Skip to content
Publicly Available Published by De Gruyter December 1, 2022

Densities and isothermal compressibilities from perturbed hard-dimer-chain equation of state: application to nanofluids

  • Mahsa Taghizadehfard , Sayed Mostafa Hosseini EMAIL logo , Mariano Pierantozzi ORCID logo and Mohammad Mehdi Alavianmehr

Abstract

Densities and isothermal compressibilities of several nanofluids were modelled using a perturbed hard-chain equation of state (EoS) by an attractive term from Yukawa tail in 273–363 K range and pressure up to 45 MPa. The nanofluids of interest comprise TiO2-Anatase (-A), TiO2-Rutile (-R), SnO2, Co3O4, CuO, ZnO, and Al2O3 as nanoparticles dispersed in ethylene glycol, water, poly ethylene glycol, ethylene glycol + water, and poly ethylene glycol + water as base fluids. The EoS was capable of estimating 1397 density data of 9 nanofluids with the overall average absolute deviations (AAD) of 0.90%. The coefficients of isothermal compressibility of 6 selected nanofluids were also predicted using the EoS with the AAD of 5.74% for 1095 data points examined. The PHDC EoS was not capable of estimating the excess volumes of 3 selected EG-, PEG-, and water-based nanofluids accurately as the relative deviations from the literature data were greater than 34%, even though the trend of results against the nanoparticle concentration was in accord with the literature. To further investigate the density prediction, we have trained a neural network with a single hidden layer and 17 neurons which was able to predict the densities of nanofluids accurately.

1 Introduction

Nanofluids are a new class of heat transfer fluids, containing a suspension of nanometer-sized solid particles dispersed in a base (pure) liquid. Nanoparticles can be dispersed in several liquids. However, the base fluids are generally chosen among one the widely used conventional heat transfer fluids such as water, ethylene glycol, poly ethylene glycol, refrigerant, or oils [1]. Materials generally used as nanoparticles include stable metals, such as Au, Ag, Cu, oxides, such as CuO, Fe2O3, Al2O3, TiO2, carbon in various forms, and other particles, such as Si compounds [2].

When a nanofluid is prepared, compared to the base fluid, some changes in its thermophysical properties such as thermal conductivity, viscosity and density such mixtures occur. Recently, nanofluids have been brought into focus due to new potential applications in cooling devices, is widely used in many engineering applications such as electronics cooling, combustion engines, nuclear reactors, the food industry, air-conditioning systems, refrigeration, antimicrobial behavior and biomedicine [1, 3, 4].

Despite the most widely studied properties of nanofluids being those related to transport properties, the accurate knowledge of other thermophysical properties such as density is crucial as it is beneficial for the study of thermal cycles and engineering systems [5, 6] where the density must be known accurately to compute matter and energy balances [7]. Although numerous efforts have been yet carried out to measure the volumetric properties nanofluids, most of which are limited to the atmospheric pressures and less attention has been paid to the volumetric measurements in the extended pressure range. Under this circumstance, the development of empirical, semi-theoretical models, accurate correlations, and equations of state (EoSs) for predicting the density data of nanofluids at various nanoparticle concentrations can be considerably useful to supplement the available high-pressure volumetric data.

Although various methods based on empirical, semi-empirical, semi-theoretical, artificial neural network, and molecular dynamics simulation [822] approaches, have been yet developed to determine the transport properties of nanofluids, less attention has been devoted to present some analytical models such as molecular-based EoSs for the volumetric properties of nanofluids in the extended pressure ranging. Along the next several paragraphs, we provide a comprehensive review for the various approaches which have been yet developed for predicting the volumetric data of nanofluids.

Most of the works carried out for the calculation of volumetric properties of nanofluid are based on the continuum model approaches [23] and the theory of heterogeneous mixtures [24] that relate linearly the density of base fluid to nanoparticle density. Regarding this, several EoSs and PρT correlations have been yet suggested to model the density data of nanofluids and nanomaterials. For instance, Sharma–Kumar [25] and Pak–Cho [26] adopted the most widely used Tammann–Tait equation [27, 28] for nanometer-sized particles to model the density of nanofluids. Vajjha et al. [29] used Pak–Cho equation to estimate the nanofluid densities however the need for accurate values of the base fluid and nanoparticle densities limits the applicability of the Pak–Cho equation. Tan and Piri [30] coupled a perturbed-chain statistical associating fluid theory (PC-SAFT) with Younge–Laplace equation [31] to represent the fluid phase equilibria in nanometer size cylindrical pores mediums. Islam et al. [32] used the modified van der Waals EoS to determine the phase behavior of nano-pore confined fluids, but this EoS was limited to only subcritical conditions. In 2017, Islam and Sun [33] developed a simple extension of Peng–Robinson (PR) EoS for nano-pore confined fluids. This model was capable of predicting the heterogeneous density of molecules inside the nano-pore structures.

Apart from EoSs and PρT correlations, Artificial Neural Network (ANN) models have recently come into focus for predicting the volumetric data of nanofluid systems. For instance, Karimi and Yousefi [34] applied the ANN model to estimate the density of four nanofluids based on EG + water and water in 273–323 K range with the average absolute deviation (AAD) of 0.13%. In 2017, the same research group [35] employed a combined ANN model with a principal component analysis technique for predicting the densities of 6 nanofluids; the authors obtained an AAD of 0.48%. Also, Alade et al. [36] recently developed an ANN model for the densities of nanofluids including carbon-based nanomaterials with 1.3 × 10−3 (g cm−3) of the root mean square error as the authors assessed the accuracy of their network using 22 experimental testing datasets.

Tammann–Tait equation is one of the most widely used PρT correlations to correlate the high-pressure volumetric data of liquids [28, 37]. Nevertheless, the use of that correlation for nanofluids is limited to the fitting of numerous coefficients to each nanoparticle concentration [29, 35]. This problem was fixed by the work of Hosseini and Alavianmehr [38] who proposed a nanoparticle concentration-dependent function to be introduced to the original Tammann–Tait equation, then Hosseini and Alavianmehr approach reduced significantly the number of fit coefficients appearing in the original scheme. The authors calculated 1215 data points for densities of 6 nanofluids using that approach in 273–363 K range and pressure up to 45 MPa with the AAD of 0.061%.

The molecular-based EoSs have been widely used by the academic researchers and engineers over the past decades for the accurate prediction of PρT data of highly non-simple fluids like molten polymers [39]and ionic liquids [40]. In the case of nanofluids, Hosseini et al. [41] examined the application of a perturbed hard-sphere (PHS) EoS for modeling the volumetric properties of some nanofluids over the pressure ranging from 0.1 to 45 MPa and a temperature range from 273 to 363 K.

In 2017, Yousefi et al. [35], coupled Tao–Mason [42], EoS with Pak–Cho equation [26] to calculate the densities of nanofluids. The authors have obtained an AAD of 1.11% for their Pak–Cho + TM EoS model. The authors also employed an ANN plus principal component analysis (PCA) technique for the density predictions with the AAD of 0.48%.

In 2018, Montazer et al. [43] applied a density correlation for carbon-based nanofluids using response surface methodology (RSM) [44, 45]. The optimal RSM model was quite accurate; the maximum absolute deviations of the predicted densities from the measured values were within 0.012%–0.009%.

This work is the continuation of our investigations on the modeling of the physiochemical properties of highly non-simple fluids, in particular, nanofluids [22, 38, 41]. The previous effort led to the application of a perturbed hard-chain EoS for modeling the volumetric properties of several molten polymers and ionic liquids [39, 46]. That EoS comprises a hard-dimer-chain reference fluid perturbed by a long-range attraction of Yukawa [47] with a variable single parameter, λ for the representation of the range of attractive forces. This paper aims to assess the capability of the perturbed hard-dimer-chain (PHDC) EoS mentioned above for the calculation of properties of nanofluids including the densities, isothermal compressibilities, and excess volumes as it would be interesting to see how accurate the model is when it is extended to the colloidal suspensions. Further, the PHDC EoS results for the base fluid and nanofluid densities are compared with the Pak–Cho + TM EoS model of Yousefi et al. [35] to take an insight into the degree of precision of our model.

In recent years, one of the most widely used techniques for the calculation of the physical properties of fluids is the neural network (NN). In fact, this technique allows estimating in a very accurate and precise way the experimental data starting from input parameters that must be properly chosen. Neural networks have been yet used for the calculation of several thermophysical properties of nanofluids, in particular, their density as mentioned in the above literature survey [3436].

In the present work, the NN method is going to be used for the density prediction of nanofluids at compressed states, i.e., the pressures above 0.1 MPa. Besides, NN method just uses the artificial intelligence-based computer algorithms rather than the physical context of materials under study such as the continuum approach [23] and the theory of heterogeneous mixtures [24] as both theories have been yet used in the literature for the calculation of densities [26, 29, 35] of nanofluids. Despite the origins of both NN method (as a computer-aided model) and continuum model approaches are so different, the two approaches should offer similar results.

2 Theoretical considerations

A summary of the theoretical consideration of PHDC EoS used in this work for the pure fluids has been provided in Appendix A as the details of that consideration are available in the literature [39, 48]. That is why the mixture version of the PHDC EoS under consideration is only mentioned herein as it is going to be extended to nanofluids for calculating their volumetric properties. In this respect, the relevant formalism can be read as:

(1) Z mix PHDC = P ρ k B T = 1 + m ̄ 4 η mix 2 η mix. 2 ( 1 η mix. ) 3 m ̄ 2 5 / 2 η mix. η mix. 2 1 η mix. 1 1 / 2 η mix ( m ̄ 2 ) 2 × 2 η mix. + η mix. 2 ( 1 η mix. ) 1 / 2 + η mix. 12 m ̄ η mix. T mix. * I ( η mix. , m ̄ ; λ = 1.8 )

and T * mix. is the reduced temperature of fluid mixture obtained from the scaling of absolute temperature in terms of non-bonded interaction energy between spheres and dimers (ε/k B) ij , viz:

(2) T mix * = T ε k B i j

where, the subscript “mix.” stands for the mixture, Z PHDC is the compressibility factor for the fluid mixture of interest, P is the pressure, and k B T is the thermal energy per segment. In Eq. (1), ρ is the number density of segments. In the context of the perturbation theory of hard-dimers [49], a segment unit is a dimer set of hard-spheres so that the reference fluid consists of pairs of dimers that form chains (e.g., tetramers) through covalent bonds. Regarding this, m stands for the number of monomer sites or that segment number. η mix is the segment packing fraction of a mixture of hard-chain defined through the following equations.

(3) η mix = i n x i m i η i

(4) m ̄ = i n x i m i

where, x i and η i are the mole fraction and packing fraction of ith component, respectively.

Also, the following geometric mean and arithmetic mean combining rules for the cross-energy term and cross-size term were also adapted, viz.:

(5) ε k B i j = ε k B i ε k B j ( 1 k i j )

(6) σ i j = 1 2 ( σ i + σ j )

where, k ij stands for a binary interaction parameter that was introduced to the crossed energy term (i.e., Eq. (5)) for understanding the measure of interplay between nanoparticle surface-base fluid and base fluid-base fluid molecules. Then two types of k ij have to be adapted for Eq. (5) as the first one is devoted to the nanoparticle surface-base fluid molecules, k np-bf and the second types is referring to the operating forces among the base fluid-base fluid molecules, k 12. Generally, the use of binary parameters in the study of homogenous mixtures not only makes it feasible to correlate their properties more accurately, also those parameters allow us to take some insights into the molecular interactions operating among the like and unlike molecules [50]. In the meanwhile, the preceding work of Hosseini et al. [41] showed that the binary parameters mentioned above can also be extended to heterogeneous mixtures such as the colloidal suspensions, in which the nanoparticles have interactive surfaces with the base fluid molecules leading to some contractive or expansive volumetric trend in the mixing functions, then the binary parameters can be highly useful to investigate those interactions.

3 Artificial neural network modeling

A neural network is a mathematical model comprising artificial neurons inspired by biological neural networks and is used to solve engineering problems of Artificial Intelligence related to different technological fields such as computer science, electronics, and physics. ANNs can be characterized [51] as models that are composed of at least two layers, an input layer and an output layer, and usually additional intermediate layers (hidden layers). The more complex the problem to be solved with the artificial neural network, the more layers are needed [52]. Each layer of the network contains a number of specialized artificial neurons. The processing of information in the neural network always follows the same procedure: the information in the form of patterns or signals is forwarded to neurons in the input layer, where it is processed. Each neuron is then allocated a weight, so that the neurons receive different importance. The weight, along with a transfer function, determines the input, where then the neuron is forwarded.

There are different transfer functions both linear and non-linear that are chosen depending on the problem being developed [53]. The most common functions used in the literature are the linear function, the hyperbolic tangent and the sigmoid.

In our case, we used the sigmoid. When considering a multi-layer neural network, the activation function ensures that the outputs are as close as possible to the experimental data. The purpose of the activation function is to modify the weights and biases to make them as close as possible.

A major reason why the sigmoid function is often used is that its output is between 0 and 1 while the inputs are between minus infinity and plus infinity. Another positive aspect of the sigmoid function is that the function is derivable with continuous derivative, which allows descent algorithms to work correctly. Furthermore, it is often used for non-linear problems such as the one under consideration.

In the next step, the activation function and a threshold value calculate and weight the output value of the neuron. Depending on the information evaluation and weighting, other neurons are connected and activated to a greater or lesser degree. Through these processes an algorithm is modeled. This model generates an output for each input. With each training, both the weighting and thus the algorithm are modified so that the network provides more accurate and better results.

4 Results and discussion

4.1 Estimation of densities and isothermal compressibilities

The values of molecular parameters, ε, σ and m appearing in the PHDC EoS have to be characterized prior to use that EoS for the estimation of densities and isothermal compressibilities of nanofluids. For the case of base fluids, ε, σ and m values were fitted against the PρT data taken from the literature [5456] and the minimization of the following objective function (OBF) based on the least-squares method:

(7) OBF = min 1 n i = 1 n ρ i Calc ρ i Exp ρ i Exp.

where, n is the number of density data used in the fit procedure. Generally, from 260 literature data points [5456] applied to 3 base fluids of interest, the minimum value of OBF was found to be 0.0070. But in case the nanoparticles of studied, the three molecular parameters (ε, σ, m) were fitted from true density data of compacted nanopowder [57, 58] over the whole temperature range of interest. Table 1 reports the required molecular parameters (ε, σ, m) as well as the molecular weights of studied systems.

Table 1:

The optimized molecular parameters along with the AADs of fits of based fluids and nanoparticles studied in this work.

Pure base fluid M W (g/mol)a ε/k B (K) σ (nm) m ΔT/K MPa ΔP/ NPc AAD (%)d AAD (%)e Ref.
EG 62.07 352.17 0.30092 2.803 283–343 0.1–45 98 0.63 3.73 [54]
Water 18.03 339.84 0.16952 4.738 280–370 0.1–50 112 0.60 10.86 [55]
PEG 60.00b 347.26 0.29951 2.717 298–328 0.1–40 50 0.28 3.84 [56]
Nanoparticle M W (g/mol)a ε/k B (K) σ (nm) m AAD (%)
TiO2-Rutile 79.93 3004.30 0.2539 2.335 1.24
TiO2-Anatase 79.93 2723.90 0.2539 2.432 1.30
SnO2 150.69 2499.90 0.2523 2.566 1.06
Co3O4 240.82 1630.90 0.2836 2.956 3.47
CuO 79.54 1849.80 0.2268 2.156 3.90
ZnO 81.40 3200.20 0.2228 2.517 2.85
Al2O3 101.96 3396.30 0.2642 2.666 3.19
  1. aMolecular weight. bMonomer Molecular weight. cNP represents the number of data points examined. d AAD = 100 / NP i = 1 NP ρ i Calc ρ i Exp. ρ i Exp. . e AAD = 100 / NP i = 1 NP k T i Calc k T i Tait. k T i Tait. .

First of all, the PHDC EoS was assessed for the densities of several base fluids comprising ethylene glycol (EG), water and poly ethylene glycol (PEG) as well as their mixtures as they are widely used for cooling purposes. Some binary interaction parameters (k 12) have also to be introduced to the model as the mixtures of base fluids mentioned above are strongly solvating systems and therefore these binary systems are non-ideal mixtures. Table 2 also compares the AADs (in %) of the calculated densities of mixed base fluids studied from the PHDC EoS (this work), and the preceding Pak–Cho + TM EoS model of Yousefi et al. [35], both of which were compared with the experimental data [5961]. As it is clear from Table 2, for the case of mixed base fluids, the accuracy of PHDC EOS (with AADs within 0.76%–3.04%) was comparable with Pak–Cho + TM EoS model [35] (with AADs within 0.97%–2.98%).

Table 2:

The AADa (in %) of calculated densities of mixed base fluids and nanofluids studied in this work, using the PHDC EoS (this work), and the method of Yousefi et al. [35] (Pak–Cho + TM EoS model), both compared with the measurements. Δx Water stands for the water mole fractions, in which the calculations were carried out.

Mixed base fluid ΔP/MPa ΔT/K Δx Water NPb k 12 PHDC EoS Yousefi et al. model Ref.
EG + water 0.1 298–328 0.00–1.00 44 −0.0334 0.76 0.97 [61]
0.1–45 278–363 0.755–0.755 55 −0.0334 2.66 2.78 [59]
PEG + water 0.1–0.1 283–313 0.00–1.00 77 −0.3584 3.04 2.98 [60]
Nanofluid ΔP/MPa  ΔT/K Δx np NPb k bf-np PHDC EoS Yousefi et al. model ANN model Ref.
TiO2-R/EG 0.1–45 283–343 0.014–0.039 30 0.65424 0.78 1.12 0.18 [62]
TiO2-A/EG 0.1–45 283–343 0.014–0.039 30 0.63813 0.76 1.11 0.22 [62]
SnO2/EG 0.1–45 283–323 0.004–0.021 270 0.62813 0.51 2.79 0.07 [54]
Co3O4/EG 0.1–45 283–323 0.0026–0.014 270 0.56212 0.90 2.98 0.06 [63]
CuO/water 0.1–45 283–323 0.002–0.012 450 0.56213 0.57 3.99 0.06 [64]
ZnO/PEG 0.1 293–318 2E−4−0.077 56 0.67855 1.40 0.11 [65]
ZnO/PEG + water 0.1 293–318 9E−5−0.0157 48 0.12940 0.74 1.62 [65]
ZnO/EG + water 0.1–45 278–363 0.009–0.038 165 0.90730 1.02 2.88 0.08 [59]
0.1 273–323 0.021–0.041 42 0.67504 0.40 1.32 [29]
Al2O3/EG + water 0.1 273–323 0.012–0.107 36 −0.92640 1.91 1.96 [29]
Overall 1397 0.90 1.99 0.07c
  1. a AAD = 100 / NP i = 1 NP ρ i Calc ρ i Exp ρ i Exp. . bNP represents the number of data points examined. cFor 1215 data points examined.

Despite Table 2 showing that the proposed PHDC model improved the results given by the Yousefi et al. model in general, in the case of ZnO/PEG the contrary occurs. For this case, it is somewhat attributed to the type of fit procedure used for each model. The PHDC EoS parameters for the base fluid (i.e., PEG) were fitted in the extended temperature and pressure ranging (as reported in Table 1) in contrast to the model of Yousefi et al. [35] that fitted λ-parameter of TM EoS by the help of several correlation coefficients at limited pressure (0.1 MPa) and temperature ranging in which the nanofluid densities were studied and then it obviously would lead to less AADs in the densities of PEG and ZnO/PEG nanofluid, consequently.

Negative values obtained for k 12 are not against the expectations of intermolecular forces and thermodynamic models [50], “in the case of strongly solvating systems (such as EG + water studied herein), the real crossed energy term is expected to be larger than the value provided by the geometric mean rule and this is why a negative k 12 is needed”. Anyhow, such binary interaction parameters sometimes do not make any sense from the molecular viewpoint as they are simply adjustable parameters.

Then, we extended the mixture version of PHDC EoS to nanofluids with the help of several combining rules as mentioned in Eqs. (1)(6). Regarding this, the binary interaction parameter appearing in Eq. (5) was fitted against the density data at 0.1 MPa and then PHDC EoS was extended to estimate the density values in wider ranges of pressure. All the calculations on the mass density of 9 nanofluids (including TiO2-R/EG, TiO2-A/EG, SnO2/EG, Co3O4/EG, CuO/water, ZnO/PEG, ZnO/PEG + water, ZnO/EG + water and Al2O3/EG + water) were carried out in 273–363 K range and pressures up to 45 MPa, then the results were compared with the experimental data [29, 54, 59, 62], [63], [64], [65]. As indicated in Table 2, for 1397 data points examined, the mean AAD of the calculated densities from the literature values was found to be 0.90%. The numerical values of k ij parameter for each nanofluid have also been included in Table 2. The sign of that parameter indicated if the operating forces between nanoparticle surface-base fluid and base fluid-base fluid molecules were of repulsive or attractive type. As can be seen from Table 2, the type of interaction of nanoparticle surface-base fluid molecule is dominantly repulsive. But in case the base fluid-base fluid molecules, an attractive interaction is the dominant type. The uncertainty (or that of error propagation) of the calculated densities of all studied systems was of the order of ±3.09%.

Table 2 also provided a summary of the AADs obtained for the densities from the PHDC EoS (present work), and those obtained from Pak–Cho + TM EoS model [35], both of which were compared with the literature data [29, 54, 59, 62], [63], [64], [65]. The present work outperforms (with the AAD of 0.90% for 1397 data points examined) Pak–Cho + TM EoS model [35] (i.e., with AAD of 1.99% for 1397 examined data points) which used the extended Tao–Mason EoS, together with the Pak-Cho equation.

Graphically, Figure 1 shows how are the nanofluid densities coming from the PHDC EoS when compared with the experimental values. It depicts the isothermal variation of densities against the pressure for some studied nanofluids including TiO2-A/EG (a-plot), SnO2/EG (b-plot) and ZnO/EG + water (c-plot). The markers represent the experimental isotherms of Cabaleiro et al. [62], Mariano et al. [54] and Cabaleiro et al. [59], the solid lines are those obtained using the PHDC EoS. As it is clear from Figure 1, the linear trends associated with isotherms of Cabaleiro et al. [59, 62] and Mariano et al. [54] can be shown well by the present model with the help of crossed interaction parameters. Figure 1 also demonstrates that the density of nanofluid decreases as the temperature increases (along the isobaric curve). As can be seen from 1(a) and 1(b) plots, the calculated nanofluid densities from PHDC EoS showed larger deviations from the measurements when compared with those reported in 1(c) for the case of EG + water-based nanofluid. Then using two adjustable parameters, k 12 and k bf-np improved the accuracy of PHDC EoS for the densities of nanofluids containing mixed base fluids.

Figure 1: 
The linearity of the nanofluid densities versus pressure for TiO2-A/EG (at x
np = 0.014) at several isotherms ((a)-plot), SnO2/EG (at x
np = 0.01045) at several isotherms ((b)-plot) and ZnO/EG + water (at x
np = 0.0183) at several isotherms ((c)-plot). (a)-plot represents the results at 283 K (◊), 313 K (∆) and 343 K (⚬). (b)-plot represents the results at 283 K (◊), 303 K (∆) and 323 K (⚬) and (c)-plot represents the results at 278 K (◊), 303 K (∆), 323 K (⚬) and 343 K (□). The markers represent the experimental isotherms of Cabaleiro et al. [62] Mariano et al. [54] and Cabaleiro et al. [59], and solid lines are coming from PHDC EoS.
Figure 1:

The linearity of the nanofluid densities versus pressure for TiO2-A/EG (at x np = 0.014) at several isotherms ((a)-plot), SnO2/EG (at x np = 0.01045) at several isotherms ((b)-plot) and ZnO/EG + water (at x np = 0.0183) at several isotherms ((c)-plot). (a)-plot represents the results at 283 K (◊), 313 K (∆) and 343 K (⚬). (b)-plot represents the results at 283 K (◊), 303 K (∆) and 323 K (⚬) and (c)-plot represents the results at 278 K (◊), 303 K (∆), 323 K (⚬) and 343 K (□). The markers represent the experimental isotherms of Cabaleiro et al. [62] Mariano et al. [54] and Cabaleiro et al. [59], and solid lines are coming from PHDC EoS.

The isothermal compressibility is calculated using the isothermal pressure derivative of density according to the following classical thermodynamic relation, viz.;

(8) k T = 1 V P V T 1 = 1 ρ P ρ T 1

Then, the values of isothermal compressibility coefficient of some selected nanofluids were predicted using the PHDC EoS. We summarized our calculation results in Table 3 for several nanofluids including TiO2-R/EG, TiO2-A/EG, SnO2/EG, Co3O4/EG, CuO/water, and ZnO/EG + water as the AAD (in %) of our estimations from those reported by Cabaleiro et al. and Mariano et al. [54, 59, 62, 63], who utilized Tammann–Tait equation to correlate the high-pressure volumetric behavior of nanofluids. We assessed the PHDC EoS by taking 1095 literature data points [54, 59, 62], [63], [64] over the pressure ranging from 0.1 to 45 MPa and temperature range from 283 to 343 K. The overall AAD for the tested data points calculated by PHDC EoS was found to be 5.74%. It should be mentioned that the uncertainty of calculated k T values by the present work was of the order of ±8.98%. The predicted values of that property have also been tabulated as supplementary material to this article.

Table 3:

AADa (in %) of predicted isothermal compressibility (κT) of some studied nanofluids using the PHDC EoS (this work), compared with those obtained from the Tait equation.

Nanofluid P/MPa T/K Δx np NPb AAD (%) Ref.
TiO2-R/EG 0.1–45 283–343 0.014–0.039 30 5.57 [62]
TiO2-A/EG 0.1–45 283–343 0.014–0.039 30 5.27 [62]
SnO2/EG 0.1–45 283–323 0.004–0.021 270 3.47 [54]
Co3O4/EG 0.1–45 283–323 0.0026–0.014 270 3.46 [63]
CuO/water 0.1–45 283–323 0.002–0.012 450 11.54 [64]
ZnO/EG + water 1–40 303–343 0.009–0.038 45 5.72 [59]
Overall 1095 5.74
  1. a AAD = 100 / NP i = 1 NP k T i Calc k T i Tait. k T i Tait. . bNP represents the number of data points examined.

As can be seen from Table 3, the AAD obtained for CuO/water nanofluid is greater than 10%. Maybe it should be searched the source of that result in Table 1, where the fit parameters along with the correlated densities and isothermal compressibilities were reported for pure water as highly associating/H-bond type fluid. As Table 3 reveals, the proposed PHDC EoS worked well for the correlation of densities of pure water with the AAD of 0.6%, but in case the isothermal compressibility, the deviation of results from Tait equation were increasing which indicates the loss of accuracy of PHDC EoS for highly associating/H-bond fluids like, indeed the present PHDC EoS was capable of correlating the density and isothermal compressibility of less associating fluids (e.g., EG containing both aliphatic and associating characters) more accurately.

To see how the PHDC EoS passes through the literature data, Figure 2 has been provided. Figure 2 depicts four isothermal compressibility versus pressure plots for TiO2-R/EG at x np = 0.039 ((a)-plot), SnO2/EG at x np = 0.01472 ((b)-plot), Co3O4/EG at x np = 0.0047 ((c)-plot), and ZnO/EG + water at x np = 0.009 ((d)-plot). The solid lines are the modeling values and markers are those reported by Cabaleiro et al. and Mariano et al. [54, 59, 62, 63]. Figure 2 shows systematic predictions for isothermal compressibility over the whole pressure range. Figure 2 also demonstrates fairly linear trend for the isothermal compressibility against the pressure (as indicated in a–d plots) for some selected EG- and EG + water-based nanofluids. The trend of isothermal compressibility in terms of the pressure is well established by the present model over different nanoparticle mole fractions up to 0.040.

Figure 2: 
Plots of isothermal compressibility versus pressure for some selected nanofluids including TiO2-R/EG at x
np = 0.039 ((a)-plot), SnO2/EG at x
np = 0.01472 ((b)-plot), Co3O4/EG at x
np = 0.0047 ((c)-plot) and ZnO/EG + water at x
np = 0.009 ((d)-plot). The solid lines are the modeling values and markers are isotherms of Cabaleiro et al., and Mariano et al. [54, 59, 62, 63], (a)-plot represents the results at 283 K (◊), 313 K (∆) and 343 K (⚬). (b)- and (c)-plots represent the results at 283 K (◊), 303 K (∆) and 323 K (⚬) and (d)-plot represents the results at 303 K (◊), 323 K (∆) and 343 K (⚬).
Figure 2:

Plots of isothermal compressibility versus pressure for some selected nanofluids including TiO2-R/EG at x np = 0.039 ((a)-plot), SnO2/EG at x np = 0.01472 ((b)-plot), Co3O4/EG at x np = 0.0047 ((c)-plot) and ZnO/EG + water at x np = 0.009 ((d)-plot). The solid lines are the modeling values and markers are isotherms of Cabaleiro et al., and Mariano et al. [54, 59, 62, 63], (a)-plot represents the results at 283 K (◊), 313 K (∆) and 343 K (⚬). (b)- and (c)-plots represent the results at 283 K (◊), 303 K (∆) and 323 K (⚬) and (d)-plot represents the results at 303 K (◊), 323 K (∆) and 343 K (⚬).

4.2 Common intersection isotherms

The present PHDC EoS has also been employed to predict some known regularities, the pressure dependence of the isotherms of isobaric expansion, α P and isothermal compressibility coefficients, κ T both of which were passing through a common intersection point. This intersection can be useful to evaluate the reliability of a given EoS for producing the equilibrium properties of matter [66].

The PHDC EoS under consideration has been tested for predicting those behaviors for a typical base fluid, EG, and nanofluid, SnO2/EG at x np = 0.021 in 0.1–350 MPa range. The results were compared with those were calculated from Tait-type equations [38]. The results for isotherms of κ T and α P were respectively shown by Figures 3(a) and 4(b) for EG and SnO2/EG nanofluid at x np = 0.021, in which the solid lines are the isotherms of our PHDC EoS and markers represent the literature isotherms [38]. As Figures 3(b) and 4(b) depict, the EoS was not able to predict the isotherms of α P versus pressure as the deviations from the results of Tait-type equation [38] were significantly large (within 46%–68%), even though neither PHDC EoS nor Tait-type equations of literature did not predict the common intersection points in the isotherms of α P. This fact can be regarded as the drawback of both PHDC EoS and Tait-type equation. But in case the pressure dependence of κ T isotherms, both PHDC EoS and Tait-type equation were able to predict the common intersection points in the isotherms of κ T , even though the trend of modeling values (solid lines) were not in accord with the literature data which reveals another drawback of the present PHDC EoS.

Figure 3: 
Isotherms of κ

T
 ((a)-plot), and α

P
 ((b)-plot) versus pressure for EG. Solid lines are calculated based on the PHDC EoS and markers are literature isotherms [38].
Figure 3:

Isotherms of κ T ((a)-plot), and α P ((b)-plot) versus pressure for EG. Solid lines are calculated based on the PHDC EoS and markers are literature isotherms [38].

Figure 4: 
Isotherms of κ

T
 ((a)-plot), and α

P
 ((b)-plot) versus pressure for SnO2/EG at x
np = 0.021. Solid lines are calculated based on the PHDC EoS and markers are literature isotherms [38].
Figure 4:

Isotherms of κ T ((a)-plot), and α P ((b)-plot) versus pressure for SnO2/EG at x np = 0.021. Solid lines are calculated based on the PHDC EoS and markers are literature isotherms [38].

4.3 Excess molar volumes

Calculating the excess properties, in particular the excess volumes, V E as a function of nanoparticle fractions, one can easily determine how the nanoparticle impacts the volumetric behavior of nanofluids. To do so, some typical calculations were performed on the excess volumes of three selected nanofluids and the results were shown in Figure 5. The results were obtained from the following equation:

(9) V m E = x np M np 1 ρ nf 1 ρ np + ( 1 x np ) M 0 1 ρ nf 1 ρ 0

where, x np represents the mole fraction of nanoparticle; M np and M 0 are the molar masses of nanoparticle and base fluid, respectively; and ρ nf, ρ np, and ρ 0 are the densities of nanofluid, nanoparticle, and base fluid, respectively. Figure 5 clearly demonstrates the sensible deviations from non-ideal liquid-solid mixture behavior for Co3O4/EG (5a), CuO/water (5b) and ZnO/PEG (5c). The solid lines are our modeling values and markers are the values obtained from the isotherms of Mariano et al. [63], Pastoriza-Gallego et al. [64] and Zafarani-Moattar et al. [65]. The excess volumetric trend of the non-ideal mixtures can be mentioned in terms of the contractive and expansive behavior of those mixtures, e.g., for the case of nanofluids, that non-ideal behavior is usually assessed against the nanoparticle concentration. As can be seen from Figure 5(a)–(c), both modeling values and literature values for the excess molar volume were approaching greater negative values as the nanoparticle concentration increased, under this circumstance the volumetric behavior of that system is then contractive in contrast to the expansive type, for which the excess molar volumes would approach greater positive values concerning the nanoparticle concentration loading. That contractive trend in the excess volumes is attributed to the strong attractive interactions between the base fluid molecules and nanoparticle surface. According to the results obtained from the PHDC EoS (solid lines), the addition of nanoparticles to the base fluid led to the excess molar volumes of “−0.12” cm3 mol−1 at most which were somewhat greater than those obtained from the experimental density data of Mariano et al. [63], Pastoriza-Gallego et al. [64], and Zafarani-Moattar et al. [65] (with V E of “−0.08 cm3 mol−1” at most). Then that is why the trend of excess volumes of PHDC EoS were found to be more contractive (i.e., towards greater negative values) than the literature values. The numerical values of the calculated excess volumes along with the relative deviations from the literature data were also reported in Table 4 for 3 selected nanofluids. As can be seen from Table 4, the present EoS modeling was not capable of estimating the excess volumes accurately, as the relative deviations were within −34.30%–51.31%.

Figure 5: 
Excess volumes against nanoparticle mole fractions for three selected nanofluids including Co3O4/EG (5a), CuO/water (5b) and ZnO/PEG (5c). The solid lines are our modeling values and markers are values obtained from the isotherms of Mariano et al. [63], Pastoriza-Gallego et al. [64] and Zafarani-Moattar et al. [65]. (5a): data were taken from 0.1 MPa at 283 (□) and 45 MPa at 323 K (▲). (5b): Data were taken from 0.1 MPa at 293 (▲) and 25 MPa at 313 K (□). (5c): Data were taken from 0.1 MPa at 293 (▲) and 0.1 MPa at 308 K (□).
Figure 5:

Excess volumes against nanoparticle mole fractions for three selected nanofluids including Co3O4/EG (5a), CuO/water (5b) and ZnO/PEG (5c). The solid lines are our modeling values and markers are values obtained from the isotherms of Mariano et al. [63], Pastoriza-Gallego et al. [64] and Zafarani-Moattar et al. [65]. (5a): data were taken from 0.1 MPa at 283 (□) and 45 MPa at 323 K (▲). (5b): Data were taken from 0.1 MPa at 293 (▲) and 25 MPa at 313 K (□). (5c): Data were taken from 0.1 MPa at 293 (▲) and 0.1 MPa at 308 K (□).

Table 4:

The numerical values of calculated excess volumes along with the relative deviations (RDs in %) from the literature data [6365] for 3 selected nanofluids at several isotherms.

T/K x np V E,lit(cm3/mol) V E, Calc. (cm3/mol) RD%
Co 3 O 4 /EG
283 K 0.0026 −0.029 −0.03359 15.84
0.0047 −0.033 −0.03997 21.11
0.0066 −0.037 −0.04728 27.80
0.0093 −0.042 −0.05803 38.16
0.0140 −0.054 −0.08005 48.25
323 K 0.0026 −0.0204 −0.01652 −19.01
0.0047 −0.027 −0.02043 −24.33
0.0066 −0.033 −0.0252 −23.65
0.0093 −0.039 −0.02989 −23.36
0.0140 −0.048 −0.03657 −23.81
ZnO/PEG
293 0.0010 −0.022 −0.01758 −20.10
0.0042 −0.028 −0.03191 13.98
0.0082 −0.036 −0.04754 32.05
0.0122 −0.043 −0.06506 51.31
308 0.0010 −0.026 −0.03 15.38
0.0042 −0.04 −0.047 17.50
0.0082 −0.058 −0.071 22.41
0.0122 −0.071 −0.099 39.44
CuO/water
293 0.0020 −0.02388 −0.01599 −33.06
0.0040 −0.02769 −0.01842 −33.49
0.0058 −0.02983 −0.02075 −30.42
0.0080 −0.03309 −0.02285 −30.95
0.0120 −0.04272 −0.02807 −34.30
313 0.0020 −0.06426 −0.05066 −21.17
0.0040 −0.07609 −0.06106 −19.75
0.0058 −0.08078 −0.07222 −10.59
0.0080 −0.0817 −0.09085 11.21
0.0120 −0.0858 −0.12561 46.40

4.4 Artificial neural network modeling results

To be able to predict the nanofluid density over a pressure range of 0.1–45 MPa, it was decided to design a network with only one hidden layer in addition to the input and output layers. The number of input parameters has been fixed in 3, namely temperature, pressure, and nanoparticle mole fraction. At this point, a clarification of the database is necessary. In fact we preferred to use only the points not at atmospheric pressure. This choice was driven by the desire to demonstrate for the first time in the literature as the pressure is one of the essential parameters for characterizing the density of nanofluids.

The training of the network was done several times by varying the number of neurons in a range between 1 and 30 recording at each step all the variations of 3 statistical parameters: AAD%, RMSE, and R 2

(10) AAD % = 1 100 i = 1 n ρ Calc ρ Exp ρ Exp

(11) RMSE = 1 N i = 1 n ρ i Exp ρ i calc 2

(12) R 2 = 1 i = 1 n ρ i dataset ρ i calc 2 i = 1 n ρ i dataset ρ ̂ 2

At the end was chosen the best neuron in the various training that has been done.

Another particular attention that was kept was to divide the database into 3 parts. In fact, one of the biggest problems of neural networks is the over-fitting. In addition, choosing networks that are not very flexible, it is possible to have excellent results that are however very compressed on the starting data. To overcome this, the database has been broken in:

  1. Training dataset: in which the network carries out its own training calculations

  2. Validation set: in which the network tests the results

  3. Test set: formed by a set of points that the network does not see while it performs the calculations. However, the results of the calculations were also tested on this dataset to verify not only the network’s ability to obtain results on the dataset, but also its ability to predict new density data.

The split across the three databases and the relative results are shown in Table 5.

Table 5:

Number of points for each database and results for the ANN architecture finally selected.

Set of data Percentage of entire dataset Number of points AAD% RMSE R 2
Training 80% 972 0.07 0.001 1
Validation 10% 121 0.09 0.001 1
Test 10% 122 0.07 0.001 1
Complete 100% 1215 0.07 0.001 1

Uncertainty can also be considered when doing network training. In this case, since the uncertainty of the data is very low, it was considered, and to reduce the impact of it during the training process, the data was normalized. This provided a more uniform distribution of the data.

As can be seen from Figure 6, increasing the number of neurons lead to the significant drop in the deviations. Given this trend we decided to choose the neuron number 17 as the minimum to which the network can reach. The minimum AAD reached by the network in the chosen neuron was 0.07 on the entire database. All other statistical parameters were given in Table 5.

Figure 6: 
AAD% values for each the entire dataset, for the training, the validation and test set, as a function of the number of neurons.
Figure 6:

AAD% values for each the entire dataset, for the training, the validation and test set, as a function of the number of neurons.

The results calculated for each fluid under consideration using the present ANN were given in Table 2. As mentioned before, the results were only for fluids with the pressure greater than atmospheric pressure.

In addition, Figure 7 represents the scatterplot between the experimental and calculated densities. As can be seen, all the points are located along the first and third quadrant bisector signifying the excellent approximation of the neural network calculation.

Figure 7: 
Calculated density values versus values in each dataset.
Figure 7:

Calculated density values versus values in each dataset.

5 Conclusions

This work discussed the details of a procedure for extending a simple and analytical EoS for estimating the densities of several nanofluids being newly came into focus as alternative heat transfer fluid suspensions The three molecular parameters to be used in the calculations were determined using the experimental density data of pure base fluids and then the results for the densities of pure base fluids are correlated as a fit procedure were used. Instead, the results for their isothermal compressibilities (being closely related to the isothermal derivative of the density) are predictive as k T data of Tait-type equation were not included in the fit procedure.

In case the nanofluids under study, the binary parameter of Eq. (5) was fitted against experimental densities at 0.1 MPa for each nanofluid, then the results obtained from the model were estimated for the wider pressure ranging. Besides, the k ij values fitted for use in Eq. (5) were in agreement with the expectations of molecular viewpoints as mentioned earlier.

The degree of accuracy of the calculations was compared with the model in literature work that was based on the EoS method. The summary of results reported in Table 2 indicated that the PHDC EoS was superior to Pak–Cho + TM EoS model.

The PHDC EoS was not capable of estimating the excess volumes of 3 selected EG-, PEG-, and water-based nanofluids accurately as the relative deviations from the literature data were greater than 34%, even though the trend of excess volumes against the nanoparticle mole fraction was in accord with the literature. In the study, the trained neural network underlines how with 3 parameters such as temperature, pressure, and nanoparticle mole fraction, it is possible to obtain excellent results in the prediction of densities of nanofluids.

Nomenclature and units

List of symbols

AAD

average absolute deviation (in %)

k B

Boltzmann’s constant (J K−1)

NP

number of data points

P

pressure (Pa)

T

absolute temperature (K)

MW

molecular weight (g mol−1)

m

segment number

R

universal gas constant (J mol−1K−1)

k ij

binary interaction parameter

Greek letters

g(σ +)

Pair radial distribution function at contact

η

Segment packing fraction

ρ

Segment density (mol m−3)

σ

Segment diameter (nm)

ε

Non-bonded interaction energy between spheres and dimmers (J)

λ

Range parameter of attractive forces

k T

isothermal compressibility coefficient

Superscripts

HDC

Hard-dimer chain reference system

Pert.

Perturbed system

HS

Hard-sphere

HD

Hard-dimer

Calc.

Calculated

Exp.

Experimental data

Tait.

refers to values from Tait equation

Subscripts

CS

Carnahan–Starling

mix.

Mixture


Corresponding author: Sayed Mostafa Hosseini, Department of Chemistry, University of Hormozgan, Bandar Abbas 71961, Iran, E-mail:

Acknowledgments

We thank the research committee of Shiraz University of Technology and University of Hormozgan.

  1. Author contribution: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.

  2. Research funding: None declared.

  3. Conflict of interest statement: The authors declare no conflicts of interest regarding this article.

Appendix A: Perturbed hard-dimer-chain EoS for pure fluid

The general frame of the perturbed hard-dimer-chain EoS for pure fluid is as:

(A1) Z PHDC = P ρ k B T = 1 + m ( Z HS 1 ) m 2 η ln g HS ( σ + ) η ( m 2 ) 2 η ln g HD ( σ + ) η 12 m η T * I ( η , m ; λ = 1.8 )

where, Z PHDC represents the reference hard-dimer-chain system perturbed by the long-range attraction with a Yukawa tail, λ for the representation of the range of attractive forces. The tail of the Lennard–Jones potential can be well-represented with λ of 1.8.

In Eq. (A1), P is the pressure, and k B T is the thermal energy per segment. The first four terms of the right hand of Eq. (A1) are coming from the thermodynamic perturbation theory of dimer fluid (TPT-D) of Wertheim [48] and Chapman et al. [67]. Also, ρ is the number density of segments, m stands for the number of monomer sites in a chain-like molecule and T *is the reduced temperature (T * = k B T/ε). g HS (σ +) and g HD (σ +) are the pair radial distribution function of hard-spheres and hard-dimers at contact, respectively, both of which were taken from literature [68, 69]. η is the segment packing fraction of hard-chain defined by:

(A2) η = m π ρ σ 3 6

In Eq. (A1), I(η,m;λ) has been previously evaluated in the literature [70, 71] for case of λ = 1.8 as a power series in segment packing fraction, η:

(A3) I ( η , m ; λ = 1.8 ) = i = 0 2 a i ( i + 1 ) η i

The values of coefficients a i have been taken from Ref. [72].

References

[1] E. E. S. Michaelides, “Thermodynamic properties,” in Nanofluidics, New York City, Springer, 2014, pp. 91–115.10.1007/978-3-319-05621-0_3Search in Google Scholar

[2] K. D. Sattler, Handbook of Nanophysics: Nanoparticles and Quantum Dots, Boca Raton, CRC Press, 2016.10.1201/9781420075458Search in Google Scholar

[3] L. Zhang, Y. Jiang, Y. Ding, M. Povey, and D. York, “Investigation into the antibacterial behaviour of suspensions of ZnO nanoparticles (ZnO nanofluids),” J. Nanoparticle Res., vol. 9, pp. 479–489, 2007. https://doi.org/10.1007/s11051-006-9150-1.Search in Google Scholar

[4] W. Yu and H. Xie, “A review on nanofluids: preparation, stability mechanisms, and applications,” J. Nanomater., vol. 2012, p. 1, 2012. https://doi.org/10.1155/2012/435873.Search in Google Scholar

[5] R. Saidur, S. Kazi, M. Hossain, M. Rahman, and H. Mohammed, “A review on the performance of nanoparticles suspended with refrigerants and lubricating oils in refrigeration systems,” Renew. Sustain. Energy Rev., vol. 15, pp. 310–323, 2011. https://doi.org/10.1016/j.rser.2010.08.018.Search in Google Scholar

[6] I. Mahbubul, R. Saidur, and M. Amalina, “Latest developments on the viscosity of nanofluids,” Int. J. Heat Mass Transfer, vol. 55, pp. 874–885, 2012. https://doi.org/10.1016/j.ijheatmasstransfer.2011.10.021.Search in Google Scholar

[7] J. J. Segovia, O. Fandiño, E. R. López, L. Lugo, M. C. Martín, and J. Fernández, “Automated densimetric system: measurements and uncertainties for compressed fluids,” J. Chem. Therm., vol. 41, pp. 632–638, 2009. https://doi.org/10.1016/j.jct.2008.12.020.Search in Google Scholar

[8] S. Sarkar and R. P. Selvam, “Molecular dynamics simulation of effective thermal conductivity and study of enhanced thermal transport mechanism in nanofluids,” J. Appl. Phys., vol. 102, p. 074302, 2007. https://doi.org/10.1063/1.2785009.Search in Google Scholar

[9] N. Sankar, N. Mathew, and C. Sobhan, “Molecular dynamics modeling of thermal conductivity enhancement in metal nanoparticle suspensions,” Int. Commun. Heat Mass Tran., vol. 35, pp. 867–872, 2008. https://doi.org/10.1016/j.icheatmasstransfer.2008.03.006.Search in Google Scholar

[10] F. Yousefi, H. Karimi, and M. M. Papari, “Modeling viscosity of nanofluids using diffusional neural networks,” J. Mol. Liq., vol. 175, pp. 85–90, 2012. https://doi.org/10.1016/j.molliq.2012.08.015.Search in Google Scholar

[11] L. S. Sundar, K. Sharma, M. Naik, and M. K. Singh, “Empirical and theoretical correlations on viscosity of nanofluids: a review,” Renew. Sustain. Energy Rev., vol. 25, pp. 670–686, 2013. https://doi.org/10.1016/j.rser.2013.04.003.Search in Google Scholar

[12] S. Atashrouz, G. Pazuki, and Y. Alimoradi, “Estimation of the viscosity of nine nanofluids using a hybrid GMDH-type neural network system,” Fluid Phase Equilib., vol. 372, pp. 43–48, 2014. https://doi.org/10.1016/j.fluid.2014.03.031.Search in Google Scholar

[13] M. Ariana, B. Vaferi, and G. Karimi, “Prediction of thermal conductivity of alumina water-based nanofluids by artificial neural networks,” Powder Technol., vol. 278, pp. 1–10, 2015. https://doi.org/10.1016/j.powtec.2015.03.005.Search in Google Scholar

[14] A. Vatani, P. L. Woodfield, and D. V. Dao, “A survey of practical equations for prediction of effective thermal conductivity of spherical-particle nanofluids,” J. Mol. Liq., vol. 211, pp. 712–733, 2015. https://doi.org/10.1016/j.molliq.2015.07.043.Search in Google Scholar

[15] W. Cui, Z. Shen, J. Yang, and S. Wu, “Molecular dynamics simulation on flow behaviors of nanofluids confined in nanochannel,” Case Stud. Therm. Eng., vol. 5, pp. 114–121, 2015. https://doi.org/10.1016/j.csite.2015.03.007.Search in Google Scholar

[16] N. Zhao, X. Wen, J. Yang, S. Li, and Z. Wang, “Modeling and prediction of viscosity of water-based nanofluids by radial basis function neural networks,” Powder Technol., vol. 281, pp. 173–183, 2015. https://doi.org/10.1016/j.powtec.2015.04.058.Search in Google Scholar

[17] M. K. Meybodi, S. Naseri, A. Shokrollahi, and A. Daryasafar, “Prediction of viscosity of water-based Al2O3, TiO2, SiO2, and CuO nanofluids using a reliable approach,” Chemom. Intell. Lab. Syst., vol. 149, pp. 60–69, 2015. https://doi.org/10.1016/j.chemolab.2015.10.001.Search in Google Scholar

[18] M. H. Esfe, M. Afrand, W.-M. Yan, and M. Akbari, “Applicability of artificial neural network and nonlinear regression to predict thermal conductivity modeling of Al2O3–water nanofluids using experimental data,” Int. Commun. Heat Mass Tran., vol. 66, pp. 246–249, 2015. https://doi.org/10.1016/j.icheatmasstransfer.2015.06.002.Search in Google Scholar

[19] E. Ahmadloo and S. Azizi, “Prediction of thermal conductivity of various nanofluids using artificial neural network,” Int. Commun. Heat Mass Tran., vol. 74, pp. 69–75, 2016. https://doi.org/10.1016/j.icheatmasstransfer.2016.03.008.Search in Google Scholar

[20] G. A. Longo, C. Zilio, L. Ortombina, and M. Zigliotto, “Application of artificial neural network (ANN) for modeling oxide-based nanofluids dynamic viscosity,” Int. Commun. Heat Mass Tran., vol. 83, pp. 8–14, 2017. https://doi.org/10.1016/j.icheatmasstransfer.2017.03.003.Search in Google Scholar

[21] F. Jabbari, A. Rajabpour, and S. Saedodin, “Thermal conductivity and viscosity of nanofluids: a review of recent molecular dynamics studies,” Chem. Eng. Sci., vol. 174, pp. 67–81, 2017. https://doi.org/10.1016/j.ces.2017.08.034.Search in Google Scholar

[22] S. M. Hosseini, M. M. Alavianmehr, and J. Moghadasi, “On the rough hard-sphere-based model for transport properties of nanofluids,” Fluid Phase Equilib., vol. 458, pp. 186–193, 2018. https://doi.org/10.1016/j.fluid.2017.11.019.Search in Google Scholar

[23] E. E. Michaelides, “Transport properties of nanofluids. A critical review,” J. Non-Equilibrium Thermodyn., vol. 38, pp. 1–79, 2013. https://doi.org/10.1515/jnetdy-2012-0023.Search in Google Scholar

[24] J. W. Gibbs, “On the equilibrium of heterogeneous substances,” Am. J. Sci., vol. 3, pp. 441–458, 1878. https://doi.org/10.2475/ajs.s3-16.96.441.Search in Google Scholar

[25] U. D. Sharma and M. Kumar, “Effect of pressure on nanomaterials,” Phys. B Condens. Matter, vol. 405, pp. 2820–2826, 2010. https://doi.org/10.1016/j.physb.2010.04.005.Search in Google Scholar

[26] B. C. Pak and Y. I. Cho, “Hydrodynamic and heat transfer study of dispersed fluids with submicron metallic oxide particles,” Exp. Heat Tran., vol. 11, pp. 151–170, 1998. https://doi.org/10.1080/08916159808946559.Search in Google Scholar

[27] G. Tammann, “The dependence of the volume of solutions on pressure,” Z. Phys. Chem. Stoechiom. Verwandtschafts, vol. 17, pp. 620–636, 1895. https://doi.org/10.1515/zpch-1895-1738.Search in Google Scholar

[28] J. H. Dymond and R. Malhotra, “The Tait equation: 100 years on,” Int. J. Thermophys., vol. 9, pp. 941–951, 1988. https://doi.org/10.1007/bf01133262.Search in Google Scholar

[29] R. S. Vajjha, D. K. Das, and B. M. Mahagaonkar, “Density measurement of different nanofluids and their comparison with theory,” Pet. Sci. Technol., vol. 27, pp. 612–624, 2009. https://doi.org/10.1080/10916460701857714.Search in Google Scholar

[30] S. P. Tan and M. Piri, “Equation-of-state modeling of confined-fluid phase equilibria in nanopores,” Fluid Phase Equilib., vol. 393, pp. 48–63, 2015. https://doi.org/10.1016/j.fluid.2015.02.028.Search in Google Scholar

[31] J. S. Rowlinson and B. Widom, Molecular Theory of Capillarity, Mineola, Dover Publications, 2013.Search in Google Scholar

[32] A. W. Islam, T. W. Patzek, and A. Y. Sun, “Thermodynamics phase changes of nanopore fluids,” J. Nat. Gas Sci. Eng., vol. 25, pp. 134–139, 2015. https://doi.org/10.1016/j.jngse.2015.04.035.Search in Google Scholar

[33] A. W. Islam and A. Y. Sun, “A theory-based simple extension of Peng–Robinson equation of state for nanopore confined fluids,” J. Pet. Explor. Prod. Technol., vol. 7, pp. 1197–1203, 2017. https://doi.org/10.1007/s13202-016-0306-y.Search in Google Scholar

[34] H. Karimi and F. Yousefi, “Application of artificial neural network–genetic algorithm (ANN–GA) to correlation of density in nanofluids,” Fluid Phase Equilib., vol. 336, pp. 79–83, 2012. https://doi.org/10.1016/j.fluid.2012.08.019.Search in Google Scholar

[35] F. Yousefi and Z. Amoozandeh, “A new model to predict the densities of nanofluids using statistical mechanics and artificial intelligent plus principal component analysis,” Chin. J. Chem. Eng., vol. 25, pp. 1273–1281, 2017. https://doi.org/10.1016/j.cjche.2016.10.016.Search in Google Scholar

[36] I. O. Alade, M. O. Oyedeji, M. A. A. Rahman, and T. A. Saleh, “Predicting the density of carbon-based nanomaterials in diesel oil through computational intelligence methods,” J. Therm. Anal. Calorim., vol. 147, pp. 8699–8707, 2022. https://doi.org/10.1007/s10973-021-11124-3.Search in Google Scholar

[37] D. H. Kumar, H. E. Patel, V. R. R. Kumar, T. Sundararajan, T. Pradeep, and S. K. Das, “Model for heat conduction in nanofluids,” Phys. Rev. Lett., vol. 93, p. 144301, 2004. https://doi.org/10.1103/physrevlett.93.144301.Search in Google Scholar PubMed

[38] S. M. Hosseini and M. M. Alavianmehr, “New version of Tammann-Tait equation: application to nanofluids,” J. Mol. Liq., vol. 220, pp. 404–408, 2016. https://doi.org/10.1016/j.molliq.2016.04.088.Search in Google Scholar

[39] S. M. Hosseini, M. M. Alavianmehr, and J. Moghadasi, “A perturbed hard-dimer chain equation of state for polymer melts,” J. Non-Cryst. Solids, vol. 362, pp. 195–200, 2013. https://doi.org/10.1016/j.jnoncrysol.2012.11.033.Search in Google Scholar

[40] S. M. Hosseini, A. Mulero, and M. M. Alavianmehr, “Predictive methods and semi-classical equations of state for pure ionic liquids: a review,” J. Chem. Therm., vol. 130, pp. 47–94, 2019. https://doi.org/10.1016/j.jct.2018.09.022.Search in Google Scholar

[41] S. M. Hosseini, M. M. Alavianmehr, and J. Moghadasi, “Application of perturbed hard-sphere equation of state to the study of volumetric properties of nano-fluids,” Fluid Phase Equilib., vol. 423, pp. 181–189, 2016. https://doi.org/10.1016/j.fluid.2016.04.026.Search in Google Scholar

[42] F. M. Tao and E. A. Mason, “Statistical‐mechanical equation of state for nonpolar fluids: prediction of phase boundaries,” J. Chem. Phys., vol. 100, pp. 9075–9087, 1994. https://doi.org/10.1063/1.466713.Search in Google Scholar

[43] E. Montazer, E. Salami, H. Yarmand, et al.., “Development of a new density correlation for carbon-based nanofluids using response surface methodology,” J. Therm. Anal. Calorim., vol. 132, pp. 1399–1407, 2018. https://doi.org/10.1007/s10973-018-6978-4.Search in Google Scholar

[44] S. J. Kalil, F. Maugeri, and M. I. Rodrigues, “Response surface analysis and simulation as a tool for bioprocess design and optimization,” Process Biochem., vol. 35, pp. 539–550, 2000. https://doi.org/10.1016/s0032-9592(99)00101-6.Search in Google Scholar

[45] A. Sadollah, A. Ghadimi, I. H. Metselaar, and A. Bahreininejad, “Prediction and optimization of stability parameters for titanium dioxide nanofluid using response surface methodology and artificial neural networks,” Sci. Eng. Compos. Mater., vol. 20, pp. 319–330, 2013. https://doi.org/10.1515/secm-2013-0017.Search in Google Scholar

[46] S. M. Hosseini, M. M. Alavianmehr, and J. Moghadasi, “Density and isothermal compressibility of ionic liquids from perturbed hard-dimer-chain equation of state,” Fluid Phase Equilib., vol. 356, pp. 185–192, 2013. https://doi.org/10.1016/j.fluid.2013.07.030.Search in Google Scholar

[47] J. Chang and S. I. Sandler, “An equation of state for the hard-sphere chain fluid: theory and Monte Carlo simulation,” Chem. Eng. Sci., vol. 49, pp. 2777–2791, 1994. https://doi.org/10.1016/0009-2509(94)e0097-a.Search in Google Scholar

[48] M. S. Wertheim, “Thermodynamic perturbation theory of polymerization,” J. Chem. Phys., vol. 87, pp. 7323–7331, 1987. https://doi.org/10.1063/1.453326.Search in Google Scholar

[49] V. Srivastava and K. Khanna, “Equation of state of long chain molecules and rings,” Mol. Phys., vol. 100, pp. 311–316, 2002. https://doi.org/10.1080/00268970110091664.Search in Google Scholar

[50] G. M. Kontogeorgis and G. K. Folas, Thermodynamic Models for Industrial Applications: From Classical and Advanced Mixing Rules to Association Theories, Hoboken, Wiley, 2009.10.1002/9780470747537Search in Google Scholar

[51] M. Kubat, “Neural networks: A comprehensive foundation by simon haykin, vol. 13, Macmillan, 1994,” Knowl. Eng. Rev., vol. 13, pp. 409–412, 1999.10.1017/S0269888998214044Search in Google Scholar

[52] K. Hornik, M. Stinchcombe, and H. White, “Multilayer feedforward networks are universal approximators,” Neural Networks, vol. 2, pp. 359–366, 1989. https://doi.org/10.1016/0893-6080(89)90020-8.Search in Google Scholar

[53] S. Sharma, S. Sharma, and A. Athaiya, “Activation functions in neural networks,” Towards Data Science, vol. 6, pp. 310–316, 2017. https://doi.org/10.33564/ijeast.2020.v04i12.054.Search in Google Scholar

[54] A. Mariano, M. J. Pastoriza-Gallego, L. Lugo, A. Camacho, S. Canzonieri, and M. M. Piñeiro, “Thermal conductivity, rheological behaviour and density of non-Newtonian ethylene glycol-based SnO2 nanofluids,” Fluid Phase Equilib., vol. 337, pp. 119–124, 2013. https://doi.org/10.1016/j.fluid.2012.09.029.Search in Google Scholar

[55] E. Lemmon, Thermophysical Properties of Fluid Systems, NIST Chemistry WebBook, NIST Standard Reference Database Number 69, 2005. Available at: http://webbook.nist.gov.Search in Google Scholar

[56] A. C. Colín, R. G. Rubio, and A. Compostizo, “Equation of state of a blend of water-soluble polymers: poly(propylene glycol)+poly(ethylene glycol),” Polymer, vol. 41, pp. 7407–7414, 2000. https://doi.org/10.1016/s0032-3861(00)00087-2.Search in Google Scholar

[57] P. O. Ifeacho, Semi-conducting Metal Oxide Nanoparticles from a Low-Pressure Premixed H_1tn2/O_1tn2/Ar Flame: Synthesis and Characterization, Göttingen, Cuvillier, 2008.Search in Google Scholar

[58] H. S. Nalwa, Handbook of Nanostructured Materials and Nanotechnology, Five-Volume Set, Amsterdam, Elsevier Science, 1999.Search in Google Scholar

[59] D. Cabaleiro, M. J. Pastoriza-Gallego, M. M. Piñeiro, and L. Lugo, “Characterization and measurements of thermal conductivity, density and rheological properties of zinc oxide nanoparticles dispersed in (ethane-1, 2-diol+water) mixture,” J. Chem. Therm., vol. 58, pp. 405–415, 2013. https://doi.org/10.1016/j.jct.2012.10.014.Search in Google Scholar

[60] G. A. Rodríguez, A. R. Holguín, F. Martínez, M. Khoubnasabjafari, and A. Jouyban, “Volumetric properties of (PEG 400+ water) and (PEG 400+ ethanol) mixtures at several temperatures and correlation with the Jouyban-Acree model,” Rev. Colomb. Ciencias Quím. Farm., vol. 41, pp. 187–202, 2012.Search in Google Scholar

[61] E. A. Muller and P. Rasmussen, “Densities and excess volumes in aqueous poly(ethylene glycol) solutions,” J. Chem. Eng. Data, vol. 36, pp. 214–217, 1991. https://doi.org/10.1021/je00002a019.Search in Google Scholar

[62] D. Cabaleiro, M. J. Pastoriza-Gallego, C. Gracia-Fernández, M. M. Piñeiro, and L. Lugo, “Rheological and volumetric properties of TiO2-ethylene glycol nanofluids,” Nanoscale Res. Lett., vol. 8, p. 286, 2013. https://doi.org/10.1186/1556-276x-8-286.Search in Google Scholar

[63] A. Mariano, M. J. Pastoriza-Gallego, L. Lugo, L. Mussari, and M. M. Piñeiro, “Co3O4 ethylene glycol-based nanofluids: thermal conductivity, viscosity and high pressure density,” Int. J. Heat Mass Transfer, vol. 85, pp. 54–60, 2015. https://doi.org/10.1016/j.ijheatmasstransfer.2015.01.061.Search in Google Scholar

[64] M. J. Pastoriza-Gallego, C. Casanova, J. L. Legido, and M. M. Piñeiro, “CuO in water nanofluid: influence of particle size and polydispersity on volumetric behaviour and viscosity,” Fluid Phase Equilib., vol. 300, pp. 188–196, 2011. https://doi.org/10.1016/j.fluid.2010.10.015.Search in Google Scholar

[65] M. T. Zafarani-Moattar and R. Majdan-Cegincara, “Effect of temperature on volumetric and transport properties of nanofluids containing ZnO nanoparticles poly(ethylene glycol) and water,” J. Chem. Therm., vol. 54, pp. 55–67, 2012. https://doi.org/10.1016/j.jct.2012.03.010.Search in Google Scholar

[66] R. Privat, F. Gaillochet, and J.-N. Jaubert, “Testing the ability of various equations of state to reproduce high-pressure isotherm crossings in the (α, P) plane,” Fluid Phase Equilib., vol. 327, pp. 45–57, 2012. https://doi.org/10.1016/j.fluid.2012.05.008.Search in Google Scholar

[67] W. G. Chapman, G. Jackson, and K. E. Gubbins, “Phase equilibria of associating fluids,” Mol. Phys., vol. 65, pp. 1057–1079, 1988. https://doi.org/10.1080/00268978800101601.Search in Google Scholar

[68] N. F. Carnahan and K. E. Starling, “Equation of state for nonattracting rigid spheres,” J. Chem. Phys., vol. 51, pp. 635–636, 1969. https://doi.org/10.1063/1.1672048.Search in Google Scholar

[69] J. Chang and S. I. Sandler, “The correlation functions of hard‐sphere chain fluids: comparison of the Wertheim integral equation theory with the Monte Carlo simulation,” J. Chem. Phys., vol. 102, pp. 437–449, 1995. https://doi.org/10.1063/1.469421.Search in Google Scholar

[70] J. Gross and G. Sadowski, “Application of perturbation theory to a hard-chain reference fluid: an equation of state for square-well chains,” Fluid Phase Equilib., vol. 168, pp. 183–199, 2000. https://doi.org/10.1016/s0378-3812(00)00302-2.Search in Google Scholar

[71] I. Nezbeda, “Statistical thermodynamics of interaction-site molecules,” Mol. Phys., vol. 33, pp. 1287–1299, 1977. https://doi.org/10.1080/00268977700101081.Search in Google Scholar

[72] S. M. Hosseini, M. M. Papari, and J. Moghadasi, “Density and isothermal compressibility of ionic liquids from perturbed hard-sphere chain equation of state,” J. Mol. Liq., vol. 174, pp. 52–57, 2012. https://doi.org/10.1016/j.molliq.2012.07.007.Search in Google Scholar

Received: 2022-06-30
Revised: 2022-10-22
Accepted: 2022-10-25
Published Online: 2022-12-01
Published in Print: 2023-01-27

© 2022 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 25.4.2024 from https://www.degruyter.com/document/doi/10.1515/jnet-2022-0046/html
Scroll to top button