Next Article in Journal
Graded 2D/3D Perovskite Hetero-Structured Films with Suppressed Interfacial Recombination for Efficient and Stable Solar Cells via DABr Treatment
Previous Article in Journal
Computational and Preclinical Prediction of the Antimicrobial Properties of an Agent Isolated from Monodora myristica: A Novel DNA Gyrase Inhibitor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Phase Transitions Equilibria of Five Dichlorinated Substituted Benzenes

by
Ana R. R. P. Almeida
*,
Bruno D. A. Pinheiro
and
Manuel J. S. Monte
*
Research Centre in Chemistry (CIQUP), Institute of Molecular Sciences (IMS), Department of Chemistry and Biochemistry (DQB), Faculty of Sciences, University of Porto (FCUP), Rua do Campo Alegre, 4169-007 Porto, Portugal
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(4), 1590; https://doi.org/10.3390/molecules28041590
Submission received: 16 December 2022 / Revised: 31 January 2023 / Accepted: 1 February 2023 / Published: 7 February 2023
(This article belongs to the Section Physical Chemistry)

Abstract

:
This work reports an experimental study aiming to determine the thermodynamic properties of five chlorinated compounds with environmental impact. The vapor pressures of the crystalline phases of three isomers of dichlorobenzoic acid (2,4-, 2,5-, and 2,6-) and 2,6-dichlorobenzonitrile were measured at several temperatures using the Knudsen effusion technique. Another technique (a static method based on capacitance diaphragm manometers) allowed the measurement of the vapor pressures of both the crystalline and liquid phases of 2,4-dichlorobenzonitrile between 303.0 and 380.0 K. This latter technique also enabled the measurement of sublimation vapor pressures of 2,6-dichlorobenzonitrile over a larger range interval of temperatures, T = 328.7 and 391.8 K. The standard molar enthalpy, entropy, and Gibbs energy of sublimation (for all the compounds studied) and vaporization (for 2,4-dichlorobenzonitrile) were derived, at reference temperatures, from the experimental vapor pressure results. The temperatures and enthalpies of fusion and the isobaric heat capacities of the five crystalline-substituted benzenes were determined using differential scanning calorimetry. The contributions of the three substituents (-COOH, -CN, and -Cl) to the sublimation thermodynamic properties of the compounds studied were discussed.

1. Introduction

Knowledge of organic compounds’ thermodynamic properties and their systematization in a convenient format is essential for several scientific, technological, and industrial applications. The measurement and prediction of these properties have received close attention since they are indispensable for optimizing processes and applications in chemistry, biotechnology, and materials science, including developing new technical equipment, sustainable energy use, and the safety assessment, behavior, and fate of pesticides and other pollutants in the environment [1,2,3,4]. Among these properties, vapor pressure occupies a prominent role due to the diversity of thermodynamic properties determined from its knowledge at different temperatures. The quantification of this fundamental physical property is often required for the industrial separation/distillation processes [5,6,7,8], in planning and monitoring activities concerning air quality in domestic environments and workplaces [9], in estimating exposure limits to volatile toxic substances [10], and in the study of precursors for thin-film deposition—a methodology used to produce high-performance hybrid organic–inorganic materials that have appeared as the next generation in photovoltaic technology [11,12].
The volatility of a material at a given temperature is adequately described by its vapor pressure (at that temperature), which is related to the Gibbs energy of vaporization or sublimation that depends both on enthalpic and entropic factors.
Information on enthalpies of sublimation or vaporization is very important for several studies, such as:
-
The conversion of the gaseous phase enthalpies of formation into the condensed phases, or vice-versa;
-
Estimating enthalpies of intermolecular hydrogen bonds formed in some crystalline materials, as we have investigated before (substituted benzoic acids [13], imidazoles [14], and benzamides [15]);
-
Understanding the behavior of moldy/musty contaminants in cork and improving processes for their removal [4];
-
The evaluation of the thermodynamic stability of compounds used in several chemical industries such as pharmaceuticals [16], fragrances and flavors [17], and agrochemicals [18].
Furthermore, the entropic contribution to Gibbs energies of sublimation or vaporization frequently helps in rationalizing the relationships between the crystalline or liquid structures predicting the behavior of these phase transitions.
Vapor pressure data are also needed to satisfy safety regulations aimed at assessing the environmental impact of chemical compounds and granting permission for their use [19]. Together with other properties such as water solubility, vapor pressure results allow the calculation of distribution coefficients, Henry’s constants, and Gibbs energies of hydration [20], which play an important role in designing new models that aim to simulate the evolution of the distribution of pollutants in the environment, particularly pesticides.
The present work consists mainly of a volatility and phase transitions study of the herbicide dichlobenil (2,6-DCBN) and other four halogen-substituted benzenes that are used in pesticide synthesis or result in their degradation as noxious compounds: 2,4-dichlorobenzonitrile (2,4-DCBN), 2,4-dichlorobenzoic acid (2,4-DCBA), 2,5-dichlorobenzoic acid (2,5-DCBA), and 2,6-dichlorobenzoic acid (2,6-DCBA), whose structural formula is depicted in Figure 1. Dichlobenil is frequently used as a broad-spectrum contact weedkiller that has been mostly applied for controlling weeds and grasses in agricultural, industrial, and residential areas. It is also used to remove tree roots and prevent their growth in sewers [21,22,23]. Moreover, this selective and systemic herbicide is an inhibitor of the biosynthesis of cellulose and does not influence photosynthesis or cell respiration [22]. The occurrence of halogenated benzoic acids in natural environments is due predominantly to the use of herbicides and the partial degradation of spilled polychlorinated biphenyls (PCBs) [24]. 2,4-dichlorobenzoic acid is one of the degradation products of chlorfenvinphos, an organophosphorus compound, insecticide, and acaricide [25], and of the fungicide propiconazole [26]. 2,5-dichlorobenzoic acid (2,5-DCBA) is used as an intermediate in the synthesis of methyl 2,5-dichlorobenzoate [27], a plant growth regulator and antifungal agrochemical [28]. It is also involved in the production of the herbicides chloramben (3-amino-2,5-dichlorobenzoic acid) and dinoben (2,5-dichloro-3-nitrobenzoic acid) [29]. 2,6-dichlorobenzoic acid (2,6-DCBA) was reported as a plant growth regulator useful for weed control [30] and is one of the three potential metabolites produced during the degradation of the herbicide dichlobenil [31]. In addition, this compound is used as an inhibitor of an enzyme involved in breast and ovarian cancer as well as in the pathology of Alzheimer’s disease [30].
The thermodynamic study of these compounds is part of a project that has as its goal the experimental determination and prediction of physicochemical properties related to the environmental mobility of organic compounds with hazardous activity. Due to the absence of experimental results of thermodynamic properties for many organic compounds (due to the high cost of samples, the difficulties regarding the limitations of the experimental apparatus, or the eventual compound decomposition) and the need to screen the reliability of the data reported in the literature, several models have been proposed for predicting results, such as empirical or semi-empirical correlation models [32,33], the quantitative structure–property relationship (QSPR) [34], artificial neural networks (ANNs) [35], and also group-contribution methods [36,37,38,39]. A user-friendly group contribution model for predicting standard Gibbs energies and enthalpies of sublimation of substituted benzenes has been proposed by us in a previous study [40]. This method considers the fusion temperatures of benzene derivatives and the group contribution of their different substituents [40], as well as meaningful interactions in the ortho and para positions [40]. To enlarge the original database that supports this model, the results of more substituted benzenes have been determined or collected from the literature, allowing us to estimate the contributions of missing groups in that database, such as acetyl, acetoxy, and acetamido substituents [41,42,43]. The results of the standard Gibbs energies and enthalpies of sublimation of the five dichlorine-substituted benzenes, experimentally determined in this work, were compared with the values estimated through the predictive equations that support this model.

2. Results and Discussion

2.1. Thermodynamic Properties of Sublimation and Vaporization

Table S1 of the Supplementary Materials reports the sublimation and vaporization vapor pressures of the compounds studied in this work using the different experimental methods referred to above (Knudsen effusion and static methods). The effusion vapor pressures included in this table are the means of the results derived from the individual effusion cells at a certain temperature. The detailed experimental results obtained from each effusion cell are summarized in Table S2. The experimental (p,T) data were fit by the truncated form of the Clarke–Glew equation [44], Equation (1):
R ln p p o = Δ cd g G m o θ θ + Δ cd g H m o θ 1 θ 1 T + Δ cd g C p , m o θ θ T 1 + ln T θ
where po is a selected reference pressure (po = 105 Pa in this work), p is the vapor pressure at the temperature T, θ is a reference temperature (in this work, unless stated otherwise, θ = 298.15 K), and R is the molar gas constant. The differences between gaseous and condensed phases of standard Gibbs energy, enthalpy, entropy (calculated using Equation (2)), and the isobaric heat capacity, are represented, respectively, by Δ cd g G m o , Δ cd g H m o , Δ cd g S m o , and Δ cd g C p , m o . Table 1 reports, for each compound studied, the values of these properties and their associated uncertainties for three different temperatures (θ = mean temperature of the experiments, θ = temperature of the triple point of 2,4-DCBN, and θ = 298.15 K). The vapor pressure results calculated from Equation (1) for the three different temperatures are also reported in this table.
Δ cd g S m o ( θ ) =   Δ cd g H m o ( θ ) Δ cd g G m o ( θ ) θ

Isobaric Heat Capacities

When accurate experimental sublimation, or vaporization, vapor pressures are measured over a wide temperature range (larger than ~50 K), the fit of Equation (1) to the experimental values frequently yields consistent values of Δ cd g C p , m o ( θ ) . The value of Δ cr g C p , m o ( θ ) of the 2,6-DCBN studied in the crystalline phase, between T = 328.7 and 391.8 K, using the static method, and the value of Δ l g C p , m o ( θ ) of the 2,4-DCBN studied in the liquid phase, between T = 333.3 and 380.8 K, were derived from those fittings to the (p,T) experimental results. Since the sublimation temperature ranges worked in this study using the Knudsen effusion apparatus were insufficiently large (due to the apparatus’ limitations), the values of Δ cr g C p , m o ( θ ) could not be calculated through the fittings referred to above. So, they were calculated as Δ cr g C p , m o ( θ ) = C p , m o ( g ) C p , m o ( cr ) , where C p , m o ( g ) and C p , m o ( cr ) are, respectively, the gas and crystalline isobaric molar heat capacities.
The values of the gas heat capacity at the temperature of 298.15 K were determined in this work for the five compounds using theoretical calculations. These results, reported in Table 2, were derived from statistical thermodynamics by applying the Gaussian 03 software package [45], using the vibrational frequencies from G3(MP2)//B3LYP calculations, scaled by a factor of (0.960 ± 0.022) [46]. For comparison reasons, the values estimated using the group contribution method proposed by Domalski and Hearing [47] are also listed in this table and are in close agreement with those calculated using our calculations.
Table 2 also includes the results of the crystal heat capacity determined in this study using differential scanning calorimetry, as well as the results estimated using the group contribution methods proposed by Domalski and Hearing [47] and by Acree Jr. and Chickos [48]. The DSC results referred to above were determined at different temperatures, enabling the fit to a third-order polynomial in temperature: C p , m o ( cr ) /J·K−1·mol−1 = a + b(T/K) + c(T2/K) + d(T3/K). The parameters of this equation are summarized in Table S3 of the Supplementary Materials for the five compounds studied. In all cases, the correlation coefficients were >0.999.
The value of Δ cr g C p , m o ( θ ) /J·K−1·mol−1 = (−26.4 ± 5.5), derived from the regression of Equation (1) to the measured crystalline vapor pressure data of 2,6-DCBN, was used for the temperature adjustments of the thermodynamic properties of this compound. This result is in good agreement with the value Δ cr g C p , m o ( θ ) /J·K−1·mol−1 = (−26.2 ± 4.8) that is derived through the difference between the results of the theoretical value of C p , m o ( g ) and the result of C p , m o ( cr ) determined in this work.
Table S4 of the Supplementary Materials includes the values of C p , m o ( g ) , C p , m o ( cr ) , Δ cr g C p , m o ( θ ) , and their associated uncertainties, as well as the results of these properties estimated using various methods [49,50], namely the ones described above.
Table 2. Gaseous and crystalline isobaric heat capacities at constant pressure and at T = 298.15 K.
Table 2. Gaseous and crystalline isobaric heat capacities at constant pressure and at T = 298.15 K.
MethodCompound
2,4-DCBA2,5-DCBA2,6-DCBA2,4-DCBN2,6-DCBN
C p , m o ( g ) /J·K−1·mol−1
G3(MP2)B3LYP a,b160.0 ± 4.8160.0 ± 4.8161.4 ± 4.8141.6 ± 4.2141.6 ± 4.2
Domalski and Hearing [47] c158.7 ± 4.0 d158.7 ± 4.0 d158.7 ± 4.0 d140.5 ± 4.0140.5 ± 4.0
C p , m o ( c r ) /J·K−1·mol−1
DSC (this work)180.9 ± 2.0179.5 ± 2.2182.8 ± 1.8168.9 ± 2.5167.8 ± 2.4
Domalski and Hearing [47] c173.0 ± 4.0173.0 ± 4.0173.0 ± 4.0161.1± 4.0 e161.1± 4.0 e
Acree Jr. and Chickos [48] f171.5 ± 17.0171.5 ± 17.0171.5 ± 17.0160.7 ± 17.0160.7 ± 17.0
a Estimated standard uncertainties: u( C p , m o ( g ) /J·K−1·mol−1) = 0.03·[ C p , m o ( g ) /J·K−1·mol−1]. b Using the scaling factor 0.96 [46]. c Group contribution method proposed by Domalski and Hearing [47]. d Calculated considering the value C p , m o ( g ) /J·K−1·mol−1 = 143 of 4-chlorobenzoic acid taken from the literature [51], and the group contribution values proposed by Domalski and Hearing [47]. e Calculated considering the value C p , m o ( cr ) /J·K−1·mol−1 = 165.13 for 4-nitrobenzonitrile taken from the literature [52] and the group contribution values proposed by Domalski and Hearing [47]. f Group contribution method proposed by Acree Jr. and Chickos [48].

2.2. Thermodynamic Properties of Fusion

The averages of the onset temperatures of fusion and the molar enthalpy and entropy of fusion of the five substituted benzenes determined using DSC (Netzsch, Selb, Germany) are reported in Table 3, together with the available literature results. Our results are in excellent agreement with the ones reported in the literature for 2,5-DCBA, 2,6-DCBA, and 2,6-DCBN. This table also describes the temperature of the triple point and the enthalpy of fusion determined indirectly using the static method for 2,4-DCBN.

2.3. Volatility Evaluation

Figure 2 presents the plots of the sublimation vapor pressures against reciprocal temperatures for 2,4-DCBA, 2,5-DCBA, and 2,6-DCBA. Figure 3 shows the crystalline vapor pressures of 2,6-DCBN determined through the Knudsen effusion and static methods, and Figure 4 illustrates the phase diagram representation of the (p,T) results of 2,4-DCBN in the neighborhood of the triple point. To the best of our knowledge, no sublimation vapor pressure data have been reported before for these five compounds.
The volatility of a crystalline compound at a certain temperature, T, decreases with increasing values in Δ cr g G m o ( T ) . Thus, considering the second member of Equation (2), the volatility decreases with increasing Δ cr g H m o ( T ) and decreasing values of the product [T· Δ cr g S m o ( Τ ) ]. The contribution of these enthalpic and entropic terms for the values of Δ cr g G m o ( 298.15   K ) of the compounds studied are easily observed in Figure 5, showing that the volatility is mostly conditioned by enthalpic factors, particularly strong for the compounds forming intermolecular hydrogen bonds. The thermodynamic properties of sublimation determined for the five compounds, reported in Table 1, and represented in Figure 5, define the following order of volatility:
2,4-DCBA < 2,5-DCBA < 2,6-DCBA < 2,6-DCBN < 2,4-DCBN
Only the crystal structures of 2,6-DCBA, 2,4-DCBA, and 2,6-DCBN were found in the literature [30,59,60]. The OH⋯O hydrogen bonds (HB) forming centrosymmetric dimers in the crystalline phases of 2,4-DCBA and 2,6-DCBA are stronger than the intermolecular interactions Cl⋯Cl or H⋯N existing in the crystals of 2,6-DCBN [60]. This explains its lower volatility compared to the studied substituted benzonitriles. Although the crystalline structures of 2,5-DCBA and 2,4-DCBN were not found in the literature, the different volatility between both compounds suggests the occurrence of OH⋯O HB dimers in the crystal pattern of the substituted benzoic acid, similarly to what happens in the crystalline packing of 2,4- and 2,6-DCBA.
Concerning the volatility differences between the dichlorobenzoic acids investigated, it is possible to demonstrate that the presence of two chlorine atoms in ortho position in the 2,6-DCBA molecules significantly increases its volatility. The crystalline structure reported in the literature for this compound [30], represented schematically in Figure 6, suggests the establishment of intermolecular OH⋯Cl contacts (l = 3.233 A°) that are absent in the crystal pattern of 2,4-DCBA [59]. These interactions may be responsible for weakening the hydrogen-bonded (OH⋯O) dimers that are formed between molecules of this compound, thus boosting its volatility. The larger volatility of this isomer is due, essentially, to its smaller enthalpic contribution.
Due to the greater influence of its enthalpic factor, 2,6-DCBN is significantly less volatile than 2,4-DCBN, which means that the interactions between the 2,6 isomer molecules are stronger than those present in 2,4-DCBN. The published crystalline structure data of 2,6-DCBN indicate the formation of strong (short) intermolecular contacts between halogen atoms X⋯X and and H⋯N [60], which do not exist in other chlorinated benzonitriles such as 2-chlorobenzonitrile or 2,4,6-trichlorobenzonitrile, where the dominant intermolecular interactions are formed between the halogen atom and the nitrile group X⋯N [61,62]. However, since the 2,4-DCBN structure was not found in the literature, there is not enough information to draw a firm conclusion about the type and strength of the intermolecular interactions between its molecules.

2.4. Estimation of Sublimation Properties of Substituted Benzenes

A previously reported method for predicting standard Gibbs energies and enthalpies of sublimation of substituted benzenes, which considers the influence of 30 different substituent groups, is described by Equations (S2) and (S3), respectively [40]. The influence of each substituent was initially estimated, considering the sublimation results of the parameters of ca. 240 substituted benzenes. This database has been updated introducing experimental sublimation results of more benzene derivatives that confirmed those predictive equations. It already contains about 300 compounds including three new substituent groups that were absent in the original database (acetamido, acetoxy, and acetyl groups [41,42,43]). The estimation equations and the meaning of their parameters are described in the Supplementary Materials. The results of Δ cr g G m o ( 298.15   K ) and Δ cr g H m o ( 298.15   K ) of the five compounds studied, estimated using Equations (S2) and (S3) (described in the Supplementary Materials), respectively, are in very good agreement with those determined in this work through the vapor pressure measurements. The experimental and estimated results are reported in Table S5 of the Supplementary Materials.

3. Experiment

3.1. Materials and Purity Control

Details on the degree of purity, the purification and analysis methods, and the water content of the five compounds studied are reported in Table S6 (Supplementary Materials).
The degree of purity of the purified samples (after sublimation under reduced pressure) was checked by gas–liquid chromatography, performed using an Agilent 4890D chromatograph (Santa Clara, CA, USA), equipped with a non-polar capillary HP-5 column and a flame ionization detector (FID). Dimethylformamide was the solvent, and nitrogen was the carrier gas. The final mass fraction purities of the compounds studied were ≥0.999.
Karl Fischer’s coulometric titration was employed to determine the water content of the samples using a Methrom titration system (Herisau, Switzerland) that comprises an 831 Coulometer equipped with a generator electrode (without a diaphragm) and using HYDRANAL™ (Honeywell, Fluka) as the reagent.

3.2. Thermal Analysis

3.2.1. Differential Scanning Calorimetry

Fusion Properties

To check the absence of eventual phase transitions in the compounds’ crystalline phase and determine their temperatures (onset) and enthalpies of fusion, the heat flux Netzsch calorimeter (model 204 F1 Phoenix, Selb, Germany) was used. This equipment is outfitted with a τ-sensor and an intracooler system capable of operating in the temperature range 193−873 K. Four independent runs were carried out for each compound, with samples sealed in hermetically sealed aluminum crucibles (concavus Al, 40 μL, Ø 5 mm, Netzsch, Selb, Germany). The samples were scanned at 2.0 K·min−1 from T/K = 298 to a temperature 15 to 20 K higher than their fusion temperature; different scans were performed through two heating–cooling cycles. Controlled nitrogen fluxes were used as a purge and as a protective gas (40 cm3·min−1 and 20 cm3·min−1), respectively, preventing the calorimeter contamination in the event of crucible hermeticity failure.
The temperature and heat flow scales of the calorimeter were calibrated using four substances from Netzsch’s kit [63]: indium (I-2803), tin (S-2776), bismuth (B-3067), and adamantane (A-2838), and using another six high-purity reference materials [64,65,66,67,68]: benzoic acid, o-terphenyl, naphthalene, cyclohexane, triphenylene, and biphenyl. The degree of purity of these calibrants, and the values of the enthalpy and temperature of fusion, are reported in Table S7 (Supplementary Materials). The standard uncertainties derived from the calibration results are u(T/K) = 0.46 and u( Δ cr l H m o ( T fus ) /kJ·mol−1) = 0.21.
Before melting, no crystalline phase transitions or signals of decomposition were observed under the experimental conditions used. The onset temperatures of fusion, Tfus, and the molar enthalpies Δ cr l H m o ( T fus ) and entropies Δ cr l S m o ( T fus ) of fusion of the five compounds studied are reported in Table S8, together with the available literature results.

Crystalline Heat Capacities

Using Netzsch’s procedures and recommendations, in accordance with the ASTM E 1269, DIN 51 007, and ISO 11357-4 standards, the crystalline heat capacity measurements were performed in the present work in the temperature range of 286 to 370 K for 2,4-DCBA, 2,5-DCBA, 2,6-DCBA, and 2,6-DCBN, and between 282 and 317 K for 2,4-DCBN. Each heat capacity measurement involves three runs: blank, calibration (sapphire), and sample studied. The temperature program used contained an isothermal step at the initial temperature (25 min), a temperature ramp at 10 K⋅min–1, and a second isothermal step at the final temperature (25 min), using nitrogen as purge and protective gas, at flow rates of 30 cm3·min−1 and 50 cm3·min−1, respectively.
To check the reliability of the experimental technique, heat capacity experiments were performed using benzoic acid (NIST, 0.99996) and synthetic sapphire (α-Al2O3, 0.9999, disk of 0.25 mm diameter, NETZSCH) as reference material, in the temperature interval T/K = 281–370. The relative percentage error of our calibration measurements in comparison with those reported in the literature [69] is less than 2%. The comparison of the molar heat capacities of benzoic acid with the reference data reported in the literature [69] is shown in Figure S1 of the Supplementary Materials. All samples were placed in the aluminum crucibles (referred to above) with their lids on (unsealed) and were weighed (8–16 mg) with a precision of ±0.1 μg on a Mettler Toledo balance (model UMT2, Columbus, Ohio, United Sates).

3.3. Vapor Pressure Measurements

3.3.1. Knudsen Mass-Loss Effusion Method

The vapor pressures of the crystalline phases of the dichlorinated benzoic acids and of 2,6-DCBN were measured at different temperatures using a Knudsen mass-loss effusion apparatus that allows the operation of nine effusion cells simultaneously. The setup and procedure of the effusion technique have been detailed, explained, and tested before [70].
During a typical experiment, the effusion cells (with different areas of the effusion orifices) are introduced in cylindrical holes inside three aluminum blocks, which are maintained at a constant temperature (that is different for the three blocks). The dimensions of the effusion orifices [71], made on platinum foil of 0.0125 ± 0.001 mm thickness, are specified in the Supplementary Materials.
For each effusion experiment, the mass loss of the samples 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 the temperature T, the vapor pressure p of the crystalline sample contained in each effusion cell was calculated using Equation (3):
p =   m A o w o t 2 π R T M 0.5
where M is the molar mass of the effusing vapor and R is the molar gas constant (R = 8.3144598 J·K−1·mol−1 [72]). The standard uncertainties of the temperatures and vapor pressure measurements were estimated as u(T/K) = 0.01 and u(p/Pa) = 0.02.

3.3.2. Static Method Based on Capacitance Diaphragm Manometers

The vapor pressures of both the crystalline and liquid phases of 2,4-DCBN and of the crystalline phase of 2,6-DCBN, over a wider temperature interval than the one determined using the Knudsen effusion method, were measured using a static apparatus based on capacitance diaphragm manometers, that was tested and fully described before [73]. The measurements were taken using an MKS Baratron diaphragm capacitance manometer (model 631A11TBFP) operating at a self-controlled temperature (Tgauge = 473 K), which is suitable for measuring pressures in the range of 3 to 1.3 × 103 Pa [74]. The expression U(p/Pa) = 0.1 + 0.0050 (p/Pa) describes the expanded uncertainty (0.95 confidence level, k = 2) of the pressure results, and the standard uncertainty of the temperature measurements is estimated to be u(T/K) = 0.01. Prior to the vapor pressure measurements, the samples were fully degassed under reduced pressure inside the apparatus, allowing for the eventual removal of any trace of volatile impurities.

4. Conclusions

Relevant thermodynamic properties of phase transitions of 2,4-, 2,5-, and 2,6-dichlorobenzoic acids as well as of 2,4- and 2,6-dichlorobenzonitriles were determined in this work.
-
The temperatures and molar enthalpies of fusion of the five compounds studied and their crystalline isobaric molar heat capacities were determined using DSC.
-
The enthalpies, entropies, and Gibbs energies of sublimation of all the compounds and of the vaporization of 2,4-DCBN were derived through vapor pressure measurements (at different temperatures) and the phase diagram representation of the (p,T) results of the latter compound, including its triple point coordinates, were reported.
-
The evaluation of the enthalpic and entropic contributions to the volatility of the compounds studied was discussed.
-
The contributions of -COOH, -CN, and -Cl substituents to the sublimation properties of the substituted benzenes studied were confirmed accordingly to our estimation model.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28041590/s1: Source, purity, and methods of purification and analysis of the compounds studied. Reference materials used in the calibration of the DSC calorimeter. DSC results: temperatures, molar enthalpies, and entropies of fusion of the compounds studied. Vapor pressure results. Detailed results from each effusion orifice of the sublimation of the compounds studied. Parameters of Equation (S1) of the temperature dependence of the crystalline isobaric heat capacities of 2,4-DCBA, 2,5-DCBA, 2,6-DCBA, 2,6-DCBN, and 2,4-DCBN determined using DSC. Gaseous, crystalline, and sublimation heat capacities at constant pressure and at T = 298.15 K. Estimation of sublimation properties of substituted benzenes. References [40,46,47,48,49,50,51,52,53,54,55,56,57,58,63,64,65,66,67,68,69] are cited in the supplementary materials.

Author Contributions

Conceptualization, A.R.R.P.A. and M.J.S.M.; methodology, A.R.R.P.A. and B.D.A.P.; software, A.R.R.P.A. and B.D.A.P.; validation, A.R.R.P.A. and M.J.S.M.; formal analysis, A.R.R.P.A. and M.J.S.M.; investigation, A.R.R.P.A., B.D.A.P. and M.J.S.M.; writing—original draft preparation, A.R.R.P.A. and M.J.S.M.; writing—review and editing, A.R.R.P.A. and M.J.S.M.; supervision, A.R.R.P.A. and M.J.S.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Fundacão para a Ciência e Tecnologia (FCT) (funded by national funds through the FCT/MCTES (PIDDAC)) to CIQUP, Faculty of Sciences, University of Porto (Project UIDB/00081/2020) and IMS—Institute of Molecular Sciences (LA/P/0056/2020)). A.R.R.P.A. is financed by national funds through the FCT-I.P., in the framework of the execution of the program contract provided in paragraphs 4, 5, and 6 of art. 23 of Law no. 57/2016 of 29 August, as amended by Law no. 57/2017 of 19 July. B.D.A.P. also thanks FCT for the award of a PhD research grant (UI/BD/152715/2022).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Kabo, G.J.; Blokhin, A.V.; Paulechka, E.; Roganov, G.N.; Frenkel, M.; Yursha, I.A.; Diky, V.; Zaitsau, D.; Bazyleva, A.; Simirsky, V.V.; et al. Thermodynamic properties of organic substances: Experiment, modeling, and technological applications. J. Chem. Thermodyn. 2019, 131, 225–246. [Google Scholar] [CrossRef]
  2. Kontogeorgis, G.M.; Dohrn, R.; Economou, I.G.; de Hemptinne, J.-C.; ten Kate, A.; Kuitunen, S.; Mooijer, M.; Žilnik, L.F.; Vesovic, V. Industrial requirements for thermodynamic and transport properties: 2020. Ind. Eng. Chem. Res. 2021, 60, 4987–5013. [Google Scholar] [CrossRef] [PubMed]
  3. White, M.A.; Kahwaji, S.; Freitas, V.L.S.; Siewert, R.; Weatherby, J.A.; Ribeiro da Silva, M.D.M.C.; Verevkin, S.P.; Johnson, E.R.; Zwanziger, J.W. The Relative Thermodynamic stability of diamond and graphite. Angew. Chem. Int. Ed. 2021, 60, 1546–1549. [Google Scholar] [CrossRef] [PubMed]
  4. Almeida, A.R.R.P.; Pinheiro, B.D.A.; Lima, C.F.R.A.C.; Santos, A.F.L.O.M.; Ferreira, A.C.S.; Paz, F.A.A.; Monte, M.J.S. Thermodynamic properties of moldy-musty contaminants of wine. J. Chem. Eng. Data 2019, 64, 4741–4753. [Google Scholar] [CrossRef]
  5. Barbeiro, L.B.; Falleiro, R.M.M.; Meirelles, A.J.A. Vapour pressure and VLE data of fatty compounds. J. Chem. Thermodyn. 2021, 159, 106469. [Google Scholar] [CrossRef]
  6. Rordorf, B.F.; Geoffroy, A.; Szelagiewicz, M.; Marti, E. Vapor Pressure methods for industrial applications. In Thermal Analysis: Volume 1: Theory Instrumentation Applied Sciences Industrial Applications; Wiedemann, H.G., Ed.; Birkhäuser: Basel, Switzerland, 1980. [Google Scholar]
  7. Dauber, T.E. Strengths and weaknesses of predictive methods for estimating thermophysical properties. J. Chem. Eng. Data 1996, 41, 5345–5350. [Google Scholar]
  8. Mathias, P.M.; Schiller, M. Analysis of vapor pressures using family variations: A case study of hexadecanol isomers. Ind. Eng. Chem. Res. 2022, 61, 12239–12248. [Google Scholar] [CrossRef]
  9. Fonseca, J.M.S.; Gushterov, N.; Dohrna, R. Vapour pressures of selected organic compounds down to 1mPa, using mass-loss Knudsen effusion method. J. Chem. Thermodyn. 2014, 73, 148–155. [Google Scholar] [CrossRef]
  10. Martin Tischer, M.; Roitzsch, M. Estimating inhalation exposure resulting from evaporation of volatile multicomponent mixtures using different modelling approaches. Int. J. Environ. Res. Public Health 2022, 19, 1957. [Google Scholar] [CrossRef]
  11. Costa, J.C.S.; Azevedo, J.; Araújo, J.P.; Santos, L.M.N.B.F.; Mendes, A. High purity and crystalline thin films of methylammonium lead iodide perovskites by a vapor deposition approach. Thin Solid Film. 2018, 664, 12–18. [Google Scholar] [CrossRef]
  12. Kunte, G.V.; Ail, U.; Ajikumar, P.K.; Tyagi, A.K.; Shivashankar, S.A.; Umarji, A.M. Estimation of vapour pressure and partial pressure of subliming compounds by low-pressure thermogravimetry. Bull. Mater. Sci. 2011, 34, 1633–1637. [Google Scholar] [CrossRef]
  13. Almeida, A.R.R.P.; Monte, M.J.S. Vapor pressures of four methyl esters of substituted benzoic acids. The intermolecular hydrogen bond OH···O. J. Chem. Eng. Data 2016, 61, 1012–1020. [Google Scholar] [CrossRef]
  14. Almeida, A.R.R.P.; Monte, M.J.S. Vapour pressures of 1-methyl derivatives of benzimidazole, pyrazole and indole. The energy of the intermolecular hydrogen bond NH⋯N. J. Chem. Thermodyn. 2014, 77, 46–53. [Google Scholar] [CrossRef]
  15. Almeida, A.R.R.P.; Monte, M.J.S. Thermodynamic study of benzamide, N-methylbenzamide, and N,N-dimethylbenzamide: Vapor pressures, phase diagrams, and hydrogen bond enthalpy. J. Chem. Eng. Data 2010, 55, 3507–3512. [Google Scholar] [CrossRef]
  16. Almeida, A.R.R.P.; Cunha, A.F.G.; Matos, M.A.R.; Morais, V.M.F.; Monte, M.J.S. Thermodynamic properties of the methyl esters of p-hydroxy and p-methoxy benzoic acids. J. Chem. Thermodyn. 2014, 78, 43–57. [Google Scholar] [CrossRef]
  17. Almeida, A.R.R.P.; Freitas, V.L.S.; Campos, J.I.S.; Ribeiro da Silva, M.D.M.C.; Monte, M.J.S. Volatility and thermodynamic stability of vanillin. J. Chem. Thermodyn. 2019, 128, 45–54. [Google Scholar] [CrossRef]
  18. Almeida, A.R.R.P.; Pinheiro, B.D.A.; Lobo Ferreira, A.I.M.C.; Monte, M.J.S. Thermodynamic stability of Fenclorim and Clopyralid. Molecules 2022, 27, 39. [Google Scholar] [CrossRef]
  19. Available online: http://ec.europa.eu/enterprise/sectors/chemicals/reach/index_en.htm (accessed on 18 October 2022).
  20. Monte, M.J.S.; Almeida, A.R.R.P. Estimations of the thermodynamic properties of halogenated benzenes as they relate to their environment mobility. Chemosphere 2017, 189, 590–598. [Google Scholar] [CrossRef]
  21. Holtze, M.S.; Sørensen, S.R.; Sørensen, J.; Aamand, J. Microbial degradation of the benzonitrile herbicides dichlobenil, bromoxynil and ioxynil in soil and subsurface environments-insights into degradation pathways, persistent metabolites and involved degrader organisms. Environ. Pollut. 2008, 154, 155–168. [Google Scholar] [CrossRef]
  22. Frková, Z.; Badawi, N.; Johansen, A.; Schultz-Jensen, N.; Bester, K.; Reinhold Sørensen, S.; Karlson, U.G. Degradation of three benzonitrile herbicides by Aminobacter MSH1 versus soil microbial communities: Pathways and kinetics. Pest. Manag. Sci. 2014, 70, 1291–1298. [Google Scholar] [CrossRef]
  23. Li, W.-K.; Zhang, H.-X.; Shi, Y.-P. Simultaneous determination of bifenox, dichlobenil and diclofop methyl by hollow carbon nanospheres enhanced magnetic carboxylic multi-walled carbon nanotubes. Anal. Chim. Acta 2018, 1011, 40–49. [Google Scholar] [CrossRef] [PubMed]
  24. Van der Woude, B.J.; Gottschal, J.C.; Prins, R.A. Degradation of 2,5-dichlorobenzoic acid by Pseudomonas aeruginosa JB2 at low oxygen tensions. Biodegradation 1995, 6, 39–46. [Google Scholar] [CrossRef]
  25. Klamertha, N.; Gernjakb, W.; Malato, S.; Aguërac, A.; Lendla, B. Photo-Fenton decomposition of chlorfenvinphos: Determination of reaction pathway. Water Res. 2009, 43, 441–449. [Google Scholar] [CrossRef] [PubMed]
  26. United States Environment Protection Agency (USEPA). Prevention, pesticides and toxic substances. In Registration Eligibility Decision (RED) for Propiconazole: Case No. 3125; EPA 738R-06-027; United States Environment Protection Agency (USEPA): Washington, DC, USA, 2006. [Google Scholar]
  27. Babar, T.M.; Qadeer, G.; Rama, N.H.; Ruzickand, A.; Padelkova, Z. Methyl 2,5-dichlorobenzoate. Acta Cryst. 2008, E64, o1970. [Google Scholar] [CrossRef]
  28. European Food Safety Authority. Conclusion on pesticide peer reviev: Conclusion regarding the peer review of the pesticide risk assessment of the active substance 2,5-dichlorobenzoic acid methylester. Eur. Food Saf. Auth. Sci. Rep. 2008, 180, 1–50. [Google Scholar]
  29. Available online: https://www.lookchem.com/2-5-Dichlorobenzoic-acid/ (accessed on 21 November 2022).
  30. Pinkus, A.G.; Kautz, J.A.; Vajjula, P.A. Crystal structures of 2,6- and 3,5-dichlorobenzoic acids: Nonbonded Cl⋯Cl contacts. J. Chem. Crystallogr. 2003, 33, 181–186. [Google Scholar] [CrossRef]
  31. Fava, L.; Orrù, M.A.; Crobe, A.; Caracciolo, A.B.; Bottonia, P.; Funaria, E. Pesticide metabolites as contaminants of groundwater resources: Assessment of the leaching potential of endosulfan sulfate, 2,6-dichlorobenzoic acid, 3,4-dichloroaniline, 2,4-dichlorophenol and 4-chloro-2-methylphenol. Microchem. J. 2005, 79, 207–211. [Google Scholar] [CrossRef]
  32. Ambrose, D. The correlation and estimation of vapour pressures IV. Observations on Wagner’s method of fitting equations to vapour pressures. J. Chem. Thermodyn. 1986, 18, 45–51. [Google Scholar] [CrossRef]
  33. An, H.; Yang, W.M. A new generalized correlation for accurate vapor pressure prediction. Chem. Phys. Lett. 2012, 543, 188–192. [Google Scholar] [CrossRef]
  34. Godavarthy, S.S.; Robinson, R.L.; Gasem, K.A.M. SVRC–QSPR model for predicting saturated vapor pressures of pure fluids. Fluid Phase Equilibria 2006, 246, 39–51. [Google Scholar] [CrossRef]
  35. Yaffe, D.; Cohen, Y. Neural network-based temperature-dependent quantitative structure property relations (QSPRs) for predicting vapor pressure of hydrocarbons. J. Chem. Inf. Model. 2001, 41, 463–477. [Google Scholar]
  36. Ruzicka, V. Estimation of vapor pressures by a group-contribution method. Ind. Eng. Chem. Fundam. 1983, 22, 266–267. [Google Scholar] [CrossRef]
  37. Chickos, J.S.; Hesse, D.G.; Liebman, J.F. A group additivity approach for the estimation of heat capacities of organic liquids and solids at 298 K. Struct. Chem. 1993, 4, 261–269. [Google Scholar] [CrossRef]
  38. Pankow, J.F.; Asher, W.E. SIMPOL.1: A simple group contribution method for predicting vapor pressures and enthalpies of vaporization of multifunctional organic compounds. Atmos. Chem. Phys. 2008, 8, 2773–2796. [Google Scholar] [CrossRef]
  39. Gharagheizi, F.; Ilani-Kashkouli, P.; Acree, W.E.; Mohammadi, A.H.; Ramjugernat, D. Improved prediction of vapor pressure for pure liquids and solids from the PR+COSMOSAC equation of state. Ind. Eng. Chem. Res. 2015, 54, 10115–10125. [Google Scholar]
  40. Monte, M.J.S.; Almeida, A.R.R.P. A new approach for the estimation of sublimation enthalpies and vapor pressures of crystalline benzene derivatives. Struct. Chem. 2013, 24, 2001–2016. [Google Scholar] [CrossRef]
  41. Almeida, A.R.R.P.; Sousa, C.A.D.; Santos, L.M.N.B.F.; Monte, M.J.S. Thermodynamic properties of sublimation of the ortho and meta isomers of acetoxy and acetamido benzoic acids. J. Chem. Thermodyn. 2015, 86, 6–12. [Google Scholar] [CrossRef]
  42. Almeida, A.R.R.P.; Monte, M.J.S. Vapour pressures and phase transition properties of four substituted acetophenones. J. Chem. Thermodyn. 2017, 107, 42–50. [Google Scholar] [CrossRef]
  43. Almeida, A.R.R.P.; Monte, M.J.S. The influence of the halogen atoms and acetyl group on vapour pressures and related properties of the p-haloacetophenones. J. Chem. Thermodyn. 2016, 92, 118–125. [Google Scholar] [CrossRef]
  44. Clarke, E.C.W.; Glew, D.N. Evaluation of thermodynamic functions from equilibrium constants. Trans. Faraday Soc. 1966, 62, 539–547. [Google Scholar] [CrossRef]
  45. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Montgomery, J.A., Jr.; Vreven, T.; Kudin, K.N.; Burant, J.C.; et al. Gaussian 03, Revision, C.01; Gaussian Inc.: Wallingford, CT, USA, 2004. [Google Scholar]
  46. 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. Available online: http://cccbdb.nist.gov/vibscalejust.asp (accessed on 2 December 2022).
  47. 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. [Google Scholar] [CrossRef]
  48. 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. [Google Scholar] [CrossRef]
  49. 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. [Google Scholar] [CrossRef]
  50. 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. [Google Scholar] [CrossRef]
  51. Monte, M.J.S.; Santos, L.M.N.B.F.; Fonseca, J.M.S.; Sousa, C.A.D. Vapour pressures, enthalpies and entropies of sublimation of para substituted benzoic acids. J. Therm. Anal. Calorim. 2010, 100, 465–474. [Google Scholar] [CrossRef]
  52. Ribeiro da Silva, M.A.V.; Monte, M.J.S.; Rocha, I.M.; Cimas, Á. Energetic study applied to the knowledge of the structural and electronic properties of monofluorobenzonitriles. J. Org. Chem. 2012, 77, 4312–4322. [Google Scholar] [CrossRef] [PubMed]
  53. Altau, K.; Beasley, J.G.; Pine, H.J.; Crawford, R.H.; Tucker, W.T.; Brown, A.D.; Capps, J.D. Some derivatives of ethylbenzene. J. Chem. Eng. Data 1963, 8, 122. Available online: https://webbook.nist.gov/cgi/cbook.cgi?ID=C50793&Units=SI&Mask=4#Thermo-Phase (accessed on 21 November 2022). [CrossRef]
  54. Stork, G.; White, W.N. The stereochemistry of the SN2’ Reaction. II1. J. Am. Chem. Soc. 1956, 78, 4609. Available online: https://webbook.nist.gov/cgi/cbook.cgi?ID=C50306&Units=SI&Mask=4#ref-1 (accessed on 21 November 2022). [CrossRef]
  55. Plato, C. Differential scanning calorimetry as a general method for determining purity and heat of fusion of high-purity organic chemicals. Application to 64 compounds. Anal. Chem. 1972, 44, 1531–1534. [Google Scholar] [CrossRef]
  56. Donnelly, J.R.; Drewes, L.A.; Johnson, R.L.; Munslow, W.D.; Knapp, K.K.; Sovocool, G.W. Purity and heat of fusion data for environmental standards as determined by differential scanning calorimetry. Thermochim. Acta 1990, 167, 155–187. [Google Scholar] [CrossRef]
  57. Acree, W.E., Jr. Thermodynamic properties of organic compounds: Enthalpy of fusion and melting point temperature compilation. Thermochim. Acta 1991, 189, 37–56. [Google Scholar] [CrossRef]
  58. Rodante, F.; Vecchio, S.; Catalani, G.; Guidotti, M. Thermal analysis and non-isothermal kinetic study of some pesticides. Part II. Chlorinate derivatives. J. Therm. Anal. Calorim. 2000, 60, 605–622. [Google Scholar] [CrossRef]
  59. Kumar, G.N.A.; Hathwar, V.R. Quantitative investigation of halogen and hydrogen bonding in 2-chloro, 4-X-benzoic acids. ChemistrySelect 2022, 7, e202104338. [Google Scholar]
  60. Britton, D.; Noland, W.E.; Pinnow, M.J. Isomorphism and pseudosymmetry in 2,6-dichloro- and 2,6-dibromobenzonitrile. Acta Cryst. 2000, B56, 822–827. [Google Scholar] [CrossRef] [PubMed]
  61. Britton, D. o-Chloro- and o-bromobenzonitrile: Pseudosymmetry and pseudo-isostructural packing. Acta Cryst. 2007, C63, o14–o16. [Google Scholar] [CrossRef]
  62. Pink, M.; Britton, D.; Noland, W.E.; Pinnow, M.J. 2,4,6-Trichlorophenylisonitrile and 2,4,6-trichlorobenzonitrile. Acta Cryst. 2000, C56, 1271–1273. [Google Scholar] [CrossRef]
  63. NETZSCH Analyzing & Testing DSC 204 F1 Phoenix. Calibration Set. Selb, Germany, 2022. Available online: https://analyzing-testing.netzsch.com/_Resources/Persistent/3/f/6/2/3f62fce9e4fd19c3cb5d75f1514a90ef444f9c23/DSC_204_F1_Phoenix_en_web.pdf (accessed on 14 December 2022).
  64. 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. [Google Scholar] [CrossRef]
  65. Sabbah, R.; Xu-Wu, A.; 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. [Google Scholar] [CrossRef]
  66. Della Gatta, G.; Richarson, 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. [Google Scholar] [CrossRef]
  67. Roux, M.V.; Temprado, M.; Chickos, J.S.; Nagano, Y.J. Critically evaluated thermos chemical properties of polycyclic aromatic hydrocarbons. Phys. Chem. Ref. Data 2008, 37, 1855–1996. [Google Scholar] [CrossRef]
  68. 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. [Google Scholar] [CrossRef]
  69. Furukawa, G.T.; McCoskey, R.E.; King, G.J. Calorimetric properties of benzoic acid from 0o to 410o K. J. Res. Natl. Inst. 1951, 47, 256–261. [Google Scholar]
  70. 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. [Google Scholar] [CrossRef]
  71. Almeida, A.R.R.P.; Oliveira, J.A.S.A.; Monte, M.J.S. Thermodynamic study of nicotinamide, N-methylnicotinamide and N,N-dimethylnicotinamide: Vapour pressures, phase diagrams, and hydrogen bonds. J. Chem. Thermodyn. 2015, 82, 108–115. [Google Scholar] [CrossRef]
  72. 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. [Google Scholar] [CrossRef]
  73. 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. [Google Scholar] [CrossRef]
  74. 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. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Structural formula of the compounds studied in this work.
Figure 1. Structural formula of the compounds studied in this work.
Molecules 28 01590 g001
Figure 2. Plots of the sublimation vapor pressures (measured using the Knudsen effusion method) against reciprocal temperature for 2,4-DCBA; 2,5-DCBA, and 2,6-DCBA. , small effusion orifices; Δ, medium effusion orifices; , large effusion orifices.
Figure 2. Plots of the sublimation vapor pressures (measured using the Knudsen effusion method) against reciprocal temperature for 2,4-DCBA; 2,5-DCBA, and 2,6-DCBA. , small effusion orifices; Δ, medium effusion orifices; , large effusion orifices.
Molecules 28 01590 g002
Figure 3. Plot of the sublimation vapor pressures against reciprocal temperature for 2,6-DCBN: Δ, MKS, static method; ▲, KEM, Knudsen effusion method.
Figure 3. Plot of the sublimation vapor pressures against reciprocal temperature for 2,6-DCBN: Δ, MKS, static method; ▲, KEM, Knudsen effusion method.
Molecules 28 01590 g003
Figure 4. Phase diagram of 2,4-DCBN: , vaporization (MKS, static method); Δ, sublimation (MKS, static method); triple point (TP): T = 331.8 K; p = 48.2 Pa.
Figure 4. Phase diagram of 2,4-DCBN: , vaporization (MKS, static method); Δ, sublimation (MKS, static method); triple point (TP): T = 331.8 K; p = 48.2 Pa.
Molecules 28 01590 g004
Figure 5. Relation between Δ cr g G m o , Δ cr g H m o , and T Δ cr g S m o of the compounds studied.
Figure 5. Relation between Δ cr g G m o , Δ cr g H m o , and T Δ cr g S m o of the compounds studied.
Molecules 28 01590 g005
Figure 6. Schematic representation of the crystalline structure of 2,6-DCBA [30].
Figure 6. Schematic representation of the crystalline structure of 2,6-DCBA [30].
Molecules 28 01590 g006
Table 1. Standard (po = 105 Pa) thermodynamic properties of sublimation of all the compounds studied and of vaporization of 2,4-DCBN.
Table 1. Standard (po = 105 Pa) thermodynamic properties of sublimation of all the compounds studied and of vaporization of 2,4-DCBN.
ΔTθ Δ cd g G m o θ apb Δ cd g H m o θ a Δ cd g S m o θ cR2 Δ cd g C p , m o θ aσrd
KKkJ·mol−1PakJ·mol−1J·K−1.mol−1J·K−1·mol−1
2,4-DCBA
Crystalline phase (Knudsen effusion method)
335.2 to 357.2298.1546.42 ± 0.177.37 × 10−4111.4 ± 1.2217.9 ± 4.10.999720.9 ± 5.2 e0.0143
346.19 f36.03 ± 0.020.366110.4 ± 1.2
2,5-DCBA
Crystalline phase (Knudsen effusion method)
332.1 to 354.3298.1545.49 ± 0.101.07 × 10−3109.6 ± 0.8215.0 ± 2.70.999919.5 ± 5.3 e0.0093
343.20 f35.86 ± 0.020.348108.7 ± 0.8
2,6-DCBA
Crystalline phase (Knudsen effusion method)
321.1 to 343.3298.1541.56 ± 0.075.24 × 10−399.6 ± 0.7194.7 ± 2.40.999921.4 ± 5.1 e0.0088
332.20 f34.97 ± 0.010.31798.8 ± 0.7
2,6-DCBN
Crystalline phase (Knudsen effusion method)
295.3 to 319.2298.1533.73 ± 0.021.23 × 10−188.3 ± 0.6183.0 ± 2.00.999926.4 ± 5.5 g0.0095
307.27 f32.06 ± 0.010.35588.0 ± 0.6
Crystalline phase (static method)
328.7 to 391.8298.1533.51 ± 0.061.35 × 10−187.3 ± 0.7180.4 ± 2.41.000026.4 ± 5.5 h0.0046
360.24 f22.47 ± 0.0155.285.7 ± 0.1
2,4-DCBN
Crystalline phase (static method)
303.0 to 329.0298.1527.00 ± 0.041.8679.9 ± 0.6177.4 ± 2.00.999827.3 ± 4.9 e0.0110
315.99 f23.85 ± 0.0211.479.4 ± 0.6
331.83 i21.07 ± 0.0448.279.0 ± 0.6174.6 ± 0.9
Liquid phase (static method)
333.3 to 380.8298.1525.29 ± 0.043.7163.7 ± 0.5128.8 ± 1.71.000065.8 ± 8.5 h0.0036
357.08 f18.05 ± 0.0122959.8 ± 0.1
331.83 i21.07 ± 0.0348.261.5 ± 0.2121.8 ± 0.6
a Uncertainties are expressed as the expanded uncertainty (0.95 level of confidence, k = 2). b Calculated from Equation (1) for three different temperatures (θ represents the mean temperature of the experiments, or the temperature of the triple point, or the temperature 298.15 K). c Calculated using Equation (2); uncertainties calculated through the RSS method. d σr is the relative standard deviation of the fit, defined as σ r = i = 1 n ( ln p     ln p   c a l c )   i 2 /   ( n   m ) 1 / 2 . e Calculated as Δ cr g C p , m o ( θ ) = C p , m o ( g ) C p , m o ( cr ) . f Mean temperature. g Value derived from the fittings of Equation (1) to the (p,T) crystalline data of 2,6-DCBN, determined through the static method. h Adjustable parameter derived from the fittings of Equation (1) to the (p,T) data. Uncertainties are standard deviations of the least squares regressions. i Temperature of the triple point.
Table 3. Fusion properties: temperature, molar enthalpy, and molar entropy of the compounds studied.
Table 3. Fusion properties: temperature, molar enthalpy, and molar entropy of the compounds studied.
Ttp/KTfus/K a Δ cr l H m o ( T ) b/kJ·mol−1 Δ cr l S m o ( T ) b,c/J·K−1·mol−1Method/Ref.
2,4-DCBA
435.20 ± 0.8928.19 ± 0.44 a64.8 ± 1.0DSC/this work
2,5-DCBA
426.87 ± 0.8927.54 ± 0.50 a64.5 ± 1.2DSC/this work
426.65 [53]
427.65 [53]
2,6-DCBA
414.18 ± 0.8914.13 ± 0.26 a34.1 ± 0.6DSC/this work
415 ± 2 [54]
2,6-DCBN
416.46 ± 0.8926.19 ± 0.34 a62.9 ± 0.8DSC/this work
416.725.9462.3DSC/[55]
417.226.1762.7DSC/[56,57]
421.224.5658.3DSC/[58]
2,4-DCBN
331.46 ± 0.8917.61 ± 0.26 a53.1 ± 0.8DSC/this work
331.8 17.5 ± 0.452.8 ± 1.1static/this work
a Reported experimental uncertainties were determined from the combined standard uncertainties (which include the standard deviation of the mean of the four experimental runs and the standard uncertainty of the differential scanning calorimeter calibration) and the coverage factor k = 2 (0.95 level of confidence). b T represents the temperature of fusion or the temperature of the triple point (Ttp). c Uncertainties calculated through the RSS method.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Almeida, A.R.R.P.; Pinheiro, B.D.A.; Monte, M.J.S. Phase Transitions Equilibria of Five Dichlorinated Substituted Benzenes. Molecules 2023, 28, 1590. https://doi.org/10.3390/molecules28041590

AMA Style

Almeida ARRP, Pinheiro BDA, Monte MJS. Phase Transitions Equilibria of Five Dichlorinated Substituted Benzenes. Molecules. 2023; 28(4):1590. https://doi.org/10.3390/molecules28041590

Chicago/Turabian Style

Almeida, Ana R. R. P., Bruno D. A. Pinheiro, and Manuel J. S. Monte. 2023. "Phase Transitions Equilibria of Five Dichlorinated Substituted Benzenes" Molecules 28, no. 4: 1590. https://doi.org/10.3390/molecules28041590

Article Metrics

Back to TopTop