Next Article in Journal
Green Removal of DUV-Polarity-Modified PMMA for Wet Transfer of CVD Graphene
Next Article in Special Issue
Time-Dependent Flow of Water-Based CoFe2O4-Mn-ZnFe2O4 Nanoparticles over a Shrinking Sheet with Mass Transfer Effect in Porous Media
Previous Article in Journal
Self-Assembly 2D Ti3C2/g-C3N4 MXene Heterojunction for Highly Efficient Photocatalytic Degradation of Tetracycline in Visible Wavelength Range
Previous Article in Special Issue
Nanostructured AlGaAsSb Materials for Thermophotovoltaic Solar Cells Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Response Surface Methodology (RSM) on the Hybrid Nanofluid Flow Subject to a Vertical and Permeable Wedge

by
Najiyah Safwa Khashi’ie
1,*,
Iskandar Waini
1,
Mohd Fariduddin Mukhtar
1,
Nurul Amira Zainal
1,
Khairum Bin Hamzah
1,
Norihan Md Arifin
2,3 and
Ioan Pop
4
1
Fakulti Teknologi Kejuruteraan Mekanikal dan Pembuatan, Universiti Teknikal Malaysia Melaka, Hang Tuah Jaya, Durian Tunggal 76100, Melaka, Malaysia
2
Institute for Mathematical Research, Universiti Putra Malaysia (UPM), Serdang 43400, Selangor, Malaysia
3
Department of Mathematics, Faculty of Science, Universiti Putra Malaysia (UPM), Serdang 43400, Selangor, Malaysia
4
Department of Mathematics, Babeş-Bolyai University, 400084 Cluj-Napoca, Romania
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(22), 4016; https://doi.org/10.3390/nano12224016
Submission received: 11 September 2022 / Revised: 11 October 2022 / Accepted: 11 November 2022 / Published: 15 November 2022
(This article belongs to the Special Issue Thermal Properties of Nanomaterials: Fundamentals and Applications)

Abstract

:
The mixed convection flow with thermal characteristics of a water-based Cu-Al2O3 hybrid nanofluid towards a vertical and permeable wedge was numerically and statistically analyzed in this study. The governing model was constructed using physical and theoretical assumptions, which were then reduced to a set of ordinary differential equations (ODEs) using similarity transformation. The steady flow solutions were computed using the Matlab software bvp4c. All possible solutions were presented in the graphs of skin friction coefficient and thermal rate. The numerical results show that the flow and thermal progresses are developed by enhancing the controlling parameters (wedge parameter, volumetric concentration of nanoparticles, and suction parameter). Moreover, the response surface methodology (RSM) with analysis of variance (ANOVA) was employed for the statistical evaluation and conducted using the fit general linear model in the Minitab software. From the standpoint of statistical analysis, the wedge parameter and volumetric nanoparticle concentration have a considerable impact on all responses; however, the suction parameter effect is only substantial for a single response.

1. Introduction

The utilization of nanofluid as a conductive coolant is one of the most well-known methods that ensure great thermal performance at a low cost. Nanofluid is formed by the homogeneous combination of extremely small nanoscale particles and a base fluid. Shah et al. [1] analyzed the mass transport and hydro-thermal characteristics with the convective flow of a non-Newtonian micropolar fluid with copper oxide nanomaterial and a mixture of pure water and ethylene glycol subjected to an electromagnetic surface. They found that the micropolarity and electrical conducting of the nanofluidic medium play an important role in the nanofluid motion. Moreover, the flow and thermal characteristics of alumina–water nanofluid with different nanoparticle shapes (sphere, platelet, cylinder, and brick) due to a rotating disk were scrutinized by Sabu et al. [2]. They concluded that the highest drag was contributed by the platelet-shaped alumina, followed by the cylinder-, brick-, and sphere-shaped alumina. Another numerical study regarding the nanofluid flow subjected to the Riga surface in a Darcy–Forchheimer porous medium was conducted by Rasool et al. [3]. They revealed that a significant enhancement in the thermal rate could be achieved by manipulating the electromagnetic planar support and convective heating process. Hybrid nanofluids are the latest generation of heat transfer fluids that offer high heat transfer compared to conventional fluids as the hybrid nanoparticles increase the thermal conductivity of the fluids. Jana et al. [4] experimentally investigated the enhancement of thermal conductivity in hybrid nanoparticles. Suresh et al. [5] reported the advantage of hybrid nanoparticles in the enhancement of the fluid thermal conductivity, which was then continued by Takabi and Salehi [6] and Devi and Devi [7]. Several studies on hybrid nanofluid incorporated with the non-Newtonian fluid model were reported by researchers [8,9,10,11]. Nabwey and Mahdy [8] considered the micropolar hybrid nanofluid flow through a porous medium with dusty particles. Similarly, Roy et al. [9] investigated the flow of micropolar hybrid nanofluid over a shrinking sheet. The thermal performance of peristaltic flow utilizing the hybrid nanoparticles in Eyring–Powell fluid was reported by Riaz et al. [10]. Additionally, Khashi’ie et al. [11] considered the stagnation point flow of second-grade fluid containing hybrid nanoparticles towards a Riga plate. The hybrid nanofluid flow over the Riga channel with slip conditions was studied by Abbas et al. [12], while Waqas et al. [13] and Bilal et al. [14] considered the rotating disk and inclined cylinder geometries, respectively. Uysal and Korkmaz [15] and Kumar and Sarkar [16] considered the hybrid nanofluid flow in a mini channel. Further discussions regarding the thermal and flow characteristics of hybrid nanofluids were reported by Salehi et al. [17], Zainal et al. [18,19,20], Khashi’ie et al. [21,22,23,24], Waini et al. [25,26,27,28,29] and Shah et al. [30].
In recent decades, the flow through a wedge-shaped surface has gained much attention due to its extensive uses in the engineering and chemical industry, such as in the fields of geothermal energy and aerodynamics. The pioneering study in wedge flow was initiated by Falkner and Skan [31] and is known as Falkner–Skan flow. Later, the pressure gradient was considered in this model by Hartree [32] and called a Hartree pressure gradient parameter. Since then, the wedge flow with the various effect of physical parameters has been published, see Refs. [33,34,35,36,37,38,39]. Moreover, the moving wedge flow was examined by Ishak et al. [40,41,42], Khan and Pop [43], and Hedayati et al. [44]; additionally, the shrinking wedge surface was reported by Alam et al. [45], Khan et al. [46], Awaludin et al. [47], and Waini et al. [48]. In recent years, the effect of the nanoparticles on wedge flow has been reported by researchers, for example, Rashad [49], Hassan et al. [50], Ahmed et al. [51], Zaib and Haq [52], Rawat et al. [53], and Mahanthesh et al. [54].
There are numerous advantages to employing the design of experiment (DOE) in research with various factors and outcomes. Response surface methodology (RSM) is one of the design types. RSM is a statistical method widely used for modeling and analyzing processes in which the response of interest is affected by multiple variables, where the goal of the method is to maximize the response [55,56]. The primary advantage of the RSM is that it saves time and money by reducing the number of trials required. The RSM can be summarized as a method for determining how independent variables interact. Based on the dataset, an analysis of variance (ANOVA) was performed to determine whether or not the variables in the experiment were statistically significant. The application of RSM, including the ANOVA, was discussed by Mehmood et al. [57] for the rotating disk flow problem. There are also many fluid flow problems that have been reported with the RSM and statistical data analysis (see Mahanthesh and Thriveni [58], Shafiq et al. [59], Vahedi et al. [60], and Abdelmalek et al. [61]).
Hence, our main objective was to generate all available numerical solutions from the present model and conduct the statistical data analysis using response surface methodology. For the numerical solutions, the reduced system of linear equations was solved using the bvp4c solver. The selected data for the ANOVA were selected based on the central composite design in RSM. We believe that no similar work is being considered, which supports the novelty and significance of this work. From the ANOVA, the fitted model for the responses (skin friction coefficient and heat transfer rate) can be generated based on the physical factors (suction, wedge parameter, and volumetric concentration of hybrid nanoparticles). These equations can be used for practical and future applications regarding the mixed convection (opposing) flow subject to a permeable and vertical wedge.

2. Mathematical Formulation

Consider a mixed convection and steady flow of a water-based hybrid nanofluid with copper–alumina (Cu-Al2O3) nanoparticles towards a permeable wedge. The free stream flow with velocity u e x = a x m is assumed to move over the static wedge as portrayed in Figure 1, where a > 0 is a constant while m = β / 2 β is a positive constant related to the angle of the wedge, and the chosen m must be within the interval of 0 m 1 . Further, β is the Hartree pressure gradient and the total angle of the wedge (see Figure 1) is denoted as Ω = β π (Waini et al. [48], Rosca et al. [62]). Other presumptions for this physical model are:
  • The variable wall temperature is T w x = T + T 0 x / L 2 m 1 where L is a characteristic length of the wedge, T w > T   T 0 > 0 corresponds to an assisting flow (heated wedge), while T w < T   T 0 < 0 denotes an opposing flow (cooled wedge);
  • The far-field temperature T is fixed (constant);
  • Both nanoparticles and base fluid are in a thermal equilibrium state;
  • The model excludes the effect of sedimentation/aggregation since the hybrid nanofluid is in a stable synthesis.
The governing flow and energy equations are [48,62]
u   x + v y = 0 ,
u u x + v u y = u e d u e d x + μ h n f ρ h n f 2 u y 2 + β T h n f g T T cos α ,  
u T x + v T y = k h n f ( ρ C p ) h n f 2 T y 2 ,
with the boundary conditions
u = 0 , v = v w x , T = T w x ,   at   y = 0 u u e x , T T ,   as   y .
Here u x direction and v y direction are the hybrid nanofluid velocities, v w is the mass velocity, T is the temperature of the working fluid, α = Ω / 2 is the respective angle for the model and g is the gravitational acceleration [62]. The following similarity variables are introduced which complies Equation (1),
u = a x m f η , v = m + 1 a ν f 2 x m 1 2 f η + m 1 m + 1 η f η , θ η = T T T w T , η = y a m + 1 2 ν f x m 1 2 .
Hence, the respective surface mass flux velocity is
v w = m + 1 a ν f 2 x m 1 2 S ,
where S  represents the fluid suction or removal/injection. Further, by substituting Equation (5) into Equations (2)–(4), the following ODEs with the reduced BCs are obtained
μ h n f / μ f ρ h n f / ρ f f + f f + 2 m m + 1 1 f 2 + ρ β T h n f / ρ β T f ρ h n f / ρ f 2 m + 1 λ cos α θ = 0 ,
1 Pr k h n f / k f ρ C p h n f / ρ C p f θ + f θ 2 m 1 f θ = 0 ,
μ h n f / μ f ρ h n f / ρ f f + f f + 2 m m + 1 1 f 2 + ρ β T h n f / ρ β T f ρ h n f / ρ f 2 m + 1 λ cos α θ = 0 ,
where Pr = C p μ f / k f (Prandtl number), λ = G r / Re x 2 (mixed convection parameter), G r = g β T f T w x T x 3 / ν f 2 (local Grashof number) and Re x = x u e x / ν f (local Reynolds number). Further information for the mixed convection parameter is λ < 0 , λ = 0 and λ > 0 stand for an opposing, pure forced and assisting flows, respectively.
Following Takabi and Salehi [6], the correlations of hybrid nanofluid properties that were experimentally validated are shown in Table 1. These correlations are also used in many numerical studies regarding boundary layer flow. The exact properties of the pure water, Al2O3, and Cu nanoparticles for the computational analysis are listed in Table 2 [63,64]. A copper-water nanofluid model is obtained by setting ϕ Al 2 O 3 = ϕ 1 = 0 % and alumina-water nanofluid model when ϕ Cu = ϕ 2 = 0 % . Furthermore, a viscous fluid model is presentable if ϕ 1 = ϕ 2 = 0 % .
The definition of the skin friction coefficient and local Nusselt number is
C f = τ w ρ f u e 2 , N u x = x q w k f T w x T ,
where τw and qw are the wall shear stress and heat flux, respectively, defined as
τ w = μ h n f u y y = 0 , q w = k h n f T y y = 0 .
Using Equations (5), (9) and (10),
Re x 1 / 2 C f = μ h n f μ f m + 1 2 f 0 , Re x 1 / 2 N u x = k h n f k f m + 1 2 θ 0 .

3. Results and Discussion

In this section, the results are discussed based on the numerical solutions of Equations (7)–(9) through the bvp4c application in the Matlab software. The thermal and flow performances of Al2O3-Cu/water hybrid nanofluid are observed and computed for three regions: when λ > 0 (assisting flow solution), λ < 0 (opposing flow solution) and λ = 0 (pure force convection). For that reason, the effect wedge parameter m , suction S and concentration of the hybrid nanoparticles ϕ h n f are numerically studied on the skin friction coefficient and thermal rate as displayed in Figure 2, Figure 3, Figure 4, Figure 5, Figure 6 and Figure 7. The numerical solutions and appearance of dual solutions are observed within this interval 0.2 < m 0.3 , 0.05 < S 0.055 , 0 % ϕ h n f 2 % and λ c < λ 1 . The Prandtl number Pr = 6.2 is used which represents water as the carrier fluid. For the computational analysis, α = β π 2 in Equation (7) is modified in term of m such that α = m m + 1 π (angle in radians). For the model’s accurateness and validity, few solutions are validated by comparing them with existing literatures as presented in Table 3. Further, the observation of critical value λ c is necessary to find the final point of laminar flow separation. Usually, the critical value appears in the opposing flow region and beyond this value, no solution exists. Table 4 summarizes the critical values from Figure 2, Figure 3, Figure 4, Figure 5, Figure 6 and Figure 7 when different physical factors are considered. The expansion of the critical values is seen with the increment of m ,   S and ϕ h n f which reveals that all these physical parameters are beneficial in the deceleration of the boundary layer separation.
In addition to the projection of the critical values, Figure 2, Figure 3, Figure 4, Figure 5, Figure 6 and Figure 7 also exhibited the impact of the parameters on the flow behavior and thermal progress. In Figure 2 and Figure 3, the addition of wedge parameter m = 0.2 , 0.25 , 0.3 , which corresponds to the angle of the wedge α = 30 ° , 36 ° , 41.5 ° (in degree), enhances both skin friction Re x 1 / 2 C f and thermal rate Re x 1 / 2 N u x . Moreover, both Re x 1 / 2 C f and Re x 1 / 2 N u x increase as the mixed convection parameter λ 1 . Theoretically, the positive λ shows an assisting flow behavior that induces and aids fluid movement, including the active process of heat removal. Figure 4 and Figure 5 show the augmentation of Re x 1 / 2 C f and Re x 1 / 2 N u x with the increment of the suction parameter. However, the skin friction distribution was only slightly affected as compared to the heat transfer progress. As previously discussed, the limitation of suction magnitude was based on the observation of dual solutions. There is no restriction if the researchers use a higher magnitude of suction. The impact of volumetric concentration on the hybrid nanoparticles is presented in Figure 6 and Figure 7. It is worth mentioning that we considered an equal concentration of Cu and Al2O3 nanoparticles such that ϕ h n f = 0.01 , 0.02 , 0.03 corresponds to ϕ 1 = ϕ 2 = 0.005 , 0.01 , 0.015 . The results show that both skin friction coefficient and thermal rate enhance with the increment of ϕ h n f .

4. Response Surface Methodology

The experimental design for the particular set of data in the boundary layer flow problem is also important, where the researchers can estimate which parameters (factors) are influential or beneficial in optimizing the responses (skin friction coefficient/thermal rate). There are many types of experimental design available such as factorial design and response surface methodology through central composite design or Box–Behnken design. From the numerical interpretation, the suction, volumetric nanoparticles concentration, and wedge parameter affect and enhance both Re x 1 / 2 C f and Re x 1 / 2 N u x ; however, from the statistical data analysis, the most significant factor contributing to the enhancement of responses can be predicted. Table 4 displays the wedge parameter, volumetric concentration of the hybrid nanoparticles, and suction parameter as the factors, and they are symbolized as A, B, and C, respectively. The level is referred to the magnitude of each factor and is categorized as low, medium, and high. As previously stated, the controlling parameters are used within the range of 0.2 < m 0.3 , 0.01 ϕ h n f 0.03 and 0.05 < S 0.055 . The division of low, medium, and high magnitudes of the parameters is also clearly stated in Table 5. The total runs for three factors k = 3 with 5 centre points C = 5 are based on this formula R = 2 k + 2 k + C where 2 k is the factorial points, 2 k is the axial points, and C is the center points [57]. Table 6 shows the response surface methodology using a central composite design with 19 total runs when λ = 1 and Pr = 6.2 . By using the data in Table 6, the correlations between the factors m , ϕ h n f , S and responses Re x 1 / 2 C f , Re x 1 / 2 N u x can be defined by this general response surface Equation (13)
y = r 0 + r A A + r B B + r C C + r A 2 A 2 + r B 2 B 2 + r C 2 C 2 + r A B A B + r C A C A + r B C B C + ε ,
where r 0 is an intercept, r A , r B , r C is the linear effects, r A 2 , r B 2 , r C 2 is the quadratic effects and r A B , r C A , r B C is the interaction effects. Two response surface equations were considered for the two responses. The execution of data analysis was further conducted using analysis of variance (ANOVA) through the fit general linear model in statistical analysis Minitab software. The results are presented in Table 5, Table 6, Table 7, Table 8 and Table 9 and Figure 8 and Figure 9.
In the interest of producing a good model and well-fitted to the response-surface component, three main indicators need to be considered, which are the p-values (p-value < 0.001) from the analysis of variance (ANOVA) table, the value of adjusted R square (R-sq), and standard residual plot. Table 7 presents the ANOVA table to analyse the effect considered parameters such as wedge parameter (A), suction parameter (B), and volumetric concentration of nanoparticles (C) to the model of response-surface component for the skin friction coefficient, Re x 1 / 2 C f , and heat transfer rate, Re x 1 / 2 N u x . Based on the result of the p-values, it is apparent that the wedge parameter (A) and volumetric concentration of nanoparticles (C) have a significant impact on all two responses, Re x 1 / 2 C f and Re x 1 / 2 N u x . However, the model’s suction parameter (B) effect is only significant for. In addition, the symbol * in Table 7 shows that the value is too small. Model summary for Re x 1 / 2 C f and Re x 1 / 2 N u x , which include adjusted R-sq, is presented in Table 8.
This value is represented by how much the models explain the variation in response used. It was obtained that the value of adjusted R-sq for Re x 1 / 2 C f and Re x 1 / 2 N u x is 99.89% and 99.95%, respectively. Based on this result, it was shown that all models explain a very high percentage of the total variation in respective responses. The residual normal plot for both fitted models, Re x 1 / 2 C f and Re x 1 / 2 N u x , is presented in Figure 8. This result is used to evaluate the goodness of fit to the models of the response-surface component. A good model that accurately represents the relationship between behavioral input parameters and response reveals a one-to-one correlation between theoretical and observed quantiles. It was found that for both models, there is almost a one-to-one correlation between theoretical and observed quintiles. The distribution of residual from both fitted models Re x 1 / 2 C f and Re x 1 / 2 N u x is presented in Figure 9. It is shown that both fitted models’ residuals are normally distributed. Therefore, both models are well-fitted.
Table 9 presents the fitted model terms for the skin friction coefficient, Re x 1 / 2 C f , and heat transfer rate, Re x 1 / 2 N u x , to analyze the significant input parameters together with t-value and p-value with a 95% confidence interval. It was found that the wedge parameter (A) and volumetric concentration of nanoparticles (C) are significant terms affecting Re x 1 / 2 C f , with a p-value < 0.001. Whereas the wedge parameter (A), suction parameter (B), and volumetric concentration of nanoparticles (C) are significantly (p-value < 0.001) affecting Re x 1 / 2 N u x . Therefore, the corresponding fitted models for Re x 1 / 2 C f and heat transfer rate by considering the three effects can be expressed as
y skin   friction = 0.19818 + 0.19995 A + 0.00663 B + 0.04001 C 0.00428 AB   0.01084 AC 0.00024 BC 0.04114 A 2 0.00019 B 2 0.00131 C 2 ,
y heat   transfer = 0.651294 + 0.111817 A + 0.010414 B + 0.017393 C 0.001419 AB   0.002866 AC 0.000153 BC 0.010382 A 2 0.001627 B 2 0.000664 C 2 .

5. Conclusions

The flow behavior and thermal properties of Cu-Al2O3/water with mixed convection in the context of a permeable and vertical wedge were addressed and discussed in detail. The similarity transformation was used to simplify and reduce the partial differential equations into a set of ordinary (similarity) differential equations. The Matlab software, with its capable bvp4c function, was utilized to numerically compute the steady similarity solutions. The numerical solutions were then presented in the graphs of skin friction coefficient and heat transfer rate for various wedge parameters, the volumetric concentration of nanoparticles, and suction parameters. Moreover, for the statistical evaluation, the response surface methodology was used to gather the data and then analyzed using the analysis of variance (ANOVA) through the fit general linear model in the Minitab software. The following is the summary of the findings:
  • The steady flow problem was solved for three cases: assisting flow, opposing flow, and pure force convective flow. The dual solutions were observable only in the opposing flow region when appropriate parameters were used;
  • From the numerical evaluation, the addition of the wedge parameter, the volumetric concentration of nanoparticles, and the suction parameter contribute to the expansion of the critical value, which implies the delay in boundary layer separation. Furthermore, the skin friction coefficient and heat transfer process in the opposing flow region were also raised by these controlling parameters;
  • From the statistical evaluation, the two responses (heat transfer rate and skin friction coefficient) were significantly affected by the wedge and volumetric concentration of nanoparticles factors. However, the effect of the suction parameter is only relevant for the heat transfer rate and not for the skin friction coefficient;
  • Nonetheless, all models account for a significant part of the total variation in the responses. Moreover, the residuals of both fitted models were also demonstrated to be normally distributed and well-fitted.

Author Contributions

Formulation of mathematical model and methodology, N.S.K., I.W. and M.F.M.; Model validation, N.S.K. and M.F.M.; Writing, N.S.K., I.W., N.A.Z. and K.B.H.; review and editing, N.A.Z., K.B.H., N.M.A. and I.P. All authors have read and agreed to the published version of the manuscript.

Funding

The funding is from incentive grant JURNAL/2020/FTKMP/Q00050 (Universiti Teknikal Malaysia Melaka-UTeM).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

We acknowledge the grants JURNAL/2020/FTKMP/Q00050 and FRGS/1/2021/STG06/UTEM/03/1 from the Ministry of Higher Education (Malaysia) and Universiti Teknikal Malaysia Melaka.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shah, N.A.; Wakif, A.; El-Zahar, E.R.; Ahmad, S.; Yook, S.J. Numerical simulation of a thermally enhanced EMHD flow of a heterogeneous micropolar mixture comprising (60%)-ethylene glycol (EG),(40%)-water (W), and copper oxide nanomaterials (CuO). Case Stud. Therm. Eng. 2022, 35, 102046. [Google Scholar] [CrossRef]
  2. Sabu, A.S.; Wakif, A.; Areekara, S.; Mathew, A.; Shah, N.A. Significance of nanoparticles’ shape and thermo-hydrodynamic slip constraints on MHD alumina-water nanoliquid flows over a rotating heated disk: The passive control approach. Int. Commun. Heat Mass Transf. 2021, 129, 105711. [Google Scholar] [CrossRef]
  3. Rasool, G.; Shah, N.A.; El-Zahar, E.R.; Wakif, A. Numerical investigation of EMHD nanofluid flows over a convectively heated riga pattern positioned horizontally in a Darcy-Forchheimer porous medium: Application of passive control strategy and generalized transfer laws. Waves Random Complex Media 2022, 1–20. [Google Scholar] [CrossRef]
  4. Jana, S.; Salehi-Khojin, A.; Zhong, W.H. Enhancement of fluid thermal conductivity by the addition of single and hybrid nano-additives. Thermochim. Acta 2007, 462, 45–55. [Google Scholar] [CrossRef]
  5. Suresh, S.; Venkitaraj, K.P.; Selvakumar, P.; Chandrasekar, M. Synthesis of Al2O3-Cu/water hybrid nanofluids using two step method and its thermo physical properties. Colloids Surf. A Physicochem. Eng. Asp. 2011, 388, 41–48. [Google Scholar] [CrossRef]
  6. Takabi, B.; Salehi, S. Augmentation of the heat transfer performance of a sinusoidal corrugated enclosure by employing hybrid nanofluid. Adv. Mech. Eng. 2014, 6, 147059. [Google Scholar] [CrossRef]
  7. Devi, S.P.A.; Devi, S.S.U. Numerical investigation of hydromagnetic hybrid Cu-Al2O3/water nanofluid flow over a permeable stretching sheet with suction. Int. J. Nonlinear Sci. Numer. Simul. 2016, 17, 249–257. [Google Scholar] [CrossRef]
  8. Nabwey, H.A.; Mahdy, A. Transient flow of micropolar dusty hybrid nanofluid loaded with Fe3O4-Ag nanoparticles through a porous stretching sheet. Results Phys. 2021, 21, 103777. [Google Scholar] [CrossRef]
  9. Roy, N.C.; Hossain, M.A.; Pop, I. Analysis of dual solutions of unsteady micropolar hybrid nanofluid flow over a stretching/shrinking sheet. J. Appl. Comput. Mech. 2021, 7, 19–33. [Google Scholar] [CrossRef]
  10. Riaz, A.; Ellahi, R.; Sait, S.M. Role of hybrid nanoparticles in thermal performance of peristaltic flow of Eyring–Powell fluid model. J. Therm. Anal. Calorim. 2021, 143, 1021–1035. [Google Scholar] [CrossRef]
  11. Khashi’ie, N.S.; Waini, I.; Zokri, S.M.; Kasim, A.R.M.; Arifin, N.M.; Pop, I. Stagnation point flow of a second-grade hybrid nanofluid induced by a Riga plate. Int. J. Numer. Methods Heat Fluid Flow 2021, 32, 2221–2239. [Google Scholar] [CrossRef]
  12. Abbas, N.; Nadeem, S.; Malik, M.Y. Theoretical study of micropolar hybrid nanofluid over Riga channel with slip conditions. Phys. A Stat. Mech. Its Appl. 2020, 551, 124083. [Google Scholar] [CrossRef]
  13. Waqas, H.; Farooq, U.; Naseem, R.; Hussain, S.; Alghamdi, M. Impact of MHD radiative flow of hybrid nanofluid over a rotating disk. Case Stud. Therm. Eng. 2021, 26, 101015. [Google Scholar] [CrossRef]
  14. Bilal, M.; Khan, I.; Gul, T.; Tassaddiq, A.; Alghamdi, W.; Mukhtar, S.; Kumam, P. Darcy-forchheimer hybrid nano fluid flow with mixed convection past an inclined cylinder. Comput. Mater. Contin. 2021, 66, 2025–2039. [Google Scholar] [CrossRef]
  15. Uysal, C.; Korkmaz, M.E. Estimation of Entropy Generation for Ag-MgO/Water Hybrid Nanofluid Flow through Rectangular Minichannel by Using Artificial Neural Network. J. Polytech. 2019, 22, 41–51. [Google Scholar] [CrossRef]
  16. Kumar, V.; Sarkar, J. Particle ratio optimization of Al2O3-MWCNT hybrid nanofluid in minichannel heat sink for best hydrothermal performance. Appl. Therm. Eng. 2020, 165, 114546. [Google Scholar] [CrossRef]
  17. Salehi, S.; Nori, A.; Hosseinzadeh, K.; Ganji, D.D. Hydrothermal analysis of MHD squeezing mixture fluid suspended by hybrid nanoparticles between two parallel plates. Case Stud. Therm. Eng. 2020, 21, 100650. [Google Scholar] [CrossRef]
  18. Zainal, N.A.; Nazar, R.; Naganthran, K.; Pop, I. Viscous dissipation and MHD hybrid nanofluid flow towards an exponentially stretching/shrinking surface. Neural Comput. Appl. 2021, 33, 11285–11295. [Google Scholar] [CrossRef]
  19. Zainal, N.A.; Nazar, R.; Naganthran, K.; Pop, I. Unsteady MHD mixed convection flow in hybrid nanofluid at three-dimensional stagnation point. Mathematics 2021, 9, 549. [Google Scholar] [CrossRef]
  20. Zainal, N.A.; Nazar, R.; Naganthran, K.; Pop, I. Heat generation/absorption effect on MHD flow of hybrid nanofluid over bidirectional exponential stretching/shrinking sheet. Chin. J. Phys. 2021, 69, 118–133. [Google Scholar] [CrossRef]
  21. Khashi’ie, N.S.; Arifin, N.M.; Sheremet, M.; Pop, I. Shape factor effect of radiative Cu–Al2O3/H2O hybrid nanofluid flow towards an EMHD plate. Case Stud. Therm. Eng. 2021, 26, 101199. [Google Scholar] [CrossRef]
  22. Khashi’ie, N.S.; Waini, I.; Zainal, N.A.; Hamzah, K.B. Hybrid nanofluid flow past a shrinking cylinder with prescribed surface heat flux. Symmetry 2020, 12, 1493. [Google Scholar] [CrossRef]
  23. Khashi’ie, N.S.; Arifin, N.M.; Wahi, N.; Pop, I.; Nazar, R.; Hafidzuddin, E.H. Thermal marangoni flow past a permeable stretching/shrinking sheet in a hybrid Cu-Al2O3/water nanofluid. Sains Malays. 2020, 49, 211–222. [Google Scholar] [CrossRef]
  24. Khashi’ie, N.S.; Arifin, N.M.; Pop, I. Magnetohydrodynamics (MHD) boundary layer flow of hybrid nanofluid over a moving plate with Joule heating. Alex. Eng. J. 2021, 61, 1938–1945. [Google Scholar] [CrossRef]
  25. Waini, I.; Ishak, A.; Pop, I. Hybrid nanofluid flow on a shrinking cylinder with prescribed surface heat flux. Int. J. Numer. Methods Heat Fluid Flow 2021, 31, 1987–2004. [Google Scholar] [CrossRef]
  26. Waini, I.; Ishak, A.; Pop, I. Flow towards a stagnation region of a vertical plate in a hybrid nanofluid: Assisting and opposing flows. Mathematics 2021, 9, 448. [Google Scholar] [CrossRef]
  27. Waini, I.; Ishak, A.; Pop, I. MHD stagnation point flow and heat transfer towards a permeable stretching/shrinking surface in a hybrid nanofluid. Sains Malays. 2021, 50, 2819–2832. [Google Scholar] [CrossRef]
  28. Waini, I.; Ishak, A.; Pop, I. Hybrid nanofluid flow with homogeneous-heterogeneous reactions. Comput. Mater. Contin. 2021, 68, 3255–3269. [Google Scholar] [CrossRef]
  29. Waini, I.; Ishak, A.; Pop, I. Flow towards a stagnation region of a curved surface in a hybrid nanofluid with buoyancy effects. Mathematics 2021, 9, 2330. [Google Scholar] [CrossRef]
  30. Shah, N.A.; Wakif, A.; El-Zahar, E.R.; Thumma, T.; Yook, S.J. Heat transfers thermodynamic activity of a second-grade ternary nanofluid flow over a vertical plate with Atangana-Baleanu time-fractional integral. Alex. Eng. J. 2022, 61, 10045–10053. [Google Scholar] [CrossRef]
  31. Falkner, V.M.; Skan, S.W. Some approximate solutions of the boundary layer equations. Philos. Mag. 1931, 12, 865–896. [Google Scholar] [CrossRef]
  32. Hartree, D.R. On an equation occurring in Falkner and Skan’s approximate treatment of the equations of the boundary layer. Math. Proc. Camb. Philos. Soc. 1937, 33, 223–239. [Google Scholar] [CrossRef]
  33. Ahmad, R.; Khan, W.A. Numerical study of heat and mass transfer MHD viscous flow over a moving wedge in the presence of viscous dissipation and heat source/sink with convective boundary condition. Heat Transf.-Asian Res. 2014, 43, 17–38. [Google Scholar] [CrossRef]
  34. Khan, U.; Ahmed, N.; Bin-Mohsen, B.; Mohyud-Din, S.T. Nonlinear Radiation Effects on Flow of Nanofluid over a Porous Wedge in the Presence of Magnetic Field Thermophoresis coefficient Brownian motion coefficient Prandtl nu. Int. J. Numer. Methods Heat Fluid Flow 2017, 27, 48–63. [Google Scholar] [CrossRef]
  35. Ullah, I.; Shafie, S.; Khan, I. MHD heat transfer flow of Casson fluid past a stretching wedge subject to suction and injection. Malays. J. Fundam. Appl. Sci. 2017, 13, 637–641. [Google Scholar] [CrossRef] [Green Version]
  36. Khan, M.; Azam, M.; Munir, A. On unsteady Falkner-Skan flow of MHD Carreau nanofluid past a static/moving wedge with convective surface condition. J. Mol. Liq. 2017, 230, 48–58. [Google Scholar] [CrossRef]
  37. Ahmad, K.; Hanouf, Z.; Ishak, A. MHD Casson nanofluid flow past a wedge with Newtonian heating. Eur. Phys. J. Plus 2017, 132, 87. [Google Scholar] [CrossRef] [Green Version]
  38. Tlili, I.; Hamadneh, N.N.; Khan, W.A. Thermodynamic analysis of MHD heat and mass transfer of nanofluids past a static wedge with navier slip and convective boundary conditions. Arab. J. Sci. Eng. 2019, 44, 1255–1267. [Google Scholar] [CrossRef]
  39. Hashim; Khan, M.; Huda, N.U.; Hamid, A. Non-linear radiative heat transfer analysis during the flow of Carreau nanofluid due to wedge-geometry: A revised model. Int. J. Heat Mass Transf. 2019, 131, 1022–1031. [Google Scholar] [CrossRef]
  40. Ishak, A.; Nazar, R.; Pop, I. Moving wedge and flat plate in a micropolar fluid. Int. J. Eng. Sci. 2006, 44, 1225–1236. [Google Scholar] [CrossRef]
  41. Ishak, A.; Nazar, R.; Pop, I. Falkner-Skan equation for flow past a moving wedge with suction or injection. J. Appl. Math. Comput. 2007, 25, 67–83. [Google Scholar] [CrossRef]
  42. Ishak, A.; Nazar, R.; Pop, I. Moving wedge and flat plate in a power-law fluid. Int. J. Non-Linear Mech. 2011, 46, 1017–1021. [Google Scholar] [CrossRef]
  43. Khan, W.A.; Pop, I. Boundary layer flow past a wedge moving in a nanofluid. Math. Probl. Eng. 2013, 2013, 637285. [Google Scholar] [CrossRef]
  44. Hedayati, F.; Malvandi, A.; Ganji, D.D. Second-law analysis of fluid flow over an isothermal moving wedge. Alex. Eng. J. 2014, 53, 1–9. [Google Scholar] [CrossRef] [Green Version]
  45. Alam, M.S.; Khatun, M.A.; Rahman, M.M.; Vajravelu, K. Effects of variable fluid properties and thermophoresis on unsteady forced convective boundary layer flow along a permeable stretching/shrinking wedge with variable Prandtl and Schmidt numbers. Int. J. Mech. Sci. 2016, 105, 191–205. [Google Scholar] [CrossRef]
  46. Khan, U.; Ahmed, N.; Mohyud-Din, S.T.; Bin-Mohsin, B. Nonlinear radiation effects on MHD flow of nanofluid over a nonlinearly stretching/shrinking wedge. Neural Comput. Appl. 2017, 28, 2041–2050. [Google Scholar] [CrossRef]
  47. Awaludin, I.S.; Ishak, A.; Pop, I. On the stability of MHD boundary layer flow over a stretching/shrinking wedge. Sci. Rep. 2018, 8, 13622. [Google Scholar] [CrossRef] [Green Version]
  48. Waini, I.; Ishak, A.; Pop, I. MHD flow and heat transfer of a hybrid nanofluid past a permeable stretching/shrinking wedge. Appl. Math. Mech. 2020, 41, 507–520. [Google Scholar] [CrossRef]
  49. Rashad, A.M. Impact of thermal radiation on MHD slip flow of a ferrofluid over a non-isothermal wedge. J. Magn. Magn. Mater. 2017, 422, 25–31. [Google Scholar] [CrossRef]
  50. Hassan, M.; Faisal, A.; Ali, I.; Bhatti, M.M.; Yousaf, M. Effects of Cu–Ag hybrid nanoparticles on the momentum and thermal boundary layer flow over the wedge. Proc. Inst. Mech. Eng. Part E J. Process Mech. Eng. 2019, 233, 1128–1136. [Google Scholar] [CrossRef]
  51. Ahmed, N.; Tassaddiq, A.; Alabdan, R.; Adnan; Khan, U.; Noor, S.; Mohyud-Din, S.T.; Khan, I. Applications of nanofluids for the thermal enhancement in radiative and dissipative flow over a wedge. Appl. Sci. 2019, 9, 1976. [Google Scholar] [CrossRef] [Green Version]
  52. Zaib, A.; Haq, R. Magnetohydrodynamics mixed convective flow driven through a static wedge including TiO2 nanomaterial with micropolar liquid: Similarity dual solutions via finite difference method. Proc. Inst. Mech. Eng. Part C J. Mech. Eng. Sci. 2019, 233, 5813–5825. [Google Scholar] [CrossRef]
  53. Rawat, S.K.; Upreti, H.; Kumar, M. Comparative Study of Mixed Convective MHD Cu-Water Nanofluid Flow over a Cone and Wedge using Modified Buongiorno’s Model in Presence of Thermal Radiation and Chemical Reaction via Cattaneo-Christov Double Diffusion Model. J. Appl. Comput. Mech. 2020, 7, 1383–1402. [Google Scholar] [CrossRef]
  54. Mahanthesh, B.; Shehzad, S.A.; Ambreen, T.; Khan, S.U. Significance of Joule heating and viscous heating on heat transport of MoS2–Ag hybrid nanofluid past an isothermal wedge. J. Therm. Anal. Calorim. 2021, 143, 1221–1229. [Google Scholar] [CrossRef]
  55. Braima, N.; Maryam, A.N.; Odejobi, O.J. Utilization of Response Surface Methodology (RSM) in the Optimization of Crude Oil Refinery Process, New Port-Harcourt Refinery. J. Multidiscip. Eng. Sci. Technol. 2016, 3, 4361–4369. [Google Scholar]
  56. Wang, M.; Wang, J.; Tan, J.X.; Sun, J.F.; Mou, J.L. Optimization of Ethanol Fermentation from Sweet Sorghum Juice Using Response Surface Methodology. Energy Sources Part A Recovery Util. Environ. Eff. 2011, 33, 1139–1146. [Google Scholar] [CrossRef]
  57. Mehmood, T.; Ramzan, M.; Howari, F.; Kadry, S. Application of response surface methodology on the nanofluid flow over a rotating disk with autocatalytic chemical reaction and entropy generation optimization. Sci. Rep. 2021, 11, 4021. [Google Scholar] [CrossRef]
  58. Mahanthesh, B.; Thriveni, K. Nanoparticle aggregation effects on radiative heat transport of nanoliquid over a vertical cylinder with sensitivity analysis. Appl. Math. Mech. 2021, 42, 331–346. [Google Scholar] [CrossRef]
  59. Shafiq, A.; Sindhu, T.N.; Al-Mdallal, Q.M. A sensitivity study on carbon nanotubes significance in Darcy–Forchheimer flow towards a rotating disk by response surface methodology. Sci. Rep. 2021, 11, 8812. [Google Scholar] [CrossRef]
  60. Vahedi, S.M.; Ghadi, A.Z.; Valipour, M.S. Application of response surface methodology in the optimization of magneto-hydrodynamic flow around and through a porous circular cylinder. J. Mech. 2018, 34, 695–710. [Google Scholar] [CrossRef]
  61. Abdelmalek, Z.; Mahanthesh, B.; Basir, M.F.; Imtiaz, M.; Mackolil, J.; Khan, N.S.; Nabwey, H.A.; Tlili, I. Mixed radiated magneto Casson fluid flow with Arrhenius activation energy and Newtonian heating effects: Flow and sensitivity analysis. Alex. Eng. J. 2020, 59, 3991–4011. [Google Scholar] [CrossRef]
  62. Roșca, N.C.; Roșca, A.V.; Pop, I. Mixed convection flow of a hybrid nanofluid past a vertical wedge with thermal radiation effect. Int. J. Numer. Methods Heat Fluid Flow 2021, 32, 806–824. [Google Scholar] [CrossRef]
  63. Khanafer, K.; Vafai, K.; Lightstone, M. Buoyancy-driven heat transfer enhancement in a two-dimensional enclosure utilizing nanofluids. Int. J. Heat Mass Transf. 2003, 46, 3639–3653. [Google Scholar] [CrossRef]
  64. Oztop, H.F.; Abu-Nada, E. Numerical study of natural convection in partially heated rectangular enclosures filled with nanofluids. Int. J. Heat Fluid Flow 2008, 29, 1326–1336. [Google Scholar] [CrossRef]
Figure 1. The physical model.
Figure 1. The physical model.
Nanomaterials 12 04016 g001
Figure 2. Re x 1 / 2 C f towards λ when S = 0.05 , ϕ h n f = 0.02 , and different m .
Figure 2. Re x 1 / 2 C f towards λ when S = 0.05 , ϕ h n f = 0.02 , and different m .
Nanomaterials 12 04016 g002
Figure 3. Re x 1 / 2 N u x towards λ when S = 0.05 , ϕ h n f = 0.02 , and different m .
Figure 3. Re x 1 / 2 N u x towards λ when S = 0.05 , ϕ h n f = 0.02 , and different m .
Nanomaterials 12 04016 g003
Figure 4. Re x 1 / 2 C f towards λ when m = 0.3 , ϕ h n f = 0.02 , and variety S .
Figure 4. Re x 1 / 2 C f towards λ when m = 0.3 , ϕ h n f = 0.02 , and variety S .
Nanomaterials 12 04016 g004
Figure 5. Re x 1 / 2 N u x towards λ when m = 0.3 , ϕ h n f = 0.02 , and variety S .
Figure 5. Re x 1 / 2 N u x towards λ when m = 0.3 , ϕ h n f = 0.02 , and variety S .
Nanomaterials 12 04016 g005
Figure 6. Re x 1 / 2 C f towards λ when m = 0.3 , S = 0.05 , and different ϕ h n f .
Figure 6. Re x 1 / 2 C f towards λ when m = 0.3 , S = 0.05 , and different ϕ h n f .
Nanomaterials 12 04016 g006
Figure 7. Re x 1 / 2 N u x towards λ when m = 0.3 , S = 0.05 , and different ϕ h n f .
Figure 7. Re x 1 / 2 N u x towards λ when m = 0.3 , S = 0.05 , and different ϕ h n f .
Nanomaterials 12 04016 g007
Figure 8. The residual normal plot for (a) Re x 1 / 2 C f and (b) Re x 1 / 2 N u x .
Figure 8. The residual normal plot for (a) Re x 1 / 2 C f and (b) Re x 1 / 2 N u x .
Nanomaterials 12 04016 g008
Figure 9. The distribution of residual from (a) Re x 1 / 2 C f and (b) Re x 1 / 2 N u x .
Figure 9. The distribution of residual from (a) Re x 1 / 2 C f and (b) Re x 1 / 2 N u x .
Nanomaterials 12 04016 g009
Table 1. Correlations of hybrid nanofluids.
Table 1. Correlations of hybrid nanofluids.
PropertiesCorrelations
Thermal conductivity k h n f = ϕ 1 k 1 + ϕ 2 k 2 ϕ h n f 2 ϕ h n f k f + 2 ϕ 1 k 1 + ϕ 2 k 2 + 2 k f ϕ 1 k 1 + ϕ 2 k 2 ϕ h n f + ϕ h n f k f ϕ 1 k 1 + ϕ 2 k 2 + 2 k f k f
Thermal expansion ρ β T h n f = ϕ 1 ρ β T s 1 + ϕ 2 ρ β T s 2 + 1 ϕ h n f ρ β T f
Heat capacity ρ C p h n f = ϕ 1 ρ C p s 1 + ϕ 2 ρ C p s 2 + 1 ϕ h n f ρ C p f
Density ρ h n f = ϕ 1 ρ s 1 + ϕ 2 ρ s 2 + 1 ϕ h n f ρ f
Dynamic viscosity μ h n f = μ f 1 ϕ h n f 2.5 ; ϕ h n f = ϕ 1 + ϕ 2
Table 2. Thermophysical properties for H2O, Al2O3, and Cu.
Table 2. Thermophysical properties for H2O, Al2O3, and Cu.
PropertiesWaterCopperAlumina
ρ kg / m 3 997.189333970
C p J / kgK 4179385765
k W / mK 0.613040040
β T S / m 21 × 10 5 1.67 × 10 5 0.85 × 10 5
Prandtl number, Pr 6.2--
Table 3. Comparison of Re x 1 / 2 C f with different value of ϕ 2 when λ = S = 0 .
Table 3. Comparison of Re x 1 / 2 C f with different value of ϕ 2 when λ = S = 0 .
ϕ 2
PresentWaini et al. [28]Waini et al. [48]
01.29689021.2968901.296890
0.251.55384961.5538501.553850
Table 4. Summary of the critical values with different physical parameters.
Table 4. Summary of the critical values with different physical parameters.
m S ϕ h n f λ c
0.20.050.02−1.0491
0.25 −1.3561
0.3 −1.7248
0.053 −1.7435
0.055 −1.7561
0.050.01−1.6598
0.03−1.7898
Table 5. Experimental design for the factors and their levels.
Table 5. Experimental design for the factors and their levels.
FactorSymbolLevel
Low
(−1)
Medium
(0)
High
(1)
m A0.20.250.3
S B0.050.0530.055
ϕ h n f C0.010.020.03
Table 6. Response surface methodology with central composite design for the factors and responses.
Table 6. Response surface methodology with central composite design for the factors and responses.
RunsReal ValuesCoded ValuesResponses
m S ϕ h n f ABCSkin Friction CoefficientHeat Transfer Rate
10.20.0530.03−1010.0086506820.549364184
20.20.0530.01−10−1−0.1008464970.506083442
30.250.0530.020000.1981819090.651293721
40.30.050.021−100.3521591480.741277394
50.250.0530.020000.1981819090.651293721
60.30.0530.031010.3906312450.768679201
70.250.0530.020000.1981819090.651293721
80.250.050.010−1−10.1561118880.623220285
90.250.0550.0101−10.1670341940.643437992
100.30.0530.0110−10.3224947360.736190043
110.250.0530.020000.1981819090.651293721
120.250.050.030−110.2268042090.65487521
130.250.0530.020000.1981819090.651293721
140.250.0550.030110.2367624520.674480548
150.30.0550.021100.359681230.760184189
160.20.050.02−1−10−0.0545330560.515546913
170.250.0530.020000.1981819090.651293721
180.20.0550.02−110−0.0298968040.540129695
190.20.0530.03−1010.0086506820.549364184
Table 7. Analysis of variance (ANOVA) for the responses.
Table 7. Analysis of variance (ANOVA) for the responses.
SourceDFSeq SSContributionAdj SSAdj MSF-Valuep-Value
Re x 1 / 2 C f
A10.34741894.01%0.3444350.34443515024.570.000
B10.0003520.10%0.0003520.00035215.340.004
C10.0133173.60%0.0137900.013790601.550.000
A*B10.0000730.02%0.0000730.0000733.190.108
A*C10.0002380.06%0.0005490.00054923.930.001
B*C10.0000000.00%0.0000000.0000000.010.922
A*A10.0079312.15%0.0076830.007683335.160.000
B*B10.0000000.00%0.0000000.0000000.010.933
C*C10.0000080.00%0.0000080.0000080.340.574
Error90.0002060.06%0.0002060.000023
Lack-of-Fit30.0002060.06%0.0002060.000069**
Pure Error60.0000000.00%0.0000000.000000
Total180.369543100.00%
Re x 1 / 2 N u x
A10.10874596.40%0.1077180.10771838033.610.000
B10.0008680.77%0.0008680.000868306.350.000
C10.0026022.31%0.0026060.002606920.210.000
A*B10.0000080.01%0.0000080.0000082.840.126
A*C10.0000190.02%0.0000380.00003813.540.005
B*C10.0000000.00%0.0000000.0000000.030.860
A*A10.0005190.46%0.0004890.000489172.820.000
B*B10.0000130.01%0.0000120.0000124.240.070
C*C10.0000020.00%0.0000020.0000020.710.423
Error90.0000250.02%0.0000250.000003
Lack-of-Fit30.0000250.02%0.0000250.000008**
Pure Error60.0000000.00%0.0000000.000000
Total180.112801100.00%0.112801
The symbol * shows that the value is too small.
Table 8. Model summary for the responses.
Table 8. Model summary for the responses.
SR-sqR-sq (adj)R-sq (pred)
Re x 1 / 2 C f
0.004788099.94%99.89%99.19%
Re x 1 / 2 N u x
0.001682999.98%99.95%99.67%
Table 9. Fitted model terms for the responses.
Table 9. Fitted model terms for the responses.
TermCoefSE Coef95% CIT-Valuep-ValueVIF
Re x 1 / 2 C f
Constant0.198180.00195(0.19376, 0.20260)101.390.000
A0.199950.00163(0.19626, 0.20364)122.570.0001.04
B0.006630.00169(0.00280, 0.01046)3.920.0041.00
C0.040010.00163(0.03632, 0.04370)24.530.0001.04
A*B−0.004280.00239(−0.00969, 0.00114)−1.790.1081.00
A*C−0.010840.00222(−0.01586, −0.00583)−4.890.0011.06
B*C−0.000240.00239(−0.00566, 0.00517)−0.100.9221.00
A*A−0.041140.00225(−0.04622, −0.03605)−18.310.0001.04
B*B−0.000190.00225(−0.00528, 0.00489)−0.090.9331.02
C*C−0.001310.00225(−0.00639, 0.00377)−0.580.5741.04
Re x 1 / 2 N u x
Constant0.6512940.000687(0.649740, 0.652848)947.970.000
A0.1118170.000573(0.110520, 0.113114)195.020.0001.04
B0.0104140.000595(0.009068, 0.011760)17.500.0001.00
C0.0173930.000573(0.016096, 0.018690)30.330.0001.04
A*B−0.0014190.000841(−0.003322, 0.000485)−1.690.1261.00
A*C−0.0028660.000779(−0.004629, −0.001104)−3.680.0051.06
B*C−0.0001530.000841(−0.002057, 0.001750)−0.180.8601.00
A*A−0.0103820.000790(−0.012169, −0.008596)−13.150.0001.04
B*B−0.0016270.000790(−0.003413, 0.000160)−2.060.0701.02
C*C−0.0006640.000790(−0.002450, 0.001123)−0.840.4231.04
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Khashi’ie, N.S.; Waini, I.; Mukhtar, M.F.; Zainal, N.A.; Hamzah, K.B.; Arifin, N.M.; Pop, I. Response Surface Methodology (RSM) on the Hybrid Nanofluid Flow Subject to a Vertical and Permeable Wedge. Nanomaterials 2022, 12, 4016. https://doi.org/10.3390/nano12224016

AMA Style

Khashi’ie NS, Waini I, Mukhtar MF, Zainal NA, Hamzah KB, Arifin NM, Pop I. Response Surface Methodology (RSM) on the Hybrid Nanofluid Flow Subject to a Vertical and Permeable Wedge. Nanomaterials. 2022; 12(22):4016. https://doi.org/10.3390/nano12224016

Chicago/Turabian Style

Khashi’ie, Najiyah Safwa, Iskandar Waini, Mohd Fariduddin Mukhtar, Nurul Amira Zainal, Khairum Bin Hamzah, Norihan Md Arifin, and Ioan Pop. 2022. "Response Surface Methodology (RSM) on the Hybrid Nanofluid Flow Subject to a Vertical and Permeable Wedge" Nanomaterials 12, no. 22: 4016. https://doi.org/10.3390/nano12224016

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

Article Metrics

Back to TopTop