PDF (3 MB)

Apr 25, 2019 - Centro de Investigação em Quнmica (CIQUP), Department of Chemistry and Biochemistry, Faculty of Science, University of. Porto, Rua do ...
0 downloads 0 Views 3MB Size
Article Cite This: J. Chem. Eng. Data XXXX, XXX, XXX−XXX

pubs.acs.org/jced

Thermodynamic Properties of Moldy-Musty Contaminants of Wine Ana R. R. P. Almeida,† Bruno D. A. Pinheiro,† Carlos F. R. A. C. Lima,† Ana Filipa L. O. M. Santos,‡ António C. S. Ferreira,‡ Filipe A. Almeida Paz,§ and Manuel J. S. Monte*,† †

Downloaded via BUFFALO STATE on July 30, 2019 at 07:13:51 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Centro de Investigaçaõ em Química (CIQUP), Department of Chemistry and Biochemistry, Faculty of Science, University of Porto, Rua do Campo Alegre, 687, P-4169-007 Porto, Portugal ‡ Cork Supply Portugal, SA, Rua Nova do Fial, no. 102, 4535-465 São Paio de Oleiros, Portugal § CICECO - Aveiro Institute of Materials, Department of Chemistry, University of Aveiro, Campus Universitário de Santiago, 3810-193 Aveiro, Portugal S Supporting Information *

ABSTRACT: This paper reports thermodynamic properties of phase transitions of 2,4,6-trichloro and 2,4,6-tribromo anisoles and of 2,4,6-tribromophenol. The vapor pressures of both crystalline and liquid phases (including supercooled liquid) of the three compounds were measured, respectively, in the temperature ranges T = (297.1 to 368.3) K, T = (330.7 to 391.7) K, and T = (336.5 to 401.7) K, using a static method based on capacitance diaphragm manometers. Moreover, the sublimation vapor pressures of 2,4,6tribromophenol were also measured in the temperature interval (307.2 to 329.2) K, using a Knudsen mass-loss effusion technique. The standard molar enthalpies, entropies, and Gibbs energies of sublimation and of vaporization, at reference temperatures, were derived from the experimental results as well as the (p,T) values of the triple point of each compound. The temperatures and molar enthalpies of fusion of the three benzene derivatives were determined using differential scanning calorimetry and were compared with the values derived indirectly from the vapor pressure measurements. The thermodynamic results were discussed together with the available literature data for 2,4,6-trichlorophenol. To help rationalize the phase behavior of these substances, the crystallographic structure of 2,4,6-tribromophenol was determined by single crystal X-ray diffraction. water)5,9 and may exist in cork, producing cork taint in the presence of fungal strains capable of converting TCP to TCA, through an o-biomethylation reaction.11 TCP is not the most influent compound to “musty” odors in wines, but as it is one of the major forerunners of TCA, it has an important function in producing cork taint. Other musty-smelling compounds, such as 2,4,6-tribromoanisole (TBA), were also identified as contributors to off-flavors in wines.9,10,12−15 TBA can be derived from its direct precursor, 2,4,6-tribromophenol (TBP), also via omethylation by bacterial microorganisms and usually comes from sources associated with the winery environment.12 Offflavor compounds, even in very low concentrations (parts per trillion), modify the organoleptic properties of wine, strongly compromising its quality. The methodologies that have been developed and applied in the cork industry to extract (vapor,16,17 supercritical fluids,18,19 and microwave radiation20) or to eliminate (γ radiation21) these compounds from cork are not

1. INTRODUCTION Cork is a natural material obtained from the outer bark of the cork oak Quercus suber and it is an exceptional biocomposite, showing a very specific combination of properties, such as elasticity, compressibility, low permeability of liquids, acoustic insulation, low thermal conductivity, low density, and significant chemical/microbial resistance.1−3 Therefore, cork products found application in several industrial sectors, such as construction, including floors, insulation, and covering. The wine industry, however, is the main target sector of cork products through cork stoppers manufacturing. Despite the technical advantages of using cork stoppers to seal wine bottles,2 corks can transfer taint to wine due to the aroma-intense compounds present in the cork. Nevertheless, there are also several faults during wine production or storage that can be the cause of such contamination, which is erroneously attributed to cork. This off-aroma, described as “moldy-musty,” is commonly related to the chloroanisole family of compounds. 2,4,6Trichloroanisole (TCA) is considered the major contributor to the sensory deviations related to cork due to its particularly low sensory threshold.4−10 Its phenolic precursor, 2,4,6trichlorophenol (TCP), can be produced from chlorine and phenol from several sources (such as cleaning products or town © XXXX American Chemical Society

Special Issue: Celebrating Our High Impact Authors Received: January 18, 2019 Accepted: April 25, 2019

A

DOI: 10.1021/acs.jced.9b00062 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 1. Source, Purity and Methods of Purification and Analysis of the Compounds Studied compound

CASNR

source

minimum initial puritya

2,4,6-trichloroanisole 2,4,6-tribromoanisole 2,4,6-tribromophenol

87-40-1 607-99-8 118-79-6

Sigma-Aldrich Sigma-Aldrich Sigma-Aldrich

0.999 0.999 0.998

purification method

final mass fraction purity

analysis methodb

sublimation under reduced pressure

0.9992 0.9998 0.9990

GC (FID) GC (FID) GC (FID)

As stated in the certificate of analysis from the manufacturer. Gas−liquid chromatography with flame ionization detector.

a

b

yet totally efficient. Despite the dire impact that these compounds have for the wine and cork industries worldwide, some of its thermodynamic properties remain unknown. In particular, the knowledge of accurate vapor pressures and of the related enthalpies of sublimation and vaporization is needed to understand the behavior of these compounds in cork and to improve processes for their removal. There are a few results of those thermodynamic properties for TCP reported in the literature that were determined from the dependence of vapor pressure with temperature.22,23 To the best of our knowledge, no experimental vapor pressures were reported previously for TCA, TBA, and TBP. Thus, this work reports their vapor pressures at different temperatures. The calorimetric study of the three compounds, using differential scanning calorimetry, was also performed in order to determine their enthalpy and temperature of fusion. The results were discussed focusing on understanding the phase equilibrium of these substances.

and data were corrected for absorption by the multiscan semiempirical method implemented in SADABS 2016/2.28 The structure was solved using the algorithm implemented in SHELXT-2014/5,29 which allowed the immediate location of almost all of the heaviest atoms composing the molecular unit. The remaining missing and misplaced non-hydrogen atoms were located from difference Fourier maps calculated from successive full-matrix least-squares refinement cycles on F2 using the latest SHELXL from the 2018/3 release.30 All structural refinements were performed using the graphical interface ShelXle.31 Hydrogen atoms bound to carbon were placed at their idealized positions using the HFIX 43 (aromatic carbon atoms) instruction in SHELXL. These hydrogen atoms were included in subsequent refinement cycles with isotropic thermal displacement parameters (Uiso) fixed at 1.2 × Ueq of the parent carbon atoms. The hydrogen atom associated with the O−H moiety was directly located from difference Fourier maps and included in the final structural model with the O−H distance restrained to (0.95 ± 0.01), and with the isotropic thermal displacements parameter (Uiso) fixed at 1.5 × Ueq of the parent oxygen atom. The last difference Fourier map synthesis showed the highest peak (0.398 eÅ−3) and the deepest hole (−0.433 eÅ−3) located at 1.37 and 1.08 Å from H1 and C3, respectively. The Flack parameter refined to −(0.04 ± 0.03).32 Structural drawings have been created using the software package Crystal Impact Diamond.33 Crystal Data for TBP. C6H3Br3O, M = 330.81, monoclinic, space group P21, Z = 2, a = (7.579 ± 0.002) Å, (1 Å = 10−10 m), b = (4.3822 ± 0.0012) Å, c = (12.293 ± 0.003) Å, β = (97.197 ± 0.006)°, V = (405.06 ± 0.19) Å3, μ(Mo Kα) = 14.864 mm−1, Dc = 2.712 g cm−3, colorless needle with crystal size of 0.28 × 0.10 × 0.07 mm3. Of a total of 2695 reflections collected, 1404 were independent (Rint = 0.0214). Final R1 = 0.0280 [I > 2σ(I)] and wR2 = 0.0794 (all data). Data completeness to theta = 25.22°, 99.0%. CCDC 1889759. 2.3. Thermal Analysis. The heat flux calorimeter NetzschGerätebau Thermal Analyzer, model DSC 200 F3 Maia, was used to check the absence of eventual phase transitions in the crystalline phase of the three compounds studied and to determine their temperatures (onset) and enthalpies of fusion. For each compound, at least four independent runs, using samples not previously melted and sealed in hermetic aluminum crucibles, were performed through two heating−cooling cycles. The samples were scanned at 2.0 K·min−1 from T = 253 K to a temperature (20 to 25) K higher than their temperature of fusion under a controlled nitrogen flux that was used to avoid contamination of the calorimeter if the hermiticity of the crucibles fails. The calorimeter was calibrated using seven highpurity reference materials.34−38 The degree of purity and the values of the enthalpy and temperature of fusion of the calibrant samples are reported in detail in Table S1. The standard uncertainties of the calibration results are u(T/K) = 0.12 and u(Δ1crHm ° /kJ·mol−1) = 0.22. Under the experimental conditions

2. EXPERIMENT 2.1. Materials. The three compounds studied experimentally in this work were commercially acquired with certified purity degrees of 0.999 for TCA, 0.999 for TBA, and 0.998 for TBP. Considering the high original purity of the two anisoles, they were studied without further purification while TBP was purified by sublimation under reduced pressure before the experimental study. The degree of purity of the original compounds and of the purified sample of TBP was checked by gas−liquid chromatography, using an Agilent chromatograph model 4890D, equipped with an HP-5 column and a flame ionization detector (FID). The samples were dissolved in dimethylformamide, and nitrogen was used as the carrier gas. Detailed information about the source and the purity features of the compounds is gathered in Table 1. 2.2. Single-Crystal X-ray Diffraction of TBP. In order to determine the crystalline structure, crystals of this compound were obtained through recrystallization from cyclohexane. TBP was dissolved in hot solvent under magnetic stirring, and the solution was left to evaporate in air over ∼1 week. Single crystals of TBP were manually harvested from the crystallization vials and immersed in highly viscous FOMBLIN Y perfluoropolyether vacuum oil (LVAC 140/13, Sigma-Aldrich) to avoid degradation caused by the evaporation of the solvent.24 Crystals were mounted on MiTeGen MicroLoops, typically with the help of a Stemi 2000 stereomicroscope equipped with Carl Zeiss lenses. All the following uncertainties reported in this section are standard uncertainties. X-ray diffraction data were collected at (150 ± 2) K on a Bruker X8 Kappa APEX II CCD area-detector diffractometer (Mo Kα graphite monochromated radiation, λ = 0.71073 Å) controlled by the APEX3 software package25 and equipped with an Oxford Cryosystems Series 700 cryostream monitored remotely using the software interface Cryopad.26 Diffraction images were processed using the software package SAINT+,27 B

DOI: 10.1021/acs.jced.9b00062 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

09 software package.44 The optimized geometries and respective electronic energies for the compounds studied were computed at the M06-2X/6-31+G(d,p)45 and B3LYP/6-31+G(d,p) levels of theory. Frequency calculations, at the same level, were performed for all speciesno imaginary frequencies were found, confirming that the structures correspond to true minima. All calculations were performed without symmetry restrictions. The one-dimensional (1D) hindered rotor model was applied for the treatment of −OH internal rotation (C−C− O−H dihedral angle) in phenol, 2-chlorophenol, 2-bromophenol, TCP, and TBP. The respective torsional potential energy profiles were obtained by computing Eel, using M06-2X/631+G(d,p), for a fixed dihedral angle and scanning from 0° to 180° in regular step intervals. The energies of each conformation were calculated relative to the respective absolute minimum and neglecting the contribution of the zero-point energy.

used, no solid−solid phase transitions or signals of decomposition were observed in the thermograms of the three compounds. Detailed results of each experimental run are presented in the Supporting Information (Table S2). The assigned standard uncertainties of the temperature and enthalpy of fusion were calculated through the root sum squared (RSS) method by combining the expanded uncertainties of the mean of the four experimental runs with the standard uncertainties referred to above assigned to the calibration results. 2.4. Vapor Pressure Measurements. The vapor pressures of both crystalline and liquid (including supercooled liquid) phases of the three halogenated benzenes were measured at different temperatures using a static apparatus, based on capacitance diaphragm gauges, which was tested and fully described before.39 Two capacitance MKS diaphragm gauges operate at self-controlled constant temperatures: gauge I, Baratron 631A01TBEH (Tgauge = 423 K), suitable for measuring pressures in the range (0.5 to 1.3 × 102) Pa, and gauge II, Baratron 631A11TBFP (Tgauge = 473 K), which can be used for measuring pressures in the range (3 to 1.3 × 103) Pa.40 The standard uncertainty of the temperature measurements was estimated to be u(T/K) = 0.01, and the expressions U(p/Pa) = 0.01 + 0.0050 (p/Pa) for gauge I and U(p/Pa) = 0.1 + 0.0050 (p/Pa) for gauge II describe the expanded uncertainties (0.95 confidence level, k = 2) assigned to the pressure measurements. Prior to the vapor pressure measurements, the samples were completely outgassed under reduced pressure, allowing the elimination of eventual traces of volatile impurities (including water). The sublimation vapor pressures of TBP were also measured over the range (0.1−1) Pa, using a Knudsen effusion apparatus that enables the operation of nine effusion cells simultaneously; the setup and procedure of this experimental effusion technique have been entirely described and tested before.41 The effusion cells, with different areas of the effusion orifices, are introduced into cylindrical holes inside three aluminum blocks, which are maintained at a constant temperature (different for the three blocks). The effusion orifices, made with platinum foil of (0.0125 ± 0.001) mm thickness, have the following areas:42 A0(A1) = A0(A2) = A0(A3) = (0.636 ± 0.004) mm2; A0(B1) = A0(B2) = A0(B3) = (0.785 ± 0.004) mm2; and A0(C1) = A0(C2) = A0(C3) = (0.985 ± 0.004) mm2, where the uncertainties were calculated by the RSS method. The Clausing factors of the effusion orifices were calculated as w0 = 1/{1 + (l/2r)}, where l is the thickness of the platinum foil and r is the radius of the orifices, yielding the results 0.986, 0.988, and 0.989 for the orifices of the series A, B, and C, respectively. The mass loss of each sample due to the effusion process, m, was determined by weighing the respective effusion cells, before and after the effusion period, t, with an estimated uncertainty of 1 × 10−5 g. At each temperature T, the vapor pressure p of the crystalline sample contained in each effusion cell was calculated using eq 1: p=

m ij 2πRT yz0.5 jj zz A 0w0t k M {

3. RESULTS AND DISCUSSION 3.1. Crystal Structure Description of TBP. TBP was studied in detail in the solid state using single-crystal X-ray diffraction. Figure 1 depicts the molecular unit composing the

Figure 1. Schematic representation of the molecular unit composing the crystal structure of TBP. Non-hydrogen atoms are represented as thermal ellipsoids drawn at the 50% probability level and hydrogen atoms as small spheres with arbitrary radii. The atomic labeling is provided for all non-hydrogen atoms.

asymmetric unit of the crystal structure. The compound was solved in the noncentrosymmetric monoclinic P21 space group. It is noteworthy that, though the molecule is achiral, it packs in the solid state in a chiral space group. Nevertheless, the refined Flack parameter cannot unequivocally allow the conclusion that the crystal studied was, indeed, an enantiomeric pure arrangement of such molecules in the solid state. The molecular unit is rich in both donor and acceptor atoms capable of engaging in strong and directional supramolecular interactions in the solid state. The most striking one concerns the −OH group, which promotes direct connections (along the [010] direction) between adjacent molecular units leading to the formation of a C11(2) graph set motif.46 It is noteworthy that these connections are strong [dO···O = (2.832 ± 0.007) Å] and relatively directional with the ∠(OHO) interaction angle being 149° (Figure 2a). This supramolecular chain is further strengthened by the presence of weak C−H···Br contacts as shown in Figure 2b, with a dC···Br distance of (3.941 ± 0.009) Å with an interaction angle ∠(CHBr) of 169°. The combination of such interactions forms in the crystal structure a supramolecular layer placed on the bc plane of the unit cell. Interactions between such layers are further ensured by the presence of a third type of connection: Br···Br connections (dashed blue lines in Figure 2c) are markedly present in the crystal structure connecting Br1 and

(1)

where M is the molar mass of the effusing vapor, and R is the molar gas constant (8.3144598 J·K−1·mol−1 43). The standard uncertainties of the vapor pressure and temperature measurements were estimated as u(p/Pa) = 0.02 and u(T/K) = 0.01. 2.5. Computational Chemistry Calculations. All quantum chemical calculations were performed using the Gaussian C

DOI: 10.1021/acs.jced.9b00062 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 3. Schematic representation of the crystal packing of TBP viewed in perspective along the [010] direction of the unit cell. O−H··· O hydrogen bonds (dashed green lines) and weak C−H···Br (dashed gold lines) and Br···Br interactions (dashed blue lines) are also represented to emphasize the vast network of supramolecular contacts present in the compound.

Figure 2. Schematic representation of the most relevant supramolecular interactions present in the crystal structure of TBP. (a) O− H···O hydrogen bonds connecting adjacent molecular units and forming a one-dimensional supramolecular tape described by C11(2) graph set motif running parallel to the a axis of the unit cell. (b) Weak C−H···Br interactions between adjacent molecular units. (c) Weak Br···Br and Br···π interactions. Symmetry transformations pertaining to symmetry-related atoms have been omitted for clarity.

in this table. Vapor pessures of each compound at other temperatures may be calculated using the reported value of ΔgcdC°p,m and the values of any of the three sets of parameters ΔgcdG°m(θ) and ΔgcdH°m(θ).

Br2 with an interatomic distance of (3.6949 ± 0.0015) Å. It is also important to emphasize that the crystal structure of the compound (Figure 3) is also supported by the existence of additional Br···π contacts (gray dashed columns in Figure 2c) involving Br2 and the aromatic ring of the neighboring molecule [dBr···π = (3.545 ± 0.004) Å]. 3.2. Thermodynamic Properties of Sublimation and of Vaporization. 3.2.1. Standard Gibbs Energy and Enthalpy. Table 2 reports the sublimation and vaporization vapor pressures of TCA, TBA, and TBP. The experimental data of the liquid and crystalline vapor pressures were independently fit by the truncated form of the Clarke and Glew equation, eq 2:47 ij p yz Δg G ° (θ ) 1y i1 g R lnjjjj ° zzzz = − cd m Hm° (θ )jjj − zzz + Δcd θ θ T k { kp { ÄÅ ÉÑ Å Ñ i T yÑ Åi θ y g o + Δcd C p ,m(θ )ÅÅÅjjj zzz − 1 + lnjjj zzzÑÑÑ ÅÅÇk T { k θ {ÑÑÖ

g g Δcd Hm° (θ ) − Δcd Gm° (θ ) (3) θ When accurate experimental crystalline or liquid vapor pressures are available over a large temperature interval (>50 K), the fit of eq 2 to the experimental data often yields reliable values of ΔgcdC°p,m(θ). The values of Δgl C°p,m(θ) of the three compounds studied were derived from those fittings to the (p,T) liquid results. The value of ΔgcrCp,m ° (θ) of TBP was derived from the fitting of eq 2 to the combined experimental Knudsen and static (p,T) crystalline data, measured in this work. Due to the short temperature ranges of the sublimation experiments, the values of ΔgcrCp,m ° (θ) for TCA and TBA could not be derived using that procedure. Instead, they were calculated as ΔgcrCp,m ° (θ) = Cp,m ° (g) − Cp,m ° (cr), where Cp,m ° (g) and Cp,m ° (cr) are, respectively, the gas and crystalline isobaric molar heat capacities. The values of Cp,m ° (g) were determined at the temperature 298.15 K for the compounds studied in this work, and for TCP. These results were derived from statistical thermodynamics, computed by means of the Gaussian 09 software package,44 using the vibrational frequencies from B3LYP/6-31+G(d,p) calculations (scaled by a factor of 0.964248): the calculated values, in J·K−1·mol−1, are (171.9 ± 5.2) for TCA, (171.5 ± 5.1) for TBA, (147.7 ± 4.4) for TBP, and (150.8 ± 4.5) for TCP. The standard uncertainties in Cp,m ° (g) were estimated as u[Cp,m ° (g)] = 0.03·C°p,m(g), according to the study of Č ervinka et al.49 on the use of the rigid-rotor harmonic oscillator (RRHO) model for molecules containing symmetrical internal rotors, instead of the more appropriated one-dimension hindered rotor (1-DHR). g ° Δcd Sm(θ ) =

(2)

where p is the vapor pressure at the temperature T, p° is a selected reference pressure (p° = 105 Pa in this work), θ is a reference temperature (in this work, unless stated otherwise, θ = 298.15 K), and R is the gas constant. Differences in the standard Gibbs energy, enthalpy, entropy, and heat capacity at constant pressure between the gaseous and one of the condensed phases g g g are represented, respectively, by Δ cd G m° , Δ cd H m° , Δ cd S m° g (calculated through eq 3) and ΔcdCp,m ° . Table 3 reports, for each studied compound, the values of these properties and their related uncertainties for three different temperatures (θ = mean temperature of the experiments, θ = temperature of the triple point, and θ = 298.15 K). The vapor pressure results calculated from eq 2 for the three different temperatures are also reported D

DOI: 10.1021/acs.jced.9b00062 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 2. Vapor Pressure Resultsa T/K

p/Pa

100Δp/pb

T/K

100Δp/pb

p/Pa

297.11 300.13 303.10 306.06

2.39 3.37 4.62 6.49

−0.1 0.2 −1.0 1.0

309.01 311.98 315.20 317.95

TCA crystalline phase (static method) 8.77 −0.2 11.99 0.3 16.42 −1.0 21.89 0.4

319.65 322.82 325.77 328.77 331.73 334.71

35.51c 43.91c 53.84c 66.11c 80.60c 97.66

0.8 −0.4 −0.4 −0.2 0.0 −0.2

337.68 340.63 343.63 346.05 348.60 351.58

liquid phase (static method) 118.0 −0.2 142.4d 0.3 171.2d 0.3 197.1d −0.1 229.2d 0.0 272.4d 0.0

330.66 332.63 334.50 336.63 338.58

2.84 3.49 4.20 5.10 6.23

0.0 0.7 0.5 −1.1 0.0

336.57 338.61 340.57 342.54 344.50 346.48 348.50 350.43 352.41 354.42

8.89c 10.33c 11.73c 13.69c 15.49c 17.86c 20.61c 23.45c 26.97c 30.65c

0.6 0.6 −0.9 0.5 −1.0 −0.5 −0.1 −0.2 0.5 0.0

307.16 309.36 311.20 313.15

0.107 0.141 0.177 0.221

−0.8 0.0 0.7 −0.3

TBA crystalline phase (static method) 340.54 7.47 −0.7 342.55 9.15 0.5 344.50 10.86 −0.6 346.47 13.17 0.4 348.43 15.63 −0.4 liquid phase (static method) 356.36 34.85c 0.3 358.31 39.46c 0.1 360.25 44.72c 0.3 362.27 50.49 −0.2 364.24 56.78 −0.6 366.22 64.59 0.2 368.15 72.57d 0.2 370.19 81.97d 0.2 372.11 91.83d 0.3 374.05 102.3d −0.2 TBP crystalline phase (Knudsen effusion method)e 315.37 0.286 −0.4 317.19 0.358 1.0 319.15 0.440 −0.7 321.33 0.570 0.7

336.49 338.49 340.48 342.48

2.84 3.43 4.23 5.13

0.3 −0.9 0.3 0.0

344.42 346.35 348.35 350.35

344.43 346.53 348.28 350.37 352.36 354.35 356.31 358.20 360.17 362.20

10.29c 12.00c 13.67c 15.82c 18.13c 20.79c 23.67c 26.89c 30.85c 34.91c

0.3 0.3 0.6 0.3 −0.1 −0.3 −0.8 −0.9 −0.3 −1.4

364.19 366.13 368.11 370.06 372.03 374.17 376.12 378.03 380.00 381.93

crystalline phase (static method)f 6.12 −1.1 7.43 −0.2 8.94 −0.6 10.85 0.1 liquid phase (static method) 40.45c 0.3 46.05c 0.8 52.07 0.4 58.65 0.1 65.91 −0.5 75.45 −0.1 85.37d 0.4 95.95d 0.7 107.1d 0.1 119.0d −0.7

T/K

p/Pa

100Δp/pb

320.87 323.91 326.89 329.85

28.95 39.21 51.74 66.47

−0.2 1.0 0.8 −1.4 ⟨100Δp/p⟩ = ± 0.6

354.51 356.47 359.37 362.38 365.32 368.31

322.5d 359.3d 421.4d 494.3d 580.1d 672.9d

0.2 0.1 0.1 −0.2 0.3 −0.4 ⟨100Δp/p⟩ = ± 0.2

350.42 352.40 354.32 356.37

18.81 22.40 26.60 31.75

0.2 0.0 0.3 0.1 ⟨100Δp/p⟩ = ± 0.4

376.07 377.95 379.95 381.85 383.82 385.79 387.78 389.79 391.72

115.2d 127.7d 142.6d 159.5d 177.0d 195.8d 220.3d 243.6d 267.6d

0.1 −0.2 −0.4 0.4 0.0 −0.5 0.7 0.2 −0.5 ⟨100Δp/p⟩ = ± 0.4

323.20 325.15 327.33 329.21

352.41 354.40

0.700 0.857 1.100 1.330

13.16 15.84

0.5 −0.6 0.8 −0.2 ⟨100Δp/p⟩ = ± 0.5 0.4 0.8 ⟨100Δp/p⟩ = ± 0.5

383.93 385.88 387.86 389.84 391.79 393.77 395.73 397.68 399.69 401.66

134.8d 151.3d 168.2d 187.6d 209.4d 233.9d 259.0d 284.0d 316.4d 351.7d

0.3 0.7 0.1 0.0 0.3 0.6 0.3 −0.8 −0.6 −0.2 ⟨100Δp/p⟩ = ± 0.4

a The standard uncertainty of the temperature is u(T/K) = 0.01 and the expanded uncertainties (0.95 confidence level, k = 2) of the vapor pressures are U(p/Pa) = 0.01 + 0.0050 (p/Pa) for static pressures measured through gauge 1 and U(p/Pa) = 0.1 + 0.0050 (p/Pa) for static pressures measured through gauge 2. For the effusion pressures, u(p/Pa) = 0.02. bΔp = p − pcalcd where pcalcd is calculated from the Clarke and Glew eq 2 with parameters given in Table 3. cVapor pressures of the supercooled liquid. dMeasured using gauge 2. eThe reported effusion pressures are the

E

DOI: 10.1021/acs.jced.9b00062 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 2. continued mean of the values obtained using the small, medium, and large effusion orifices. fThe sample of TBP ran out during the static measurements, not allowing to determine accurately the (p,T) data between 354.40 K and the temperature of the triple point.

Table 3. Standard (p° = 105 Pa) Molar Properties of Sublimation and of Vaporization of TCA, TBA, TBP, and TCP ΔT/K

297.1 to 329.9

319.7 to 368.3

330.7 to 356.4

336.6 to 391.7

307.2 to 354.4

344.4 to 401.7

299.4 to 340.3

342.8 to 462.7

θ/K

° (θ)a/kJ· ΔgcdGm mol−1

313.48e 332.10f 298.15

23.14 ± 0.02 19.60 ± 0.02 26.09 ± 0.02

343.98e 332.10f 298.15

18.17 ± 0.02 19.60 ± 0.02 23.84 ± 0.04

343.52e 360.48f 298.15

26.32 ± 0.02 23.09 ± 0.02 35.12 ± 0.10

364.15e 360.48f 298.15

22.63 ± 0.02 23.09 ± 0.02 31.25 ± 0.10

330.78e 368.22f 298.15

30.42 ± 0.02 23.14 ± 0.06 36.89 ± 0.04

373.05e 368.22f 298.15

22.51 ± 0.02 23.14 ± 0.02 32.78 ± 0.14

319.85e 338.55f 298.15

23.58 ± 0.02 20.19 ± 0.04 27.54 ± 0.04

402.76e 338.55f 298.15

12.57 ± 0.04 20.19 ± 0.04 25.53 ± 0.20

p(θ)/Pab

ΔgcdHm ° (θ)a/kJ· mol−1

ΔgcdSm ° (θ)c/J·K−1· mol−1

R2

TCA crystalline phase (static method), this work 13.9 83.1 ± 0.4 191.3 ± 1.3 82.6 82.4 ± 0.6 189.1 ± 1.8 2.69 83.6 ± 0.6 192.9 ± 2.0 0.9999 liquid phaseh (static method), this work 1.74 × 102 59.4 ± 0.2 119.9 ± 0.6 82.6 60.0 ± 0.2 121.6 ± 0.6 6.66 62.0 ± 0.8 128.0 ± 2.7 1.0000 TBA crystalline phase (static method), this work 9.95 92.0 ± 0.2 191.2 ± 0.6 45.1 91.2 ± 0.6 188.9 ± 1.7 7.0 × 10−2 94.0 ± 1.6 197.5 ± 5.4 1.0000 liquid phaseh (static method), this work 56.7 67.8 ± 0.2 124.0 ± 0.6 45.1 68.0 ± 0.2 124.6 ± 0.6 3.4 × 10−1 72.3 ± 1.0 137.7 ± 3.4 1.0000 TBP crystalline phase (Knudsen effusion and static methods), this work 1.57 95.4 ± 0.2 196.4 ± 0.6 52.2 94.0 ± 1.2 192.4 ± 3.3 3.4 × 10−2 96.6 ± 1.0 200.3 ± 3.4 1.0000 liquid phaseh (static method), this work 70.5 71.0 ± 0.2 130.0 ± 0.5 52.2 71.3 ± 0.2 130.8 ± 0.5 1.8 × 10−1 76.0 ± 1.4 145.0 ± 4.7 1.0000 TCPj crystalline phase (static22 and transpiration23 methods) 14.1 81.7 ± 0.6 181.7 ± 1.9 76.7 81.2 ± 0.8 180.2 ± 2.4 1.50 82.2 ± 0.8 183.3 ± 2.7 0.9997 liquid phase (static22 and tranpiration23 methods) 2.3 × 103 57.3 ± 0.4 111.1 ± 1.0 76.7 63.1 ± 1.0 126.8 ± 3.0 3.37 66.7 ± 1.6 138.1 ± 5.4 0.9998

a

−ΔgcdCp,m ° (θ)/J·K−1· mol−1

σd/Pa

σrd

36.8 ± 17.8g

0.35

0.83%

57.5 ± 7.6i

0.89

0.32%

45.3 ± 17.7g

0.05

0.53%

68.2 ± 7.7i

0.48

0.46%

36.7 ± 14.4i

0.04

0.66%

66.9 ± 9.2i

0.76

0.57%

25.5 ± 17.7g

0.47

1.9%

89.5 ± 8.7i

97.6

2.2%

b

Uncertainties are expressed as the expanded uncertainty (0.95 level of confidence, k = 2). Calculated from eq 2 for three different temperatures (θ = mean temperature of the experiments, θ = temperature of the triple point, and θ = 298.15 K). cCalculated using eq 3; uncertainties calculated through the RSS method. dσ is the standard deviation of the fit, defined as σ = [∑ni=1(p − pcalcd)i2/(n − m)]1/2 where n is the number of experimental points used in the fit and m is the number of adjustable parameters of Clarke and Glew, eq 2. σr is the relative standard deviation of the fit, defined as σr = [∑ni=1(ln p − ln pcalcd)i2/(n − m)]1/2. eMean temperature. fTemperature of triple point gEstimated value. hIncluding supercooled liquid. iAdjustable parameter derived from the fittings of eq 2 to the (p,T) data. Uncertainties are standard deviations of the leastsquares regressions. jThe thermodynamic properties for TCP were derived, in this work, from the fitting of eq 2 to the crystalline and liquid (p,T) data reported in the literature.22,23

The values of Cp,m ° (cr, 298.15 K) for TCA and TBA were estimated, respectively, as C°p,m = (208.7 ± 17.0) J·K−1·mol−1 and C°p,m = (216.8 ± 17.0) J·K−1·mol−1, using the group additivity method, proposed by Acree and Chickos.50 The selected values of Cp,m ° (g), Cp,m ° (cr), and ΔgcrCp,m ° , and the respective uncertainties, are reported in bold in Table S4 (in the Supporting Information), together with those estimated using different methods.51−53 As the estimated values of C°p,m(cr) are all referred to the temperature of 298.15 K, we assumed that Δ crg C p,m ° is approximately constant inside the assigned uncertainties.

The Clarke and Glew equation (eq 2) can be replaced by eq 4, which may be more appealing for calculating vapor pressures as a function of temperature. ln(p /Pa) = A + B /(T /K) + C ln(T /K)

(4)

For T = 298.15 K and p° = 105 Pa, the parameters A, B, and C (defined in the Supporting Information) were calculated for the crystalline and liquid vapor pressure equations of the three halogenated benzenes and are presented in Table 4. The vapor equilibrium concentrations (mol·m−3), in a closed system, may F

DOI: 10.1021/acs.jced.9b00062 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 4. Equations ln(p/Pa) = A + B/(T/K) + C ln(T/K) for the Crystalline and Liquid Phases of TCA, TBA, and TBP for T = 298.15 K TCA crystalline liquid

crystalline liquid

crystalline liquid

ln[(p ± 0.008)/Pa] = 64.356 − [11374.4/(T/K)] − [4.426· ln(T/K)] ln[(p ± 0.016)/Pa] = 73.225 − [9518.8/(T/K)] − [6.916· (ln(T/K)] TBA ln[(p ± 0.040)/Pa] = 71.756 − [12930.0/(T/K)] − [5.448· (ln(T/K)] ln[(p ± 0.040)/Pa] = 83.010 − [11141.3/(T/K)] − [8.203· (ln(T/K)] TBP ln[(p ± 0.016)/Pa] = 65.163 − [12934.4/(T/K)] − [4.414· ln(T/K)] ln[(p ± 0.056)/Pa] = 82.838 − [11539.7/(T/K)] − [8.046· ln(T/K)]

Figure 5. Phase diagram of TBP. ○, vaporization; ●, vaporization (supercooled liquid); ■, sublimation (static method); ▲, sublimation (Knudsen effusion method). Triple point data determined in this work: T = 368.2 K; p = 52.2 Pa. Phase diagram of TCP:22,23 ○, vaporization (transpiration method23); ○, vaporization (static method22); Δ, sublimation (transpiration method23);▲, sublimation (static method22); ×, result related to sublimation according to Verevkin et al.23 Triple point data determined in this work using the reported vapor pressures: T = 338.6 K; p = 76.7 Pa.

be calculated using eq 5, which was derived combining eq 4 with the ideal-gas equation. ln(c /mol m−3) = A + B /(T /K) + C ln(T /K) − ln(RT /J ·mol−1)

(5)

Figures 4 and 5 show phase diagrams in the neighborhood of the triple point for the three compounds experimentally studied

enthalpies of sublimation and of vaporization for TCP, derived from those fittings at different temperatures, are reported in Table 3, together with the values determined in this work for the other three compounds. The relative deviations of the experimental vapor pressures from the values calculated using eq 2 (for the three compounds studied in this work) are plotted in Figure 6. Allot and co-workers54 also measured the standard molar enthalpy of sublimation of TBP using Calvet microcalorimetry. The result reported by these authors, ΔgcrHm ° (298.15 K) = (97.6 ± 1.1) kJ·mol−1,54 is in excellent agreement with the results derived in the present work.

Figure 4. Phase diagrams of TCA and TBA. TCA: ○, vaporization; ●, vaporization (supercooled liquid); ■, sublimation (static method). Triple point data determined in this work: T = 332.1 K; p = 82.6 Pa. TBA: ○, vaporization; ●, vaporization (supercooled liquid); ■, sublimation (static method). Triple point data determined in this work: T = 360.5 K; p = 45.1 Pa.

in this work. The temperature and pressure at each triple point were calculated from the intersection of the sublimation and vaporization vapor pressure equations. In addition, vapor pressures reported in the literature for solid and liquid 2,4,6trichlorophenol (TCP)22,23 are included in Figure 5. The literature results were fit together in this work by eq 2, enabling us to determine the temperature of the triple point (Ttr = 338.55 K) and to build the phase diagram for this compound. At this temperature, the vapor pressure result reported by Verevkin et al.23 falls in the crystalline phase, and so it was not considered in both crystalline and liquid (p,T) data fittings by the Clarke and Glew equation. The standard molar Gibbs energies and

Figure 6. Percentage of deviation of experimental vapor pressures, p, from the calculated values (eq 2), pcalcd, for TCA, TBA, and TBP. ○, sublimation vapor pressures of TCA; ●, vaporization vapor pressures of TCA. □, sublimation vapor pressures of TBA; ■, vaporization vapor pressures of TBA. Δ, sublimation vapor pressures of TBP; ▲, vaporization vapor pressures of TBP. G

DOI: 10.1021/acs.jced.9b00062 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 7. Increment scheme of the standard molar enthalpies and Gibbs energies of vaporization for TCA, TBA, TCP and TBP. All values in kJ·mol−1. a Derived in this work from the fitting of eq 2 to the liquid (p,T) results reported in the literature.22,23

Figure 8. Increment scheme of the standard molar enthalpies and Gibbs energies of sublimation of TCA, TBA, TCP and TBP. All values in kJ·mol−1. a Derived in this work from the fitting of eq 2 to the crystalline (p,T) results reported in the literature.22,23

Figure 7 describes schematically the increments of the standard Gibbs energy and enthalpy of vaporization for TCA, TBA, TCP, and TBP, at T = 298.15 K. For what follows concerning the discussion on the increments of the thermody-

namic properties, we omit the units of the values that are all in units of kJ·mol−1. This scheme highlights the excellent agreement of the increments for both properties (Δgl G°m and ° ) from the chlorinated (TCP) to the brominated (TBP) Δgl Hm H

DOI: 10.1021/acs.jced.9b00062 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

phenols {Δ(Δgl G°m) = (7.25 ± 0.24) and Δ(Δgl H°m) = (9.3 ± 2.1)}, with those from the chlorinated (TCA) to the brominated ° ) = (7.41 ± 0.11) and Δ(Δgl Hm °) = (TBA) anisoles {Δ(Δgl Gm (10.3 ± 1.3)}. Similar agreement is also observed when going from TCP to TCA {Δ(Δgl G°m) = −(1.69 ± 0.20) and Δ(Δgl H°m) = −(4.7 ± 1.8)} and from TBP to TBA {Δ(Δgl G°m) = −(1.53 ± 0.09) and Δ(Δgl Hm ° ) = −(3.7 ± 0.9)}. Increments in the enthalpy and Gibbs energy of vaporization for the parent molecules phenol and anisole show an increase of 10.8 and 5.1, respectively, when going from anisole (Δgl H°m = (46.6 ± 0.2)55 and Δgl G°m = 13.355) to phenol (Δgl H°m = 57.456 and Δgl G°m = 18.456). These differences are probably due to the presence of intermolecular hydrogen bonds in the liquid phase of phenol. However, in the case of the compounds studied in this work, the averages of these increases are only (4.2 ± 2.5) and (1.61 ± 0.26) for Δgl Hm ° and Δgl Gm ° , respectively, suggesting that the ortho-halogen atoms have some inhibition effect concerning the formation of such intermolecular H bonds. On the other hand, the increment per halogen atom observed in the enthalpy of ° = (41.2 vaporization when changing from chlorobenzene (Δgl Hm ° = (44.3 ± 0.4)56,58) is 3.1. ± 0.3)56,57) to bromobenzene (Δgl Hm This compares well with the increment of (9.3/3) = 3.1 on changing from TCP to TBP and with (10.3/3) = 3.4 related to the change TCA to TBA. Similar agreement is found concerning the standard Gibbs energy of vaporization: when going from chlorobenzene (Δgl G°m= 10.356,57) to bromobenzene (Δgl G°m= 12.956,58). The increment is 2.6, which is similar to the increments per halogen atom of (7.25/3) = 2.4 and (7.41/3) = 2.5, associated with the changes on going from TCP to TBP and TCA to TBA, respectively. In summary, this incremental analysis gives strong indication that the bulkier TCP and TBP are unable to establish intermolecular H-bonds in the liquid phase as efficiently as phenol does. Analogous calculations were performed on the sublimation results, at T = 298.15 K (Figure 8). But while the Gibbs energy increments related to the several pairs of compounds referred to above are comparable, the following enthalpic increments show some discrepancies between them: from TCP to TBP {Δ(ΔcrgHm° ) = (14.4 ± 1.3)} while from TCA to TBA {Δ(ΔgcrH°m) = (10.4 ± 1.7)}, a discrepancy of −(4.0 ± 2.1). A similar discrepancy is observed when comparing the sublimation enthalpy increment related to the change of TCP to TCA with the one related with the change from TBP to TBA. These observations indicate that either the enthalpy of sublimation of TCP is 4 kJ·mol−1 larger than expected (as it would be predicted by simple group additivity considerations) or the enthalpy of sublimation of TBP is 4 kJ·mol−1 smaller. The crystalline structural data reported in the literature for TCP59 indicates the existence of both intramolecular and intermolecular OH···Cl contacts (Figure 9). This intramolecular contact is probably weakening the intermolecular interactions between TCP molecules (note that dO···Cl is relatively large), thus diminishing the enthalpy of sublimation of this compound. To check what happens on intermolecular forces between TBP molecules, information on the crystal lattice of TBP could be helpful to rationalize the higher value of its enthalpy of sublimation. But, to the best of our knowledge, no crystal structure has been reported for this phenolic compound. The existence of intermolecular hydrogen bonds OH···O, however, was previously suggested by Faniran and Shurvell who recorded the Raman and IR spectra of crystalline TCP and TBP.60 The authors commented that, with the exception of the OH stretching region and of the C−X stretching bands, the IR

Figure 9. Schematic representation of the O−H···Cl interactions in the crystal structure of TCP (adapted from the ref 59).

spectra of the two compounds are identical and conclude that there is evidence of intermolecular hydrogen bonding OH···O in the solid state of TBP, while there are no indications of such interaction in TCP. To corroborate this information, we have decided to perform a single crystal X-ray diffraction determination of TBP. According to its crystalline structure, schematically represented in Figure 2, it is possible to observe the existence of evident intermolecular OH···O hydrogen bonds, which may justify its higher enthalpy of sublimation. 3.2.2. Standard Molar Entropy of Phase Transitions. As shown in Figure 8, the increments of ΔgcrG°m from TCP to TBP are in a good agreement with those related to the change from TCA to TBA. Likewise, the increments from TCP to TCA are close to those from TBP to TBA. Since the ΔgcrG°m follows an additive trend, the above-discussed discrepancies observed in the values of ΔgcrHm ° seem to be compensated by an equivalent change in ΔgcrSm ° . The entropies of the phase transitions are schematically reported in Figure 10 and show that the ΔgcrS°m value for TBP is larger than expected. The hindered rotor analysis of OH rotation in TCP and TBP does not reveal any significant difference in the rotational profile of the two phenols (Table S5 and Figure S1). Since there are no obvious additional molecular features in these molecules that may contribute to significant entropic differentiation in the gas phase, the most probable reason for these entropy discrepancies must be found in the condensed phases. Hence, S°m(cr) seems smaller than expected for TBP (when compared to TCP), which is consistent with a higher degree of intermolecular OH···O bonding in the crystalline phase of TBP. The negative increment on Δgl Sm ° from the phenol to the anisole derivatives suggests that the two halophenols studied present some degree of H-bonding in the liquid phase, although smaller than in liquid phenol. The increment in the entropy of vaporization on going from phenol [(Δgl S°m/J·K−1·mol−1 = (130.8 ± 0.7)] to anisole (Δgl S°m/J·K−1· mol−1 = 111.7) is −19.1 J·K−1·mol−1, which is much larger than the ones determined for the compounds studied in this work, supporting the higher degree of hydrogen bonding in the liquid phase of the phenol. Since H-bonding contributes to a decrease in the value of Sm ° (cd), we can conclude, as a rule of thumb, that the lower the value of ΔgcdSm ° of the anisole relative to the corresponding phenol, the higher the degree of intermolecular H-bonding in the condensed phase (cd) of the phenol derivative. Note that this increment is only positive for Δ(ΔgcrSm ° ) (TCP → TCA), in agreement with the absence of significant intermolecular HI

DOI: 10.1021/acs.jced.9b00062 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 10. Increment scheme of standard molar entropies of phase transitions of TCP, TCA, TBP, and TBA. All values in J·K−1·mol−1. aCalculated considering the standard Gibbs energies and enthalpies of the phase transitions derived in this work from the fitting of eq 2 to the (p,T) results reported in the literature.22,23

Table 5. Fusion Properties: Temperature, Molar Enthalpy and Molar Entropy of TCA, TBA, and TBP Ttp/K

Tfus/Ka 332.47 ± 0.13

332.10 ± 0.34c 359.44 ± 0.19 360.48 ± 0.32c 366.07 ± 0.16 368.22 ± 0.64c 367.5 366.2 368.65 368 366

ΔlcrHm ° (T)b/kJ·mol−1 TCA 22.40 ± 0.31a 22.4 ± 0.6c TBA 23.88 ± 0.55a 23.2 ± 0.6c TBP 20.8 ± 0.37a 22.7 ± 1.2c 20.9 18.52

ΔlcrSm ° (T)b,c/J·K−1·mol−1

method/ref

67.4 ± 0.9 67.5 ± 1.9

DSC/this work VP/this work

66.4 ± 1.5 64.3 ± 1.8

DSC/this work VP/this work

56.8 ± 1.0 61.6 ± 3.3

DSC/this work VP/this work DSC/61 62 63 64 65

a

Standard uncertainties calculated through the RSS method combining the expanded uncertainties of the four experimental runs (0.95 level of confidence, k = 2) with the standard uncertainties of the DSC calibration. bT represents the temperature of fusion or the temperature of the triple point (Ttp). cUncertainties calculated through the RSS method.

measurements. For TBP, however, the temperature of the triple point and the enthalpy of fusion (at Tfus) derived indirectly differ by 2.2 K and 1.9 kJ·mol−1 from those determined calorimetrically. The Tfus of TBP obtained using DSC is 1.2 K smaller than the average of the results reported in the literature, and its enthalpy of fusion is equal to the one determined by Kuramochi et al.61 and 2.3 kJ·mol−1 larger than the value reported by Acree.62

bonding verified in crystalline TCP. The more negative Δ(Δgl S°m) for TCP → TCA in comparison to the one related to TBP → TBA also makes sense considering the lower ability of TBP to H-bond in the liquid phase due to the bulkier ortho-Br atoms. 3.3. Thermodynamic Properties of Fusion. Table 5 reports the results of the temperatures (onset), enthalpies, and entropies of fusion derived from the DSC experiments determined for TCA, TBA, and TBP, together with available literature results, as well as the temperatures and the enthalpies of fusion calculated indirectly from the vapor pressure measurements. The temperatures and enthalpies of fusion of TCA and TBA, determined from DSC analysis, are in excellent agreement with the results derived through the vapor pressure

4. CONCLUSIONS The relevant conclusions of the present work are the following: • The Gibbs energies and enthalpies of sublimation and vaporization of 2,4,6-trichloroanisole, tribromoanisole, and J

DOI: 10.1021/acs.jced.9b00062 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data



tribromophenol (at T = 298.15 K) were derived through vapor pressure measurements. • The phase behavior of the three compounds and of 2,4,6trichlorophenol was studied. • The crystallographic structure of 2,4,6-tribromophenol was determined by single crystal X-ray diffraction. • There is some degree of intermolecular H-bonding in the liquid phase of TCP and TBP, although not as high as in liquid phenol due to the steric effect of the ortho-halogens. • The existence of an intermolecular OH···O bond in crystalline 2,4,6-tribromophenol and the absence of this bond in the crystals of TCP was suggested by the results of the thermodynamic properties and supported by the crystallographic data.



REFERENCES

(1) Lopes, M. H.; Barros, A. S.; Pascoal Neto, C. P.; Rutledge, D.; Delgadillo, I.; Gil, A. M. Variability of cork from Portuguese Quercussuber studied by solid state 13C-NMR and FTIR spectroscopies. Biopolymers 2001, 62, 268−277. (2) Silva, S. P.; Sabino, M. A.; Fernandes, E. M.; Correlo, V. M.; Boesel, L. F.; Reis, R. L. Cork: properties, capabilities and applications. Int. Mater. Rev. 2005, 50, 345−365. (3) Pereira, H. Cork Biology Production and Uses, 1st ed.; Elsevier: Amsterdam, The Netherlands, 2007. (4) Sefton, M. A.; Simpson, R. F. Compounds causing cork taint and the factors affecting their transfer from natural cork closures to wine − a review. Aust. J. Grape Wine Res. 2005, 11, 226−240. (5) Simpson, R. F.; Sefton, M. A. Origin and fate of 2,4,6trichloroanisole in cork bark and wine corks. Aust. J. Grape Wine Res. 2007, 13, 106−116. (6) Recio, E.; Á lvarez-Rodríguez, M. L.; Rumbero, A.; Garzón, E.; Coque, J. J. R. Destruction of chloroanisoles by using a hydrogen peroxide activated method and its application to remove chloroanisoles from cork stoppers. J. Agric. Food Chem. 2011, 59, 12589−12597. (7) Garcia, A. R.; Lopes, L. F.; Barros, R. B.; Ilharco, L. M. The problem of 2,4,6-trichloroanisole in cork planks studied by attenuated total reflection infrared spectroscopy: proof of concept. J. Agric. Food Chem. 2015, 63, 128−135. (8) Parker, M.; Capone, D. L.; Francis, I. L.; Herderich, M. J. Aroma precursors in grapes and wine: flavor release during wine production and consumption. J. Agric. Food Chem. 2018, 66, 2281−2286. (9) Silva Teixeira, C. S.; Silva Ferreira, A. C.; Cerqueira, N. M. F. S. A. Studying haloanisoles interaction with olfactory receptors. ACS Chem. Neurosci. 2016, 7, 870−885. (10) Boutou, S.; Chatonnet, P. Rapid headspace solid-phase microextraction/gas chromatographic/mass spectrometric assay for the quantitative determination of some of the main odorants causing off-flavours in wine. J. Chromatogr. A 2007, 1141, 1−9. (11) Maggi, L.; Mazzoleni, V.; Fumi, M. D.; Salinas, M. R. Transformation ability of fungi isolated from cork and grape to produce 2,4,6-trichloroanisole from 2,4,6-trichloro phenol. Food Addit. Contam., Part A 2008, 25, 265−269. (12) Chatonnet, P.; Bonnet, S.; Boutou, S.; Labadie, M.-D. Identification and responsibility of 2,4,6-tribromoanisole in musty, corked odors in wine. J. Agric. Food Chem. 2004, 52, 1255−1262. (13) Capone, D. L.; Van Leeuwen, K. A.; Pardon, K. H.; Daniel, M. A.; Elsey, G. M.; Coulter, A. D.; Sefton, M. A. Identification and analysis of 2-chloro-6-methylphenol, 2,6-dichlorophenol and indole: causes of taints and off-flavors in wines. Aust. J. Grape Wine Res. 2010, 16, 210− 217. (14) Slabizki, P.; Fischer, C.; Legrum, C.; Schmarr, H.-G. Characterization of atypical off-flavor compounds in natural cork stoppers by multidimensional gas chromatographic techniques. J. Agric. Food Chem. 2015, 63, 7840−7848. (15) Sadoughi, N.; Schmidtke, L. M.; Antalick, G.; Blackman, J. W.; Steel, C. C. Gas chromatography-mass spectrometry method optimized using response surface modeling for the quantitation of fungal offflavors in grapes and wine. J. Agric. Food Chem. 2015, 63, 2877−2885. (16) Bordado, J. C. M.; Marques, J. P. S.; Gonçalves, M. A. P; Allegro, I. M. Q. M. S.; Mesquita, A. C. A. L. C.; Filipe, R. L. New process for treating cork stoppers or planks for the reduction of strange aromas, namely 2,4,6-trichloroanisole. Patent No. WO03041927, 2003. (17) Hall, M.; Williams, J.; Byrd, N. An assessment of the effect of the ROSA treatment on the level of TCA in naturally-contaminated cork granules. Aust. N. Z. Grapegrow Winemak. 2004, 57−59. (18) Taylor, M. K.; Young, T. M.; Butzke, C. E.; Ebeler, S. E. Supercritical fluid extraction of 2,4,6-trichloroanisole from cork stoppers. J. Agric. Food Chem. 2000, 48, 2208−2211. (19) Nunes da Ponte, M. L. M.; Lopes, J. A. S.; Najdanovic-Visak, V.; Manic, M.; Mesquita, A. C. A. L. C.; Silva, R. P. M.; Allegro, I. M. Q. M. S. Method for direct treatment of cork stoppers, using supercritical fluids. Patent No. WO2010093273 (A1), 2010.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jced.9b00062. Purity degree and literature results of the fusion properties of the reference materials used in the calibration of the DSC calorimeter; temperatures, molar enthalpies, and entropies of fusion of TCA, TBA, and TBP; vapor pressure results determined using the individual effusion cells for the sublimation of TBP; isobaric gaseous, crystalline, and sublimation heat capacities, at T = 298.15 K, for TCA, TBA, TBP, and TCP; definition of parameters A, B, and C of eq 4; electronic energies for the −OH dihedral angles in the phenolic compounds studied; potential energy profile of −OH rotation in the compounds studied; optimized geometries of TCP, TBP, TCA, and TBA (PDF) Crystallographic information (CIF)



Article

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Ana R. R. P. Almeida: 0000-0002-5424-5674 Carlos F. R. A. C. Lima: 0000-0002-5065-0912 Filipe A. Almeida Paz: 0000-0003-2051-5645 Manuel J. S. Monte: 0000-0002-2210-3559 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was developed within the scope of the projects UID/ QUI/00081/2013, POCI-01-0145-FEDER-006980, and NORTE-01-0145-FEDER-000028 (Sustained Advanced Materials, SAM), awarded to CIQUP, financed by Fundaçaõ para a Ciência e Tecnologia (FCT, Portugal) and cofinanced in the framework of Operational Programme for Competitiveness and Internationalisation (COMPETE, Portugal), with community funds (FEDER, Portugal) and national funds of MEC. This work was also developed within the scope of the project CICECO-Aveiro Institute of Materials, FCT ref. UID/CTM/ 50011/2019, financed by national funds through the FCT/ MCTES. A.R.R.P.A. also thanks FCT, European Social Fund (ESF) and national funds of MEC for the award of the postdoctoral grant (SFRH/BPD/97046/2013). We are grateful to Prof. Artur Silva (CICECO) for his kind help. K

DOI: 10.1021/acs.jced.9b00062 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

(43) Mohr, P. J.; Newell, D. B.; Taylor, B. N. CODATA recommended values of the fundamental physical constants: 2014. Rev. Mod. Phys. 2016, 88, 1−73. (44) 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.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A, Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö .; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, Revision A.1; Gaussian Inc.: Wallingford, CT, 2009. (45) Zhao, Y.; Truhlar, D. G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215−241. (46) Grell, J.; Bernstein, J.; Tinhofer, G. Graph-set analysis of hydrogen-bond patterns: some mathematical concepts. Acta Crystallogr., Sect. B: Struct. Sci. 1999, 55, 1030−1043. (47) Clarke, E. C. W.; Glew, D. N. Evaluation of thermodynamic functions from equilibrium constants. Trans. Faraday Soc. 1966, 62, 539−547. (48) NIST Computational Chemistry Comparison and Benchmark Database, NIST Standard Reference Database Number 101, Release 16a; Johnson, R. D., III, Ed.; NIST Chemistry Web Book, 2013. SRD 69: http://cccbdb.nist.gov/vibscalejust.asp (accessed March 7, 2018). (49) Č ervinka, C.; Fulem, M.; Ružička, K. Evaluation of uncertainty of ideal-gas entropy and heat capacity calculations by density functional theory (DFT) for molecules containing symmetrical internal rotors. J. Chem. Eng. Data 2013, 58, 1382−1390. (50) Acree, W., Jr.; Chickos, J. S. Phase transition enthalpy measurements of organic and organometallic compounds. Sublimation, vaporization and fusion enthalpies from 1880 to 2015. Part 1. C1 - C10. J. Phys. Chem. Ref. Data 2016, 45, 033101−033565. (51) Domalski, E. S.; Hearing, E. D. Estimation of the thermodynamic properties of C-H-N-O-S-halogen compounds at 298.15 K. J. Phys. Chem. Ref. Data 1993, 22, 805−1159. (52) Monte, M. J. S.; Almeida, A. R. R. P.; Matos, M. A. R. Thermodynamic study on the sublimation of five aminomethoxybenzoic acids. J. Chem. Eng. Data 2010, 55, 419−423. (53) Chickos, J. S.; Hosseini, S.; Hesse, D. G.; Liebman, J. F. Heat capacity corrections to a standard state: a comparison of new and some literature methods for organic liquids and solids. Struct. Chem. 1993, 4, 271−278. (54) Allot, P. H.; Finch, A.; Pilcher, G.; Nunez, L.; Barral, L. Enthalpies of formation of 2,4,6-tribromophenol and of 2,4,6 tribromoaniline. J. Chem. Thermodyn. 1987, 19, 771−779. (55) Vasiltsova, T. V.; Verevkin, S. P.; Bich, E.; Heintz, A.; BogelLukasik, R.; Domanska, U. Thermodynamic Properties of Mixtures Containing Ionic Liquids. Activity Coefficients of Ethers and Alcohols in 1-Methyl-3-Ethyl-Imidazolium Bis(Trifluoromethyl-sulfonyl) Imide Using the Transpiration Method. J. Chem. Eng. Data 2005, 50, 142− 148. (56) Monte, M. J. S.; Almeida, A. R. R. P.; Liebman, J. F. Prediction of enthalpy and standard Gibbs energy of vaporization of haloaromatics from atomic properties. Chemosphere 2015, 138, 478−485. (57) Verevkin, S. P.; Emel'yanenko, V. N.; Varfolomeev, M. A.; Solomonov, B. N.; Zherikova, K. V. Vaporization enthalpies of a series of the fluoro- and chlorosubstituted methylbenzenes. Fluid Phase Equilib. 2014, 380, 67−75.

(20) Jäger, J. Microwave treatment of cork and corks. Patent No. WO9937334, 1999. (21) Gil, L. M. C. C.; Pereira, C. R. Process for eliminating/reducing compounds with a musty taste/odor in materials that are to come into contact with foodstuffs and in foods or drinks Patent No. WO/2005/ 014249, 2005. (22) Mokbel, I.; Pauchon, V.; Jose, J. Vapor pressures and sublimation pressures of 2,4-dichlorophenol, 2,6-dichlorophenol, and 2,4,6trichlorophenol. ELDATA: Int. Electron. J. Phys. Chem. Data. 1995, 11, 55. (23) Verevkin, S. P.; Emel’yanenko, V. N.; Klamt, A. Thermochemistry of chlorobenzenes and chlorophenols: ambient temperature vapor pressures and enthalpies of phase transitions. J. Chem. Eng. Data 2007, 52, 499−510. (24) Kottke, T.; Stalke, D. Crystal handling at low temperatures. J. Appl. Crystallogr. 1993, 26, 615−619. (25) APEX3, Data Collection Software, version 2016.9−0; Bruker AXS: Delft, The Netherlands, 2005. (26) Cryopad, Remote Monitoring and Control, version 1.451; Oxford Cryosystems: Oxford, United Kingdom,(2006). (27) SAINT+, Data Integration Engine v. 8.37a; Bruker AXS: Madison, Wisconsin, 1997. (28) Krause, L.; Herbst-Irmer, R.; Sheldrick, G. M.; Stalke, D. Comparison of silver and molybdenum microfocus X-ray sources for single-crystal structure determination. J. Appl. Crystallogr. 2015, 48, 3− 10. (29) Sheldrick, G. M. SHELXT - Integrated space-group and crystalstructure determination. Acta Crystallogr., Sect. A: Found. Adv. 2015, 71, 3−8. (30) Sheldrick, G. M. Crystal structure refinement with SHELXL. Acta Crystallogr., Sect. C: Struct. Chem. 2015, 71, 3−8. (31) Hübschle, C. B.; Sheldrick, G. M.; Dittrich, B. ShelXle: a Qt graphical user interface for SHELXL. J. Appl. Crystallogr. 2011, 44, 1281−1284. (32) Flack, H. D.; Bernardinelli, G. Reporting and evaluating absolutestructure and absolute-configuration determinations. J. Appl. Crystallogr. 2000, 33, 1143−1148. (33) Brandenburg, K. DIAMOND, version 3.2f; Crystal Impact GbR: Bonn, Germany, 1997. (34) Chang, S. S.; Bestul, A. B. Heat capacity and thermodynamic properties of o-terphenyl crystal, glass, and liquid. J. Chem. Phys. 1972, 56, 503−516. (35) Sabbah, R.; El Watik, L. New reference materials for the calibration (temperature and energy) of differential thermal analysers and scanning calorimeters. J. Therm. Anal. 1992, 38, 855−863. (36) Sabbah, R.; An, X.-W.; Chickos, J. S.; Planas Leitão, M. L.; Roux, M. V.; Torres, L. A. Reference materials for calorimetry and differential thermal analysis. Thermochim. Acta 1999, 331, 93−204. (37) Della Gatta, G.; Richardson, M. J.; Sarge, S. M.; Stølen, S. Standards, calibration, and guidelines in microcalorimetry part 2. Calibration standards for differential scanning calorimetry. Pure Appl. Chem. 2006, 78, 1455−1476. (38) Roux, M. V.; Temprado, M.; Chickos, J. S.; Nagano, Y. Critically evaluated thermochemical properties of polycyclic aromatic hydrocarbons. J. Phys. Chem. Ref. Data 2008, 37, 1855−1996. (39) Monte, M. J. S.; Santos, L.M.N.B.F.; Fulem, M.; Fonseca, J. M. S.; Sousa, C. A. D. New static apparatus and vapor pressure of reference materials: naphthalene, benzoic acid, benzophenone and ferrocene. J. Chem. Eng. Data 2006, 51, 757−766. (40) Freitas, V. L. S.; Monte, M. J. S.; Gomes, J. R. B.; Santos, L. M. N. B. F.; Ribeiro da Silva, M. D. M. C. Energetic studies and phase diagram of thioxanthene. J. Phys. Chem. A 2009, 113, 12988−12994. (41) Ribeiro da Silva, M. A. V.; Monte, M. J. S.; Santos, L. M. N. B. F. The design, construction and testing of a new Knudsen effusion apparatus. J. Chem. Thermodyn. 2006, 38, 778−787. (42) Almeida, A. R. R. P.; Oliveira, J. A. S. A.; Monte, M. J. S. Thermodynamic study of nicotinamide, N-methylnicotinamide and N,N-dimethylnicotinamide: vapor pressures, phase diagrams, and hydrogen bonds. J. Chem. Thermodyn. 2015, 82, 108−115. L

DOI: 10.1021/acs.jced.9b00062 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

(58) Verevkin, S. P.; Sazonova, A. Y.; Emel'yanenko, V. N.; Zaitsau, D. H.; Varfolomeev, M. A.; Solomonov, B. N.; Zherikova, K. V. Thermochemistry of halogensubstituted methylbenzenes. J. Chem. Eng. Data 2015, 60, 89−103. (59) Gonzalez Martínez, S. P.; Bernès, S. 2,4,6-Trichlorophenol. Acta Crystallogr., Sect. E: Struct. Rep. Online 2007, E63, 3947. (60) Faniran, J. A.; Shurvell, H. F. Raman and infrared spectra of 2,4,6Trichloropheno1 and 2,4,6-Tribromophenol. J. Raman Spectrosc. 1980, 9, 73−75. (61) Kuramochi, H.; Maeda, K.; Kawamoto, K. Water solubility and partitioning behavior of brominated phenols. Environ. Toxicol. Chem. 2004, 23, 1386−1393. (62) Acree, W. E. Thermodynamic properties of organic compounds: enthalpy of fusion and melting point temperature compilation. Thermochim. Acta 1991, 189, 37−56. (63) Beringer, F. M.; Brierley, A.; Drexler, M.; Gindler, E. M.; Lumpkin, C. C. Diaryliodonium salts II. The phenylation of organic and inorganic bases. J. Am. Chem. Soc. 1953, 75, 2708−2712. (64) Timmermans, J. Research in stoichiometry. I. The heat of fusion of organic compounds. Bull. Soc. Chim. Belg. 1935, 44, 17−40. Cited in NIST Chemistry Web Book, SRD 69: https://webbook.nist.gov/cgi/ cbook.cgi?ID=C118796&Mask=4#Thermo-Phase (accessed March 20, 2019). (65) Robertson, P. W. Atomic and molecular heats of fusion. J. Chem. Soc., Trans. 1902, 81, 1233,. Cited in NIST Chemistry Web Book, SRD 69: https://webbook.nist.gov/cgi/cbook.cgi?ID=C118796&Mask= 4#Thermo-Phase (accessed March 20, 2019).

M

DOI: 10.1021/acs.jced.9b00062 J. Chem. Eng. Data XXXX, XXX, XXX−XXX