(CD) bands in the spectra of TFA···DMK and TFA···DMKD6 complexes. This is the signature of strong interaction between the CH3 (CD3) and OH groups, which should be taken into account in calculations of the ν(OH) band profile. This effect has been observed for the first time in a heterogeneous complex. The calculations show that the complexes between trifluoroacetic acid and both acetone and diethyl ether have a cyclic structure with one strong hydrogen bond and one weak bond CH···O. This weak interaction seems to play an important role in the energetics of the complex formation. Although the H-bond in the TFA···DMK complex is weaker than that in TFA···DEE complex, based on structural, spectroscopic, and electronic properties, the total complexation energy at a relatively low temperature is larger in the first complex.

Supporting Information The file with supporting information contains the tables with Cartesian coordinates of atoms in the complexes and monomers studied; the wavenumbers of maxima of the bands in experimental spectra, calculated frequencies and intensities in harmonic approximation, selected frequencies and intensities calculated in anharmonic approximation; calculated energies of complexation; description of ν(OH) bands; positions and parameters of bond critical points; some aspects of spectra extraction; discussion of ways to obtain the equilibrium constant.

Acknowledgements R.E.A. is grateful to Dr. S. Melikova and Dr. P. Tolstoy for fruitful discussions. Author acknowledges a financial support of the Russian Science Foundation (Project 18-13-00050). The 16 ACS Paragon Plus Environment

Page 17 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

spectra were obtained with core facilities of RC “Geomodel” of St. Petersburg State University. The calculations have been performed using the computer resources of the Resource Centre of SPbGU (http://cc.spbu.ru).

References (1)

Bulychev, V. P.; Grigoriev, I. M.; Gromova, E. I.; Tokhadze, K. G. Study of the ν1 Band Shape of the H2O⋯HF, H2O⋯DF, and H2O⋯HCl Complexes in the Gas Phase. Phys. Chem. Chem. Phys. 2005, 7, 2266–2278.

(2)

Rekik, N.; Ghalla, H.; Hanna, G. Explaining the Structure of the OH Stretching Band in the IR Spectra of Strongly Hydrogen-Bonded Dimers of Phosphinic Acid and Their Deuterated Analogs in the Gas Phase: A Computational Study. J. Phys. Chem. A 2012, 116, 4495–4509.

(3)

Bulychev, V. P.; Svishcheva, E. A.; Tokhadze, K. G. Experimental and Theoretical Study of Absorption Spectrum of the (CH3)2CO⋯HF Complex. Influence of Anharmonic Interactions on the Frequency and Intensity of the CO and H–F Stretching Bands. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2014, 117, 679–685.

(4)

Van Hoozen, B. L.; Petersen, P. B. Origin of the Hadži ABC Structure: An Ab Initio Study. J. Chem. Phys. 2015, 143, 184305.

(5)

Rekik, N.; Oujia, B.; Wójcik, M. J. Theoretical Infrared Spectral Density of H-Bonds in Liquid and Gas Phases: Anharmonicities and Dampings Effects. Chem. Phys. 2008, 352, 65–76.

(6)

Rekik, N.; Wójcik, M. J. On the Influence of Electrical Anharmonicity on Infrared Bandshape of Hydrogen Bond. Chem. Phys. 2010, 369, 71–81.

(7)

Asfin, R. E.; Denisov, G. S.; Mielke, Z.; Tokhadze, K. G. Particular Features of the ν(OH) Absorption Band of Strongly Hydrogen-Bonded Complexes in the Gas Phase, LowTemperature Matrices, and Crystalline Films at 12–600 K. Opt. Spectrosc. 2005, 99, 56– 67. 17 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(8)

Page 18 of 36

Maréchal, Y. IR Spectra of Carboxylic Acids in the Gas Phase: A Quantitative Reinvestigation. J. Chem. Phys. 1987, 87, 6344–6353.

(9)

Florio, G. M.; Zwier, T. S.; Myshakin, E. M.; Jordan, K. D.; Sibert, E. L. Theoretical Modeling of the OH Stretch Infrared Spectrum of Carboxylic Acid Dimers Based on FirstPrinciples Anharmonic Couplings. J. Chem. Phys. 2003, 118, 1735–1746.

(10)

Emmeluth, C.; Suhm, M. A.; Luckhaus, D. A Monomers-in-Dimers Model for Carboxylic Acid Dimers. J. Chem. Phys. 2003, 118, 2242–2255.

(11)

Dreyer, J. Hydrogen-Bonded Acetic Acid Dimers: Anharmonic Coupling and Linear Infrared Spectra Studied with Density-Functional Theory. J. Chem. Phys. 2005, 122, 184306.

(12)

Petković, M.; Novak, J.; Došlić, N. Shaping the Infrared Spectrum of the Acetic Acid Dimer in the OH-Stretching Range: Multiple Conformers and Anharmonic Coupling. Chem. Phys. Lett. 2009, 474, 248–252.

(13)

Blaise, P.; Wojcik, M. J.; Henri-Rousseau, O. Theoretical Interpretation of the Line Shape of the Gaseous Acetic Acid Cyclic Dimer. J. Chem. Phys. 2005, 122, 064306.

(14)

Van Hoozen, B. L.; Petersen, P. B. Origin of the 900 cm−1 Broad Double-Hump OH Vibrational Feature of Strongly Hydrogen-Bonded Carboxylic Acids. J. Chem. Phys. 2015, 142, 104308.

(15)

Mackeprang, K.; Xu, Z.-H.; Maroun, Z.; Meuwly, M.; Kjaergaard, H. G. Spectroscopy and Dynamics of Double Proton Transfer in Formic Acid Dimer. Phys. Chem. Chem. Phys. 2016, 18, 24654–24662.

(16)

Gług, M.; Brela, M. Z.; Boczar, M.; Turek, A. M.; Boda, Ł.; Wójcik, M. J.; Nakajima, T.; Ozaki, Y. Infrared Spectroscopy and Born–Oppenheimer Molecular Dynamics Simulation Study on Deuterium Substitution in the Crystalline Benzoic Acid. J. Phys. Chem. B 2017, 121, 479–489.

(17)

Van Hoozen, B. L.; Petersen, P. B. A Combined Electronic Structure and Molecular 18 ACS Paragon Plus Environment

Page 19 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Dynamics Approach to Computing the OH Vibrational Feature of Strongly HydrogenBonded Carboxylic Acids. J. Chem. Phys. 2017, 147, 224304. (18)

Van Hoozen, B. L.; Petersen, P. B. Vibrational Tug-of-War: The PKA Dependence of the Broad Vibrational Features of Strongly Hydrogen-Bonded Carboxylic Acids. J. Chem. Phys. 2018, 148, 134309.

(19)

Flakus, H. T.; Hachuła, B. Effect of the Resonance of the C–H and O–H Bond Stretching Vibrations on the IR Spectra of the Hydrogen Bond in Formic and Acetic Acid. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2011, 79, 1276–1284.

(20)

Flakus, H. T. Effect of the Resonance of Molecular Vibrations on the Ir Spectra of Hydrogen Bonds. J. Mol. Struct. 1981, 72, 107–117.

(21)

Flakus, H. T.; Machelska, A. ‘Long-Range’ Isotopic H/D Effects in the Infrared Spectra of the Centrosymmetric Dimeric Systems of Hydrogen Bonds. J. Mol. Struct. 1998, 447, 97–109.

(22)

Ling, C.; Christian, S. D.; Affsprung, H. E. Vapor Phase Association of Trifluoroacetic Acid with Acetone and Cyclopentanone. J. Phys. Chem. 1966, 70, 901–904.

(23)

Gerasimov, I. V.; Tokhadze, K. G. Spectroscopic Determination of the Energy of Trifluoroacetic Acid Complexes Containing Proton Acceptors in the Gas Phase. J. Appl. Spectrosc. 1977, 26, 772–775.

(24)

Gerasimov, I. V.; Kul’bida, A. I.; Tokhadze, K. G.; Shraiber, V. M. Spectroscopic Investigation of Systems with a Strong Hydrogen Bond in the Gas Phase at High Temperatures. J. Appl. Spectrosc. 1980, 32, 629–633.

(25)

Wierzejewska-Hnat, M.; Mielke, Z.; Ratajczak, H. Infrared Matrix Isolation and Gas Phase Studies of Complexes between Trifluoroacetic Acid and Diethylether and Acetone. Spectrochim. Acta Part A Mol. Spectrosc. 1987, 43, 675–678.

(26)

Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Petersson, G. A.; Nakatsuji, H.; et al. Gaussian 16. 19 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 36

Gaussian, Inc.: Wallingford CT 2016. (27)

Biegler–König, F.; Schönbohm, J.; Bayles, D. AIM2000. J. Comput. Chem. 2001, 22, 545–559.

(28)

Bader, R. F. W. Atoms in Molecules. A Quantum Theory; Clarendon Press, 1994.

(29)

Nahringbauer, I.; Lundgren, J. O.; Andersen, E. K. Trifluoroacetic Acid. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 1979, 35, 508–510.

(30)

Karle, J.; Brockway, L. O. An Electron Diffraction Investigation of the Monomers and Dimers of Formic, Acetic and Trifluoroacetic Acids and the Dimer of Deuterium Acetate 1. J. Am. Chem. Soc. 1944, 66, 574–584.

(31)

Lundin, R. E.; Harris, F. E.; Nash, L. K. The Vapor Phase Association of Trifluoroacetic Acid. J. Am. Chem. Soc. 1952, 74, 4654–4656.

(32)

Taylor, M. D.; Templeman, M. B. The Vapor Phase Dissociation of Some Carboxylic Acids. III. 1 Trifluoroacetic Acid and Trifluoroacetic Acid-d 2,3. J. Am. Chem. Soc. 1956, 78, 2950–2953.

(33)

Frurip, D. J.; Curtiss, L. A.; Blander, M. Vapor Phase Association in Acetic and Trifluoroacetic Acids. Thermal Conductivity Measurements and Molecular Orbital Calculations. J. Am. Chem. Soc. 1980, 102, 2610–2616.

(34)

Lumbroso-Bader, N.; Coupry, C.; Baron, D.; Clague, D. .; Govil, G. Dimerization of Carboxylic Acids: A Vapor Phase NMR Study. J. Magn. Reson. 1975, 17, 386–392.

(35)

Kagarise, R. E. Infrared Spectrum of Trifluoroacetic Acid Vapor. J. Chem. Phys. 1957, 27, 519–522.

(36)

Sauren, H.; Winkler, A.; Hess, P. Kinetics and Energetics of Hydrogen Bond Dissociation in Isolated Acetic Acid-D1 and -D4 and Trifluoroactic Acid Dimers. Chem. Phys. Lett. 1995, 239, 313–319.

(37)

Hansen, A. S.; Kjaergaard, H. G. Dimethyl Sulfoxide Complexes Detected at Ambient Conditions. J. Phys. Chem. A 2017, 121, 6046–6053. 20 ACS Paragon Plus Environment

Page 21 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(38)

Fuson, N.; Josien, M.; Jones, E. A.; Lawson, J. R. Infrared and Raman Spectroscopy Studies of Light and Heavy Trifluoroacetic Acids. J. Chem. Phys. 1952, 20, 1627–1634.

(39)

Murty, T. S. S. R.; Pitzer, K. S. Trifluoroacetic Acid. Nature of Association in Dilute Solutions in Nonpolar Solvents. J. Phys. Chem. 1969, 73, 1426–1429.

(40)

Barceló, J. R.; Otero, C. Infra-Red Spectra of the Fluorinated Acetic Acids in Condensed and Vapour Phase. Spectrochim. Acta 1962, 18, 1231–1247.

(41)

Keller, J. W. The Formic Acid−Trifluoroacetic Acid Bimolecule. Gas-Phase Infrared Spectrum and Computational Studies. J. Phys. Chem. A 2004, 108, 4610–4618.

(42)

Hurley, M. D.; Sulbaek Andersen, M. P.; Wallington, T. J.; Ellis, D. A.; Martin, J. W.; Mabury, S. A. Atmospheric Chemistry of Perfluorinated Carboxylic Acids: Reaction with OH Radicals and Atmospheric Lifetimes. J. Phys. Chem. A 2004, 108, 615–620.

(43)

Jacquinet-Husson, N.; Armante, R.; Scott, N. A.; Chédin, A.; Crépeau, L.; Boutammine, C.; Bouhdaoui, A.; Crevoisier, C.; Capelle, V.; Boonne, C.; et al. The 2015 Edition of the GEISA Spectroscopic Database. J. Mol. Spectrosc. 2016, 327, 31–72.

(44)

Dellepiane, G.; Overend, J. Vibrational Spectra and Assignment of Acetone, Ααα Acetone-D3 and Acetone-D6. Spectrochim. Acta 1966, 22, 593–614.

(45)

Cossee, P.; Schachtschneider, J. H. Vibrational Analysis of Acetone, Acetaldehyde, and Formaldehyde. J. Chem. Phys. 1966, 44, 97–111.

(46)

Meyer, K. A. E.; Suhm, M. A. Vibrational Exciton Coupling in Homo and Hetero Dimers of Carboxylic Acids Studied by Linear Infrared and Raman Jet Spectroscopy. J. Chem. Phys. 2018, 149, 104307.

(47)

Atakhodzhaev, A. K.; Tukhvatullin, F. K.; Muradov, G.; Zhumboev, A.; Tashkenbaev, U. N. Tursunkulov, A. Raman Scattering Spectra of Liquid Trifluoroacetic Acid. Opt. Spectrosc. 1996, 80, 181–184.

(48)

Shipman, S. T.; Douglass, P. C.; Yoo, H. S.; Hinkle, C. E.; Mierzejewski, E. L.; Pate, B. H. Vibrational Dynamics of Carboxylic Acid Dimers in Gas and Dilute Solution. Phys. 21 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 36

Chem. Chem. Phys. 2007, 9, 4572. (49)

Häber, T.; Schmitt, U.; Emmeluth, C.; Suhm, M. A. Ragout-Jet FTIR Spectroscopy of Cluster Isomerism and Cluster Dynamics: From Carboxylic Acid Dimers to N2O Nanoparticles. Faraday Discuss. 2001, 118, 331–359.

(50)

Asfin, R. E.; Denisov, G. S.; Poplevchenkov, D. N.; Tokhadze, K. G.; Velikanova, T. V. IR ν(OH) Band and Dimerization of Phosphorus Acids in the Gas Phase and Solid State. Pol. J. Chem. 2002, 76, 1223–1231.

(51)

Badger, R. M.; Bauer, S. H. Spectroscopic Studies of the Hydrogen Bond. II. The Shift of the O–H Vibrational Frequency in the Formation of the Hydrogen Bond. J. Chem. Phys. 1937, 5, 839–851.

(52)

Iogansen, A. V.; Rassadin, B. V. Dependence of Intensity and Shift in the Infrared Bands of the Hydroxyl Group on Hydrogen Bonding Energy. J. Appl. Spectrosc. 1969, 11, 1318– 1325.

(53)

Ratajczak, H.; Orville-Thomas, W. J.; Rao, C. N. R. Charge Transfer Theory of Hydrogen Bonds: Relations between Vibrational Spectra and Energy of Hydrogen Bonds. Chem. Phys. 1976, 17, 197–216.

(54)

Wendler, K.; Thar, J.; Zahn, S.; Kirchner, B. Estimating the Hydrogen Bond Energy. J. Phys. Chem. A 2010, 114, 9529–9536.

(55)

Espinosa, E.; Molins, E.; Lecomte, C. Hydrogen Bond Strengths Revealed by Topological Analyses of Experimentally Observed Electron Densities. Chem. Phys. Lett. 1998, 285, 170–173.

(56)

Mata, I.; Alkorta, I.; Espinosa, E.; Molins, E. Relationships between Interaction Energy, Intermolecular Distance and Electron Density Properties in Hydrogen Bonded Complexes under External Electric Fields. Chem. Phys. Lett. 2011, 507, 185–189.

(57)

Vener, M. V.; Egorova, A. N.; Churakov, A. V.; Tsirelson, V. G. Intermolecular Hydrogen Bond Energies in Crystals Evaluated Using Electron Density Properties: DFT 22 ACS Paragon Plus Environment

Page 23 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Computations with Periodic Boundary Conditions. J. Comput. Chem. 2012, 33, 2303– 2309.

23 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 36

Figure Captions Figure 1. The optimized structures of complexes calculated at B3LYP/6-311++G(2df,2pd) level of theory: (A) the dimer of trifluoroacetic acid; (B) the TFA···DMK complex; (C) the TFA···DEE complex.

Figure 2. (A) The experimental spectra of trifluoroacetic acid at 6.5 Torr (a) and 3.2 Torr (b). The symbol M denotes the monomer bands. (B) The spectra of TFA monomer (d) and dimer (c) as the result of extraction from experimental spectra at different pressure.

Figure 3. The spectra of a mixture of TFA (4.1 Torr) with DMKD6 at total pressure of 12.3 Torr (a) and the isolated spectrum of the TFA···DMKD6 complex (b).

Figure 4. The spectra of the monomer of trifluoroacetic acid (TFA); the dimer of trifluoroacetic acid (TFA···TFA); monomers of protiated (DMK) and deuterated (DMKD6) acetones; TFA···DMK and TFA···DMKD6 complexes; monomers of protiated (DEE) and deuterated (DEED10) diethyl ethers; TFA···DEE and TFA···DEED10 complexes in the region of ν(C=O) bands.

Figure 5. The spectra of TFA···DMK and TFA···DMKD6 complexes, protiated acetone (DMK), deuterated acetone (DMKD6), TFA···DEE and TFA···DEED10 complexes, protiated diethyl ether (DEE), and deuterated diethyl ether (DEED10), in the range of ν(OH) bands of the corresponding complexes. The dashed lines (TFA···DEE, TFA···DEED10) is the most probable shape of the ν(OH) band of the TFA···DEE complex under the ν(CH) bands and the most probable shape of the ν(OH) band of TFA···DEED10 complex under the ν(CD) bands.

24 ACS Paragon Plus Environment

Page 25 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 1. The optimized structures of complexes calculated at B3LYP 6 311++G(2df,2pd) level of theory: (A) the dimer of trifluoroacetic acid; (B) the TFA···DMK complex; (C) the TFA···DEE complex.

25 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 36

Figure 2. (A) The experimental spectra of trifluoroacetic acid at 6.5 Torr (a) and 3.2 Torr (b). The symbol M denotes the monomer bands. (B) The spectra of TFA monomer (d) and dimer (c) as the result of extraction from experimental spectra at different pressure.

26 ACS Paragon Plus Environment

Page 27 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 3. The spectra of a mixture of TFA (4.1 Torr) with DMKD6 at total pressure of 12.3 Torr (a) and the isolated spectrum of the TFA···DMKD6 complex (b).

27 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 36

Figure 4. The spectra of the monomer of trifluoroacetic acid (TFA); the dimer of trifluoroacetic acid (TFA···TFA); monomers of protiated (DMK) and deuterated (DMKD6) acetones; TFA···DMK and TFA···DMKD6 complexes; monomers of protiated (DEE) and deuterated (DEED10) diethyl ethers; TFA···DEE and TFA···DEED10 complexes in the region of ν(C=O) bands.

28 ACS Paragon Plus Environment

Page 29 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 5. The spectra of TFA···DMK and TFA···DMKD6 complexes, protiated acetone (DMK), deuterated acetone (DMKD6), TFA···DEE and TFA···DEED10 complexes, protiated diethyl ether (DEE), and deuterated diethyl ether (DEED10), in the range of ν(OH) bands of the corresponding complexes. The dashed lines (TFA···DEE, TFA···DEED10) is the most probable shape of the ν(OH) band of the TFA···DEE complex under the ν(CH) bands and the most probable shape of the ν(OH) band of TFA···DEED10 complex under the ν(CD) bands.

29 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 36

TOC Graphic

30 ACS Paragon Plus Environment

Page 31 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 1. The optimized structures of complexes calculated at B3LYP 6 311++G(2df,2pd) level of theory: (A) the dimer of trifluoroacetic acid; (B) the TFA···DMK complex; (C) the TFA···DEE complex. 82x123mm (300 x 300 DPI)

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2. (A) The experimental spectra of trifluoroacetic acid at 6.5 Torr (a) and 3.2 Torr (b). The symbol M denotes the monomer bands. (B) The spectra of TFA monomer (d) and dimer (c) as the result of extraction from experimental spectra at different pressure. 82x80mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 32 of 36

Page 33 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 3. The spectra of a mixture of TFA (4.1 Torr) with DMKD6 at total pressure of 12.3 Torr (a) and the isolated spectrum of the TFA···DMKD6 complex (b). 82x59mm (600 x 600 DPI)

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 4. The spectra of the monomer of trifluoroacetic acid (TFA); the dimer of trifluoroacetic acid (TFA···TFA); monomers of protiated (DMK) and deuterated (DMKD6) acetones; TFA···DMK and TFA···DMKD6 complexes; monomers of protiated (DEE) and deuterated (DEED10) diethyl ethers; TFA···DEE and TFA···DEED10 complexes in the region of ν(C=O) bands. 82x99mm (600 x 600 DPI)

ACS Paragon Plus Environment

Page 34 of 36

Page 35 of 36 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 5. The spectra of TFA···DMK and TFA···DMKD6 complexes, protiated acetone (DMK), deuterated acetone (DMKD6), TFA···DEE and TFA···DEED10 complexes, protiated diethyl ether (DEE), and deuterated diethyl ether (DEED10), in the range of ν(OH) bands of the corresponding complexes. The dashed lines (TFA···DEE, TFA···DEED10) is the most probable shape of the ν(OH) band of the TFA···DEE complex under the ν(CH) bands and the most probable shape of the ν(OH) band of TFA···DEED10 complex under the ν(CD) bands. 169x80mm (600 x 600 DPI)

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

TOC Graphic 82x44mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 36 of 36