SciELO - Scientific Electronic Library Online

 
vol.10 issue1Determination of shear velocity in a mild-sloping open channel flowKinematic wave hydrologic model of the Turbio River basin, Guanajuato, Mexico author indexsubject indexsearch form
Home Pagealphabetic serial listing  

Services on Demand

Journal

Article

Indicators

Related links

  • Have no similar articlesSimilars in SciELO

Share


Ingeniería agrícola y biosistemas

On-line version ISSN 2007-4026Print version ISSN 2007-3925

Ing. agric. biosist. vol.10 n.1 Chapingo Jan./Jun. 2018  Epub Aug 24, 2018

https://doi.org/10.5154/r.inagbi.2017.10.014 

Scientific article

Classical thermodynamic-based models for predicting physical properties of a biodiesel fuel

José E. Benavides-Fajardo1 

Mauricio Romero-Bastida1 

Felipe A. Perdomo2 

María del Carmen Núñez-Santiago3 

1Instituto Politécnico Nacional, Estudios de Posgrado e Investigación de ESIME-Culhuacán. Av. Santa Ana, núm. 1000, San Francisco Culhuacán, Delegación Coyoacán, Ciudad de México, C. P. 04430, MÉXICO.

2Universidad de Bogotá Jorge Tadeo Lozano, Departamento de Ingeniería Química. Carretera 4, núm. 22-61, Bogotá, C. P. 110311, COLOMBIA.

3Instituto Politécnico Nacional, Centro de Desarrollo de Productos Bióticos. Carretera Yautepec-Jojutla km 6.5, Calle Ceprobi, núm. 8, Col. San Isidro, Yautepec, Morelos, C. P. 62570, MÉXICO.


Abstract

Objective:

To obtain a model based on the classical Peng-Robinson equation of state (PR-EOS) to evaluate the initial thermodynamic expressions (first- and second-derivative properties) in biodiesel.

Methodology:

A modified temperature dependence was used to transform the volume of the Peng-Robinson cubic equation of state to predict the thermophysical properties of biodiesel. The fuel studied is composed of five fatty acid methyl esters (methyl palmitate, stearate, oleate, linoleate and linolenate), which are the primary constituents of biodiesel.

Results:

The results showed that the approach presented in this work can improve the prediction of second-order properties (isentropic bulk modulus, heat capacities and speed of sound) if the accuracy of the primary properties is maintained (vapor pressure and liquid density).

Study limitations:

Biodiesel is highly corrosive and is used in mixtures with other fuels such as gasoline; however, the model is only applicable to the properties of biodiesel and not mixtures.

Originality:

Two concepts in the Peng-Robinson equation are used: the α(T) function and volume transformation. The second concept was implemented because second-order thermodynamic properties need accurate densities and derivatives with respect to the total volume. Additionally, the α(T) function was selected through a systematic search among a wide range of reported correlations.

Conclusions:

The proposed thermodynamic model can predict, with only a few experimental data, properties that have an impact on the representation of spray, atomization and combustion events in diesel engines.

Keywords biofuel; Peng-Robinson equation; simulation; second-order properties

Resumen

Objetivo:

Obtener un modelo basado en la ecuación de estado clásica de Peng-Robinson (EEC- PR) para evaluar las expresiones termodinámicas iniciales (propiedades de primera y segunda derivadas) en biodiésel.

Metodología:

Se presentó una dependencia de temperatura modificada para la transformación del volumen de la ecuación cúbica de estado de Peng-Robinson, esto con el fin de predecir las propiedades termofísicas del biodiesel. El combustible estudiado está compuesto por cinco ésteres metílicos de ácidos grasos (palmitato, estereato, oleato, linoleato y linolenato de metilo), que son los constituyentes primarios del biodiesel.

Resultados:

Los resultados mostraron que el enfoque presentado en este trabajo puede mejorar la predicción de las propiedades de segundo orden (el módulo de compresibilidad isentrópica, las capacidades caloríficas y la velocidad del sonido) si se mantiene la precisión de las propiedades primarias (presión de vapor y densidad del líquido).

Limitaciones del estudio:

El biodiesel es altamente corrosivo, por lo que se emplea en mezclas con otros carburantes como las gasolinas; sin embargo, el modelo sólo es aplicable en las propiedades de biodiesel y no en mezclas.

Originalidad:

Se utilizan dos conceptos en la ecuación de Peng-Robinson: la función α(T) y la transformación del volumen. El segundo concepto se implementó porque las propiedades termodinámicas de segundo orden necesitan densidades y derivadas precisas respecto del volumen total. Adicionalmente, la función α(T) se seleccionó a través de una búsqueda sistemática entre una amplia gama de correlaciones reportadas.

Conclusiones:

El modelo termodinámico propuesto puede predecir, con sólo pocos datos experimentales, propiedades que tienen impacto en la representación de eventos de aspersión, atomización y combustión en motores diésel.

Palabras clave biocombustible; Ecuación de Peng-Robinson; simulación; propiedades de segundo orden

Introduction

Biomass and biofuels in particular are nowadays considered as promising and sustainable alternative fuels due to their renewability, reduced toxic-gas emissions, and biodegradability. Direct use of vegetable oils in diesel engines has not been entirely successful due to poor properties such as low volatility, reactivity of the unsaturated molecules, and high viscosity, which can lead to coking on engine injectors, carbon deposits, and plugged orifices during prolonged engine performance. Transesterification reaction is the most often used method to improve these properties.

Any triglyceride feedstock present in the vegetable oils and animal fats (including algal lipids) can be transesterified to produce the commonly named biodiesel, which is defined as an oxygenated fuel comprised of monoalkyl esters of long chain fatty acids and is designated by the American Society of Testing Materials (ASTM, 2005) as B100.

The most common alcohol used in biodiesel production is methanol, and it is therefore said that biodiesel is comprised of a mixture of fatty acid methyl esters (FAMEs). In addition, FAME distribution in biodiesel depends directly on fatty acid profiles that were present in the source material employed in their synthesis. The most common fatty esters contained in biodiesel (Figure 1) are those obtained from palmitic, stearic, oleic, linoleic and linolenic acid. This holds for most biodiesel feedstocks, such as soybean, sunflower, canola, palm and peanut oils (Knothe, 2010).

Figure 1 Predominant fatty acid methyl ester (FAME) compounds in the most common biodiesel feedstocks. 

Several studies have been aimed at establishing a heuristic relation between fuel properties and fatty acid molecular structure and composition (Schönborn, Landommatos, Williams, Allan, & Rogerson, 2009). Accordingly, the variability of the biodiesel feedstock (presence or absence of any FAME as well as the amount) will generate differences in biodiesel molecular structure. These differences lead to distinctive patterns in fuel energy release rates and pollutant formation. The experimental measure of such properties for all types and possibilities of biodiesel fuel blends is both impractical and expensive. The experimental data needed to develop thermodynamic models are, however, very scarce for complex and large molecules such as biodiesel ones, and experimental measures are extremely costly and time consuming. Thermodynamic models afford a simple and immediate way to predict the physical properties of biodiesel fuels (Csernica & Hsu, 2011; Huber, Lemmon, Kazakov, Ott, & Bruno, 2009; Olivera, Ribeiro, Queimada, & Couthinho, 2011; Olivera, Teles, Queimada, & Coutinho, 2009). However, we consider that a molecular-based cubic equation of state (EOS) is necessary for a better approach to thermodynamical properties.

The Statistical Associating Fluid Theory (SAFT) approach (Chapman, Gubbins, Jackson, & Radosz, 1990; Perdomo, Perdomo, Millán, & Aragón, 2014) provides an EOS that takes into account the anisotropies of the fluid, both due to molecular shape and directional forces. The SAFT-VR approach (Gil-Villegas, Galindo, Whitehead, Jackson, & Burgess, 1997), wherewith fluids are modeled as chains of tangentially bonded hard spheres interacting with potentials of variable range, has also been used to model the vapor-liquid equilibria of three FAME biodiesel compounds (Perdomo & Gil-Villegas, 2010), modelling each FAME as a chain of spherical segments that can associate due to the presence of short-ranged attractive sites. These sites, as well as the intermolecular interaction among three monomer segments, were modeled by square-well potentials of variable range.

The same approach was implemented to model biodiesel blends using a corresponding mixing rule in the SAFT-VR approach and a cross-binary parameter among species for the unlike energy parameter, making it possible to determine isochoric heat capacities, speed of sound as well as the phase diagram of binary and ternary FAME mixtures (Perdomo & Gil-Villegas, 2011). In spite of the aforementioned rigorous approaches, classical EOS have been successfully applied in the chemical industry to describe the thermodynamic properties of a wide range of pure fluids and mixtures. Although it is known that classical EOS fail in the description of second-derivative thermodyanamic properties and are thus usually applied only for phase equilibrium calculations, their simplicity and wide applicability have justified continued efforts in their improvement.

In this study we propose a model based on the classical Peng-Robinson equation of state (PR-EOS) (Peng & Robinson, 1976) in order to evaluate the departure thermodynamic expressions (first- and second-derivative properties) in biodiesel fuels. To achieve this goal, the herein developed approach uses two concepts that have been commonly applied to the PR-EOS: the temperature dependent function α(T) and the translated volume. In the first case an exponential α(T) function, proposed by Gasem, Gao, Pan, and Robinson (2001), was implemented in order to improve vapor pressure prediction in pure FAMEs. It is important to highlight that the selection of the temperature function was the result of a systematic search among a wide range of correlations reported in the literature. The translated volume concept was implemented due to the fact that second-order thermodynamic properties need accurate densities and derivatives with respect to the total volume, and the original α(T) function has no significant influence on the description of volumetric properties.

Materials and methods

Model development

The methodology used for modeling biodiesel systems is based on the classical volume-translated PR-EOS (VTPR-EOS) (Equation 1) (Ahlers & Gmejling, 2001; González-Pérez, Coquelet, Paricard, & Chapoy, 2017; Peng & Robinson, 1976).

P= RTv+c-b-av+c+v+c+b+b(v+c-b) (1)

Where

aT=aTc,Pcα(T) (2)

bT=b(Tc,Pc)

and

aTc,Pc=0.45724R2Tc2Pc (3)

bTc,Pc=0.0778RTcPc

T c , P c , R, and v are the critical temperature, the critical pressure, the universal gas constant, and the molar volume respectively. The translation parameter c is defined as the difference between calculated and experimental volume at the considered reduced temperature (T r ) and is obtained via the Rackett equation (Rackett, 1970):

cTc,Pc=c1RTcPc(c2Zc-c3) (4)

where Z c = P c / Rρ c T c is a critical compressibility factor. Constants c 1 , c 2 , and c 3 have to be optimized for a particular biofuel mixture. The α-function is a temperature-dependent model function that has been the subject of several studies in order to improve the precision of EOS. Many proposed α-functions use either a high-order polynomial in the acentric factor or temperature to correlate vapor pressures more precisely. For this study, the α-function proposed by Gasem et al. (2001) was selected:

αTr,  ω=expA+BTr1-TrC+Dω+Eω2 (5)

where ω is the acentric factor that characterizes the non-sphericity of molecular interactions. Constants A, B, C, D, and E also have to be optimized for the considered biodiesel compounds.

In order to estimate ω, the group contribution method proposed by Constantinou, Gani, and O’Connell (1995), which does not need other primary properties for its determination, was implemented. Table 1 shows the primary and secondary groups needed to represent FAME biodiesel constituents. Once the framework is settled (groups and occurrences), the ω is calculated as follows (Constantinou et al., 1995):

Table 1 Group contribution information for estimating the acentric factor of the five biodiesel compounds considered in this work.  

Group ω 1i,2j MEC 16:0 MEC 18:0 MEC 18:1 MEC 18:2 MEC 18:3
Ni Mj Ni Mj Ni Mj Ni Mj Ni Mj
First order
CH3 0.29602 2 2 2 2 2
CH2 0.14691 13 15 13 11 9
CH=CH 0.25224 0 0 2 4 6
CH2COO 0.75574 1 1 1 1 1
Second order
CH2-CHm=CHn -0.0115 0 0 1 2 0

Source: Constantinou et al. (1995).

ω=aωlnNiω1i+AωMjω2j+cω1/bω (6)

where a ω = 0.4085, b ω = 0.5050, and c ω = 1.1507 are universal constants, ω 1i is the contribution of the type i first-order group which occurs N i times and ω 2j is the contribution of type j second-order group, which occurs M j times. Both contributions are involved, A ω = 1 (Constantinou et al., 1995).

Fitting data

The complete parameter set θ ≡ (A, B, C, D, E, c 1 , c 2 , c 3 ) was determined for each of the most commercially-used FAMEs (Table 2) (Hoekman, Broch, Robbins, Ceniceros, & Natarajan, 2012), of which methyl palmitate (MEC 16:0) and methyl stearate (MEC 18:0) are saturated, and methyl oleate (MEC 18:1), methyl linoleate (MEC 18:2) and methyl linolenate (MEC 18:3) are unsaturated. For this, the proposed α(T) function was fitted to the theoretical vapor pressure values of the experimental liquid-vapor equilibrium data corresponding to the FAMEs studied by Rose and Supina (1961). The critical region was excluded from the parameter regression.

Table 2 Parameter set θ for the five fatty acid methyl esters considered. 

MEC 16:0 MEC 18:0 MEC 18:1 MEC 18:2 MEC 18:3
A 1.9904 2.6131 2.3646 4.4464 1.9303
B 0.9278 0.9356 0.0043 0.5957 0.0014
C 0.0906 0.6382 0.2285 0.6195 0.0017
E 0.4376 0.2930 0.3522 0.9063 0.7524
D 0.0778 -0.5060 -0.0027 -1.1680 -0.0383
c1 -0.3921 -0.2942 -0.4221 -0.2302 -0.2881
c2 1.6488 1.7821 1.3134 1.9822 1.7439
c3 0.0981 0.0146 0.0154 0.0321 0.0048

In classical optimization procedures, the relevant parameters are optimized either for liquid-vapor equilibrium or second-derivative properties, but not both simultaneously. Due to the known difficulty to calculate second-derivative properties by classical EOS, because of their direct dependence on the molecular structure of the fluid, the following relative least square objective function was used (Liang, Maribo-Mogensen, Thmosen, Yan, & Kontogeorgis, 2012):

minFθ=i=1NePv,icalcT, θ-Pv,iexpPv,iexp2+i=1NevicalcT, θ-viexpviexp2   (7)

where N e are the number of experimental data points for vapor pressures Pv,icalc and Pv,iexp (calculated and experimental), while v i calc and v i exp are the speed of sound values, both parameters at a given temperature. The inclusion of this last property in the optimization function is necessary to correctly calculate second-derivative properties for FAME mixtures.

To determine an optimal multiparameter set, a hybrid minimization method was implemented, combining global-search and local-search algorithms. Additionally, an iterative procedure was used to obtain the most adequate values: 1) the values of the pure component parameters were estimated with the original values (c 1 , c 2 , c 3) proposed by Dietrichs, Rarey, and Mehling (2006), 2) a regression was performed on the coefficients A to E in the modified α(T) function, 3) the pure component parameters were estimated with new values (c 1 , c 2, c3), and 4) steps 1 to 3 were repeated until the desired convergence was obtained.

The simplex method was used to reduce the convergence domain, and then the Levenberg-Marquardt pseudo-second order method (Nocedal & Wright, 2000) was applied until the function gradient attained its minimum value, that is:

|Fθ|0 (8)

In order to verify the optimal parameter fitting, the saturation vapor pressure curves, representing P vs T and the Clausius-Clapeyron model, were constructed for each FAME.

Modeling thermophysical properties in biodiesel systems

This section presents the expressions and numerical procedures for calculating the key properties required in the study of fuel quality, combustion, and injection system design in engines.

Boiling point: at atmospheric pressure and called normal boiling point, it is the basis for predicting the critical properties and those dependent on temperature, such as vapor pressure, density, latent heat of vaporization, viscosity, and surface tension, which are required for biodiesel combustion modeling (Yuan, Hansen, & Zhang, 2003), but the wide temperature ranges required for these properties make it difficult to experimentally measure them. Therefore, the accurate boiling points of the pure fatty acid esters and their biodiesel mixtures are useful for predicting fuel properties and are consequently important for combustion modeling.

Vapor pressure in pure biodiesel compounds: a simple procedure can be implemented in order to satisfy the liquid-vapor equilibrium criteria and estimate the vapor pressure at some constant temperature. First, it is necessary to estimate an initial approximation of the pressure value P k with a lower value than the critical pressure, and then recalculate it using the recurrence relation:

Pk+1=PkϕLT,   PkϕVT,    Pk (9)

where ϕ is the fugacity coefficient. Further details are given in Appendix 1.

The vapor pressure in the fuel mixtures is also useful to assess the risk level in the different biodiesel production stages, as well as in its storage and transport.

Liquid heat capacities: it is a measure of the amount of energy required by a mol (or mass unit) of a substance to raise its temperature by a unit degree. It is of great importance in the design of industrial processes and engineering calculations (i.e., distillation, heat exchanger designs, estimation of reaction enthalpies as a function of temperature, thermodynamics of combustion analysis, etc.). Ignition and combustion in diesel engines are strongly influenced by the fuel spray, and this in turn is affected by heat capacities and heat of vaporization.

The heat capacities (Equation 10) of biodiesel systems can be calculated by two methodologies (Diedrichs, Rarey, & Mehling, 2006). The first one proceeds from the thermodynamic expressions that give a relation between measurable quantities and the entropy after applying the appropriate relations:

Cp=Cpid-R+TPTVVTP+T22blnv+c-1-2bv+c+1+2b (10)

where Cpid is the heat capacity under ideal conditions. The second methodology to calculate the heat capacities stems from its thermodynamical definition, i.e. from the change of enthalpy with the temperature (Appendix 2, Equation 29). Although straightforward, the results of this method are similar to those obtained with Equation 10; therefore, the latter ones are those herein reported.

Speed of sound and isentropic bulk modulus. Differences in physical properties between biodiesel and petroleum-based diesel fuel may change the engine's fuel injection timing and combustion characteristics (Schönborn et al., 2009; Tat & Van Gerpen, 2003) by altering both the combustion timing and rate. The properties that have the greatest effect on fuel injection timing are the speed of sound (v s ; Liang et al., 2012) and the isentropic bulk modulus. In this work the values of the aforementioned key properties of biodiesel components and blends at temperature and pressure ranges most commonly encountered in injection system conditions are predicted.

vs=CpCvPvT   (11)

Once v s is obtained, the isentropic bulk modulus (β) is readily given by the following expression (Tat & Van Gerpen, 2003):

β=ρvs2 (12)

Multicomponent biodiesel system

In order to predict key physico-chemical properties in FAME mixtures, the classical one-fluid theory mixing rules defined by the following set of equations was implemented (Diedrichs et al., 2006; Peng & Robinson, 1976):

am=ijxixjaij

bm=ixibi (13)

cm=ixici

where

aij=aiaj (14)

The coefficients a i , b i , and c i are calculated from Equations 2, 3 and 4, and x i is the molar fraction in the liquid phase of component i in the considered mixture. To calculate vapor-liquid equilibria, specifically the bubble (or dew) point pressure (P*), the fugacity coefficient for component i in a FAMEs mixture is written as (Peng & Robinson, 1976):

lnϕi^=bibmZ+cm-1-lnZ+cm-1-Aeosm22Beosm2jxjaijam-bibmxlnZ+cm+1+2BeosmZ+cm+1-2Beosm (15)

The molar fraction in the vapor phase of component i is expressed in terms of T, P*, and x i as:

yiT,P*,  xi=κixi (16)

where к i is the distribution coefficient denoted by:

κi=ϕ^iL T,  P*,  xiϕ^iVT,  P*,  yi (17)

The P* was obtained through a conventional Rachford-Rice procedure, without any assumption about the ideality in the FAME mixtures, which has been indeed made in several investigations (Yuan, Hansen, & Zhang, 2005); for this, the following equation was solved:

fψ=1-i=1cκiT, P*, xixi0 (18)

where ψ represents the vaporized fraction in the system, which can take any value between 0 to 1. The iteration procedure was started by estimating the bubble point as:

P0*=i=1cxiPisat (19)

where P i sat is calculated with the procedure explained in Appendix 1.

All calculations were done in MatLab language Ver. 8.5 (Math Works Inc. Natick, Massachusetts, USA).

Results and discussion

Table 3 shows the parameters of the Clausius-Clapeyron relation (logP=α+γT) calculated from the model proposed in the fitting data section for the five methyl esters considered in this work, along with the corresponding experimental values reported previously (Scott, MacMillan, & Melvin, 1952). For all FAMEs, excellent agreement was obtained with the reported experimental data in the considered temperature range. Besides corroborating the relevance of the modified PR-EOS for the considered FAMEs, the presented results also foreshadow the correctness of the latent heat of vaporization value (Appendix 3, Equation 25, Table 5).

Table 3 Clausius-Clapeyron relation parameters of fatty acid methyl esters (FAMEs) for the biodiesel compounds studied in this work, compared with the experimental data of Scott et al. (1952) with the corresponding deviation percent. 

FAME α1 γ r2 Experimental data Calculated data Deviation percent
MEC 16:0 9.3431 -4,199.03 0.999 80,219.1 73,562 1.1
MEC 18:0 9.5402 -4,496.95 0.999 85,703.8 74,673 5.7
MEC 18:1 9.3945 -4,395.05 0.997 84,446.8 76,077 7.4
MEC 18:2 9.4120 -4,397.86 0.999 83,610.4 77,697 7.7
MEC 18:3 9.5566 -4,472.14 0.999 84,573.4 73,800 7.2

1α and γ: Clausius-Clapeyron constants.

Table 4 Values of the constants cp0 to cp7 for the fatty acid methyl esters. 

MEC 16:0 MEC 18:0 MEC 18:1 MEC 18:2 MEC 18:3
cp0 120.529 247.115 90.2385 190.986 79.5913
cp1 0.0801627 -0.0916606 0.14118 0.020213 0.214648
cp2 345.62 276.94 234.797 437.371 290.379
cp3 2,952.37 556.17 613.529 3,052.11 1,213.24
cp4 289.038 408.997 335.768 287.222 81.4323
cp5 734.653 1,311.85 1,405.31 746.631 578.752
cp6 301.639 472.702 431.66 321.956 474.881
cp7 1,593.55 2,825.71 2,867.76 1,624.33 2,799.79

Source: Huber et al. (2009).

Table 5 Gas-phase standard enthalpies at 298.15 K for the biodiesel compounds studied in this work. 

FAME (hv (Kcal·mol-1)
MEC 16:0 -167.8
MEC 18:0 -177.3
MEC 18:1 -147.2
MEC 18:2 -116.1
MEC 18:3 - 85.6

FAME: fatty acid methyl esters. Source: Osmond et al. (2007).

A good prediction in the vapor pressure and, consequently, in the heat of vaporization, is also necessary in order to have reasonable accuracy in second-derivative properties because it is directly related to the change in heat capacity in the vaporization region (Equation 20).

ΔhvT=RT22ln Pvap1T2=ΔCp=ΔCpv-Cpl (20)

It is important to mention that the herein proposed model is the result of a systematic testing of various α(T) functions previously reported in the literature and reviewed in Gasem et al. (2001). They found good approximations for the vapor pressure data obtained. However, they did not obtain a similar agreement for second-derivative properties, where high deviations in the speed of sound were remarkable.

Figure 2 presents the speed of sound prediction for the five considered FAMEs when only the α(T) proposed by Soave is implemented without any volume correction. As can be readily seen, the predicted values are in strong disagreement with the experimental ones, in spite of good prediction in the vapor pressure. Additionally, for saturated FAMEs, the presence of double bonds in the molecular structure renders the prediction very difficult, especially at low temperatures.

Figure 2 Predicted speed of sound (vs) (lines) for all fatty acid methyl esters employing the Soave α(T) function. 

In order to address the aforementioned issues, the volume-translated modification was introduced to generate dependence in the vapor pressure curve with respect to the volume ((P/(ρ), thus obtaining an agreement in both sets of variables (single and double derivatives) without losing predictive capability in the prediction of vapor pressure values.

Figures 3 and 4 present the predicted speed of sound and bulk modulus values, respectively. In spite of the deviations, especially in the high temperature regime, the improvement in the calculation of these properties is remarkable when the translated volume modification is incorporated, in comparison with those presented in Figure 2, with the advantage that the previously obtained parameter set compensates the predictions of the liquid vapor equilibrium and second-derivative properties. Thus, the key properties of the biodiesel fuel components can be predicted with the same set of values for all temperature and pressure regimes considered. This is a practical advantage of our proposal worth stressing.

Figure 3 Speed of sound (v) vs temperature (T) calculated for all considered fatty acid methyl esters, using the volume translated parameters. The inset presents the same data, but on a finer scale, in the high temperature region. 

Figure 4 Isentropic bulk modulus (β) vs temperature (T) calculated for all fatty acid methyl esters, using the volume translated parameter. 

It can be seen that, for the speed of sound, the average deviations of the theoretical values from the experimental ones for the FAMEs with high carbon number are of 14 and 21 % in the low and high temperature regimes, respectively. These results are clearly superior compared to those in Figure 2, wherein the corresponding deviations in the low and high temperature regimes are of 35 and 38 %. Also, in Figure 3, a more uniform behavior is obtained for all FAMEs with high carbon numbers.

For the bulk modulus the aforementioned behavior is not observed, and there is an increasing deviation, from 17 to 29 %, of the theoretical values for all FAMEs as temperature increases.

Once the optimum parameter set values were determined, the properties of real commercial biodiesels were calculated. For this, a single Van der Waals one-fluid theory mixing rule (Equations 13 and 14) was implemented to take into account the mixing effects when dealing with a multicomponent FAMEs system. Biodiesel from palm, soy and lard was analyzed; its composition was taken from Goodrum (2002), and the P* results are reported in Figure 5, where the good agreement between our predictions (once the Rachord-Rice criteria are incorporated, Equation 18) and the experimental data taken from Goodrum (2002) can be readily seen. According to these results, it can be inferred that a single simple mixing rule is sufficient to consider the mixing effects in long-chained methyl esters.

Figure 5 Predicted (lines) bubble point values vs inverse temperature (1/T) for the commercial biodiesels obtained from palm oil, lard and soybean oil. Circles and squares correspond to predicted values from the Antoine equation reported in Yuan et al. (2005), and triangles to experimental values reported by Goodrum (2002)

The proposed model attains an appreciable improvement in the prediction of second-derivative properties like heat capacity (isochoric and isobaric), bulk modulus and speed of sound. It is important to remark that the accuracy of the prediction of liquid-vapor properties is kept with the same set of fitted parameters. In addition, the model depends on very limited experimental information (only pure FAMEs critical properties) and does not need other primary properties for its determination. Therefore, it can become a valuable tool for both the design of biodiesel injection systems and the interpretation of the combustion process.

Conclusions

The obtained results reveal that classical EOS with only a single α(T) modification cannot describe with acceptable accuracy both kinds of physico-chemical properties. However, fitting pure component parameters can significantly improve the speed of sound and bulk modulus predictions if experimental data are taken into account in the objective function to be optimized.

In the case of FAME mixtures (real biodiesel systems) a single mixing rule was sufficient to obtain very good agreement with experimental results already reported in the literature. Thus, in this work we have presented a model of predictive nature, based on physical principles, that allows the evaluation, with the same parameter set, of all physico-chemical properties necessary to assess the performance and quality of biofuels in processes such as atomization and combustion in diesel engines. In general, the proposed method turns out to be good for static properties but it is not so easy to apply for dynamic properties. However, we consider it a good approach since it affords a way to obtain a qualitative agreement of the speed of sound with available experimental data.

Acknowledgements

The first author thanks the “Programa Institucional de Formación de Investigadores" of the Instituto Politécnico Nacional, Mexico, the second author the Consejo Nacional de Ciencia y Tecnología, Mexico for financial support granted, and the third author the research department of Fundación Universitaria “Los Libertadores,” Colombia, for partially supporting this research.

References

Ahlers, J., & Gmehling, J. (2001). Development of an universal group contribution equation of state: I. prediction of liquid densities for pure compounds with a volume translated Peng-Robinson equation of state. Fluid Phase Equilibria, 191(1-2), 177-188. doi: 10.1016/S0378-3812(01)00626-4 [ Links ]

American Society of Testing Materials (ASTM). (2005). Standard specification for diesel fuel blend stock (B100) for middle distillate fuels, Designation: D 6751-03a. Philadelphia: Author. [ Links ]

Chapman, W. G., Gubbins, K. E., Jackson, G., & Radosz, M. (1990). New reference equation of state for associating fluids. Industrial & Engineering Chemistry Research, 29(8), 1709-1721. doi: 10.1021/ie00104a021 [ Links ]

Constantinou, L., Gani, R., & O’Connell, J. P. (1995). Estimation of the acentric factor and the liquid molar volume at 298 K using a new group contribution method. Fluid Phase Equilibria , 103(1), 11-22. doi: 10.1016/0378-3812(94)02593-P [ Links ]

Csernica, S. N., & Hsu, J. T. (2011). Reverse-phase ternary phase diagram, tie lines, and plait point for commercial biodiesel-glycerol-methanol. Industrial Engineering Chemistry Research, 50(2), 1012-1016. doi: 10.1021/ie101262u [ Links ]

Diedrichs, A., Rarey, J., & Mehling, J. G. (2006). Prediction of liquid heat capacities by the group contribution equation of state vtpr. Fluid Phase Equilibria , 248(1), 56-69. doi: 10.1016/j.fluid.2006.07.009 [ Links ]

Gasem, K. A. M., Gao, W., Pan, Z., & Robinson, R. L. (2001). A modified temperature dependence for the Peng-Robinson equation of state. Fluid Phase Equilibria , 181(1-2), 113-1245. doi: 10.1016/S0378-3812(01)00488-5 [ Links ]

Gil-Villegas, A., Galindo, A., Whitehead, P. J., Jackson, G., & Burgess, A. N. (1997). Statistical associating fluid theory for chain molecules with attractive potentials of variable range. The Journal of Chemical Physics, 106(10), 4168-4186. doi: 10.1063/1.473101 [ Links ]

González-Pérez, A., Coquelet, C., Paricaud, P., & Chapoy, A. (2017). Comparative study of vapour-liquid equilibrium and density modelling of mixtures related to carbon capture and storage with the SRK, PR, PC-SAFT and SAFT-VR Mie equations of state for industrial uses. Fluid Phase Equilibria , 440, 19-35. doi: 10.1016/j.fluid.2017.02.018 [ Links ]

Goodrum, J. W. (2002). Biomass & Bioenergy. Industrial Engineering Chemistry Research , 22(3), 205-211. [ Links ]

Hoekman, S. K., Broch, A., Robbins, C., Ceniceros, E., & Natarajan, M. (2012). Review of biodiesel composition, properties, and specifications. Research and Science Energy Reviews, 16(1), 143-169. doi: 10.1016/j.rser.2011.07.143 [ Links ]

Huber, M. L., Lemmon, E. W., Kazakov, A., Ott, L. S., & Bruno, T. J. (2009). Model for the thermodynamic properties of a biodiesel fuel. Energy & Fuels, 23(7), 3790-3797. doi: 10.1021/ef900159g [ Links ]

Knothe, G. (2010). Biodiesel and renewable diesel: A comparison. Progress in Energy and Combustion, 36(3), 364-373. doi: 10.1016/j.pecs.2009.11.004 [ Links ]

Liang, X., Maribo-Mogensen, B., Thmosen, K., Yan, W., & Kontogeorgis, G. M. (2012). Approach to improve speed of sound calculation within pc-saft framework. Industrial Engineering Chemical Research, 51(45), 14903-14914. doi: 10.1021/ie3018127 [ Links ]

Nocedal, J., & Wright, S. J. (2000). Numerical optimization. New York: Springer Series. [ Links ]

Olivera, M. B., Ribeiro, V., Queimada, A. J., & Coutinho, J. A. P. (2011). Modeling phase equilibria relevant to biodiesel production: A comparison of gE models, cubic EoS, EoS-gE and association EoS. Industrial & Engineering Chemical Research, 50(4), 2348-2358. doi: 10.1021/ie1013585 [ Links ]

Olivera, M. B., Teles, A. R. R., Queimada, A. J., & Coutinho, J. A. P. (2009). Phase equilibria of glycerol containing systems and their description with the cubic-plus-association (CPA) equation of state. Fluid Phase Equilibria , 280(1-2), 22-29. doi: 10.1016/j.fluid.2009.03.011 [ Links ]

Osmont, A., Catoire, L., & Gkalp, I. (2007). Thermochemistry of methyl and ethyl esters from vegetable oils. International Journal of Chemical Kinetics, 39(9), 481-491. doi: 10.1002/kin.20264 [ Links ]

Peng, D. Y., & Robinson, D. B. (1976). A new two-constant equation of state. Industrial & Engineering Chemical Fundamentals, 15(1), 59-64. doi: 10.1021/i160057a011 [ Links ]

Perdomo, F. A., & Gil-Villegas, A. (2011). Predicting thermophysical properties of biodiesel fuel blends using the saft-vr approach. Fluid Phase Equilibria , 306(1), 124-128. doi: 10.1016/j.fluid.2011.02.021 [ Links ]

Perdomo, F. A., & Gil-Villegas, A. (2010). Molecular thermodynamic of biodiesel fuel compounds. Fluid Phase Equilibria , 293(2), 182-189. doi: 10.1016/j.fluid.2010.03.011 [ Links ]

Perdomo, F. A., Perdomo, L., Millán, B., & Aragón, J. L. (2014). Design of improvement of biodiesel fuel blends by optimization of their molecular structures and compositions. Chemical Engineering Research and Design, 92(8), 1482-1494. doi: 10.1016/j.cherd.2014.02.011 [ Links ]

Rackett, H. G. (1970). Equation of state for satured liquids. Journal of Chemistry Engineering Data, 15(4), 514-517. doi: 10.1021/je60047a012 [ Links ]

Rose, A., & Supina, W. R. (1961). Vapor pressure and vapor-liquid equilibrium data for methyl esters of the common saturated normal fatty acids. Journal of Chemical & Engineering Data, 6(2), 173-179. doi: 10.1021/je60010a003 [ Links ]

Schönborn, A., Ladommatos, N., Williams, J., Allan, R., & Rogerson, J. (2009). The influence of molecular structure of fatty acid monoalkyl esters on diesel combustion. Combustion and Flame, 156(7), 1396-1412. doi: 10.1016/j.combustflame.2009.03.011 [ Links ]

Scott, T. A., MacMillan, D., & Melvin, E. H. (1952). Vapor pressures and distillation of methyl esters of some fatty acids. Industrial and Engineering Chemistry, 44(1), 172-175. doi: 10.1021/ie50505a049 [ Links ]

Tat, M. E., & Van Gerpen, J. H. (2003). Speed of sound and isentropic bulk modulus of alkyl monoesters at elevated temperatures and pressures. Journal of the American Oil Cchemists’ Society, 80(12), 1249-1256. doi: 10.1007/s11746-003-0850-6 [ Links ]

Yuan, W., & Hansen, A. C. (2009). Computational investigation of the effect of biodiesel fuel properties on diesel engine NOx emissions. International Journal of Agricultural & Biological Engineering, 2(2), 41. doi: 10.3965/j.issn.1934-6344.2009.02.041-048 [ Links ]

Yuan, W., Hansen, A. C., & Zhang, Q. (2005). Vapor pressure and normal boiling point predictions for pure methyl esters and biodiesel fuels. Fuel, 84(7-8), 943-950. doi: 10.1016/j.fuel.2005.01.007 [ Links ]

Yuan, W., Hansen, A. C., & Zhang, Q. (2003). Predicting the physical properties of biodiesel for combustion modeling. Transactions of the American Society of Agricultural Engineering, 46(6), 1486-1493. Retrieved from https://elibrary.asabe.org/azdez.asp?AID=15631&T=2Links ]

Appendix. Calculation details

A.1 Vapor pressure in pure biodiesel compounds

In a pure substance, vapor pressure is given by the solution of liquid-vapor equilibria criteria, which are determined by the following conditions:

TL=TV

PL=PV (21)

ϕLT,P=ϕVT,P

where ( represents the fugacity coefficient, and for VTPR-EOS it is given by (Peng & Robinson, 1976):

lnϕ=Z+Ceos-1-lnZ+Ceos-Beos-Aeos22BeoslnZ+Ceos+(1+2)BeosZ-Ceos+(1-2)Beos (22)

with

Aeos=aTc,PcαTr,ω,A, B, C, D, EPR2T2

Beos=bTc,PcPRT

Ceos=cTc,PcPRT

Z=PρRT (23)

The above expressions are used in the recurrence relation (Equation 9) in the “Modeling thermophysical properties in biodiesel systems” section.

A.2 Liquid heat capacities

The calculations of this parameter start from the following relations:

Cp-Cv=TVTPPTV (24)

and

Cv-Cvid=TV2PT2VdV (25)

where Cvid is the ideal gas isochoric heat capacity and is correlated with the ideal gas isobaric heat capacity Cpid (Diedrichs et al., 2006) by the relation:

Cvid=Cpid-R (26)

To estimate the ideal gas heat capacity in biodiesel compounds, the “Planck-Einstein” expression fitted for several FAMEs and given by Huber et al. (2009) was implemented:

Cp,iid=cp0Tcp1+cp2cp3T2expcp3T1-expcp3T2+cp4cp5T2expcp5T1-expcp5T2+cp6cp7T2expcp7T1-expcp7T2 (27)

where the T is in units of K and Cpid is in units of J·mol-1·K-1. The values of constants cp 0 to cp 7 for the FAMEs considered in this work are given in Table 4.

To evaluate heat capacity predictions for the selected α-function the manipulation of Equation 25 leads to Equation 10 in the main text. In order to explicitly evaluate the required derivatives in Equation 24, the following relation was used:

VTP=-PTVPVT (28)

The second way to compute the heat capacities is from their thermodynamical definition; i.e. from the change of enthalpy with the temperature (Diedrichs et al., 2006):

Cp=hTPhT+δT-hT-δT2δT (29)

with

hT=hidT+hT,PD (30)

where h D is the departure enthalpy, given by the difference between its value for a given substance as measured in an experimental setup and its value calculated for an ideal gas h id, both at a specified T and P (Diedrichs et al., 2006):

hDRT=Z-1+1RTVTPTV-Pdv (31)

The ideal gas enthalpy is only a function of temperature and is calculated as:

hidT=hidT0+T0TCpidT (32)

To estimate (arbitrarily) the h id (T 0 ) for the herein studied FAMEs, the standard heat of formation in the ideal gas state of Osmont, Catoire, and Gkalp (2007), which was determined by means of quantum chemistry calculations, was used.

A.3 Latent heat of vaporization

For biodiesel combustion modeling, a key input parameter is the latent heat of vaporization (enthalpy of vaporization, Δh v ). This property has been studied in order to determine its influence on NOx emissions (Yuan & Hansen, 2009). Also, Δh v is an important parameter for studying the fuel spray process in a diesel engine. From the Clausius-Clapeyron equation, a quantitative connection between the slope of the vapor pressure curve and the enthalpy of vaporization is obtained:

dPsatdT=hvTvg-vl (33)

The PVT behavior of the vapor and liquid may be described in terms of compressibility factor as:

Pρl,gRT=Zl,g (34)

Combining Equations 33 and 34 obtains the well-known relation, in terms of reduced properties:

hvRTc=-Zg-Zld ln Prd1/Tr (35)

The gas-phase standard enthalpies obtained at reference temperature (298.15 K) for the five FAMEs are given in Table 5.

Nomenclature

a,b

Peng-Robinson constants

c

Volume translation parameter

C p

Heat capacity at constant pressure

C v

Heat capacity at constant volume

Mj

Number of contributions of second-order group

Ne

Number of experimental data point for vapor pressure

Ni

Number of contributions of first-order group

P

Pressure

Pisat

Vapor pressure in pure biodiesel compounds

Pvap

Vaporization pressure

P0 *

Vapor pressure in the bubble point

R

Universal gas constant

T

Temperature

Tc, Pc, Zc

Conditions at critical point

Tr, Pr

Temperature and Pressure reduced

v

Molar volume

xi

Molar fraction in the liquid phase of component i

yi

Molar fraction in the vapor phase of component i

Z

Compressibility factor

Greek symbols

α(T)

Temperature-dependent model function

α, γ

Clausius-Clapeyron constants

β

Isentropic bulk modulus

Δhv

Enthalpy of vaporization

ΔHid

Ideal gas enthalpy

ϕ

Fugacity coefficient

κi

Distribution coefficient of component i

ns

Speed of sound

ρ

Density

ω

Acentric factor

ω1i

Contribution of the type first-order group

ω2j

Contribution of the type second-order group

ψ

Vaporized fraction of biodiesel

Received: October 23, 2017; Accepted: February 02, 2018

*Corresponding author: mnunez@ipn.mx, tel. (+52) 55 57 29 60 00, ext. 82509.

Creative Commons License Este es un artículo publicado en acceso abierto bajo una licencia Creative Commons