SciELO - Scientific Electronic Library Online

 
vol.68 número6First principle investigation of physical properties of MNiBi: (M = Sc, Y) half-Heusler compoundsPrediction of equations of state of molecular liquids by an artificial neural network índice de autoresíndice de assuntospesquisa de artigos
Home Pagelista alfabética de periódicos  

Serviços Personalizados

Journal

Artigo

Indicadores

Links relacionados

  • Não possue artigos similaresSimilares em SciELO

Compartilhar


Revista mexicana de física

versão impressa ISSN 0035-001X

Rev. mex. fis. vol.68 no.6 México Nov./Dez. 2022  Epub 31-Jul-2023

https://doi.org/10.31349/revmexfis.68.061701 

Research

Thermodynamics and Statistical Physics

Standard enthalpies of formation of 3-hydroxyphthalic anhydride

M. López-Ortegaa 

M. A. García-Castroa  * 

J. A. Galicia-Aguilara 

R. J. Aranda-Garcíaa 

H. S. Ramos-Hernándeza 

F. Díaz-Sáncheza 

a Facultad de Ingeniería Química de la Benemérita Universidad Autónoma de Puebla, 18 Sur y Av. San Claudio Ciudad Universitaria, Puebla, Pue, 72570, México.


Abstract

In this investigation thermochemical properties of 3-hydroxyphthalic anhydride were determined. Fusion enthalpy, fusion temperature and molar heat capacity in solid phase were obtained by differential scanning calorimetry. By the Knudsen effusion method, the molar enthalpy of sublimation at 298.15 K was obtained. By thermogravimetric data, it was possible to determine the molar enthalpy of sublimation and the molar enthalpy of vaporization. On the other hand, the molar enthalpy of formation in solid phase at 298.15 K was determined by combustion calorimetry. The molar enthalpy of formation in gas phase at 298.15 K was calculated from molar enthalpy of sublimation at 298.15 K and molar enthalpy of formation in solid phase at 298.15 K. This experimental value was compared both theoretical enthalpy of formation in gas phase by using Gaussian methods at the G3 and G4 levels and with an estimated value in relation to phthalic anhydride.

Keywords: 3-hydroxyphthalic anhydride; thermogravimetry; Knudsen effusion; calorimetry combustion; enthalpy of formation

1 Introduction

Phthalic anhydride derivatives have been applied in the chemical industry as plasticizers and heat resistant polymers such as polyesters, polyimides, and poly(ether imide)s [1]. One of its derivatives is the 3-hydroxyphthalic anhydride (3-HPA), which exhibits a hydroxyl group substituent in position 3, Fig. 1.

Figure 1 Chemical composition of 3-hydroxyphthalic anhydride (3-HPA). 

In recent publications it has been shown that 3-HPA can serve as a raw material to obtain drugs with potential applications, such as, anticancer [2,3], antimicrobial and in the treatment of diabetes [4,5], herpes [6-8], HIV [9,10]. For the design and production of new drugs, it is necessary to have exact and precise values of various properties of the compounds involved in the manufacturing processes [11]. Among these, the thermochemical properties, have an important role. For example, the knowledge of the standard molar enthalpies of formation of all reactants, it allows the determination of reaction heat. The standard molar enthalpy of formation in the solid and gas phase, allows the calculation of enthalpic increments related to the contribution of functional groups concerning a base chemical structure. Therefore, it was decided to determine the main thermochemical properties of 3-HPA, whose results are reported in the present work. The temperature and enthalpy of fusion and molar heat capacity in solid phase, were obtained by differential scanning calorimetry (DSC). The molar enthalpy of sublimation at T = 298.15 K was determined by Knudsen effusion method and by thermogravimetric analysis (TGA). In this last method, as 3-HPA, sublimates and vaporizes, it was possible to obtain the molar enthalpy of sublimation and vaporization at 298.15 K, using data of process of sublimation (from 400.15 K to 470.15 K) and using data of process of vaporization (from 476.15 K to 486.15 K). The values of molar enthalpy of sublimation at 298.15 K calculated by different ways, were compared. The standard molar enthalpy of formation in the solid phase was determined from standard molar enthalpy of combustion obtained by combustion calorimetry. The standard molar enthalpy of formation in gas phase at 298.15 K, Δf H m(g), obtained from the standard molar enthalpy of sublimation and formation in the solid phase at 298.15 K was compared to the theoretical enthalpy of formation in gas phase determined from Gaussian methods at the G3 and G4 levels. Finally, Δf H m(g), standard molar enthalpy of formation in gas phase of 3-HPA was estimated from the contribution of the OH functional group related to phthalic anhydride (base chemical structure). Similar works have been done in other research [11-13], allowing their use in the prediction of data for related structures.

2 Experimental

2.1 Materials and purity control

Compound 3-HPA was acquired from Sigma-Aldrich, the purity of the indicated molar fraction was > 0.98, so it was recrystallized with ethyl acetate. Once recrystallized and purified, fusion temperature and enthalpy of fusion were determined by DSC in a Perkin Elmer DSC7 device at a heating rate of 1.0 Kmin-1 and a nitrogen flow of 30.0 cm3min-1. The DSC device was calibrated for both temperature and heat flow using an Indium sample with enthalpy of 28.6 Jg-1 and temperature of 429.75 K [14]. The 3-HPA purity was calculated using the fractional fusion technique, methodology described in other works [15,16]. High purity nitrogen (x = 0.99997) was supplied by Infra Co. The heat capacity was obtained at a heating rate of 10.0 K min-1 and nitrogen flow of 30.0 cm3min-1 in a temperature range from 293.15 to 412.15 K. The calibration was performed with aluminum oxide as a reference, using the two steps method [17]. The device shown a Cp for the aluminum oxide of 78.58 Jmol-1K-1 at 300 K, 88.71 Jmol-1K-1 at 350 K, and 95.54 Jmol-1K-1 at 400 K, these values were as reported in the literature (79.43 Jmol-1K-1 at 300 K, 88.87 Jmol-1K-1 at 350 K and 96.10 Jmol-1K-1 at 400 K) [14]. Table I shows the commercial sources, initial and final purities of compounds used in this work.

Table I Supplier, initial and final purities of the samples used m this work 

Chemical name CAS number Source Initial
mole-
fraction
purity
Purification
method
Final
mole-
fraction
purity
3-HP A 37418-88-5 Sigma-
Aldrich
>0.98 Recrystallization 0.9972±
0.0012
Indium 7440-74-6 NIST 0.999999
Aluminium oxide 1344-28-1 NIST 0.9995
Pyrene 129-00-0 J.T.
Baker
>0.99 Sublimation 0.9996±
0.0003
Phenanthrene 85-01-8 J.T.
Baker
> 0.99 Sublimation 0.9997±
0.0001
Ferrocene 102-54-5 Sigma-
Aldnch
>0.99 Sublimation 0.9997±
0.0003
Anthracene 120-12-7 Sigma-
Aldrich
>0.98 Sublimation 0.9998±
0.0003

2.2 Combustion calorimetry

The combustion energy was determined by combustion calorimetry, in an isoperibolic calorimeter equipped with a semi-micro combustion bomb. This technique has been described elsewhere [18]. The calorimeter device was calibrated with benzoic acid as calorimetric standard (NIST Standard Material Reference 39j), its certified massic energy of combustion of Δc u = —(26434.0 ± 3.0) Jg-1 (where the uncertainty is the standard deviation of the mean) was corrected using the equation provided by Coops [19]. The energy equivalent was obtained from its complete oxidation, ϵ(calor) = (1281.2 ± 0.8) JK-1, the uncertainty is twice to the standard deviation of the mean of six calibration runs, data of calibration experiments was shown in previous work [20]. The combustion of 3-HPA was carried out under the same experimental conditions as those mentioned above for the calibration procedure, with a sample weight of 40 mg and a high purity oxygen inlet pressure of 3.04 MPa (x = 0.99999, supplied by Airgas Co). The combustion chamber was purged with oxygen for five minutes to ensure an oxidizing atmosphere; no nitric acid corrections were made. In each experiment 0.1 mL of water was added to the bomb, whose volume is 0.022 L. Cotton thread fuse was used, with a composition C1.000H1.742O0.921 and a combustion energy of Δc u = —(146945.2 ± 4.2) Jg-1 (the uncertainty is the same as standard deviation of the mean) [21]. In each experiment an ignition unit (Parr 2901) was used with an energy of 4.184 J.

The temperature measurement was carried out using a thermistor (5610 Hard Scientific), the resistance was registered with a digital multimeter HP 34420 A. Table II shows the physical properties of the materials used in the combustion experiment for this work.

Table II Physical properties at = 0.1 MPa 

Compound Mg mol-1 a ρg cm-3 -δu/δpTJ g-1MPa-1 Cpcr298.15KJ g-1K-1
3-HPA 164.116 1.624±0.060 b 0.200 c 1.136 ±0.025 d
Benzoic acid 122.123 1.320 c 0.115 c 1.209 c
Cotton 28.502 1.500 c 0.289 c 1.674 c

a Based on the 2013 IUPAC recommendation [22].

b Calculated using Advanced Chemistry Development (ACD/Labs) Software VI 1.02.

c Estimated value in reference at T=298.15 K [23].

d Experimental average value from two experiments using DSC device. Its uncertainty corresponds to expanded uncertainty with a level of confidence of approximately 95%. Include the contributions from the calibration and u(T)=0.l K. The experiments were realized under average atmospheric pressure (78.8 kPa), u(p) = 1 kPa.

The standard mass energy of combustion of 3-HPA was determined by (1):

Δcu(3-HPA)=ΔUIBP+ΔUcorr-mΔcu(cotton)m(3-HPA), (1)

where ΔUIBP is the energy of the isothermal bomb process, included the ignition energy ΔUign, ΔUcorr is the correction to standard state (at p = 0.1 MPa of the pure species), mΔc u(cotton) is the combustion energy of the fuse and m(3 — HPA) is the compound’ mass.

2.3 Thermogravimetry using the Langmuir method

The enthalpy of sublimation was determined from Langmuir equation, according to the procedure described by Price [24]:

dmdt1A=pα(M/2πRT)1/2, (2)

where (dm/dt)(1/A) is the rate of mass loss per area of vaporization A, p is the vapor pressure at absolute temperature T, M is the molar mass, R is the gas constant, and α is the vaporization coefficient (assumed to be 1 under vacuum conditions or macromolecules). To simplify the equation, it can be expressed as:

p=kυ, (3)

where k=(2πR)1/2α and υ=(dm/dt)(1/A)(T/M)1/2. According to the Clausius-Clapeyron’s equation, it is possible to determine the enthalpy of phase change from the vapor pressure. Combining the Clausius-Clapeyron and Langmuir equations, it is possible to write (4):

lnv=C-ΔαβHmR1T, (4)

where C groups the involved constants and ΔαβHm is the enthalpy of phase change at average temperature. In the TGA, if the mass loss is carried out at lower than melting temperature, this allows the enthalpy of sublimation to be evaluated. Otherwise, if the loss of mass occurs at higher than melting temperature, this allows the enthalpy of vaporization to be evaluated. In this work it was possible to calculate both the vaporization enthalpy and sublimation enthalpy from the same TGA thermogram of compound 3-HPA, at average temperature Tm and to 298.15 K. The TGA was carried out in a simultaneous TGA/DSC TA Instruments SDT Q600, the device’s features have been described elsewhere [25]. The calibration procedure was performed in mass with a standard provided and certified by the National Institute of Standards and Technology (NIST) as (315.162 ± 0.0048) mg. In addition, it was calibrated in temperature and heat flow with high purity Indium, NIST certified temperature of (429.7485 ± 0.00034) K and enthalpy of fusion of (28.51 ± 0.19) Jg-1. The calibrating constant was 1.0050, determined from the enthalpy of fusion. The experimental heating rate in TGA/DSC was 10.0 K min-1 and the nitrogen flow of 100.0 cm3min-1. The heating rate was determined from different tests of heating rates and nitrogen flow. These tests were performed with pyrene and phenanthrene calculating vaporization enthalpy, whose experimental values were ΔlgHm(298.15K)=(89.5±1.4) kJmol-1 and ΔlgHm(298.15K)=(79.5±1.4) kJmol-1, they agreed with the results available in the literature [26-34].

2.4 Knudsen effusion method

According to the Knudsen effusion equation it is possible to evaluate the vapor pressure from rate of mass loss:

p=ΔmΔtw0A02πRTM1/2, (5)

where Δmt is the rate of mass loss at temperature T in the aluminum cell, w0 is the Clausing probability factor, A0 is the effusion area, R is the ideal gas constant, T is the temperature at which the loss of mass takes place and M is the molar mass. Using the above equation and the Clausius-Clapeyron equation it is possible to evaluate the enthalpy of sublimation:

lnv=lnB-ΔcrgHmR1T, (6)

where v is (Δmt)T1/2, ln B includes the integration constant and the (1/A0w0)(2πR/M), ΔcrgHm is the enthalpy of sublimation at average temperature, the details of experimental method has been described elsewhere [11]. The enthalpy of sublimation was determined from the mass loss (Δm) as a function of time (Δt) at constant temperature, in a range temperature from 332.60 K to 342.28 K with 2.0 K steps and 10-6 Torr vacuum pressure. The experiments were carried out in pierced aluminum cells, with a silver disk. The dimensions of disk and effusion hole are: Cell A (diameter: 1.345 mm, A0: 1.42 mm2, thickness: 0.150 mm, w0 factor: 0.9228), Cell B (diameter: 1.031 mm, A0: 0.835 mm2, thickness: 0.135 mm, w0 factor: 0.9106) and Cell C (diameter: 0.842 mm, A0: 0.557 mm2, thickness: 0.135 mm, w0 factor: 0.8927). The pumping system consists of two VARIAN vacuum pumps (model DS 102 and model V70D). The temperature was registered by a thermistor Hart Scientific model A1143-01, connected to a multimeter (Agilent model 34420A). The validation of the method was carried out by determining the sublimation enthalpy of two standards, ferrocene and anthracene, whose experimental values were ΔcrgHm(298.15K)=(73.4±2.2) kJmol-1 and ΔcrgHm(298.15K)=(104.2±3.2) kJmol-1, they agreed with the results available in the literature [26,35-44].

2.5 Computational details

The Gaussian —3 and —4 theories were occupied to determine the energies at 0 K and the enthalpies at 298.15 K of the compounds studied from ab initio molecular orbital calculations and from Gaussian 09 [45-48]. In turn, the enthalpies of formation in gas phase at 298.15 K were calculated from these energies using the G3 and G4 methods and applying the considerations made by Nicolaides [49]. The methodology has already been used and validated by our research group [11].

3 Results and discussion

The molar fraction after purification of 3-HPA by recrystallization increased to a value greater than 0.99 in DSC. Hermetic aluminum cells were utilized for the enthalpy and melting temperature due to the loss of mass at lower temperatures of melting temperature. It is worth nothing that only the melting temperature of 3-HPA is reported in the literature, but not the other thermochemical properties. Table III resumes the melting temperatures, ranging around 471.15 K, the data reported in this work agrees with the values reported by other authors.

Table III. Melting temperature of 3-HPA from literature and this work 

Purity Tfus (K) Δcr1HmTfuskJ mol-1 Cp,mcr,298.15KJmol-1K-1 Method Reference
NA 471.15-472.15 NA NA NA 50
NA 472.15-473.15 NA NA NA 51
NA 472.15-474.15 NA NA NA 52
NA 471.15-473.15 NA NA NA 53
0.9972±0.0012 474.16±0.45 36.40±0.21 186.4±4.2 DSC Tins work

The uncertainties correspond to the expanded uncertainty with a level of confidence of approximately 95%, including uncertainty of calibration and u(T) = 0.1 K. The experiments were made under average atmospheric pressure (78.8 kPa). u(P) = 1 kPa.

The heat capacity as a function of the temperature was determined in the range from 293.15 K to 412.15 K from DSC data. The numerical coefficients of (7) were obtained from a polynomial regression applied to a set of Cp,m(cr, Jmol-1K-1) vs T(K) data. The experiments were carried out in duplicate. Table IV shows three of six values of specific energies of combustion at 298.15 K, experimentally obtained.

Cp,m(cr)=3.3677×10-7T4-4.5504×10-4T3+2.2903×10-1T2-5.0404×101T+4.2541×103(r2=0.9999). (7)

Table IV Combustion experiments for 3-HPA: at 298.15 K and p° = 0.1 MPa 

Exp. 1 Exp. 2 Exp. 3
m (3-HPA)/g 0.0390125 0.0382843 0.0387241
m (cotton)/g 0.0005212 0.0006911 0.0006135
m (platinum)/g 0.2317748 0.2314025 0.2310838
Ti/K 298.1549 298.156 298.154
Tf/K 298.7846 298.781 298.7789
ΔTcorr/K 0.053441 0.058209 0.054869
ΔTc/K 0.576259 0.566791 0.570031
ε(calor)(-ΔTc)/kJ -0.73830 -0.72617 -0.73032
ε(cont)(-ΔTc)/kJ -0.00067 -0.00066 -0.00066
ΔUign/kJ 0.00418 0.00418 0.00418
ΔUIBP/kJ -0.73479 -0.72265 -0.7268
(-ΔUcorr)/kJ 0.00081 0.00080 0.00080
(-mΔcuo)(cotton)/kJ 0.00883 0.01171 0.0104
(-mΔcuo)(3-HPA)/kJ 0.72515 0.71014 0.7156
Δcuo(3-HPA)/kJ g-1 -18.5876 -18.5491 -18.4794
Average value<-18.5270±0.0160>

m is the mass corrected for buoyancy using densities listed in Table 2 : ΔTc is the corrected temperature rise calculated by ΔTc=Tf-Ti- ΔTcorr, ε(cont) is energy equivalent of the contents of the bomb calculated by ε(cont) (-ΔTc)= εi(cont)(Ti-298.15K)+ εf(cont)(298.15K-Tf+ ΔTcorr), ΔUign is the ignition energy, ΔUIBP the energy of the isothermal bomb process, which was calculated by ΔUIBP=[ε(calor)(-ΔTc)+εi(cont)-(Ti-298.15 K)+εf(cont) (298.15 K - TfTcorr)+ΔUign], εi(cont) and εf(cont) the energy equivalents of the bomb contents m the initial and final states., respectively. ΔUcorr is the correction to standard state and Δcuo (3-HPA) is the 3-Hydroxyphthalic anhydride mass energy of combustion. Its uncertainty corresponds to expanded uncertainty with a confidence level 95 % approximately.

Table V resumes the energy and enthalpy of combustion and the enthalpy of formation in solid phase of 3-HPA. The enthalpy of combustion Δc Hm at T = 298.15 K was calculated from the energy of combustion Δc Um at the same conditions. The enthalpy of formation in solid phase of 3-HPA, Δf Hm(cr), was determined from Δc Hm at T = 298.15 K and from enthalpies of formation of the products of the idealized reaction, Eq. (8). The utilized values of the enthalpies of formation in standard state at T = 298.15 K were —(285.83 ± 0.04) kJmol-1 for H2O (l) and —(939.51 ± 0.13) kJmol-1 for CO2 (g) [54].

C8H4O4(cr)+7O2(g)=8CO2(g)+2H2O(l). (8)

Table V Standard molar energy and enthalpy of combustion and standard molar enthalpy of formation in solid phase at 298.15 K 

ΔcUmokJmol-1 ΔcHmokJmol-1 ΔfHmokJmol-1
-3040.6±5.6 a -3038.1±5.6 a -681.6±5.7 b

a The uncertainties correspond to the expanded uncertainty with confidence level of 95% approximately, they include the calibration contributions from benzoic acid and energy of combustion of cotton thread.

b The uncertainty corresponds to the expanded uncertainty with confidence level of 95 % approximately, it includes the uncertainties of standard enthalpy of formation of H2O (l) and CO2 (g).

The rate of mass loss as a function of temperature (dm/dt) was determined by using TGA. Two thermal processes, from 400.15 K to 470.15 K and from 476.15 K to 486.15 K, corresponding to the sublimation and vaporization processes, were observed. In the first temperature range (sublimation process), the enthalpy of sublimation was calculated at average temperature of Tm = 435.15 K, Table VI. This latter value was corrected to 298.15 K according to (9).

ΔcrgHm(298.15  K)/kJmol-1=ΔcrgHm(Tm/K)+ΔcrgCp,mTm/K-298.15  K, (9)

Table VI Sublimation and vaporization enthalpies for 3-HPA 

ΔαβHm(Tm/K)kJmol-1 Interval of
T/K
Tm/K ΔαβHm(298.15K)kJmol-1 Method
102.9±0.4 a c 400.15-470.15 425.15 16.8±0.4 a d TGA, sublimation process
103.0±1.8 a c 332.60-342.28 336.42 104.1±3.6 a d Knudsen efussion
62.8±1.3 b c 476.15-486.15 481.15 75.2 ±2.6 b d TGA, vaporization process

a Sublimation enthalpy.

b Vaporization enthalpy.

c The weighted average value (μ) and its standard deviation (σ) were calculated as μ = Σ(xii2)/Σ(1/σi2) and σ2 = N[1/Σ(1/σi2] respectively, where xi is each one of the N enthalpies of sublimation and vaporization data, and σi is its respective standard deviation. The experiments of sublimation and vaporization were realized under average atmospheric pressure (78.8 kPa), u(P) = 1 kPa.

d The uncertainties presented corresponds to twice the combined standard.

where ΔcrgCp,m(kJmol-1K-1)=0.75+0.15[Cp,m(cr,298.15K)/Jmol-1K-1]/1000. The heat capacity for solid phase is shown in Table III.

In turn, this sublimation enthalpy was compared with that obtained via Knudsen effusion, where solid-gas equilibrium can be guaranteed. The effusion process was performed in a range temperature from 332.60 K to 342.28 K, Table VI. The enthalpy of sublimation at Tm = 336.42 K was corrected to 298.15 K through the Eq. (9). The values of the sublimation enthalpies calculated by using both methods, are shown in Tables VI and VII. On the other hand, the enthalpy of vaporization at average temperature of 481.15 K, ΔlgHm(481.15K), was determined in the second temperature range of TGA thermogram from 476.15 K to 486.15 K (vaporization process). The vaporization enthalpy at average temperature was corrected at 298.15 K, according to Eq. (10). This value is shown in the Table VI.

ΔlgHm(298.15K)/kJmol-1=ΔlgHm(481.15K)+ΔlgCp,m(481.15K-298.15K), (10)

where

ΔlgCp,m(kJmol-1K-1)=10.58+0.26[Cp,m(l,298.15K)/Jmol-1K-1]/1000

Cp,m(l,298.15K)=(Cp,m(g,298.15K)+10.58)/0.74

Table VII Comparison of enthalpies of sublimation ΔCIgHm (298.15 K) 

Δcr1Hm(298.15K)kJmol-1 ΔlgHm(298.15K)kJmol-1
TGA
ΔcrgHm(298.15K)kJmol-1
Knudsen efussion
ΔcrgHm(298.15K)kJmol-1
TGA
Δcr1Hm298.15K+ΔlgHm(298.15K)kJmol-1
29.5±0.5 75.2±2.6 104.1±3.6 106.8±0.8 104.7±2.6

a Value calculated by equation:

Δcr1Hm298.15K=Δcr1HmTfus/K+ΔcrgCp,mTfus/K-298.15-ΔlgCp,mTfus/K-298.15 where ΔcrgCp,mkJ  mol-1K-1=0.75+0.15Cp,mcr,298.15K/J  mol-1K-1/1000 and ΔlgCp,mkJ  mol-1K-1=10.58+0.26Cp,m1, 298.15K/J  mol-1K-1/1000.

All the uncertainties correspond to twice the combined standard.

The heat capacity for gas phase at Jmol-1 K-1 was evaluated taking into the vibrational frequencies determined from B3LYP /6-311G++** theory level, Eq. (11).

Cp,m(g)=1.7233×10-9(T)4-2.6011×10-6(T)3+1.0857×10-3(T)2+2.9765×10-1(T)+2.2131×101(r2=0.9999). (11)

From Table VIII, comparing the enthalpy of sublimation at 298.15 K obtained by Knudsen effusion and thermogravimetric data, at lower temperatures of the melting temperature, it can be observed that there is a difference between both of 2.7 kJmol-1. Likewise, it is observed that by adding the vaporization enthalpy with the fusion enthalpy at 298.15 K, the estimated value of the sublimation enthalpy presents a difference of 0.6 relative to the value obtained by Knudsen effusion technique. This procedure was validated with pyrene (with a difference of 0.3 kJmol-1) and phenanthrene (with a difference of 0.7 kJmol-1).

Table VIII Standard molar enthalpies of formation and sublimation of 3-HPA at 298.15 K 

ΔfH  m0(cr)kJmol-1 ΔcrgHm(298.15K)kJmol-1 ΔfH  m0(g)kJmol-1
-681.6±5.7 104.1±3.6 -577.5±6.7

The uncertainty corresponds to the expanded uncertainty with a level of confidence of approximately 95%.

The enthalpy of formation in gas phase was determined from Eq. (12) and was summarized in Table VIII.

ΔfHm(g,298.15K)=ΔfHm(cr,298.15K)+ΔcrgHm(298.15K). (12)

The theoretical enthalpy of formation in gas phase of 3-HPA at 298.15 K was determined using the energies obtained by G3 and G4 methods [47,55-57]. These energies were applied in atomization and isodesmic reactions and are shown in Tables IX and X.

Table IX Energies calculated at 0 K from G3 and G4 method for reference compounds used in this work a 

Compound G3 G4
E0 H298 E0 H298
C -37.827717 -37.825356 -37.834168 -37.831808
H -0.501003 -0.498642 -0.501420 -0.499060
0 -75.030991 -75.02863 -75.045501 -75.043141
3-HPA -607.835396 -607.825585 -607.943830 -607.934210
Benzene -232.052179 -232.04673 -232.093978 -232.088577
Methane -40.457626 -40.453810 -40.465299 40.461483
Ethane -79.723397 -79.718911 -79.738111 -79.733661
Ethene -78.507414 -78.503411 -78.521873 -78.517874
Maleic anhydride -379.086863 -379.080673 -379.155839 -379.149685
Methanol -115.629212 -115.624922 -115.651768 -115.64749
Phthalic anhydride -532.63707 -532.628329 -532.732303 -532.723658
Succinic anhydride -380.298147 -380.291243 -380.367586 -380.360719

a hartrees (Eh). 1 Eh,=2625.46 kJ mol-1

Table X Theoretical enthalpy of formation in gas phase for 3-HPA and reference compounds at 298.15 K a b 

Compound G3 G4 Exp. c
ΔfHo(g) ΔfHo(g) ΔfHo(g)
3-HPA -577.50 (0.00) -570.43 (-7.07) -577.50 d
Benzene 85.33 (-2.91) 84.74 (-2.32) 82.42
Methane -75.93 (1.04) -74.68 (-0.21)) -74.89
Ethane -85.26(1.41) -83.16 (-0.69) -83.85
Ethene 51.67 (0.80) 52.20 (0.27) 52.47
Maleic anhydride -406.10 (5.69) -402.09(1.68) -400.41
Methanol -201.35 (-0.32) -200.95 (-0.72) -201.67
Phthalic anhydride 384.70(13.01) -379.29 (7.60) -371.69
Succinic anhydride -531.24 (3.34) -526.18 (-1.72) -527.90

a Values in kJ moL-1. b The values in parentheses are the differences between the experimental value and the calculated. c Values obtained from [58]. d Value determined in this work.

The atomization reaction (i) and the isodesmic reactions (from ii to v) for 3-HPA (C8H4O4) occupied in the present research are expressed below:

(i) C8H4O4= 8C(g)+4H(g)+4O(g)

(ii) 3-HPA + ethene + methane = benzene + maleic anhydride + methanol

(iii) 3-HPA + methane = methanol + phthalic anhydride

(iv) 3-HPA + ethane + methane = benzene + succinic anhydride + methanol

(v) 3-HPA + 8 methane = ethene + ethanol + dimethyl ether + acetone

Table XI shows theoretical enthalpies of formation in gas phase of 3-HPA at 298.15 K, evaluated from both atomization reaction and isodesmic reactions [57], more approximate to the experimental value determined at the G3 level, the values in parentheses are the differences between experimental and theoretical results, all values are in kJmol-1.

Table XI Comparison between experimental and calculated enthalpy of formation in gas phase at 298.15 K for the 3-HPA 

G3 G4 G3 G4
Atomization reaction Isodesmic reaction Exp.
-577.50 (0.00) -570.43 (-7.07) i -576.88 (-0.62) -571.85 (-5.65) -577.50
ii -565.85 (-11.65) -563.34 (-15.16)
iii -579.84 (2.34) -574.29 (-3.21)
iv -577.59 (0.09) -567.76 (-9.74)

Finally, another option to estimate the molar enthalpy of formation in gas phase of 3-HPA was from the replacement of a proton with an OH group relative to phthalic anhydride. To this end it was necessary to calculate the contribution of this group taking into the enthalpic change from benzoic acid [58] to 2-hydroxy benzoic acid [59] and from methyl benzoate [60] to methyl 2-hydroxybenzoate [61], Figure 2. As can be seen, the average value of the contribution of the OH group of these two routes was —(202.2 ± 3.9) kJmol-1. Taking this value and the enthalpy of formation in gas phase at 298.15 K of phthalic anhydride [62], the estimate enthalpy of formation in gas phase at 298.15 K of 3-HPA was calculated, giving a value of ΔfHmm(g)=-(573.2±6.3) kJmol-1. It can be observed that there is a difference between the molar enthalpy in gas phase at 298.15 K, estimated and experimental, of 4.3 kJmol-1.

Figure 2 Increments enthalpics, ΔΔf Hm(g), related to the alcohol functional group. 

4 Conclusions

To conclude, thermochemical properties of 3-hydroxyphthalic anhydride, were obtained from differential scanning calorimetry, thermogravimetric analysis, and Knudsen effusion. The melting temperature was within the range reported by other authors. The enthalpy of sublimation determined by thermogravimetric data and Knudsen effusion were compared, getting a difference of 2.7 kJmol-1 and 0.6 kJmol-1, in sublimation and vaporization process. The theorical and estimate standard molar enthalpy of formation in gas phase of the 3-HPA showed excellent concordance regarding to the experimental value.

Acknowledgments

We are grateful for the facilities provided by the Departamento de Química (CINVESTAV), and by VIEP project number 100420555-VIEP2021 “Mecanosíntesis y estudio termoquímico de la o-diimida diácida derivada del anhídrido ftálico”.

References

1. U. Edlund and A. C. Albertsson, Polyesters based on diacid monomers, Adv. Drug Delivery Rev. 55 (2003) 585, https://doi.org/10.1016/S0169-409X(03)00036-X. [ Links ]

2. B. Jiang et al., Development of dual and selective degraders of cyclin-dependent kinases 4 and 6, Angew. Chem. Int. Ed. 58 (2019) 6321, https://doi.org/10.1002/anie.201901336. [ Links ]

3. C. Steinebach et al., Homo-PROTACs for the chemical knockdown of cereblon, ACS Chem. Biol. 13 (2018) 2771, https://doi.org/10.1021/acschembio.8b00693. [ Links ]

4. N. S. Gray et al., Compounds for tau protein degradation, WO Pat. 2019014429; PCT Int. Appl. (2019). [ Links ]

5. E. Hyeon et al., Synthesis of Phthalimide Derivatives as Potential PPAR-γ Ligands, Mar. Drugs. 14 (2016) 112, doi: 10.3390/md140601123. [ Links ]

6. S. Jiang et al., Virucidal and antibacterial activities of 3-HP-β LG. Molecular Approaches to the Control of Infectious Diseases, Int. Congr. Ser. 97 (1997) 327. [ Links ]

7. A. R. Neurath, N. Strick, and Li YY, 3-Hydroxyphthaloyl β-lactoglobulin. III. Antiviral activity against herpesviruses, Antivir. Chem. Chemoth. 9 (1998) 177, https://doi.org/10.1177/095632029800900209. [ Links ]

8. A. Oevermann, M. Engels, U. Thomas, and A. Pellegrini, The antiviral activity of naturally occurring proteins and their peptide fragments after chemical modification, Antivir. Res. 59 (2003) 23, https://doi.org/10.1016/S0166-3542(03)00010-X. [ Links ]

9. Li M, J. Duan, Qiu J, Yu F, Che X, S. Jiang, and Li L, 3-Hydroxyphthalic anhydride-modified human serum albumin as a microbicide candidate against HIV type 1 entry by targeting both viral envelope glycoprotein gp120 and cellular receptor CD4, AIDS. Res. Hum. Retrov. 29 (2013) 1455, https://doi.org/10.1089/aid.2013.0043. [ Links ]

10. Li L, Qiu J, Lu L, An S, P. Qiao, S. Jiang, and Liu S, 3-Hydroxyphthalic anhydride-modified human serum albumin as a microbicide candidate inhibits HIV infection by blocking viral entry, J. Antimicrob. Chemoth. 68 (2013) 573, https://doi.org/10.1093/jac/dks458. [ Links ]

11. M. A. García-Castro, P. Amador, J. M. Hernández-Pérez, A. E. Medina-Favela, and H. Flores, Experimental and computational thermochemistry of 3- and 4-nitrophthalic anhydride, J. Phys. Chem. A 118 (2014) 3820, https://doi.org/10.1021/jp5003929. [ Links ]

12. M. A. García-Castro et al., J. Chem. Thermodyn. 127 (2018) 117, https://doi.org/10.1016/j.jct.2018.07.026. [ Links ]

13. K. Salas-López, P. Amador, A. Rojas, F. J. Melendez, and H. Flores, Experimental and theoretical thermochemistry of the isomers 3- and 4-nitrophthalimide, J. Phys. Chem. A 121 (2017) 5509, https://doi.org/10.1021/acs.jpca.7b02508. [ Links ]

14. R. Sabbah, Xu-wu, J. S. Chickos, M. L. Planas Leitão, M. V. Roux, and L. A. Torres, Reference materials for calorimetry and differential thermal analysis, Thermochim. Acta. 331 (1999) 93, https://doi.org/10.1016/S0040-6031(99)00009-X. [ Links ]

15. C. Plato, and A. R. Glasgow, Differential scanning calorimetry as a general method for determining the purity and heat of fusion of high-purity organic chemicals. Application to 95 compounds, J. Anal. Chem. 41 (1969) 330, https://doi.org/10.1021/ac60271a041. [ Links ]

16. M. E. Brown, Determination of purity by differential scanning calorimetry (DSC), J. Chem. Educ. 56 (1979) 310, https://doi.org/10.1021/ed056p310. [ Links ]

17. P. Góralski, M. Tkaczyk, and M. Chorazewski, Heat capacities of α, w-dichloroalkanes at temperatures from 284.15 K to 353.15 K and a group additivity analysis, J. Chem. Eng. Data. 48 (2003) 492, https://doi.org/10.1021/je020042y. [ Links ]

18. E. A. Camarillo, and H. Flores, Construction, calibration and testing of a micro-combustion calorimeter, J. Chem. Thermodyn. 38 (2006) 1269, https://doi.org/10.1016/j.jct.2006.03.001. [ Links ]

19. J. Coops, R. S. Jessup, K. G. van Nes, F. D. Rossini, Experimental Thermochemistry (Interscience, NY, 1956). [ Links ]

20. K. Salas-López et al., Thermochim. Acta . 697 (2021) 178861, https://doi.org/10.1016/j.tca.2021.178861. [ Links ]

21. S. P. Verevkin, Thermochemistry of amines: strain in six-membered rings from experimental standard molar enthalpies of formation of morpholines and piperazines, J. Chem. Thermodyn. 30 (1998) 1069, https://doi.org/10.1006/jcht.1998.0371. [ Links ]

22. J. Meija et al., Atomic weights of the elements 2013 (IUPAC Technical Report), Pure Appl. Chem. 88 (2016) 265, https://doi.org/10.1515/pac-2015-0305. [ Links ]

23. W. D. Good, and N. K. Smith, Enthalpies of combustion of toluene, benzene, cyclohexane, cyclohexene, methylcyclopentane, 1-methylcyclopentene, and n-hexane, J. Chem. Eng. Data 14 (1969) 102, https://doi.org/10.1021/je60040a036. [ Links ]

24. D. M. Price, Vapor pressure determination by thermogravimetry, Thermochim. Acta 367 (2001) 253, https://doi.org/10.1016/S0040-6031(00)00676-6. [ Links ]

25. M.T. Vieyra-Eusebio, and A. Rojas, Vapor Pressures and Sublimation Enthalpies of Nickelocene and Cobaltocene Measured by Thermogravimetry, J. Chem. Eng. Data 56 (2011) 5008, https://doi.org/10.1021/je200815v. [ Links ]

26. F. Ramos, J. M. Ledo, H. Flores, E. A. Camarillo, J. Carvente, and M. P. Amador, Evaluation of sublimation enthalpy by thermogravimetry: Analysis of the diffusion effects in the case of methyl and phenyl substituted hydrantoins, Thermochim. Acta 655 (2017) 181 , https://doi.org/10.1016/j.tca.2017.06.024. [ Links ]

27. J. J. H. Haftka, J. R. Parsons, and H. A. J Govers, Supercooled liquid vapor pressures and related thermodynamic properties of polycyclic aromatic hydrocarbons determined by gas chromatography, J. Chromatrogr. A 1135 (2006) 91, https://doi.org/10.1016/j.chroma.2006.09.050. [ Links ]

28. K. Nass, D. Lenoir, and A. Kettrup, Calculation of the Thermodynamic Properties of Polycyclic Aromatic Hydrocarbons by an Incremental Procedure, Angew. Chem. Int. Ed. Engl. 34 (1995) 1735, https://doi.org/10.1002/anie.199517351. [ Links ]

29. F. S. Mortimer, and R. V. Murphy, The Vapor Pressures of Some Substances Found in Coal Tar. Ind. Eng. Chem. 15 (1923) 1140, https://doi.org/10.1021/ie50167a012. [ Links ]

30. J. S. Chickos, and W. Hanshaw, Vapor Pressures and Vaporization Enthalpies of the n-Alkanes from C21 to C30 at T = 298.15 K by Correlation Gas Chromatography, J. Chem. Eng. Data 49 (2004) 77, https://doi.org/10.1021/je0301747. [ Links ]

31. J. S. Chickos, and W. Hanshaw, Vapor Pressures and Vaporization Enthalpies of the n-Alkanes from C31 to C38 at T = 298.15 K by Correlation Gas Chromatography, J. Chem. Eng. Data 49 (2004) 620, https://doi.org/10.1021/je030236t. [ Links ]

32. J. S. Chickos, D. Hesse, S. Hosseini, G. Nichols, and P. Webb, Sublimation enthalpies at 298.15 K using correlation gas chromatography and differential scanning calorimetry measurements, Thermochim. Acta 313 (1998) 101, https://doi.org/10.1016/S0040-6031(97)00432-2. [ Links ]

33. K. Ruzicka, and V. Majer, Simultaneous Treatment of Vapor Pressures and Related Thermal Data Between the Triple and Normal Boiling Temperatures for n-Alkanes C31 to C38, J. Phys. Chem. Ref. Data 23 (1994) 1, https://doi.org/10.1063/1.555942. [ Links ]

34. J. S. Chickos, S. Hosseini, and D. G. Hesse, Determination of vaporization enthalpies of simple organic molecules by correlations of changes in gas chromatographic net retention times, Thermochim. Acta 249 (1995) 41, https://doi.org/10.1016/0040-6031(95)90670-3. [ Links ]

35. W. Acree, and J. S. Chickos, Phase Transition Enthalpy Measurements of Organic and Organometallic Compounds. Sublimation, Vaporization and Fusion Enthalpies From 1880 to 2010, J. Phys. Chem. Ref. Data 45 (2016) 033101, https://doi.org/10.1063/1.4948363. [ Links ]

36. A. Rojas, and E. Orozco, Measurement of the enthalpies of vaporization and sublimation of solids aromatic hydrocarbons by differential scanning calorimetry, Thermochim. Acta 405 (2003) 93, https://doi.org/10.1016/S0040-6031(03)00139-4. [ Links ]

37. R. Sabbah, A. Xu-wu, J. S. Chickos, M. L. Planas Leitāo, M. V. Roux, and L. A. Torres, Reference materials for calorimetry and differential thermal analysis. Thermochim. Acta 331 (1999) 93, https://doi.org/10.1016/S0040-6031(99)00009-X. [ Links ]

38. A. Rojas, and M. T. Vieyra-Eusebio, Enthalpies of sublimation of ferrocene and nickelocene measured by calorimetry and the method of Langmuir, J. Chem. Thermodyn. 43 (2011) 1738, https://doi.org/10.1016/j.jct.2011.06.001. [ Links ]

39. R. S. Bradley, and T. G. Cleasby, The vapour pressure and lattice energy of some aromatic ring compounds, J. Chem. Soc. 169 (1953) 1690, https://doi.org/10.1039/JR9530001690. [ Links ]

40. C. G. De Kruif, Enthalpies of sublimation and vapour pressures of 11 polycyclic hydrocarbons, J. Chem. Thermodyn. 12 (1980) 243, https://doi.org/10.1016/0021-9614(80)90042-7. [ Links ]

41. L. Malaspina, R. Gigli, and G. Bardi, Microcalorimetric determination of the enthalpy of sublimation of benzoic acid and anthracene, J. Chem. Phys. 59 (1973) 387, https://doi.org/10.1063/1.1679817. [ Links ]

42. P. C. Hansen, and C. A. Eckert, An improved transpiration method for the measurement of very low vapor pressures, J. Chem. Eng. Data 31 (1986) 1, https://doi.org/10.1021/je00043a001. [ Links ]

43. V. P. Klochkov, The vapor pressure of some aromatic compounds, Zh. Fiz. Khim 32 (1985) 1177. [ Links ]

44. K. V. Zherikova, and S. P. Verevkin, Ferrocene: Temperatura adjustments of sublimation and vaporization enthalpies, Fluid Phase Equilibr. 472 (2018) 196, https://doi.org/10.1016/j.fluid.2018.05.004. [ Links ]

45. L.A. Curtiss, K. Raghavachari, P.C. Redfern, V. Rassolov, and J.A. Pople, Gaussian-3 (G3) theory for molecules containing first and second-row atoms, J. Chem. Phys. 109 (1998) 7764, https://doi.org/10.1063/1.477422. [ Links ]

46. L.A. Curtiss, P.C. Redfern, and K. Raghavachari, Gaussian-4 theory, J. Chem. Phys. 126 (2007) 084108, https://doi.org/10.1063/1.2436888. [ Links ]

47. W.J. Hehre, L. Radom, P.V.R. Schleyer, and J.A. Pople, Ab Initio, Molecular Orbital Theory, (Wiley, New York, 1986). [ Links ]

48. M.J. Frisch et al., Gaussian 09, Revision A.02, Gaussian, Inc., Wallingford, CT, 2009. [ Links ]

49. A. Nicolaides, A. Rauk, M. N. Glukhovtsev, and L. Radom, Heats of Formation from G2, G2(MP2), and G2(MP2,SVP) Total Energies, J. Phys. Chem. 100 (1996) 17460, https://doi.org/10.1021/jp9613753. [ Links ]

50. D. S. Pratt, and G. A. Perkins, Phthalic acid derivatives; constitution and color, v.1 some derivatives of tetrachloro- and tetraiodo phthalimides, J. Am. Chem. Soc. 40 (1918) 198, https://doi.org/10.1021/ja02234a023. [ Links ]

51. E. L. Eliel, A. W. Burgstahler, D. E. Rivard, and L. Haefele, The Methyl Esters of 3-Hydroxyphthalic Acid. Selective Reduction of Monomethyl Phthalates with Lithium Aluminum Hydride, J. Am. Chem. Soc. 77 (1955) 5092, https://doi.org/10.1021/ja01624a042. [ Links ]

52. J. A. H. Näsman, A Versatile Synthetic Route to 3-Hydroxyphtalic Anhydride, Syntesis 8 (1985) 788, 10.1055/s-1985-31350. [ Links ]

53. R. J. Petfield, and E. D. Amstutz, Halogen reactivities. VI. The reactivities of severalL α-bromofurans. The isolation of 2-methoxyfuran, J. Org. Chem. 19 (1954) 1944, https://doi.org/10.1021/jo01377a011. [ Links ]

54. J. D. Cox, D. D. Wagman, and V. A. Medvedev, CODATA key values for thermodynamics (Hemisphere Pub. Corp. 1989). [ Links ]

55. R. Notario, O. Castano, J. Abboud, R. Gomperts, L.M. Frutos, and R. Palmeiro, Organic thermochemistry at high Ab initio levels. 1. A G2(MP2) and G2 study of cyclic saturated and unsaturated hydrocarbons (including aromatics). J. Org. Chem. 64 (1999), 9011, https://doi.org/10.1021/jo990898e. [ Links ]

56. R. Notario, O. Castano, R. Gomperts, L.M. Frutos, and R. Palmeiro, Organic thermochemistry at high AB initio levels. 3. A G3 study of cyclic saturated and unsaturated hydrocarbons (including aromatics). J. Org. Chem. 65 (2000) 4298, https://doi.org/10.1021/jo000089r. [ Links ]

57. K. Raghavachari, and B.B. Stefanov, L.A. Curtiss, Accurate thermochemistry for larger molecules: Gaussian-2 theory with bond separation energies. J. Chem. Phys. 106 (1997) 6764, https://doi.org/10.1063/1.473659. [ Links ]

58. J. B. Pedley, Thermochemical data and structures of organic compounds, (Thermodynamics Research Center, 1994). [ Links ]

59. R. Sabbah, and T. H. D. Le, Ètude thermodynamique des trois isomères de l’acide hydroxybenzoïque, Can. J. Chem. 71 (1993) 1378, https://doi.org/10.1139/v93-178. [ Links ]

60. W. V. Steele, R. D. Chirico, A. B. Cowell, S. E. Knipmeyer, and A. Nguyen, Thermodynamic Properties and Ideal-Gas Enthalpies of Formation for Methyl Benzoate, Ethyl Benzoate, (R)-(+)-Limonene, tert-Amyl Methyl Ether, trans-Crotonaldehyde, and Diethylene Glycol, J. Chem. Eng. Data 47 (2002) 667, https://doi.org/10.1021/je0100847. [ Links ]

61. J. M. Ledo, Experimental and theoretical study of methyl n-hydroxybenzoates, J. Chem. Thermodyn. 124 (2018) 1, https://doi.org/10.1016/j.jct.2018.04.011. [ Links ]

62. V. L. S. Freitas, C. P. F. Santos, M. D. M. C. Ribeiro da Silva, and M. A. V. Ribeiro da Silva, The effect of ketone groups on the energetic properties of phthalan derivatives, J. Chem. Thermodyn. 96 (2016) 74, https://doi.org/10.1016/j.jct.2015.12.018. [ Links ]

Received: January 01, 2022; Accepted: February 09, 2022

Creative Commons License This is an open-access article distributed under the terms of the Creative Commons Attribution License