Next Article in Journal
PRG3 and PRG5 C-Termini: Important Players in Early Neuronal Differentiation
Next Article in Special Issue
CADMA-Chem: A Computational Protocol Based on Chemical Properties Aimed to Design Multifunctional Antioxidants
Previous Article in Journal
The Potential Role of the Methionine Aminopeptidase Gene PxMetAP1 in a Cosmopolitan Pest for Bacillus thuringiensis Toxin Tolerance
Previous Article in Special Issue
The Mayo Clinic Florida Microdosimetric Kinetic Model of Clonogenic Survival: Application to Various Repair-Competent Rodent and Human Cell Lines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Research of a Thermodynamic Function (px)T, x0: Temperature Dependence and Relation to Properties at Infinite Dilution

1
College of Chemistry and Chemical Engineering, Ocean University of China, Qingdao 266100, China
2
Institute of Chemical Engineering, Guangdong Academy of Science, Guangzhou 510665, China
3
College of Chemistry and Chemical Engineering, Qingdao University, Qingdao 266071, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(21), 12998; https://doi.org/10.3390/ijms232112998
Submission received: 12 September 2022 / Revised: 12 October 2022 / Accepted: 24 October 2022 / Published: 27 October 2022
(This article belongs to the Special Issue Feature Papers in Physical Chemistry and Chemical Physics 2022)

Abstract

:
In this work, we propose the idea of considering ( p x ) T ,   x 0 as an infinite dilution thermodynamic function. Our research shows that ( p x ) T , x 0 as a thermodynamic function is closely related to temperature, with the relation being simply expressed as: ln ( p x ) T ,   x 0 = A T + B . Then, we use this equation to correlate the isothermal vapor–liquid equilibrium (VLE) data for 40 systems. The result shows that the total average relative deviation is 0.15%, and the total average absolute deviation is 3.12%. It indicates that the model correlates well with the experimental data. Moreover, we start from the total pressure expression, and use the Gibbs–Duhem equation to re-derive the relationship between ( p x ) T , x 0 and the infinite dilution activity coefficient ( γ ) at low pressure. Based on the definition of partial molar volume, an equation for ( p x ) T , x 0 and gas solubility at high pressure is proposed in our work. Then, we use this equation to correlate the literature data on the solubility of nitrogen, hydrogen, methane, and carbon dioxide in water. These systems are reported at temperatures ranging from 273.15 K to 398.15 K and pressures up to 101.325 MPa. The total average relative deviation of the predicted values with respect to the experimental data is 0.08%, and the total average absolute deviation is 2.68%. Compared with the Krichevsky–Kasarnovsky equation, the developed model provides more reliable results.

1. Introduction

Knowledge of the phase-equilibrium behavior of solutions at infinite dilution is important not only for phase-equilibrium calculations of a dilute solution, but also for high-concentration solutions. Indeed, the phase-equilibrium properties of the dilute region are widely used in the process design of unit operations such as distillation, extraction, and absorption [1,2,3]. Hence, the study of infinite dilution thermodynamic properties is a crucial topic. There are two functions used to describe thermodynamic properties at infinite dilution, i.e., the infinite dilution activity coefficient ( γ ) and Henry’s constant ( H i ). The former can be used to calculate relative volatility, partition coefficients, and phase-equilibrium data [4,5,6,7]. The latter is used to describe the partitioning capacity of a compound in the gas/water phase, mainly calculating the solubility of the gas in liquids [8,9,10].
Several experimental techniques are available to determine γ , among which the static differential technique and differential ebulliometry are common means. In the static differential method, the pressure difference between two static units of pure solvent and dilute solution is measured in a constant temperature bath. Then, the difference is used to plot the composition-pressure diagram of the liquid phase at equilibrium to obtain ( p x ) T , x 0 . However, a differential ebulliometer is used to measure the change in the boiling point of a solvent to obtain ( T x ) p , x 0 . In these methods, γ is calculated from the relationship derived from Gautreaux and Coates [11]. Earlier studies have indicated that accurate data from these two methods can be obtained if strictly considering the liquid and vapor hold-ups in various parts of the systems [12,13]. Experimental data for Henry’s constant can be calculated using independently measured solubility and partial pressure [10]. Moreover, it can be obtained by the static differential technique. Ayuttaya et al. [14] reported a method for determining H i with a differential static cell. The results suggested that the pressure accuracy influenced the uncertainty of the determined Henry’s constant.
As one of the thermodynamic properties at infinite dilution, the value of γ is closely related to temperature. Currently, the theoretical model of γ and temperature is derived from the relationship between the activity coefficient and the partial molar excess function, which is expressed as: ln γ i = Δ h i E , R T Δ s i E , R . Assuming infinite dilution molar entropy ( Δ h E , ) and infinite dilution molar enthalpy ( Δ s E , ) are constants within a certain temperature range, the above relationship expresses a two-parameter model [7,15,16]. The subsequent three-parameter model [17], four-parameter model [18], and five-parameter model with the introduction of molecular descriptors Xi [19] were developed from this equation. Notably, although the introduction of Xi improves the correlation accuracy, this equation contains five variables, which complicates the relation between γ and temperature.
The pressure effect has a negligible effect on H i when the pressure is low. In this case, H i is closely related to the temperature. Considerable research efforts have been devoted to the two-parameter model [20,21]. The results of the correlation between H i and temperature were not satisfactory for systems with a wide range of temperatures. When considering the change in dissolution heat with temperature, the relationship between H i and temperature was developed into a three-parameter model and a four-parameter model [22]. Although the correlation accuracy of these two models is high over a wide temperature range, the relationship between H i and temperature becomes complicated due to too many parameters, which is not favorable for engineering applications. However, the effect of pressure on H i at high pressures is not negligible. Hence, it is necessary to study H i at high pressures. Thus far, much work has focused on γ and H i , where both γ and H i can be obtained by experimental determination of ( p x ) T , x 0 .
In this study, we consider ( p x ) T , x 0 as an infinite dilution thermodynamic function. With this viewpoint, we propose a novel correlation model of ( p x ) T , x 0 with temperature. Based on the calculation results of 40 systems, the novel model is proved to be objective and rational. The outstanding feature of this model is that there are only two adjustable parameters, which can be more easily adopted in engineering applications. Furthermore, as a thermodynamic property, ( p x ) T , x 0 can be used to calculate other thermodynamic properties of infinite dilution, such as γ and H i . Considering the relationship between ( p x ) T , x 0 and H i at high pressures, a new model is proposed to calculate the solubility of the gas in water. The main feature of this model is that the solubility at a high pressure can be expressed as the property ( v p ) T , x 0 of the pure solvent and the property ( p x ) T , x 0 of the mixture at a low pressure, which constitutes a substantial addition to our understanding of an infinitely diluted solution from a thermodynamic viewpoint.
The rest of this paper is organized as follows. Section 2 gives details of the theory regarding the relationship between ( p x ) T , x 0 and other infinite dilution thermodynamic functions. Section 3 discusses the source of ( p x ) T , x 0 . We propose a model to describe the relationship between ( p x ) T , x 0 and temperature. Then, all proposed models are validated using literature data. Finally, our conclusion is stated in Section 4.

2. Theory

2.1. Definition of ( p x ) T , x 0

For a binary isothermal vapor–liquid equilibrium system, ( p x ) T , x 0 is given by the partial derivative of p with respect to x , where p is the total pressure and x is the molar fraction of solute in the liquid phase. It is notable that x tends to zero. For convenience, the term x in the latter is equivalent to x 1 , and the solute is denoted as component “1”, while the solvent is component “2”. It can not only describe the nonideality of the solution but also reflect the interaction between the solute and solvent at infinite dilution.

2.2. Relationship between ( p x ) T , x 0 and γ

For a binary system, the total pressure equation [23] at a low pressure can be described as Equation (1):
p = x 1 γ 1 p 1 s + x 2 γ 2 p 2 s
where p 1 s is the saturated vapor pressure of the solute, p 2 s is the saturated vapor pressure of the solvent. It is important to emphasize that this equation applies to a low pressure. At a constant temperature, the activity coefficients γ 1 and γ 2 are only a function of the composition. For binary systems, the relationship between the two components is described as:
x 1 + x 2 = 1
where x 1 is the molar fraction of component 1 in the liquid phase, x 2 is the molar fraction of component 2 in the liquid phase. Therefore, taking the partial derivative of x 1 at isothermal conditions, we find:
( p x 1 ) T = ( γ 1 x 1 ) T p 1 s x 1 + γ 1 p 1 s ( γ 2 x 1 ) T p 2 s x 1 γ 2 p 2 s + ( γ 2 x 1 ) T p 2 s
According to the expression of the Gibbs–Duhem equation for the binary system:
x 1 dln γ 1 + x 2 dln γ 2 = h E R T 2 d T + V E R T d p
where h E is the excess enthalpy, V E is the excess volume, R is the universal gas constant, while the temperature remains a constant, d T = 0 . The liquids can be considered as an ideal solution at infinite dilution [24,25], V E = 0 . It is convenient to rewrite Equation (4) as:
x 1 dln γ 1 + x 2 dln γ 2 = 0
Dividing both sides by d x 1 , Equation (5) can be represented as follows:
x 1 dln γ 1 d x 1 + x 2 dln γ 2 d x 1 = 0
The above equation is equivalent to:
x 1 γ 1 d γ 1 d x 1 + x 2 γ 2 d γ 2 d x 1 = 0
From Equation (7), as x 1 approaches zero, x 2 γ 2 d γ 2 d x 1 0 . Based on the normalized description of the activity coefficient: as the molar fraction of component 2 approaches 1, its activity coefficient converges to 1. Thus, when x 1 0 , ( γ 1 x 1 ) T p 1 s x 1 0 , ( γ 2 x 1 ) T 0 , substituting Equation (7) into Equation (3), it becomes:
( p x 1 ) T , x 1 0 = γ 1 p 1 s p 2 s
Notice that Equation (8) is consistent with the formula reported by Gautreaux and Coates at a low pressure [11]. The difference is that the starting point in our work is the expression of the total pressure at a low pressure. The physical meaning of the derivation in our work is clear and its assumptions are accessible. This is because V E R T is usually considered as a negligible factor in many studies, and Equation (5) is a common formula that can be used to examine activity coefficient experimental data for thermodynamic consistency [26,27,28].

2.3. Relationship between ( p x ) T , x 0 and H i

Henry’s law is usually described as a limit law, that is:
H 1 = lim x 1 0 f 1 L x 1
When considering an ideal solution based on Raoult’s law to define the activity coefficient, the liquid phase fugacity can be expressed as:
f 1 L = x 1 γ 1 f 1
where f 1 is the fugacity of the pure liquid at solution temperature and pressure. Thus, combining Equations (9) and (10), that is:
H 1 = lim x 1 0 γ 1 f 1
At low pressures, f 1 is replaced by the saturated vapor pressure of the liquid p 1 s . Hence, Equation (11) becomes [7,29]:
H 1 = γ 1 p 1 s
Thus, the expression of the relationship between ( p x ) T , x 0 and H 1 is obtained directly from Equations (8) and (12):
( p x 1 ) T , x 1 0 = H 1 p 2 s
Obviously, Equation (13) indicates that there is a simple relationship between ( p x ) T , x 0 and H 1 . In particular, when the solvent saturation vapor pressure p 2 s is low, Equation (10) becomes:
( p x 1 ) T , x 1 0 = H 1
This reveals that ( p x ) T , x 0 is Henry’s coefficient in some special conditions. Therefore, it is appropriate to consider it as a thermodynamic property alone.

2.4. Relationship between ( p x ) T , x 0 and Solubility at High Pressures

The above consideration represents the case of a low pressure. However, the influence of the pressure effect on Henry’s constant cannot be ignored when the pressure is high. Therefore, it is necessary to extend the expressions of ( p x 1 ) T , x 1 0 and H 1 under high pressures.
From strict thermodynamic relations, we can obtain:
( ln f 1 L p ) T , x = v ¯ 1 R T
where v ¯ 1 is the partial molar volume of the solute in liquids. Substituting this result into Equation (9), we get:
( ln H 1 p ) T = v ¯ 1 R T
where v ¯ 1 is the partial molar volume of the solute at infinite dilution. Assuming that f 1 is proportional to x 1 at constant temperature and pressure, a more general form of Henry’s law can be obtained:
ln f 1 x 1 = ln H 1 ( p 0 ) + p 0 p v ¯ 1 d p R T
where p 0 is the reference pressure, p 0 = 101.325 kPa, H 1 ( p 0 ) is the Henry’s constant at 101.325 kPa. By definition, v ¯ 1 can be written as:
v ¯ 1 = ( p V ) T , p , n 2
Thus, the partial molar volume, as defined by Equation (18), can be evaluated using the following transformation:
v ¯ 1 = ( p / n 1 ) T , V , n 2 ( p / V ) T , n 1 , n 2
where ( p V ) T , n 1 , n 2 describes the properties of solvents, which can be obtained by the equation of state. Since V changes near the saturated volume at infinite dilution, ( p n 1 ) T , V , n 2 can be written as ( p n 1 ) T , x 1 0 . Thus:
( p n 1 ) T , x 1 0 = ( p x 1 ) T , x 1 0 ( x 1 n 1 ) T , x 1 0
Substituting Equation (19) to Equation (20) gives:
v ¯ 1 = ( p x 1 ) T , x 1 0 ( v p ) T , x 1 0
where ( p x 1 ) T , x 1 0 represents the properties of the solute at a low pressure. Generally, it is difficult to determine v ¯ 1 by experiment. In this work, v ¯ 1 is expressed as the properties of the solvent at a high pressure and the solute at a low pressure, which provides a new idea for obtaining v ¯ 1 .
Hence, an equation for ( p x ) T , x 0 and solubility at high pressures can be written as:
ln f 1 x 1 = ln H 1 ( p 0 ) 1 R T ( p x 1 ) T , x 1 0 p 0 p ( v p ) T , x 1 0 d p
Considering the variation in the solute activity coefficient with molar composition, we added the activity coefficient for correction. Therefore, the final equation can be written as follows:
ln f 1 x 1 = ln H 1 ( p 0 ) 1 R T ( p x 1 ) T , x 1 0 p 0 p ( v p ) T , x 1 0 d p + ln γ 1

2.5. Relationship between ( p x ) T , x 0 and ( T x ) p , x 0

Differential ebulliometry [30] uses ( T x ) p , x 0 , while the differential statics method [31] uses ( p x ) T , x 0 to calculate γ . Similarly, regarding the properties of an infinitely diluted solution, there should be a certain relationship between ( T x ) p , x 0 and ( p x ) T , x 0 .
According to the Gibbs phase rule, pressure is a function of temperature and composition for a binary system:
p ( T , x ) = 0
d p = ( p T ) x d T + ( p x ) T d x
When d p is equal to zero at a constant pressure, we get:
( p T ) x ( d T ) p + ( p x ) T ( d x ) p = 0
Notably, ( p T ) x converges to d p 2 s d T as x approaches zero. Hence, Equation (26) yields the desired result:
( p x ) T , x 0 = ( d p 2 s d T ) x 0 ( T x ) p , x 0
Equation (27) is consistent with the descriptions of Van Ness and Abbott [32], which reflects the strict relationship between ( p x ) T , x 0 and ( T x ) p , x 0 . Several additional points warrant explanation. In general, ( d p 2 s d T ) x 0 > 0 , ( p x ) T , x 0 , and ( T x ) p , x 0 have opposite signs. The available experimental data show that ( T x ) p , x 0 < 0 ; thus, ( p x ) T , x 0 should be positive. In addition, it is clear from Equation (27) that ( p x ) T , x 0 and ( T x ) p , x 0 can be extrapolated from each other if ( d p 2 s d T ) x 0 is accurate. Generally, ( T x ) p , x 0 is relatively easy to measure [30], and ( p x ) T , x 0 can be obtained by determining ( T x ) p , x 0 using Equation (27). Therefore, as an important thermodynamic property, ( p x ) T , x 0 can be measured not only directly but also indirectly by ( T x ) p , x 0 .

3. Results and Discussion

3.1. Source of ( p x ) T , x 0

We select the isothermal vapor–liquid equilibrium (VLE) data [33,34,35,36,37] with the pressure as a function of the molar fraction for regression at low concentration ranges ( x 1 < 0.05 ), typically 3–7 points. As shown in Figure 1, the VLE data of water in the tert-amyl alcohol system are regressed using the least squares method, where x represents the molar fraction of water in tert-amyl alcohol. It was found that p is linearly related to x , which can be written as Equation (28). More interestingly, Equation (28) can also be applied to other systems. Thus, the slope a is ( p x ) T , x 0 at a constant temperature.
p = a x + b

3.2. Correlation of ( p x ) T , x 0 with Temperature

As previously noted, there is a simple relationship between ( p x ) T , x 0 and H i in Equations (13) and (14). In particular, when p 2 s is very low, ( p x ) T , x 0 is equal to H i .
Several excellent studies which describe the relationship between H i and temperature are all two-parameter models [7,20,38], with the relations being expressed as follows:
ln H 1 = c T + d
Therefore, we propose a relationship between ( p x ) T , x 0 and temperature which can be written as follows:
ln ( p x ) T ,     x 0 = A T + B
To prove the correctness of Equation (29), acetaldehyde–water and water–tert-amyl-alcohol systems are correlated by this relationship, as illustrated in Figure 2. The results show that the correlation coefficient R 2 > 0.99 , which indicates a good linear relationship between ln ( p x ) T ,   x 0 and the inverse of temperature.
Forty groups of isothermal VLE data containing 156 values of ( p x ) T , x 0 are used to test the correlation model in this work, including 20 groups of organic–aqueous systems, 5 groups of aqueous–organic systems, and 15 groups of organic–organic systems. The correlation of ( p x ) T , x 0 and temperature is performed using Equation (30). The average deviations of model parameters and experimental data of the correlation result equation are listed in Table 1. We use two deviations to obtain conclusions about the accuracy of the given model [39], where the average absolute deviation (A.A.D. (%)) is calculated based on Equation (31) and the average relative deviation (A.R.D. (%)) is calculated based on Equation (32).
A . A . D . ( % ) = 100 N p i = 1 N p [ | ( p x ) T ,     x 0 , exp . ( p x ) T ,     x 0 , cal . | ( p x ) T ,     x 0 , exp . ] i
A . R . D . ( % ) = 100 N p i = 1 N p [ ( p x ) T ,     x 0 , exp . ( p x ) T ,     x 0 , cal . ( p x ) T ,     x 0 , exp . ] i
Satisfactory results are obtained from the correlation of aqueous solution systems of 2-butanone and ethyl acrylate, and the values of A.A.D. (%) for these systems are less than 1%. The calculation results of the model are accurate for the correlation of small molecule solution systems, especially for aqueous solutions of organic compounds. However, the largest deviation is found for the methyl-acetate–ethanol, with an A.A.D. (%) of 6.69%. Such deviation occurs because the mole fraction of methyl acetate in ethanol is close to 0.05 and the solution concentration is high. Hence, a more accurate value of ( p x ) T , x 0 can be obtained as x approaches 0, when the correlation between ( p x ) T , x 0 and temperature improves.

3.3. Calculate ( p x ) T , x 0 from ( T x ) p , x 0

In addition to the correlation model, we can also determine ( T x ) p , x 0 by using ( p x ) T , x 0 . The isobaric VLE data [33,34,35] of eight systems are selected to calculate their ( T x ) p , x 0 at different temperature and pressure in this work, and ( p x ) T , x 0 is calculated from Equation (27), where p s is calculated by the Antoine equation with parameters from the NIST database (https://webbook.nist.gov/, accessed on 28 March 2020). The results indicate that the calculation values are in good agreement with the values of ( p x ) T , x 0 obtained from isothermal VLE data, as shown in Table 2. It is noteworthy that for systems with low relative volatilities [40], ( p x ) T , x 0 is not easy to determine, but ( T x ) p , x 0 can be well measured by differential ebulliometry. In this case, an accurate value of ( p x ) T , x 0 can be calculated from ( T x ) p , x 0 .

3.4. Use ( p x ) T , x 0 to Calculate Other Infinite Dilution Properties

As a thermodynamic property, the value of ( p x ) T , x 0 can be calculated from γ and H i as shown in Equations (8) and (13). In this section, these two infinite dilution thermodynamic functions are calculated by ( p x ) T , x 0 from isothermal VLE data. Moreover, the accuracy of Equations (22) and (23) are verified using high-pressure solubility data for different systems.

3.4.1. Using ( p x ) T , x 0 to Calculate γ

We use isothermal VLE data [33,34,35] with the pressure as a linear function of the molar fraction to calculate ( p x ) T , x 0 , and then obtain γ from Equation (8). The results are shown in Table 3. According to the results of the calculations, most systems are well adapted to the literature, while a few systems such as nitromethane–water show significant differences. This result may be related to the accuracy of the γ experimental value; many researchers reported that γ measured by different methods showed significant differences [18,41]. Moreover, Sherman et al. [42] evaluated the database of γ for non-electrolytes in water and found that the accuracy of the database for measured values is estimated at 10% for γ < 1000 .

3.4.2. Using ( p x ) T , x 0 to Calculate H i

We calculate ( p x ) T , x 0 from the isothermal VLE data [33,34,35] with the pressure as a linear function of the molar fraction, and we use Equation (13) to obtain H i . The calculation results are shown in Table 4. As shown, the calculated results are in good agreement with the literature data, which illustrates the accuracy of the relationship between ( p x ) T , x 0 and H i . It can be seen that the saturation vapor pressure of water at 288.15 K is very low, close to zero. Meanwhile, the value of Henry’s constant for acetaldehyde in water is approximately equal to ( p x ) T , x 0 . This proves that ( p x ) T , x 0 is reasonable as a thermodynamic property.

3.5. Use ( p x ) T , x 0 to Calculate the Solubility of the Gas at High Pressure

As a traditional model, the Krichevsky–Kasarnovsky (K–K) equation is often used to express the effect of high pressures on gas solubility [46], given by:
ln f 1 x 1 = ln H 1 ( p 0 ) + v ¯ 1 ( p p 0 ) R T
where p 0 is the reference pressure, and H 1 ( p 0 ) is Henry’s constant at p 0 .
In this work, we deduce formulas for solubility and ( p x ) T , x 0 at high pressure, as shown in Equations (22) and (23). In both formulae, ( v p ) T is the property of pure water, which can be calculated by the equations of IAPWS-IF97 [47]. In Figure 3, we compared the relationship between solubility and fugacity using different models for four aqueous systems. The blue solid lines indicate the best fits to the experimental data (“this work 2”), while pink dotted lines represent K–K equation results. It can be seen that both the proposed model and the K–K equation can describe the solubility of hydrogen and carbon dioxide well in water at high pressures. In particular, the model in this work is superior to the K–K equation for the nitrogen and methane systems. Moreover, Figure 3 shows that the model with the activity coefficient in this work, as illustrated Equation (23), can better describe the effect of high pressures on solubility, especially for systems with pressures up to 101.325 MPa. This suggests that the solubility of gas in water increases at high pressures, and the activity coefficient of the solute in water cannot be ignored.
In addition, four typical systems are selected for comparison with the proposed model, with temperatures ranging from 273.15 K to 398.15 K and pressures up to 101.325 MPa. The average deviations of the calculated results concerning experimental data in this work are listed in Table S1.
The results show that the A.A.D. (%) of the model with the activity coefficient for 13 groups is as low as 2.68%, and the model prediction of this paper is more accurate compared with that of the K–K equation. In this study, we find that the activity coefficient of the solute is closely related to the composition. The parameters between the activity coefficient and the composition are shown in Table 5. It is worth noting that for hydrogen and nitrogen systems, the relationship between the activity coefficient and the composition can be simply expressed as a linear relationship. However, for systems with relatively strong interactions with water, such as carbon dioxide and methane systems, it can be expressed as a quadratic functional relationship.
.

4. Conclusions

In this paper, we propose the idea of considering ( p x ) T , x 0 as a thermodynamic function, which reflects the interaction of the solute and solvent in the dilute region. Based on this idea, a model is proposed to describe the relationship between ( p x ) T , x 0 and temperature in our work. We find that ln ( p x ) T , x 0 is linearly related to the reciprocal of temperature. The values of parameters A and B are reported based on isothermal VLE data from 40 groups of small-molecule systems. The results show that the total A.R.D. (%) of the calculated results compared with the experimental data is 0.15%. This suggests that the proposed model matches the experimental data well.
In addition, we re-derive the relationship between ( p x ) T , x 0 and γ at low pressures, in terms of the Gibbs–Duhem equation. We also describe the relationship between ( p x ) T , x 0 and other infinite dilution thermodynamic functions. Based on these relationships, we can obtain the value of ( p x ) T , x 0 from γ or H i data, or by using isothermal VLE data and differential ebulliometry. Furthermore, we put forward a new solubility model considering the relationship between ( p x ) T , x 0 and H i at high pressures. The parameters of the proposed model are determined according to the experimental data of four systems. Comparing the calculated results with the experimental data, the A.A.D. (%) of the proposed model is 2.68%, and the A.R.D. (%) of the proposed model is only 0.08%. Compared to the classical K–K model, the new model can correlate the solubility data more accurately. Overall, these results give a comprehensive understanding of infinite dilution properties, which can facilitate improvements in promising applications for chemical process design.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms232112998/s1. Table S1. Calculated results of the proposed model [46,48,49,50,51].

Author Contributions

Conceptualization, J.Z. and Y.H.; data curation, W.S.; formal analysis, J.Z. and Y.W.; funding acquisition, L.W.; methodology, J.Z., Y.H., and X.C.; resources, W.Z. and Y.W.; validation, Q.S. and X.C.; visualization, W.Z.; writing—original draft, J.Z.; writing—review and editing, X.C., Y.W., and L.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Nature Science Fund Program of China (21776264).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data generated or analyzed during this study are available within the article or upon request from the corresponding author.

Acknowledgments

My sincere appreciation goes to Hu and Wu who provided scientific guidance.

Conflicts of Interest

The authors declare no conflict of interests.

Nomenclature

A ,   B , C , D , E , a , b , c , d the model constants for the model
f 1 L fugacity of component 1 in the liquid phase [kPa]
H i Henry’s constant of solute i in the solvent [kPa]
h E excess enthalpy [kJ]
Δ h i E , infinitely dilute molar enthalpy [kJ]
N p the number of experimental points
n 1 the molar composition of component 1 [mol]
p total pressure [kPa]
p s saturated vapor pressure [kPa]
R universal gas constant [J·mol−1·K−1]
Δ s i E , infinitely dilute molar entropy [kJ·mol−1·K−1]
T temperature [K]
V volume of a system [m3]
V E excess volume [m3]
v partial molar volume [cm3·mol−1]
v ¯ 1 partial molar volume of component 1 [cm3·mol−1]
x i liquid phase mole fraction of component i
γ the activity coefficient of component i at infinite dilution

References

  1. Eckert, C.A.; Newman, B.A.; Nicolaides, G.L.; Long, T.C. Measurement and Application of Limiting Activity Coefficients. AIChE J. 1981, 27, 33–40. [Google Scholar] [CrossRef]
  2. Chen, Z.; Zhang, H.; Li, H.; Xu, Y.; Shen, Y.; Zhu, Z.; Gao, J.; Ma, Y.; Wang, Y. Separation of N-Heptane and Tert-Butanol by Ionic Liquids Based on Cosmo-Sac Model. Green Energy Environ. 2021, 6, 380–391. [Google Scholar] [CrossRef]
  3. Lei, Z.G.; Wang, H.Y.; Zhou, R.Q.; Duan, Z.T. Influence of Salt Added to Solvent on Extractive Distillation. Chem. Eng. J. 2002, 87, 149–156. [Google Scholar] [CrossRef]
  4. Bao, J.B.; Han, S.J. A Method for the Estimation of Infinite Dilution Activity Coefficient for a Real Solution by Monte Carlo Simulation. Acta Phys.-Chi. Sin. 1995, 11, 568–571. [Google Scholar]
  5. Brouwer, T.; Kersten, S.R.A.; Bargeman, G.; Schuur, B. Trends in Solvent Impact on Infinite Dilution Activity Coefficients of Solutes Reviewed and Visualized Using an Algorithm to Support Selection of Solvents for Greener Fluid Separations. Sep. Purif. Technol. 2021, 272, 118727. [Google Scholar] [CrossRef]
  6. Paduszyński, K.; Królikowska, M. Interactions between Molecular Solutes and Task-Specific Ionic Liquid: Measurements of Infinite Dilution Activity Coefficients and Modeling. J. Mol. Liq. 2016, 221, 235–244. [Google Scholar] [CrossRef]
  7. Reza, J.; Trejo, A. Temperature Dependence of the Infinite Dilution Activity Coefficient and Henry’s Law Constant of Polycyclic Aromatic Hydrocarbons in Water. Chemosphere 2004, 56, 537–547. [Google Scholar] [CrossRef]
  8. Li, J.Q.; Shen, C.Y.; Wang, H.M.; Han, H.Y.; Zheng, P.C.; Xu, G.H.; Jiang, H.H.; Chu, Y.N. Dynamic Measurements of Henry’s Lawconstant of Aromatic Compounds Using Proton Transfer Reaction Mass Spectrometry. Acta Phys.-Chi. Sin. 2008, 24, 705–708. [Google Scholar]
  9. Murad, S.; Gupta, S. A Simple Molecular Dynamics Simulation for Calculating Henry’s Constant and Solubility of Gases in Liquids. Chem. Phys. Lett. 2000, 319, 60–64. [Google Scholar] [CrossRef]
  10. Schwardt, A.; Dahmke, A.; Köber, R. Henry’s Law Constants of Volatile Organic Compounds between 0 and 95 °C–Data Compilation and Complementation in Context of Urban Temperature Increases of the Subsurface. Chemosphere 2021, 272, 129858. [Google Scholar] [CrossRef]
  11. Gautreaux, M.F.; Coates, J. Activity Coefficients at Infinite Dilution. AIChE J. 1955, 1, 496–500. [Google Scholar] [CrossRef]
  12. Raal, J.D.; Ramjugernath, D. Rigorous Characterization of Static and Dynamic Apparatus for Measuring Limiting Activity Coefficients. Fluid Phase Equilibria 2001, 187, 473–487. [Google Scholar] [CrossRef]
  13. Tabai, S.; Rogalski, M.; Solimando, R.; Malanowski, S.K. Activity Coefficients of Chlorophenols in Water at Infinite Dilution. J. Chem. Eng. Data 1997, 42, 1147–1150. [Google Scholar] [CrossRef]
  14. Ayuttaya, P.C.N.; Rogers, T.N.; Mullins, M.E.; Kline, A.A. Henry’s Law Constants Derived from Equilibrium Static Cell Measurements for Dilute Organic-Water Mixtures. Fluid Phase Equilibria 2001, 185, 359–377. [Google Scholar] [CrossRef]
  15. Singh, S.; Bahadur, I.; Naidoo, P.; Redhi, G.; Ramjugernath, D. Application of 1-Butyl-3-Methylimidazolium Bis (Trifluoromethylsulfonyl) Imide Ionic Liquid for the Different Types of Separations Problem: Activity Coefficients at Infinite Dilution Measurements Using Gas-Liquid Chromatography Technique. J. Mol. Liq. 2016, 220, 33–40. [Google Scholar] [CrossRef]
  16. Mgxadeni, N.; Mmelesi, O.; Kabane, B.; Bahadur, I. Influence of Hydrogen Bond Donor on Zinc Chloride in Separation of Binary Mixtures: Activity Coefficients at Infinite Dilution. J. Mol. Liq. 2022, 351, 118596. [Google Scholar] [CrossRef]
  17. Bernauer, M.; Dohnal, V. Temperature Dependences of Limiting Activity Coefficients and Henry′s Law Constants for N-Methylpyrrolidone, Pyridine, and Piperidine in Water. Fluid Phase Equilibria 2009, 282, 100–107. [Google Scholar] [CrossRef]
  18. Dohnal, V.; Fenclová, D.; Vrbka, P. Temperature Dependences of Limiting Activity Coefficients, Henry’s Law Constants, and Derivative Infinite Dilution Properties of Lower (C1–C5) 1-Alkanols in Water. Critical Compilation, Correlation, and Recommended Data. J. Phys. Chem. Ref. Data 2006, 35, 1621–1651. [Google Scholar] [CrossRef]
  19. Xi, L.; Sun, H.; Li, J.; Liu, H.; Yao, X.; Gramatica, P. Prediction of Infinite-Dilution Activity Coefficients of Organic Solutes in Ionic Liquids Using Temperature-Dependent Quantitative Structure–Property Relationship Method. Chem. Eng. J. 2010, 163, 195–201. [Google Scholar] [CrossRef]
  20. Staudinger, J.; Roberts, P.V. A Critical Review of Henry’s Law Constants for Environmental Applications. Crit. Rev. Environ. Sci. Technol. 1996, 26, 205–297. [Google Scholar] [CrossRef]
  21. Chen, F.; Freedman, D.L.; Falta, R.W.; Murdoch, L.C. Henry’s Law Constants of Chlorinated Solvents at Elevated Temperatures. Chemosphere 2012, 86, 156–165. [Google Scholar] [CrossRef]
  22. Heron, G.; Christensen, T.H.; Enfield, C.G. Henry’s Law Constant for Trichloroethylene between 10 and 95 °C. Environ. Sci. Technol. 1998, 32, 1433–1437. [Google Scholar] [CrossRef]
  23. Prausnitz, J.M.; Lichtenthaler, R.N.; Azevedo, E. Molecular Thermodynamics of Fluid-Phase Equilibria, 3rd ed.; Pearson Education: London, UK, 1998. [Google Scholar]
  24. Sandler, S.I. Chemical, Biochemical, and Engineering Thermodynamics, 5th ed.; John Wiley & Sons: Hoboken, NJ, USA, 2017. [Google Scholar]
  25. Tschoegl, N. The Ideal Solution. In Fundamentals of Equilibrium and Steady-State Thermodynamics; Elsevier: Pasadena, CA, USA, 2000; pp. 104–112. [Google Scholar]
  26. Jackson, P.L.; Wilsak, R.A. Thermodynamic Consistency, Tests Based on the Gibbs-Duhem Equation Applied to Isothermal, Binary Vapor-Liquid-Equilibrium Data-Data Evaluation and Model Testing. Fluid Phase Equilibria 1995, 103, 155–197. [Google Scholar] [CrossRef]
  27. Wisniak, J.; Apelblat, A.; Segura, H. An Assessment of Thermodynamic Consistency Tests for Vapor-Liquid Equilibrium Data. Phys. Chem. Liq. 1997, 35, 1–58. [Google Scholar] [CrossRef]
  28. Stevenson, F.D.; Sater, V.E. Local Thermodynamic Consistency of Vapor-Liquid Equilibrium Data for Binary and Multicomponent Systems. AIChE J. 1966, 12, 586–588. [Google Scholar] [CrossRef]
  29. Brockbank, S.A.; Russon, J.L.; Giles, N.F.; Rowley, R.L.; Wilding, W.V. Critically Evaluated Database of Environmental Properties: The Importance of Thermodynamic Relationships, Chemical Family Trends, and Prediction Methods. Int. J. Thermophys. 2013, 34, 2027–2045. [Google Scholar] [CrossRef]
  30. Raal, J.D.; Ramjugernath, D. 13 Measurement of Limiting Activity Coefficients: Non-Analytical Tools. Exp. Thermodyn. 2005, 7, 339–357. [Google Scholar]
  31. Pividal, K.A.; Birtigh, A.; Sandler, S.I. Infinite Dilution Activity-Coefficients for Oxygenate Systems Determined Using a Differential Static Cell. J. Chem. Eng. Data 1992, 37, 484–487. [Google Scholar] [CrossRef]
  32. Van Ness, H.C.; Abbott, M.M. Classical Thermodynamic s of Nonelectrolyte Solutions; McGraw-Hill: New York, NY, USA, 1982. [Google Scholar]
  33. Gmehling, J.; Onken, U. Vapor-Liquid Equilibrium Data Collection: Aqueous-Organic Systems; DECHEMA: Frankfurt, Germany, 1977. [Google Scholar]
  34. Gmehling, J.; Onken, U.; Arlt, W. Vapor-Liquid Equilibrium Data Collection: Aqueous-Organic Systems (Supplement 1); DECHEMA: Frankfurt, Germany, 1981. [Google Scholar]
  35. Gmehling, J.; Onken, U. Vapor-Liquid Equilibrium Data Collection: Organic Hydroxy Compounds: Alcohols; DECHEMA: Frankfurt, Germany, 1986. [Google Scholar]
  36. Gmehling, J.; Onken, U.; Weidlich, U.; Rareynies, J.R. Vapor-Liquid Equilibrium Data Collection: Organic Hydroxy Compounds: Alcohols and Phenols; DECHEMA: Frankfurt, Germany, 1990. [Google Scholar]
  37. Gmehling, J.; Onken, U.; Arlt, W. Vapor-Liquid Equilibrium Data Collection: Aldehydes and Ketones, Ethers; DECHEMA: Frankfurt, Germany, 1979. [Google Scholar]
  38. Zin, R.M.; Coquelet, C.; Valtz, A.; Abdul Mutalib, M.I.; Sabil, K.M. Measurement of Henry’s Law Constant and Infinite Dilution Activity Coefficient of Isopropyl Mercaptan and Isobutyl Mercaptan in (Methyldiethanolamine (1) + Water (2)) with W 1 = 0.25 and 0.50 at Temperature of (298 to 348) K Using Inert Gas Stripping Method. J. Chem. Thermodyn. 2016, 93, 193–199. [Google Scholar]
  39. Valderrama, J.O.; Alvarez, V.H. Correct Way of Reporting Results when Modelling Supercritical Phase Equilibria using Equations of State. Can. J. Chem. Eng. 2008, 83, 578–581. [Google Scholar] [CrossRef]
  40. Raal, J.D. Characterization of Differential Ebulliometers for Measuring Activity Coefficients. AIChE J. 2000, 46, 210–220. [Google Scholar] [CrossRef]
  41. Kojima, K.; Zhang, S.J.; Hiaki, T. Measuring Methods of Infinite Dilution Activity Coefficients and a Database for Systems Including Water. Fluid Phase Equilibria 1997, 131, 145–179. [Google Scholar] [CrossRef]
  42. Sherman, S.R.; Trampe, D.B.; Bush, D.M.; Schiller, M.; Eckert, C.A.; Dallas, A.J.; Li, J.J.; Carr, P.W. Compilation and Correlation of Limiting Activity Coefficients of Nonelectrolytes in Water. Ind. Eng. Chem. Res. 1996, 35, 1044–1058. [Google Scholar] [CrossRef]
  43. Tiegs, D.; Gmehling, J.; Menke, J.; Schiller, M. Activity Coefficients at Infinite Dilution; DECHEMA: Frankfurt, Germany, 1986. [Google Scholar]
  44. Gupta, A.K.; Teja, A.S.; Chai, X.S.; Zhu, J.Y. Henry’s Constants of N-Alkanols (Methanol through N-Hexanol) in Water at Temperatures between 40 °C and 90 °C. Fluid Phase Equilibria 2000, 170, 183–192. [Google Scholar] [CrossRef]
  45. Benkelberg, H.J.; Hamm, S.; Warneck, P. Henry’s Law Coefficients for Aqueous Solutions of Acetone, Acetaldehyde and Acetonitrile, and Equilibrium Constants for the Addition Compounds of Acetone and Acetaldehyde with Bisulfite. J. Atmos. Chem. 1995, 20, 17–34. [Google Scholar] [CrossRef]
  46. Krichevsky, I.R.; Kasarnovsky, J.S. Thermodynamical Calculations of Solubilities of Nitrogen and Hydrogen in Water at High Pressures. J. Am. Chem. Soc. 2002, 57, 2168–2171. [Google Scholar] [CrossRef]
  47. Kretzschmar, H.J.; Wagner, W. Properties of Water and Steam Based on the Industrial Formulation Iapws-If97, 2nd ed.; Springer: Heidelberg, Germany, 2014. [Google Scholar]
  48. Fernández-Prini, R.; Alvarez, J.L.; Harvey, A.H. Henry’s Constants and Vapor–Liquid Distribution Constants for Gaseous Solutes in H2O and D2O at High Temperatures. J. Phys. Chem. Ref. Data 2003, 32, 903–916. [Google Scholar]
  49. Wiebe, R.; Gaddy, V.L. The Solubility of Carbon Dioxide in Water at Various Temperatures from 12 to 40° and at Pressures to 500 Atmospheres. Critical Phenomena. J. Am. Chem. Soc. 2002, 62, 815–817. [Google Scholar]
  50. Dhima, A.; de Hemptinne, J.C.; Jose, J. Solubility of hydrocarbons and CO2 mixtures in water under high pressure. Ind. Eng. Chem. Res. 1999, 38, 3144–3161. [Google Scholar]
  51. O’Sullivan, T.D.; Smith, N.O. Solubility and partial molar volume of nitrogen and methane in water and in aqueous sodium chloride from 50 to 125.deg. and 100 to 600 atm. J. Phys. Chem. 2002, 74, 1460–1466. [Google Scholar]
Figure 1. Relationship between total pressure and liquid phase composition x of water in a tert-amyl alcohol system at 303.32 K.
Figure 1. Relationship between total pressure and liquid phase composition x of water in a tert-amyl alcohol system at 303.32 K.
Ijms 23 12998 g001
Figure 2. Relationship between ( p x ) T , x 0 and temperature in different systems.
Figure 2. Relationship between ( p x ) T , x 0 and temperature in different systems.
Ijms 23 12998 g002
Figure 3. Solubility versus fugacity calculated by different models for nitrogen and methane systems. (a) nitrogen, 373.15 K; (b) methane, 375.65 K; (c) hydrogen, 298.15 K; (d) carbon dioxide, 323.15 K; this work 1: calculated from Equation (22); this work 2: calculated from Equation (23).
Figure 3. Solubility versus fugacity calculated by different models for nitrogen and methane systems. (a) nitrogen, 373.15 K; (b) methane, 375.65 K; (c) hydrogen, 298.15 K; (d) carbon dioxide, 323.15 K; this work 1: calculated from Equation (22); this work 2: calculated from Equation (23).
Ijms 23 12998 g003
Table 1. The values of the parameters and calculated results of the proposed model.
Table 1. The values of the parameters and calculated results of the proposed model.
SoluteSolventTemperature RangeNpParametersA.A.D. (%)A.R.D. (%)
KAB
MethanolWater[298.15, 343.15]5−4739.8314.423.840.11
2-PropanolWater[298.15, 353.15]3−6309.0520.161.260.01
Tert-butyl alcoholWater[283.15, 348.15]7−3771.9112.255.190.19
2-PentanolWater[343.15, 363.15]3−5310.7518.023.080.05
PhenolWater[323.15, 363.14]5−5437.3614.065.850.22
2-ButanoneWater[323.15, 343.15]3−2308.439.220.100.00
AcroleinWater[273.15, 326.55]4−3539.7013.575.990.26
3-MethylpyridineWater[308.15, 338.15]4−3025.8011.902.340.03
2-EthylbutylamineWater[298.15, 313.15]3−6884.9324.260.830.00
N, N-Dimethyl-tert-butylamineWater[293.15, 318.05]5−6096.8021.323.440.08
Ethyl acetateWater[298.15, 343.15]5−4983.9918.405.50−0.07
Ethyl acrylateWater[333.20, 353.20]3−1793.239.670.520.00
Tert-Butyl ethyl etherWater[293.15, 313.15]3−3979.7317.361.551.54
NitromethaneWater[294.15, 323.15]4−3576.8512.243.251.10
N, N-Diethyl isopropyl amineWater[283.15, 313.15]4−6750.6823.215.510.17
DiethylamineWater[311.50, 329.95]3−3238.1010.053.530.07
OxiraneWater[283.15, 298.15]3−3813.1815.140.920.00
AcetaldehydeWater[283.15, 313.15]7−3791.9014.182.350.04
ButyraldehydeWater[323.15, 353.15]5−4643.3517.083.260.06
1,4-DioxaneWater[293.15, 343.15]5−4080.8311.846.370.25
WaterPropylene carbonate[288.15, 300.15]3−11,493.7737.213.230.07
Water2-Butoxyethanol[358.15, 371.19]4−6324.3117.982.300.02
Water2-Pentanol[343.15, 363.15]3−4480.9313.071.670.01
Water2-Methyl-2-butanol[283.3, 343.25]4−5180.7515.033.620.09
WaterFurfural[310.95, 366.45]3−3330.4710.103.090.05
MethanolTetrachloromethane[303.15, 353.15]6−3588.4113.225.220.17
MethanolBenzene[293.15, 318.15]3−2310.708.555.540.18
MethanolHexanes[308.15, 348.15]4−4626.6116.232.540.04
TetrachloromethaneEthanol[293.15, 338.15]3−3349.4910.681.000.01
ChloroformEthanol[308.15, 328.15]5−3882.6711.951.710.02
AcetoneEthanol[305.15, 321.15]3−2775.278.691.700.01
Methyl acetateEthanol[323.15, 353.15]4−4557.5714.016.690.23
Ethyl acetateEthanol[313.15, 343.15]3−3634.7410.774.280.10
Ethyl acetate2-Methoxyethanol[343.15, 363.15]3−4390.6012.834.560.16
Benzene1-Propanol[308.18, 338.15]5−4067.3212.521.540.02
BenzeneIsopropanol[303.15, 333.15]3−6344.9819.280.07−0.06
N-methylanilineEthylene glycol[368.15, 418.15]3−4732.6311.123.080.06
2-Butanone2-Methoxyethanol[343.15, 363.15]3−5783.9716.801.270.01
Iso-butanolDimethyl sulfoxide[353.15, 378.15]3−4779.2112.096.480.24
ToluenePentanol[303.15, 383.15]5−3621.2010.104.140.09
Average 3.210.15
Table 2. Calculate the value of ( p x ) T , x 0 from ( T x ) p , x 0 .
Table 2. Calculate the value of ( p x ) T , x 0 from ( T x ) p , x 0 .
SoluteSolventTp
( T x ) p ,   x 0
( d p s d T ) x 0
( p x ) T ,   x 0 , cal .
( p x ) T ,   x 0 , ref .
Ref
KkPaKkPa/KkPakPa
MethanolWater372.88101.33 −156.004.05 560.33 563.37 [33]
EthanolWater373.13101.33 −300.634.05 1088.23 1113.56 [34]
1-PropanolWater323.5913.17 −244.111.01 151.99 149.96 [33]
2-PropanolWater342.3330.40 −443.331.01 580.59 574.51 [33]
AcetoneWater330.3420.27 −510.341.01 421.51 421.51 [34]
AcetoneWater356.1359.78 −492.312.03 1043.65 1035.54 [33]
AcetoneWater363.7180.05 −480.953.04 1304.05 1287.84 [33]
AcetoneWater373.13101.33 −454.404.05 1643.49 1668.82 [33]
Methanol1,4-Dioxane374.20101.33 −394.553.04 1220.97 1235.15 [35]
MethanolBenzene353.13101.33 −239.873.04 748.79 756.90 [35]
MethanolBenzene327.4745.60 −288.512.03 463.06 452.92 [35]
MethanolBenzene340.5869.91 −261.892.03 592.75 594.78 [35]
Benzene1-Propanol370.35101.33 −123.684.05 476.23 470.15 [35]
Table 3. Calculation of γ at different temperature by using ( p x ) T , x 0 .
Table 3. Calculation of γ at different temperature by using ( p x ) T , x 0 .
SoluteSolventT
( p x ) T ,   x 0 , VLE
γ VLE   a
( p x ) T ,   x 0 , cor .
γ cor .   b
γ ref .
Ref
Katmatm
MethanolWater323.150.771.630.761.611.64[18]
EthanolWater298.280.365.040.354.945.22[18]
Tert-butyl alcoholWater298.150.6312.160.6512.4111.91[43]
NitromethaneWater294.151.1230.341.0829.2534.80[43]
NitromethaneWater296.151.1227.341.1628.2333.21[43]
NitromethaneWater323.153.1727.383.2127.7629.10[43]
ButyraldehydeWater343.1533.9339.9334.4340.5139.25[43]
TetrachloromethaneEthanol293.150.474.430.494.584.90[43]
TetrachloromethaneEthanol338.152.153.272.243.383.73[43]
ChloroformEthanol308.150.511.620.531.641.67[43]
ChloroformEthanol318.150.771.760.771.751.70[43]
ChloroformEthanol328.151.091.801.131.841.55[43]
AcetoneEthanol305.150.671.940.671.942.28[43]
AcetoneEthanol313.150.821.780.851.832.21[43]
AcetoneEthanol321.151.061.761.061.762.14[43]
Ethyl acetateEthanol313.150.452.500.432.442.77[43]
a Calculation of γ from ( p x ) T , x 0 of gas–liquid equilibrium data. b Calculation of γ from ( p x ) T , x 0 of the correlation model.
Table 4. Calculation of H i at different temperatures using ( p x ) T , x 0 .
Table 4. Calculation of H i at different temperatures using ( p x ) T , x 0 .
SoluteSolventT
p 2 s
( p x ) T ,   x 0 , VLE
H 1 , VLE
( p x ) T ,     x 0 , cor .
H 1 , cor .
H 1 , ref .
Ref
KkPakPakPakPakPakPa
MethanolWater323.1512.3078.0890.3877.0689.3681.41[33,44]
MethanolWater338.1525.43152.76178.19153.86179.29177.56[34,44]
EthanolWater298.153.2336.7139.9435.8939.1141.32[33,44]
AcetoneWater298.153.06134.36137.42128.60131.66141.68[33,45]
AcetoneWater303.154.12168.80172.92161.21165.33172.70[34,45]
AcetaldehydeWater288.150.71294.54295.25290.08290.79288.29[33,45]
AcetaldehydeWater293.152.98355.17358.15369.99372.97364.43[33,45]
OxiraneWater293.152.33865.54867.87860.94863.27-[33]
Methyl acetateWater308.155.61753.33758.94748.58754.19-[33]
Tert-butyl ethyl etherWater293.412.924499.674502.594481.384484.30-[33]
Tert-butyl ethyl etherWater313.157.3810,580.5510,587.9310,534.6810,542.06-[33]
Dimethyl isopropyl amineWater293.152.67203.96206.63202.97205.64-[33]
Water2-Pentanol343.1512.99101.03114.02101.84114.83-[34]
Benzene1-Propanol318.159.5777.7787.3477.7387.29-[35]
2-Butanone2-Methoxyethanol343.1514.4996.46110.9594.58109.08-[35]
Table 5. Parameters of correlation activity coefficient and solubility calculation deviation.
Table 5. Parameters of correlation activity coefficient and solubility calculation deviation.
SoluteTNpParameters aA.A.D. (%)A.R.D. (%)
KCDE
Hydrogen273.158-61.41-2.95−0.74
Hydrogen298.158-89.05-2.48−0.87
Hydrogen323.158-99.9-3.08−1.53
Hydrogen373.156-96.67-4.16−2.77
Nitrogen298.158-151.86-2.69−0.06
Nitrogen323.158-272.26-4.58−0.13
Nitrogen348.158 293.08 2.762.63
Nitrogen373.157-177.98-6.664.50
Carbon dioxide313.157−4771.16170.63−1.580.510.01
Carbon dioxide323.158−3851.95122.85−1.021.350.01
Carbon dioxide344.157−3217.5383.57−0.573.640.08
Methane324.656−20,808.6558.9−0.020.260.00
Methane375.656−10,593.63−17.44−0.160.900.00
Methane398.156−2050.41−46.150.080.660.01
Average 2.680.08
a For hydrogen and nitrogen systems, ln γ 1 = D x 1 ; for carbon dioxide and methane systems, ln γ 1 = C x 1 2 + D x 1 + F .
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zheng, J.; Chen, X.; Wang, Y.; Sun, Q.; Sun, W.; Wu, L.; Hu, Y.; Zhang, W. Research of a Thermodynamic Function (px)T, x0: Temperature Dependence and Relation to Properties at Infinite Dilution. Int. J. Mol. Sci. 2022, 23, 12998. https://doi.org/10.3390/ijms232112998

AMA Style

Zheng J, Chen X, Wang Y, Sun Q, Sun W, Wu L, Hu Y, Zhang W. Research of a Thermodynamic Function (px)T, x0: Temperature Dependence and Relation to Properties at Infinite Dilution. International Journal of Molecular Sciences. 2022; 23(21):12998. https://doi.org/10.3390/ijms232112998

Chicago/Turabian Style

Zheng, Jiahuan, Xia Chen, Yan Wang, Qichao Sun, Wenting Sun, Lianying Wu, Yangdong Hu, and Weitao Zhang. 2022. "Research of a Thermodynamic Function (px)T, x0: Temperature Dependence and Relation to Properties at Infinite Dilution" International Journal of Molecular Sciences 23, no. 21: 12998. https://doi.org/10.3390/ijms232112998

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop