Reactivity and Clustering between Carboxylic Acids and Esters

Feb 7, 2019 - Lorenzo Spada , Luca Evangelisti , Weixing Li , Ramona Orlacchio , and Walther Caminati. J. Phys. Chem. A , Just Accepted Manuscript...
0 downloads 0 Views 299KB Size
Subscriber access provided by WEBSTER UNIV

A: Spectroscopy, Molecular Structure, and Quantum Chemistry

Reactivity and Clustering between Carboxylic Acids and Esters: Rotational Study of Formic Acid – Isopropylformate Lorenzo Spada, Luca Evangelisti, Weixing Li, Ramona Orlacchio, and Walther Caminati J. Phys. Chem. A, Just Accepted Manuscript • DOI: 10.1021/acs.jpca.8b11294 • Publication Date (Web): 07 Feb 2019 Downloaded from http://pubs.acs.org on February 8, 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 18 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

Carboxylic Acids, Reactivity with Alcohols and Clustering with Esters: A Rotational Study of Formic Acid – Isopropylformate Lorenzo Spada, Luca Evangelisti, Weixing Li, Ramona Orlacchio, Walther Caminati* Dipartimento di Chimica “G. Ciamician” dell’Università, Via Selmi 2, I-40126 Bologna, Italy.

Corresponding Author *Walther Caminati, Email: [email protected]

ACS Paragon Plus Environment

1

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 18

ABSTRACT. The rotational spectrum of the 1:1 complex formic acid – isopropylformate (FAIPF) has been first observed when trying to assign the pulsed jet Fourier transform microwave (FTMW) spectrum of the adduct formic acid-isopropanol, by expanding a binary mixture of HCOOH and isopropanol in He. The strong FTMW of isopropylformate, formed by the esterification reaction, was observed instead. However, when HCCOH was in excess in the binary mixture, it was possible to observe and assign the rotational spectrum of FA-IPF. Later on a much intense spectrum of FA-IPF was obtained, when combining FA with IPF. Finally, the spectra of five isotopologues of the most stable conformer of formic acid - isopropylformate have been observed by means of rotational spectroscopy in supersonic expansion. Some of them, HCOOH-(CH3)2CHOOCD and HCOOH-(CH3)2CDOOCH have been synthesized in the MW cavity by using DCOOH or (CH3)2CDOH as precursors in the esterification process. In the observed isomer of the complex, the two subunits are linked to each other by a standard O-H···O and a weak C-H···O hydrogen bond. The dissociation energy has been estimated to be 34.1 kJ·mol-1.

ACS Paragon Plus Environment

2

Page 3 of 18 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 Formic acid (FA) is one of the most abundant carboxylic acids in the troposphere,1 and it is involved in several reactions taking place there.2-4 It has been observed that the dynamics of these reactions is changed when FA forms molecular adducts with other molecules.2 For this reason the investigation of the viability of FA to form molecular adducts with other molecules has been studied for a number of organic molecules. Probably, rotational spectroscopy in supersonic expansions is one of the best techniques to characterize molecular adducts and study thei energetics and structural features, in an environment free from solvent and matrix effects. Most of the complexes of FA investigated by FTMW spectroscopy are with other carboxylic acids,5-12 but rotational studies of the adducts of FA with water,13 carbon dioxide,14 formaldehyde,15 dimethylether,16 trimethylamine,17 pyridine,18 formamide,19 cyclobutanone20 and difluoromethane21 have been performed, showing that each of these complexes is stabilized by at least one hydrogen bond either of the O-H···O or of O-H···N type, resulting in estimated dissociation energies from 20 up to 60 kJ/mol. Up to now, no microwave data are available on the interactions between esters and carboxylic acids. The formers constitute a class of organic compounds which has been detected in the atmosphere22 and whose oxidation products generated by the reaction with Cl• and OH• have been reported recently for isopropylformate (IPF) and tert-butylformate23. In particular, apart from the formic acid concentration from other sources, the atmospheric oxidation process of IPF by OH• gives rise to a molar yield of 15-20% of FA among the other products.

ACS Paragon Plus Environment

3

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 18

Below we report the results of the investigation of the FA-IPF molecular complex by means of rotational spectroscopy in supersonic expansions, describing the hydrogen bond features linking the two subunits and providing chemical and energetic information on them. We have to outline, however, that we came to this investigation accidentally. When trying to assign the pulsed jet Fourier transform microwave (FTMW) spectrum of the adduct FAisopropanol by expanding a binary mixture of HCOOH and isopropanol in He, we discovered that isopropylformate was formed by the esterification reaction instead.24 However, when HCCOH was in excess in the binary mixture, it was possible to observe and assign the rotational spectrum of FA-IPF.

Experimental Section The rotational spectra of FA-IPF and of four monodeuterated isotopologues (HCOODMe2CHOCH=O, DCOOH- Me2CHOCH=O, HCOOH- Me2CDOCH=O and HCOOHMe2CHOCD=O) have been recorded using the COBRA-type25 pulsed supersonic-jet Fouriertransform microwave (FTMW) spectrometer,26 described elsewhere,27 working in the 6-18.5 GHz frequency region. Helium at stagnation pressure of 0.3 MPa was flown over IPF (Aldrich 98%, cooled at 273 K) and FA (Aldrich 98%), and expanded through the solenoid valve (General Valve, Series 9, nozzle diameter 0.5 mm) into the Fabry-Pérot-type cavity. The same procedure has been followed for the other isopologues whose respective monomers have been prepared either by direct H→D exchange of FA with D2O (obtaining HCOOD), by esterification of FA with Me2CDOH (commercial sample, giving Me2CDOCH=O) and of DCOOH with isopropanol (Me2CHOCD=O), or by using commercial sample (DCOOH, Aldrich, 95% wt).

ACS Paragon Plus Environment

4

Page 5 of 18 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 rest frequency of each transition is calculated as the arithmetic mean of the frequencies of the two Doppler components. The estimated accuracy of the frequency measurements is better than 3 kHz, resolution is better than 7 kHz.

Theoretical calculations Four isomers of FA-IPF, within 1000 cm-1, have been calculated at the MP2/6-311++G(d,p)28 level, using the GAUSSIAN09 package29, resulting from a search based on chemical intuition. Their shapes are shown at the top of Table 1. Three of them (I, III and IV) have the same seven member ring motif, containing two hydrogen bonds (O-H···O and C-H···O) connecting the two units. They differ from each other on the orientation of the isopropyl group. Isomer II displays a nine member ring, still containing the O-H···O and one C-H···O hydrogen bonds linking the two subunits. The hydrogen atom involved in the C-H···O weak hydrogen bonds belong the C atom bearing the two methyl groups. Also a weaker C-H···O interaction exists, involving one hydrogen of the methyl group closer to the CO group. In Table 1 we report the rotational constants and the electric dipole moment components. At the bottom of the Table, we list the relative energies, the relative energies including zero point corrections (harmonic approximation), the relative energies taking into account the BSSE corrections, and the dissociation energies (ED) for the four isomers. Table 1. MP2/6-311++G(d,p) spectroscopic parameters, equilibrium and zero point corrected energies of the four isomers of FA-IPF. I

A/MHz

4167

II

1984

III

1962

IV

3465

ACS Paragon Plus Environment

5

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 18

B/MHz 554 982 920 614 C/MHz 522 761 699 589 3.0 1.7 0.9 3.0 a|/D 0.4 0.2 2.9 0.4 |b|/D 0.1 0.7 0.1 0.0 |c|/D a -1 0 28 E/cm 574 832 b -1 0 9 E0 /cm 451 849 c -1 0 180 E0+BSSE /cm 588 892 ED/kJ·mol-1 27.4 25.3 20.4 16.8 aAbsolute energy = -496.319702 E . bAbsolute energy with zero point correction = -496.163677 h Eh. cAbsolute energy zero point and BSSE corrected = -496.160455 Eh. However, in the following section “Structural Information”, we used the MP2/maug-cc-pvtzdH31 geometry, which resulted more useful in performing a fit for a partial r0 structure for the assigned conformer.

Rotational spectrum The first spectral search was dedicated to the strong μa-type transitions of isomer (I), the most stable one according to the ab initio calculations. After the identification of the 808←707 which was observed at 8642.3 MHz (see Figure 1), other 45 μa-type R-band and 15 much weaker μband μc-type transitions have been measured.

ACS Paragon Plus Environment

6

Page 7 of 18 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 rotational transition 808 ← 707 for the most abundant isotopologue of the observed FA-IPF adduct. Later on, the rotational spectra of the single deuterated HCOOD- Me2CHOCH=O, DCOOHMe2CHOCH=O, HCOOH- Me2CDOCH=O and HCOOH- Me2CHOCD=O isotopologues have been collected and assigned. The respectively transitions have been fitted by using the Pickett’s SPFIT program,32 within the Ir-representation of Waltson’s S reduction,33 whose results are reported in Table 2. Relaxation processes upon supersonic expansion of the other isomers towards the most stable one take place when inter-conversion barriers are smaller than 2kT.34 This can explain the lack of the experimental detection of their rotational transitions.

Table 2. The experimental spectroscopic parameters of the parent and of the FA monodeuterated species of FA-IPF.

ACS Paragon Plus Environment

7

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 18

HCOOH-

HCOOD-

DCOOH-

HCOOH-

HCOOH-

(CH3)2CHOOCH

(CH3)2CHOOCH

(CH3)2CHOOCH

(CH3)2CHOOCD

(CH3)2CDOOCH

A/MHz

4174.24937(34)a

4141.5962(29)

4173.49669(70)

4119.0108(37)

4114.7670(11)

B/MHz

557.62958(11)

554.86463(12)

544.77041(12)

557.68340(15)

554.38589(14)

C/MHz

524.973222(85)

522.182867(97)

513.56410(11)

524.32535(13)

522.65806(15)

DJ/kHz

0.05231(246)

DJK/kHz

1.0156(83)

[0.05231]b

0.04908(43)

[0.05231]

0.05122(49)

[1.0156]

0.979(18)

[1.0156]

0.889(51)

d1/Hz

-3.58(18)

[-3.58]

[-3.58]

[-3.58]

[-3.58]

d2/Hz σ/kHzc

-1.51(14)

[-1.51]

[-1.51]

[-1.51]

[-1.51]

1.8

2.5

1.4

1.4

Nd

60

2.1

17

36

12

24

aErrors

in parentheses in units of the last digit. bData in brackets fixed to the corresponding values of the parent species. cRMS error of the fit. dNumber of fitted lines.

Structural Information The substitution coordinates (rs) of four hydrogen atoms (H3, H8, H18 and H19 in Figure 2) have been calculated according to the Kraitchman’s equations35 thanks to the availability of the five isotopologue sets of rotational constants.

ACS Paragon Plus Environment

8

Page 9 of 18 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 2. Sketch of the observed isomer of FA-IPF showing atom numbering and the principal axes of inertia. The obtained data are summarized in Table 3, and there compared with those resulting from MP2/maug-cc-pVTZ-dH36 ab initio calculations. This level of calculation, in contrast with the MP2/6-311++G(d,p) one, allows a satisfactory reproduction of the planar moment of inertia of the observed isotopologues for isomer I. Table 3. rs and re (MP2/maug-cc-pVTZ-dH) coordinates of the substituted atoms in the principal axis system of the parent species. a/Å H3 H8 H18 H19 aError

rs ±0.511*i ±1.9761(8) ±2.0864(7) ±4.6265(3)

re -0.5788 2.0039 -1.9815 -4.6280

b/Å rs ±1.203(1)a ±0.591(3) ±0.891(2) ±0.09(2)

c/Å re -1.2006 0.6816 0.8880 0.0088

rs ±0.422(4) ±1.188(1) ±0.408(4) ±0.12(1)

re 0.3952 -1.1095 -0.3948 0.0303

in parenthesis are in units of the last digit.

However, substitution coordinates of hydrogen atoms involved in hydrogen bond should be considered with caution, because of the inverse Ubbelohde effect, as discussed in ref.s [13-15, 17-19] makes these values not very reliable.

The Ubbelohde effect was first discussed by Ubbelohde and Gallagher,37 who observed that the H  D isotopic substitution of an hydrogen atom involved in a hydrogen bond increases the length of the hydrogen bond in a crystal. Such effect characterizes also the double hydrogen bond of the dimers of carboxylic acids in the gas phase (see, for example, Ref. [9]), and it is related to the double minimum potential associated with the concerted proton transfer taking place in these bimolecules. Vice versa, in molecular adducts made of two subunits held together by a single hydrogen bond (O-H···O, O-H···N, etc., see), a shrinkage of the O···O distance is observed. In this case the finding is called inverse Ubbelhode effect.38

ACS Paragon Plus Environment

9

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 18

This effect is evident for hydrogen H3, for which an imaginary value is obtained for the acoordinate. This is related to a very small value of this coordinate, and the mass effects of the H→D substitution are overwhelmed by the shortening of the C2···O17 distance. The shortening is proofed by a reduction of the planar moment of inertia Paa of -0.261 uÅ2 in going from the parent to the HCOOH-(CH3)2CHOOCD species, while the “rigid” model would imply an increase of 0.350 uÅ2. A decrease of ~2 mÅ of the C2···O17 distance would account for such a reduction of Paa. Such an effect is reasonably present also in the case of the O1-H18···O15 Hbond, but it is hidden by the high value of the a-coordinate of atom H18. The refinement of the observed isomer structure has been performed by fitting the O1-O15 distance (rO1O15) and the C2-O1-O15 angle (∠C2O1O15) of Figure 2, while keeping the remaining geometrical parameters at their ab initio values. The obtained results are reported in Table 4 where they are compared with the equilibrium values together with the derived hydrogen bond features.

Table 4. The re and r0 values of the fitted parameters together with the derived hydrogen bond ones. re rO1O15/Å 2.726 114.9 C2O1O15/° Hydrogen Bond Parameters rH18O1/Å 1.734 rH3O17/Å 2.386 O1H18O15/° 179.9 126.7 C2H3O17/°

r0 2.763(9) 113.2(6) 1.772 2.363 179.9 128.6

Bonding Energy of FA-IPF

ACS Paragon Plus Environment

10

Page 11 of 18 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

According to the pseudo diatomic approximation,39 since the stretching motion leading to the dissociation of the molecular complex is almost parallel to the a-axis, the force constants relative to that mode can be estimated as: ks  16π 4 ( D RCM ) 2 [4 B 4  4C 4  ( B  C ) 2  ( B  C ) 2 ] /(hDJ )

(1)

where μD is the pseudo diatomic reduced mass, RCM (= 4.78 Å) is the distance between the centers of the mass of the two subunits, and B, C and DJ are the spectroscopic parameters reported in Table 2. The resulting value of ks = 17.8 N m-1, which corresponds to a harmonic stretch fundamental ν = 100 cm-1. Within the Lennard-Jones type potential, the bonding energy of the complex at the zero point, can be estimated by applying the equation:40 2 EB  ks RCM / 72

(2)

The obtained bonding energy is EB=34.1 kJ·mol-1 which is close to the predicted one (27.4 kJ·mol-1) but larger than that one found in FA-formaldehyde15 (19.2 kJ·mol-1), a complex which has the same seven member ring motif, containing two hydrogen bonds (O-H···O and C-H···O), connecting the two units. The r0 structural features of the two hydrogen bonds of the two adducts are in Table 5. Table 5. The r0 values of the fitted parameters together with the derived hydrogen bond ones are compared to those of FA-formaldehyde. FA partner rO-H···O/Å IPF 1.772 formaldehyde15 1.745

rC-H···O/Å O-H···O 2.363 179.6 2.485 179.1

C-H···O 128.6 125.0

ED 34.1 19.2

Although, the two units are linked by the same hinge in both complexes, their dissociation energies calculated with the pseudodiatomic model results quite different from each other. We generally quote an uncertainty of 20% on these data. The model assumes the two

ACS Paragon Plus Environment

11

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 18

component molecules to be rigid, and the dissociation motion to take place along the a-axis. These two requirements seem to be better satisfied for FA-Formaldehyde. However, the value obtained for FA-IPF is in a satisfactory agreement with the ab initio value, suggesting that the electron donor properties of the two methyl groups of IPF can play a role on the dissociation energy.

Conclusions The rotational spectra of five isotopologues of FA-IPF prove that the two subunits are linked together by one classical O-H···O and one weak C-H···O hydrogen bonds, inserted in a seven membered ring: the same features found in the FA-Formaldehyde complex. A shortening of about 2 mÅ in the C2···O17 distance has been observed upon H→D substitution of the hydrogen atom contained in the weak C-H···O hydrogen bond (reverse Ubbelohde effect). The magnitude of the obtained dissociation energy (34.1 kJ·mol-1 or 27.4 kJ·mol-1 from pseudodiatomic approximation or at MP2/6-311++G(d,p) level calculations, respectively) suggests that if the concentration of IPF is not negligible, the formation of the FA-IPF complex could play a role in the atmospheric chemistry since its value is similar to that one found for a single FA-water complex (40.3 kJ·mol-1 calculated at the MP2 6-311++G(3df ,2p) level).41 In fact, by means of the formation of strong or relatively strong hydrogen bonded clusters, atmospheric reactions such as the H-abstraction of carboxylic proton in the formic acid + OH· reaction2 or in the formation of sulfuric acid by hydrolysis of SO3 42 are suggested to be catalyzed.

ACS Paragon Plus Environment

12

Page 13 of 18 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

ASSOCIATED CONTENT Supporting Information. 1) Completion of Reference 29. 2) Tables of the transition frequencies. 3) Table with the rotational constants and optimized geometry of the most stable species of FA-IPF. This material is available free of charge via the Internet at http://pubs.acs.org. Author Contributions The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. All authors contributed equally. ACKNOWLEDGMENT We thank the Italian MIUR (PRIN project 2010ERFKXL_001) and the University of Bologna (RFO) for financial support. L. E. was supported by a Marie Curie fellowship PIOF-GA-2012328405. W. L. thanks the China Scholarships Council (CSC) for financial support. REFERENCES 1)

Chebbi, A. and Carlier, P. Carboxylic Acids in the Troposphere, Occurrence, Sources, and Sinks: A Review. Atmospheric Environment 1996, 30, 4233-4249.

2)

Iuga, C.; Alvarez-Idaboy J. R.; Vivier-Bunge A. Mechanism and Kinetics of the WaterAssisted Formic Acid + OH Reaction under Tropospheric Conditions. J. Phys. Chem. A 2011, 115, 5138–5146.

3)

Jolly, G. S.; McKenney, D. J.; Singleton, D. L.; Paraskevopoulos, G.; Bossard, A. R. Rate Constant and Mechanism for the Reaction of Hydroxyl Radical with Formic Acid. J. Phys. Chem. 1986, 90, 6557-6562.

4)

Singleton, D. L.; Paraskevopoulos, G., Irwin, R. S.; Jolly, G. S.; McKenneys, D. J. Rate and Mechanism of the Reaction of Hydroxyl Radicals with Formic and Deuterated Formic Acids. J. Am. Chem. Soc. 1988, 110, 7786-7790.

ACS Paragon Plus Environment

13

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

5)

Page 14 of 18

Martinache, L.; Kresa, W.; Wegener, M.; Vonmont, U.; Bauder, A. Microwave Spectra and Partial Substitution Structure of Carboxylic Acid Bimolecules. Chem. Phys. 1990, 148, 129-140.

6)

Daly, A. M.; Douglass, K. O.; Sarkozy, L. C.; Neill, J. L.; Muckle, M. T.; Zaleski, D. P.; Pate, B. H.; Kukolich, S. G. Microwave Measurements of Proton Tunneling and Structural Parameters for the Propiolic Acid-Formic Acid Dimer. J. Chem. Phys. 2011, 135, 154304, 1-12.

7)

Tayler, M. C. D.; Ouyang, B.; Howard, B. J. Unraveling the Spectroscopy of Coupled Intramolecular Tunneling Modes: A Study of Double Proton Transfer in the Formic-Acetic Acid Complex. J. Chem. Phys. 2011, 134, 054316, 1-9.

8)

Evangelisti, L.; Ecija, P.; Cocinero, E. J.; Castaño, F.; Lesarri, A.; Caminati, W.; Meyer, R. Proton Tunneling in Heterodimers of Carboxylic Acids: A Rotational Study of the Benzoic Acid-Formic Acid Bi-Molecule. J. Phys. Chem. Lett. 2012, 3, 3770–3775.

9)

Feng, G.; Gou, Q.; Evangelisti, L.; Xia, Z.; Caminati, W. Conformational Equilibria in Carboxylic Acid Bimolecules: a Rotational Study of Acrylic Acid-Formic Acid. Phys. Chem. Chem. Phys. 2013, 15, 2917–2922.

10)

Gou, Q.; Feng, G.; Evangelisti, L.; Caminati, W. Conformers of Dimers of Carboxylic Acids in the Gas Phase: A Rotational Study of Difluoroacetic Acid-Formic Acid. Chem. Phys. Lett. 2014, 591, 301–305.

11)

Evangelisti, L.; Feng, G.; Gou, Q.; Caminati, W. The Rotational Spectrum of Formic Acid···Fluoroacetic Acid. J. Mol. Spectrosc. 2014, 299, 1-5.

12)

Li, W.; Evangelisti, L.; Gou, Q.; Caminati, W.; Meyer, R. The Barrier to Proton Transfer in the Dimer of Formic Acid: A Pure Rotational Study. Angew. Chem. Int. Ed. 2019, 58, 859865.

13)

Priem, D.; Ha, T.-K.; Bauder, A. Rotational Spectra and Structures of Three HydrogenBonded Complexes between Formic Acid and Water. J. Chem. Phys. 2000, 113, 169-175.

14)

Vigorito, A.; Gou, Q.; Calabrese, C.; Melandri, S.; Maris, A.; Caminati, W. How CO2 Interacts

with

Carboxylic

Acids:

A

Rotational

Study

of

Formic

Acid-CO2.

ChemPhysChem, 2015, 16, 2961-2967.

ACS Paragon Plus Environment

14

Page 15 of 18 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

15)

Gou, Q.; Favero, L. B.; Bahamyirou, S. S.; Xia, Z. N.; Caminati, W. Interactions between Carboxylic Acids and Aldehydes: A Rotational Study of HCOOH-CH2O. J. Phys. Chem. A 2014, 118, 10738-10741.

16)

Evangelisti, L.; Spada, L.; Li, W.; Ciurlini, A.; Grabow, J.-U.; Caminati, W. Shape of the Adduct Formic Acid − Dimethyl Ether: A Rotational Study. J. Phys. Chem. A 2016, 120, 2863−2867.

17)

Mackenzie, R. B.; Dewberry, C. T.; Leopold, K. R. The Trimethylamine-Formic Acid Complex: Microwave Characterization of a Prototype for Potential Precursors to Atmospheric Aerosol. J. Phys. Chem. A 2016, 120, 2268-2273.

18)

Spada, L.; Gou, Q.; Giuliano, B. M.; Caminati, W. Interactions between Carboxylic Acids and Heteroaromatics: A Rotational Study of Formic Acid−Pyridine. J. Phys. Chem. A 2016, 120, 5094–5098.

19)

Daly, A. M.; Sargus, B. A.; Kukolich, S. G. Microwave Spectrum and Structural Parameters for the Formamide-Formic Acid Dimer. J. Chem. Phys. 2010, 133, 174304.

20)

Evangelisti, L.; Spada, L.; Li, W.; Blanco, S.; Lopez, J. C.; Lesarri, A.; Grabow, J.-U.; Caminati, W. A Butterfly Motion of Formic Acid and Cyclobutanone in the 1:1 Hydrogen Bonded Molecular Cluster. Phys. Chem. Chem. Phys. 2017, 19, 204–209.

21)

Jin, Y.; Wang, J.; Gou, Q.; Xia, Z.; Feng, G. Shape and non-Bonding Interactions in the Formic Acid-Difluoromethane Complex by Rotational Spectroscopy, Spectroch. Acta A: Mol. Biomol. Spectrosc. 2019, 206, 185-189.

22)

Legreid, G.; Lööv, J. B.; Staehelin, J.; Hueglin, C.; Hill, M.; Buchmann, B.; Prevot, A. S. H.; Reimann, S. Oxygenated Volatile Organic Compounds (OVOCs) at an Urban Background Site in Zürich (Europe): Seasonal Variation and Source Allocation, Atmos. Environ. 2007, 41, 8409–8423.

23)

Pimentel, A. S.; Tyndall, G. S.; Orlando, J. J.; Hurley, M. D.; Wallington, T. J.; Sulbaek Andersen, M. P.; Marshall, P.; Dibble, T. S. Atmospheric Chemistry of Isopropyl Formate and tert-Butyl Formate. Int. J. Chem. Kinet. 2010, 42, 479–498.

ACS Paragon Plus Environment

15

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

24)

Page 16 of 18

Evangelisti, L.; Spada, L.; Li, W.; Vazart, F.; Barone, V.; Caminati, W. The Borderline between Reactivity and Pre-reactivity of Binary Mixtures of Gaseous Carboxylic Acids and Alcohols. Angew. Chem. Int. Ed. 2017, 56, 3872–3875.

25)

Balle, T. J.; Flygare, W. H. Fabry-Perot Cavity Pulsed Fourier Transform Microwave Spectrometer with a Pulsed Nozzle Particle Source. Rev. Sci. Instrum. 1981, 52, 33−45.

26)

Grabow, J.-U.; Stahl, W.; Dreizler, H. A Multioctave Coaxially Oriented Beam-Resonator Arrangement Fourier-Transform Microwave Spectrometer. Rev. Sci. Instrum. 1996, 67, 4072−4084.

27)

Caminati, W.; Millemaggi, A.; Alonso, J. L.; Lesarri, A.; Lopez, J. C.; Mata, S. Molecular Beam Fourier Transform Microwave Spectrum of the Dimethylether-Xenon Complex: Tunnelling Splitting and

131Xe

Quadrupole Coupling Constants. Chem. Phys. Lett. 2004,

392, 1−6. 28)

Møller, C.; Plesset, M. S. Note on an Approximation Treatment for Many-Electron Systems. Phys. Rev. 1934, 46, 618.

29)

Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; et al. Gaussian09 Revision D.01; Gaussian Inc.: Wallingford, CT, 2013.

30)

Boys, S. F.; Bernardi, F. The Calculations of Small Molecular Interactions by the Differences of Separate Total Energies. Some Procedures with Reduced Errors. Mol. Phys. 1970, 19, 553−566.

31)

Papajak, E.; Leverentz, H. R.; Zheng, J.; Truhlar, D. G. Efficient Diffuse Basis Sets: ccpVxZ+ and maug-cc-pVxZ. J. Chem. Theory Comput. 2009, 5, 1197-1202.

32)

Pickett, H. M. The Fitting and Prediction of Vibration-Rotation Spectra with Spin Interactions. J. Mol. Spectrosc. 1991, 148, 371−377.

33)

Watson, J. K. G. In Vibrational Spectra and Structure; Durig, J. R., Ed.; Elsevier: New York/Amsterdam, 1977; Vol. 6, pp 1−89.

34)

Ruoff, R. S.; Klots, T. D.; Emilson, T.; Gutowski, H. S. Relaxation of Conformers and Isomers in Seeded Supersonic Jets of Inert Gases. J. Chem. Phys. 1990, 93, 3142−3150.

ACS Paragon Plus Environment

16

Page 17 of 18 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

35)

Kraitchman, J. Determination of Molecular Structure from Microwave Spectroscopic Data. Am. J. Phys. 1953, 21, 17−25.

36)

Fornaro, T., Biczysko, M., Bloino, J., Barone, V. Reliable Vibrational Wavenumbers for C=O and N–H Stretchings of Isolated and Hydrogen-Bonded Nucleic Acid Bases, Phys. Chem. Chem. Phys. 2016, 18, 8479-8490.

37)

Ubbelohde A. R.; Gallagher, K. J. Acid–Base Effects in Hydrogen Bonds in Crystals.Acta Crystallogr. 1955, 8, 71–83.

38)

See, for example: Tang, S.; Majerz, I.; Caminati, W. Sizing the Ubbelohde effect: The Rotational Spectrum of tert-Butylalcohol Dimer. Phys. Chem. Chem. Phys. 2011, 13, 91379139.

39)

Millen, D. J. Determination of Stretching Force Constants of Weakly Bound Dimers from Centrifugal Distortion Constants. Can. J. Chem. 1985, 63, 1477−1479.

40)

Novick, S. E.; Harris, S. J.; Janda, K. C.; Klemperer, W. Structure and Bonding of KrClF: Intermolecular Force Fields in van der Waals Molecules. Can. J. Phys. 1975, 53, 2007−2015.

41)

Priem, D.; Ha, T.-K.; Bauder, A. Rotational Spectra and Structures of Three HydrogenBonded Complexes between Formic Acid and Water. J. Chem. Phys. 2000, 113, 169-175.

42)

Torrent-Sucarrat M., Francisco J. S., Anglada J. M. Sulfuric Acid as Autocatalyst in the Formation of Sulfuric Acid. J. Am. Chem. Soc. 2012, 134, 20632−20644.

ACS Paragon Plus Environment

17

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 18 of 18

TOC Graphics

ACS Paragon Plus Environment

18