Volume 7, Issue 4 e70112
RESEARCH ARTICLE
Open Access

Numerical Study of Radiative and Dissipative MHD Casson Nanofluid Over a Cone With High-Order Chemical Reaction

L. Joseph Sademaki

L. Joseph Sademaki

Department of Applied Science and Social Studies, Arusha Technical College, Arusha, Tanzania

Department of Mathematics and Statistics, CNMS, the University of Dodoma, Dodoma, Tanzania

Contribution: Conceptualization, Writing - original draft, Methodology, Software, Data curation, Writing - review & editing, ​Investigation, Formal analysis, Validation

Search for more papers by this author
B. Prabhakar Reddy

B. Prabhakar Reddy

Department of Mathematics and Statistics, CNMS, the University of Dodoma, Dodoma, Tanzania

Contribution: ​Investigation, Writing - original draft, Writing - review & editing, Visualization, Methodology, Supervision, Resources

Search for more papers by this author
P. M. Matao

Corresponding Author

P. M. Matao

Department of Mathematics and Statistics, CNMS, the University of Dodoma, Dodoma, Tanzania

Correspondence: P. M. Matao ([email protected])

Contribution: Conceptualization, ​Investigation, Writing - original draft, Writing - review & editing, Methodology, Formal analysis, Supervision, Project administration, Visualization, Resources

Search for more papers by this author
First published: 05 April 2025

Funding: The authors received no specific funding for this work.

ABSTRACT

This study investigated the dissipative effects on time-dependent Casson nanofluid motion over a cone, considering variable heat source/absorption and higher-order reacting species. Water ethylene glycol was employed as the Casson base fluid. The dimensional PDEs were transformed into dimensionless PDEs by fitting non-dimensional parameters and solved using an effectual Galerkin finite element method (GFEM). The impact of physical parameters on momentum, energy, and concentration profiles is analyzed via graphical representations. The wall friction, thermal, and solutal transport rates are tabularly detailed. It was detected that increasing the absorbency parameter, Eckert number, thermal radiation, and thermal generation improves fluid velocity. Conversely, intensifying the magnetic field, Prandtl number and inclination angle reduces fluid velocity. The nanofluid temperature declines with a mounted Prandtl number and nanoparticle volume fraction, and the opposite effect is perceived with increased Eckert, Dufour, and Soret numbers. Wall friction intensifies with rising porosity, magnetic field strength, Casson parameter, and diffusive parameters, while it diminishes with higher nanoparticle volume fraction. The findings distinctly indicate that Ag WEG $$ Ag- WEG $$ nanofluid exhibits superior effectiveness in enhancing thermal and mass exchanges compared to Al 2 O 3 WEG $$ {Al}_2{O}_3- WEG $$ nanofluid. Furthermore, a comparative analysis agrees with earlier findings. This current model problem finds application across various scientific, engineering, and technological domains, including energy production, space exploration, food preservation, agricultural product manufacturing, materials processing, astrophysical phenomena, biomedical procedures, and enhanced oil recovery.

1 Introduction

Nanotechnology has fascinated inventors and investigators for its substantial use in industrial disciplines. For instance, in the pharmacological field, cancer patients are cured by nano-liquid-based machinists who embrace diverse energies and medications. Cooling and heating methods such as lessening receivable heat from electronic processing units, regulating the temperature of atomic reactors, regulating vehicle radiators, and managing heat flow in thermal valves employ nanofluids. These vital properties, together with numerous manufacturing and domestic applications, have captivated researchers and engineers in recent days. Nanofluids engrossed in pure fluids have a propensity to uplift their heat efficiency. A nanofluid comprises elements with nano-scale-sized particles. This impression originated with Choi [1] when he induced nanoelements in a pure fluid, which was called “nanofluid.” The nanoparticles comprise metals, metallic oxides, carbon nanotubes, and carbides. Nanoparticles have noteworthy industrial applications like sunblock for cars that are extra repellant to radiant emissions, car bumpers that have lesser weight, artificial bones that are stronger than nauseating sartorial, and improved sturdiness of several sports items such as balls [2-4].

A Casson fluid is a non-Newtonian fluid in which shear stress is nonlinearly related to velocity gradients. The stress–strain relationship in the Casson fluid is guided by a nonlinear Casson constitutive equation originally formulated by Casson [5]. This fluid has many practical applications and is a critical non-Newtonian fluid that is widely used in fields such as biomechanics, plastics, and metallurgy. Its significance spans food processing, metallurgy, drilling, and biotechnology, which has gathered momentous curiosity from investigators [6-8]. The Casson fluid model is predominantly effective in mathematically representing the property of fluids with non-zero plastic dynamic viscidness, relating it to a near estimation of blood. This model is also pertinent in the exploration of fluids encompassing plasma and proteins and in numerous products such as coal-water assortments, coats, artificial greases, and biotic fluids like tomato sauce, honey, potage, gelatine, and body fluid. For its highly practical applications, studies [9-15] explored further on the properties of MHD Casson fluids, reporting significant results. Reddy et al. [16] analytically analyzed the entropy study of two-phase radiative nanofluid transport in a bioconvective non-Newtonian model via HAM. Shamshuddin et al. [17] explored the impact of ternary nanofluid integration under thermal physical influences with the Cattaneo-Christov model.

The fluid movement past/over a cone has established much courtesy owing to countless applications concerning thermal transmission. It comes across several industrial uses, in addition to countless natural conditions such as medication and biomedical systems, energy preservation systems, aerospace sciences, geological mechanisms, cosmological physics, micro-inverter chips, interstellar technology, the regulatory mechanism of machine oil, and atomic energy control systems. In recent explorations, natural convective flows over a cone are the most exciting disciplines deliberated by investigators under the dynamics of fluid motion. Merk and Prins [18, 19] exhibited the overall connections for analogous results on a procedure occurring at an immobile temperature and explained the results obtained. Bilal et al. [20] considered the influence of numerous physical conditions on Maxwell nanofluid transport through a gyrating cone. Hussain et al. [21] outlined the thermal efficiency of multiple-based nanofluids across a modified cone. Hussain et al. [22] examined the enhancement of heat transport on a 3D gyrating cone of a hybrid nanofluid. Sayed and Fathy [23] reported the thermal transport consequences of a nanofluid past an upright cone, where the solution of the flow was obtained numerically. Additional relevant results on the MHD fluid flow via a cone with different flow geometries can be captured in references [24-27].

The order reactive species is termed to be one and consistent when its rate is directly correlated with reactant concentration and it proceeds entirely within a single-phase volume. Furthermore, the diffusion of species generated or consumed as a result of the reaction may be proportionate to the change in the nth-order concentration amid the interface and surrounding fluid, where n is a positive integer. Meteorological phenomena, haystack burning, milk spray drying, suspended catalyst reactors, and thermal cooling structures are all impacted by convective thermal and solutal transport with chemical reactions [28-30]. With such a significant impact on engineering and industrial applications, Swain et al. [31] outlined the outcomes of higher-order reacting species and slip frontier settings on hydro-magnetic nanofluid transport via numerical solutions. Seth et al. [32] simulated the model of conjugate heating effects on hydro-magnetic free convective Casson fluid movement with nth-order reactive species in permeable media. The upshots of MHD radiative flow nanofluid with inconstant chemical reaction past an upright plate were reported by Rahman and Uddin [33]. Studies [34-39] included the consequences of chemical reactions on MHD nanofluid transport.

Mass and energy fluxes arise as direct consequences of gradients in temperature and concentration, respectively. Thermo-diffusion, commonly referred to as the Soret effect, is driven by temperature gradients, whereas the diffusion-thermo phenomenon, known as the Dufour effect, is instigated by concentration gradients. These effects gain prominence under flow conditions where density variations are present. Moreover, convection plays a pivotal role in heat transfer processes, significantly impacting practical applications like regulating systems, metallurgical mechanisms in the steel factory, and the operation of heat exchangers. Vijay and Sharma [40] explained the cross-diffusive upshots on the hydromagnetic transport of a hybrid nanofluid across a slowing spinning disk, where solutions of the flow were numerically obtained using the BVP Midrich scheme. Ullah et al. [41] performed a comprehensive analysis of entropy generation in the context of double-diffusive dissipative magnetohydrodynamic (MHD) Jeffrey nanofluid transport, incorporating the effects of non-linear slip conditions and activation energy. Sathyanarayana and Goud [42] examined the influence of double-diffusive aspects on reacting hydromagnetic nanofluid transport across an upright cone entrenched in an absorbent material via the Runge–Kutta approach allied with the shooting technique. Ilango and Lakshminarayana [43] discussed the consequences of cross-diffusive effects on dissipative hydro-magnetic Casson nanofluid past a widening surface in a permeable material subjected to an induced magnetic domain. Further detailed scrutiny on cross-diffusive effects can be found in references [44-47].

The influence of viscous dissipation is paramount in areas characterized by high current densities and flows involving either high velocities or elevated viscosities. Viscous dissipation becomes increasingly significant in such contexts, as internal friction produces considerable heat, particularly in regions subjected to intense shear. The effects of dissipation are crucial in the broad scope of technological and engineering practices, embracing aerospace machines, jet propulsion structures, space dynamics, automotive engines, and other advanced mechanical systems. Zainal et al. [48] provided a thorough of dissipative consequences on the transport behavior of MHD hybrid nanofluids past an exponentially widening or contracting sheet. Hazarika et al. [49] scrutinized the thermophoresis upshot on a dissipative motion of a hydromagnetic fluid across a continuous surface in a permeable domain employing Cu , Fe 3 O 4 , and Ag $$ Cu,{Fe}_3{O}_4,\mathrm{and}\ Ag $$ nanoparticles. Muntazir et al. [50] discussed the radiative and dissipative hydromagnetic nanofluid transport across an absorbent elongating sheet considering Cu and Al 2 O 3 $$ Cu\ \mathrm{and}\ {Al}_2{O}_3 $$ nanoparticles. Where the transformed equations of the problem were cracked through a semi-analytical tactic utilizing the differential transformation method. Prabhakar Reddy and Sademaki [9] conducted a study of dissipative and radiative consequences on a hydromagnetic Casson fluid transport past an upright porous plate via finite difference method (FDM). The studies [51-55] merit recognition for their significant contributions to the understanding and explanation of dissipation effects in MHD fluid streams.

The investigation of inconsistent heat source and absorption in hydro-magnetic fluid transport is a fundamental insight into the thermal behavior of conducting fluids under the influence of magnetic domains. Non-uniform heat source or absorption pertains to spatial distinctions in thermal generation or consumption within the fluid, a phenomenon that is critically important across numerous manufacturing and engineering applications. This variation significantly enhances temperature distribution and transport properties, playing a pivotal role in systems where precise thermal regulation is essential. Applications span from fission reactors and astrophysical phenomena to electronic device regulators, metal processing, polymer synthesis, and other technologies reliant on controlled thermal transmission. Mythili et al. [56] presented the upshots of higher-order reactive species and nonuniform heat source/absorption on hydromagnetic Casson fluid transport across an upright cone via the FDM. Prabhavathi et al. [57] scrutinized the heat and solutal transport features of a hydro-magnetic nanofluid past an upright cone entrenched in a permeable membrane, incorporating variable heat source and absorption via a numerical approach. Konda et al. [58] explored the upshot of a non-uniform heat generation and consumption on the hydro-magnetic transport of Williamson nanofluids with thermal effects in a penetrable medium. Shi et al. [59] inspected the radiative characteristics of hydro-magnetic Maxwell-nanofluid transport, considering non-uniform heat generation and consumption. Additional significant findings regarding the consequences of non-uniform heat source/sink on nanofluid flow are documented in studies [60-62].

Upon reviewing the existing literature, it is evident that no study has addressed the combined consequences of energy dissipation, non-uniform heat source/sink, double diffusion, and surface porosity on the unsteady MHD transport of a reacting Casson nanofluid across a cone. To fill this research gap, the current scrutiny aims to investigate the dissipative behavior of time-dependent MHD Casson nanofluid transport past a cone, incorporating the influence of a non-uniform heat source/sink and higher-order reactions of chemical species. The constructed structure of highly coupled nonlinear partial differential equations (PDEs) has been computationally executed using the GFEM. This computational technique extends the prior work of Mythili et al. [56] by integrating novel consequences of magnetic intensity, radiative properties, and the unique characteristics of nanofluids. The proposed model has broad applicability across various scientific, engineering, and technological domains, including energy production, space exploration, food preservation, agricultural product manufacturing, materials processing, astrophysical phenomena, biomedical procedures, and enhanced oil recovery. This study addresses the following research questions:
  1. How does viscous dissipation and radiation influence nanofluid momentum, temperature, and concentration under the presence of an applied magnetic field?
  2. What are the effects of combined physical flow parameters on the wall friction coefficient rate, thermal and mass transfer performance of Casson nanofluids under the subjected boundary constraint in the presence of an external magnetic field?
  3. Does the inclusion of nanoparticles enhance fluid thermal efficiency?
  4. How can the finite element method (FEM) be utilized to model and predict the behavior of Casson nanofluid under realistic industrial and practical conditions?

2 Model Description

Let the contemplation of two-dimensional, time-dependent, natural convective transport of an incompressible flow Casson nanofluid over a cone with a half angle drenched with absorbent material is depicted in Figure 1. The fluid motion is persuaded by a consistent crosswise magnetic field of strength B 0 $$ {B}_0 $$ . External voltage is zero and the magnetic Reynolds number is considered very low, rendering the induced magnetic field and Hall effects insignificantly consequential. In the fluid flow regime, the viscous dissipation effects are taken into account. The cone is oriented with its axis of symmetry directed vertically, and its vertex precisely located at the origin of the coordinate system. The x-axis is defined along the cone's surface, with the y-axis oriented orthogonally. The temperature and concentration at the wall y = 0 $$ {y}^{\prime }=0 $$ at any time τ , $$ {\tau}^{\prime }, $$ denoted as T w $$ {T}_w^{\prime } $$ and C w $$ {C}_w^{\prime } $$ at any instantaneous moment are presumed to surpass their corresponding ambient values, T $$ {T}_{\infty}^{\prime } $$ and C $$ {C}_{\infty}^{\prime } $$ in the vicinity of the cone. The permeability of the material is presumed to be constant. The flow is exposed to a non-uniform heat generation/absorption, higher reacting order of species, and the cross-diffusion consequences. The rheological equation of the state for the isotropic and incompressible flow of a Casson fluid is given by
τ = τ 0 + μ α * $$ \tau ={\tau}_0+{\mu \alpha}^{\ast } $$
or
τ ij = 2 μ B + p y 2 π e ij , π > π c 2 μ B + p y 2 π c e ij , π < π c $$ {\tau}_{ij}=\left\{\begin{array}{l}2\left({\mu}_B+\frac{p_y}{\sqrt{2\pi }}\right){e}_{ij},\kern1.25em \pi >{\pi}_c\\ {}2\left({\mu}_B+\frac{p_y}{\sqrt{2{\pi}_c}}\right){e}_{ij},\kern0.75em \pi <{\pi}_c\end{array}\right. $$
where τ $$ \tau $$ is the shear stress, τ 0 $$ {\tau}_0 $$ is the Casson yield stress, μ $$ \mu $$ is the dynamic viscosity, α * $$ {\alpha}^{\ast } $$ is the shear rate, π = e ij e ij $$ \pi ={e}_{ij}{e}_{ij} $$ and e ij $$ {e}_{ij} $$ is the ( i , j ) th $$ {\left(i,j\right)}^{th} $$ component of the deformation rate, π $$ \pi $$ is the product of the component of deformation rate with itself, π c $$ {\pi}_c $$ is the critical value of this product based on the non-Newtonian fluid model, μ B $$ {\mu}_B $$ is the plastic dynamic viscosity of the non-Newtonian fluid and p y $$ {p}_y $$ is the yield stress of the fluid.
e ij = 1 2 v i x j + v j y i $$ {e}_{ij}=\frac{1}{2}\left(\frac{\partial {v}_i}{\partial {x}_j}+\frac{\partial {v}_j}{\partial {y}_i}\right) $$
Details are in the caption following the image
Geometrical insight of model flow.
The basic 2D equations governing the conservation of continuity, momentum, energy, and mass diffusion for Casson fluid using the above assumptions along with Boussinesq's approximations can be written following Mythili et al. [56] as:
u r x + v r y = 0 $$ \frac{\partial \left({u}^{\prime }{r}^{\bullet}\right)}{\partial {x}^{\prime }}+\frac{\partial \left({v}^{\prime }{r}^{\bullet}\right)}{\partial {y}^{\prime }}=0 $$ (1)
u τ + u u x + v u y = μ nf 1 + 1 α 2 u y 2 σ nf ρ nf B 0 2 u + μ nf k u + g ρ β θ nf T T ( cos β ) + g ρ β C nf C C ( cos β ) $$ {\displaystyle \begin{array}{ll}\frac{\partial {u}^{\prime }}{\partial {\tau}^{\prime }}& +{u}^{\prime}\frac{\partial {u}^{\prime }}{\partial {x}^{\prime }}+{v}^{\prime}\frac{\partial {u}^{\prime }}{\partial {y}^{\prime }}={\mu}_{nf}\left(1+\frac{1}{\alpha}\right)\frac{\partial^2{u}^{\prime }}{\partial {y^{\prime}}^2}-\frac{\sigma_{nf}}{\rho_{nf}}{B_0}^2{u}^{\prime}\\ {}& \kern1em +\frac{\mu_{nf}}{k^{\prime }}{u}^{\prime }+g{\left({\rho \beta}^{\theta}\right)}_{nf}\left({T}^{\prime }-{T}_{\infty}^{\prime}\right)\left(\cos \beta \right)\\ {}& \kern1em +g{\left({\rho \beta}^C\right)}_{nf}\left({C}^{\prime }-{C}_{\infty}^{\prime}\right)\left(\cos \beta \right)\end{array}} $$ (2)
T τ + u T x + v T y = k nf ρ c p nf 2 T y 2 + q ρ c p nf + D m k T c s c p 2 C y 2 + μ nf ρ c p nf 1 + 1 α u y 2 1 ( ρ c ) nf q r y + μ nf k ρ nf 1 + 1 α u 2 $$ {\displaystyle \begin{array}{ll}\frac{\partial {T}^{\prime }}{\partial {\tau}^{\prime }}& +{u}^{\prime}\frac{\partial {T}^{\prime }}{\partial {x}^{\prime }}+{v}^{\prime}\frac{\partial {T}^{\prime }}{\partial {y}^{\prime }}=\frac{k_{nf}}{{\left(\rho {c}_p\right)}_{nf}}\frac{\partial^2{T}^{\prime }}{\partial {y^{\prime}}^2}+\frac{q^{{\prime\prime\prime} }}{{\left(\rho {c}_p\right)}_{nf}}\\ {}& \kern1em +\frac{D_m{k}_T}{c_s{c}_p}\frac{\partial^2{C}^{\prime }}{\partial {y^{\prime}}^2}+\frac{\mu_{nf}}{{\left(\rho {c}_p\right)}_{nf}}\left(1+\frac{1}{\alpha}\right){\left(\frac{\partial {u}^{\prime }}{\partial {y}^{\prime }}\right)}^2\\ {}& \kern1em -\frac{1}{{\left(\rho c\right)}_{nf}}\left(\frac{\partial {q}_r}{\partial {y}^{\prime }}\right)+\frac{\mu_{nf}}{k{\rho}_{nf}}\left(1+\frac{1}{\alpha}\right){u^{\prime}}^2\end{array}} $$ (3)
C τ + u C x + v C y = D nf 2 C y 2 k 1 * C C n + D m k T T m 2 T y 2 $$ {\displaystyle \begin{array}{ll}\frac{\partial {C}^{\prime }}{\partial {\tau}^{\prime }}& +{u}^{\prime}\frac{\partial {C}^{\prime }}{\partial {x}^{\prime }}+{v}^{\prime}\frac{\partial {C}^{\prime }}{\partial {y}^{\prime }}={D}_{nf}\frac{\partial^2{C}^{\prime }}{\partial {y^{\prime}}^2}\\ {}& \kern1em -{k}_1^{\ast }{\left({C}^{\prime }-{C}_{\infty}^{\prime}\right)}^n+\frac{D_m{k}_T}{T_m}\frac{\partial^2{T}^{\prime }}{\partial {y^{\prime}}^2}\end{array}} $$ (4)
The suitable boundary conditions for velocity, temperature, and concentration profiles based on the physical model assumptions displayed in Figure 1 and Equations (14) are comparable to those outlined by Mythili et al. [56],
τ 0 ; u = 0 , v = 0 , T = T , C = C for all x and y τ > 0 ; u = 0 , v = 0 , T = T w , C = C w at y = 0 u = 0 , T = T , C = C at x = 0 u 0 , T T , C C as y = 0 $$ \left.\begin{array}{l}{\tau}^{\prime}\le 0;{u}^{\prime }=0,{v}^{\prime }=0,{T}^{\prime }={T}_{\infty}^{\prime },{C}^{\prime }={C}_{\infty}^{\prime}\kern0.75em \mathrm{for}\ \mathrm{all}\kern0.5em {x}^{\prime }\ \mathrm{and}\kern0.5em {y}^{\prime}\\ {}{\tau}^{\prime }>0;\kern0.5em {u}^{\prime }=0,{v}^{\prime }=0,{T}^{\prime }={T}_w^{\prime },{C}^{\prime }={C}_w^{\prime}\kern0.75em \mathrm{at}\kern1.75em {y}^{\prime }=0\\ {}\kern2.75em {u}^{\prime }=0,\kern0.75em {T}^{\prime }={T}_{\infty}^{\prime },\kern0.75em {C}^{\prime }={C}_{\infty}^{\prime}\kern2.5em \mathrm{at}\kern2em {x}^{\prime }=0\\ {}\kern2.75em {u}^{\prime}\to 0,{T}^{\prime}\to {T}_{\infty}^{\prime },{C}^{\prime}\to {C}_{\infty}^{\prime}\kern2.25em \mathrm{as}\kern1.75em {y}^{\prime }=0\end{array}\right\} $$ (5)
The thermal radiative flux q r , $$ {q}_r, $$ ascertained using the Rosseland approximation, and is derived by
q r = 4 σ s 3 k T 4 y $$ {q}_r=-\frac{4{\sigma}^s}{3{k}^{\bullet }}\frac{\partial {T}^4}{\partial y} $$ (6)
where σ s $$ {\sigma}^s $$ is the Stefan-Boltzmann constant and k $$ {k}^{\bullet } $$ is the captivation constant. Allowing minor temperature differences concerning the fluid temperature T $$ T $$ and free-stream temperature T $$ {T}_{\infty } $$ to expanded T 4 $$ {T}^4 $$ as the linear function fetching by the Taylor series about T $$ {T}_{\infty } $$ and then ignoring from 2nd order terms onwards results in:
T 4 4 T 3 T 3 T 4 $$ {T^{\prime}}^4\cong 4{T^{\prime}}_{\infty}^3T-3{T^{\prime}}_{\infty}^4 $$ (7)
Using Equations (6 and 7), Equation (3) yields to:
T τ + u T x + v T y = k nf ρ c p nf 2 T y 2 + q ρ c p nf + D m k T c s c p 2 C y 2 + μ nf ρ c p nf 1 + 1 α u y 2 1 ρ c p f 16 σ s T 3 3 k 2 T y 2 + μ nf k ρ nf 1 + 1 α u 2 $$ {\displaystyle \begin{array}{ll}\frac{\partial {T}^{\prime }}{\partial {\tau}^{\prime }}& +{u}^{\prime}\frac{\partial {T}^{\prime }}{\partial {x}^{\prime }}+{v}^{\prime}\frac{\partial {T}^{\prime }}{\partial {y}^{\prime }}=\frac{k_{nf}}{{\left(\rho {c}_p\right)}_{nf}}\frac{\partial^2{T}^{\prime }}{\partial {y^{\prime}}^2}+\frac{q^{{\prime\prime\prime} }}{{\left(\rho {c}_p\right)}_{nf}}\\ {}& \kern1em +\frac{D_m{k}_T}{c_s{c}_p}\frac{\partial^2{C}^{\prime }}{\partial {y^{\prime}}^2}+\frac{\mu_{nf}}{{\left(\rho {c}_p\right)}_{nf}}\left(1+\frac{1}{\alpha}\right){\left(\frac{\partial {u}^{\prime }}{\partial {y}^{\prime }}\right)}^2\\ {}& \kern1em -\frac{1}{{\left(\rho {c}_p\right)}_f}\left(\frac{16{\sigma}^s{T^{\prime}}^3}{3{k}^{\bullet }}\frac{\partial^2{T}^{\prime }}{\partial {y^{\prime}}^2}\right)+\frac{\mu_{nf}}{k{\rho}_{nf}}\left(1+\frac{1}{\alpha}\right){u^{\prime}}^2\end{array}} $$ (8)
The non-uniform heat sink/source q $$ {q}^{{\prime\prime\prime} } $$ is modeled as Pal and Mandal [63].
q = k * Gr T 1 2 l * 2 A * l * ν * Gr T 1 2 T T u + B * T T $$ {q}^{{\prime\prime\prime} }=\frac{k^{\ast }{\left({Gr}_T\right)}^{\frac{1}{2}}}{l^{\ast 2}}\left(\frac{A^{\ast }{l}^{\ast }}{\nu^{\ast }{\left({Gr}_T\right)}^{\frac{1}{2}}}\left({T}^{\prime }-{T}_{\infty}^{\prime}\right)u+{B}^{\ast}\left({T}^{\prime }-{T}_{\infty}^{\prime}\right)\right) $$

According to Ramadhyani et al. [64], the heat flux q $$ {q}^{{\prime\prime\prime} } $$ can be computed from the electrical power P $$ P $$ and thermal transport surface area S : q = P / S $$ {S}^{\bullet }:{q}^{{\prime\prime\prime} }=P/{S}^{\bullet } $$ It should be pointed out that the instance A * > 0 ; B * > 0 $$ {A}^{\ast }>0;{B}^{\ast }>0 $$ matches to inner heat source and the occasion A * < 0 ; B * < 0 $$ {A}^{\ast }<0;{B}^{\ast }<0 $$ links to inner thermal consumption.

Where u $$ {u}^{\prime } $$ is the fluid velocity, T $$ {T}^{\prime } $$ is the temperature of the fluid, C $$ {C}^{\prime } $$ is the nanofluid concentration of nanofluid, ρ nf $$ {\rho}_{nf} $$ is the nanofluid density, μ nf $$ {\mu}_{nf} $$ is the dynamic viscosity of nanofluid, σ nf $$ {\sigma}_{nf} $$ is the electrical conductivity of nanofluids, β nf θ $$ {\beta}_{nf}^{\theta } $$ is the heat expansivity coefficient, β nf C $$ {\beta}_{nf}^C $$ is the coefficient of volumetric expansivity, g $$ g $$ is the gravitational acceleration, ρ c p nf $$ {\left(\rho {c}_p\right)}_{nf} $$ is the specific heat capacity of the nanofluid, k nf $$ {k}_{nf} $$ is the thermal conductivity of nanofluid, D nf $$ {D}_{nf} $$ is the mass diffusivity, γ k $$ {\gamma_k}^{\prime } $$ is the chemical reaction parameter, D m $$ {D}_m $$ is the chemical molecular diffusion, T m $$ {T}_m $$ is the fluid mean temperature, k T $$ {k}_T $$ is the thermal diffusion rate, β $$ \beta $$ is the cone apex semi-angle.

Acquaint with the non-dimension quantities and properties:
ξ = x L 1 , η = y L 1 G r t 1 4 , u = u L 1 ν f G r t 1 2 , v = v L 1 ν f G r t 1 5 , t = ν f τ G r t 1 2 L 1 2 , θ = T T T w T , C = C C C w C , G r t = g β θ f T w T L 1 3 ν f 2 , G r c = g β C f C w C L 1 3 ν f 2 , S c = ν f D nf , D u = k T D m C w C ν f c s c p T w T , k 1 = k 1 * C w C n 1 L 1 2 ν f G r t 1 2 , S r = k T D m T w T C w C , N 1 = G r c G r t E c = G r t 1 2 c p f T w T L 1 2 , S r = D m k T T w T v f T m C w C , P r = μ c p f k f , M = σ f B 0 2 L 1 2 ρ nf G r t 1 2 , K A = k G r t 1 2 L 1 2 , R = r L 1 ; r = x 1 sin β $$ \left.\begin{array}{l}\xi =\frac{x^{\prime }}{L_1},\eta =\frac{y^{\prime }}{L_1}{\left(G{r}_t\right)}^{\frac{1}{4}},u=\frac{u^{\prime }{L}_1}{\nu_f}{\left(G{r}_t\right)}^{-\frac{1}{2}},v=\frac{v^{\prime }{L}_1}{\nu_f}{\left(G{r}_t\right)}^{-\frac{1}{5}},t=\frac{\nu_f{\tau}^{\prime }{\left(G{r}_t\right)}^{\frac{1}{2}}}{{L_1}^2},\theta =\frac{\left({T}^{\prime }-{T}_{\infty}^{\prime}\right)}{\left({T}_w^{\prime }-{T}_{\infty}^{\prime}\right)},\\ {}C=\frac{\left({C}^{\prime }-{C}_{\infty}^{\prime}\right)}{\left({C}_w^{\prime }-{C}_{\infty}^{\prime}\right)},G{r}_t=\frac{{\left(g{\beta}^{\theta}\right)}_f\left({T}_w^{\prime }-{T}_{\infty}^{\prime}\right){L_1}^3}{{\nu_f}^2},G{r}_c=\frac{{\left(g{\beta}^C\right)}_f\left({C}_w^{\prime }-{C}_{\infty}^{\prime}\right){L_1}^3}{{\nu_f}^2},{S}_c=\frac{\nu_f}{D_{nf}},\\ {}{D}_u=\frac{k_T{D}_m\left({C}_w^{\prime }-{C}_{\infty}^{\prime}\right)}{\nu_f{c}_s{c}_p\left({T}_w^{\prime }-{T}_{\infty}^{\prime}\right)},{k}_1=\frac{k_1^{\ast }{\left({C}_w^{\prime }-{C}_{\infty}^{\prime}\right)}^{n-1}{L_1}^2}{\nu_f}{\left(G{r}_t\right)}^{-\frac{1}{2}},{S}_r=\frac{k_T{D}_m\left({T}_w^{\prime }-{T}_{\infty}^{\prime}\right)}{\left({C}_w^{\prime }-{C}_{\infty}^{\prime}\right)},{N}_1=\frac{G{r}_c}{G{r}_t}\ \\ {}{E}_c=\frac{{\left(G{r}_t\right)}^{\frac{1}{2}}}{{\left({c}_p\right)}_f\left({T}_w^{\prime }-{T}_{\infty}^{\prime}\right){L_1}^2},{S}_r=\frac{D_m{k}_T\left({T}_w^{\prime }-{T}_{\infty}^{\prime}\right)}{v_f{T}_m\left({C}_w^{\prime }-{C}_{\infty}^{\prime}\right)},{P}_r=\frac{{\left(\mu {c}_p\right)}_f}{k_f},M=\frac{\sigma_f{B_0}^2{L_1}^2}{\rho_{nf}}{\left(G{r}_t\right)}^{-\frac{1}{2}},\\ {}\ {K}_A=\frac{k{\left(G{r}_t\right)}^{-\frac{1}{2}}}{{L_1}^2},{R}^{\bullet }=\frac{r^{\bullet }}{L_1};{r}^{\bullet }={x}_1\sin \beta \end{array}\right\} $$ (9)
where u , v , θ , t , C , L 1 , Gr t , Gr c , P r , R , N 1 , r , M , E c , α , S c , S r , k 1 , K A $$ u,v,\theta, t,C,{L}_1,{Gr}_t,{Gr}_c,{P}_r,{R}^{\bullet },{N}_1,{r}^{\bullet },M,{E}_c,\alpha, {S}_c,{S}_r,{k}_1,{K}_A $$ are dimensionless velocities in x and y $$ {x}^{\prime }\ \mathrm{and}\ {y}^{\prime } $$ directions, dimensionless temperature, dimensionless time, dimensionless concentration, thermal Grashof number, solutal Grashof number, Prandtl number, dimensionless local radius, dimensionless ratio due to buoyance force, local radius of the cone, Magnetic parameter, energy dissipation, Casson parameter, Schmidt number, thermos-diffusion parameter, reactive species parameter, and porosity parameter.
Utilizing the dimensionless parameters, the governing equations reduce to:
u ξ + v η + u ξ = 0 $$ \frac{\partial u}{\partial \xi }+\frac{\partial v}{\partial \eta }+\frac{u}{\xi }=0 $$ (10)
u t + u u ξ + v u η = 1 J 0 1 + 1 α 2 u y 2 J 2 M + 1 K A u + J 1 θ ( cos β ) + J 2 N 1 C ( cos β ) $$ {\displaystyle \begin{array}{ll}\frac{\partial u}{\partial t}& +u\frac{\partial u}{\partial \xi }+v\frac{\partial u}{\partial \eta }=\frac{1}{J_0}\left(1+\frac{1}{\alpha}\right)\frac{\partial^2u}{\partial {y}^2}-{J}_2\left(M+\frac{1}{K_A}\right)u\\ {}& \kern1em +{J}_1\theta \left(\cos \beta \right)+{J}_2{N}_1C\left(\cos \beta \right)\end{array}} $$ (11)
T t + u T ξ + v T η = J 3 λ 1 1 P r 2 θ η 2 1 P r A * u + B * θ + J 4 E c 1 + 1 α u η 2 J 3 P r 1 + 4 3 N r θ + D u 2 ψ η 2 $$ {\displaystyle \begin{array}{ll}\frac{\partial T}{\partial t}& +u\frac{\partial T}{\partial \xi }+v\frac{\partial T}{\partial \eta }={J}_3\left[{\lambda}_1\frac{1}{P_r}\frac{\partial^2\theta }{\partial {\eta}^2}-\frac{1}{P_r}\left({A}^{\ast }u+{B}^{\ast}\theta \right)\right]\\ {}& \kern1em +{J}_4{E}_c\left(1+\frac{1}{\alpha}\right){\left(\frac{\partial u}{\partial \eta}\right)}^2\\ {}& \kern1em -\frac{J_3}{P_r}\left(1+\frac{4}{3{N}_r}\right)\theta +{D}_u\frac{\partial^2\psi }{\partial {\eta}^2}\end{array}} $$ (12)
C t + u C ξ + v C η = 1 S c 2 C η 2 k 1 C n + S r 2 θ η 2 $$ \frac{\partial C}{\partial t}+u\frac{\partial C}{\partial \xi }+v\frac{\partial C}{\partial \eta }=\frac{1}{S_c}\left(\frac{\partial^2C}{\partial {\eta}^2}\right)-{k}_1{C}^n+{S}_r\left(\frac{\partial^2\theta }{\partial {\eta}^2}\right) $$ (13)

And the variables appearing in Equations (10-13) are demarcated as follows in Table 1.

TABLE 1. Nanofluid variables used in the flow.
S/N Variable description
1. J 0 = ( 1 ϕ ) 2.5 ( 1 ϕ ) + ϕ ρ s ρ f $$ {J}_0={\left(1-\phi \right)}^{2.5}\left[\left(1-\phi \right)+\phi \left(\frac{\rho_s}{\rho_f}\right)\right] $$
2. J 1 = ( 1 ϕ ) + ϕ ρ β T s ρ β T f ( 1 ϕ ) + ϕ ρ s ρ f $$ {J}_1=\frac{\left(1-\phi \right)+\phi \left(\frac{{\left({\rho \beta}^T\right)}_s}{{\left({\rho \beta}^T\right)}_f}\right)}{\left(1-\phi \right)+\phi \left(\frac{\rho_s}{\rho_f}\right)} $$
3. J 2 = ( 1 ϕ ) + ϕ ρ β C s ρ β C f ( 1 ϕ ) + ϕ ρ s ρ f $$ {J}_2=\frac{\left(1-\phi \right)+\phi \left(\frac{{\left({\rho \beta}^C\right)}_s}{{\left({\rho \beta}^C\right)}_f}\right)}{\left(1-\phi \right)+\phi \left(\frac{\rho_s}{\rho_f}\right)} $$
4. J 3 = ( 1 ϕ ) + ϕ ρ β C s ρ β C f $$ {J}_3=\left(1-\phi \right)+\phi \left(\frac{{\left({\rho \beta}^C\right)}_s}{{\left({\rho \beta}^C\right)}_f}\right) $$
5. J 4 = 1 ( 1 ϕ ) + ϕ ρ c p s ρ c p f $$ {J}_4=\frac{1}{\left(1-\phi \right)+\phi \left(\frac{{\left(\rho {c}_p\right)}_s}{{\left(\rho {c}_p\right)}_f}\right)} $$
The transformed constraints are as follows:
t 0 : u = 0 , v = 0 , θ = 0 , C = 0 for all ξ , η t > 0 : u = 0 , v = 0 , θ = 1 , C = 1 at η = 0 u = 0 , θ = 0 , C = 0 at ξ = 0 u 0 , θ 0 , C 0 at η 0 $$ \left.\begin{array}{l}t\le 0:\kern0.5em u=0,\kern0.5em v=0,\kern0.5em \theta =0,\kern0.75em C=0\kern1em \mathrm{for}\ \mathrm{all}\kern0.75em \xi, \eta \\ {}t>0:\kern0.5em u=0,\kern0.5em v=0,\kern0.5em \theta =1,\kern1em C=1\kern1.75em \mathrm{at}\kern0.75em \eta =0\\ {}\kern3em u=0,\theta =0,\kern0.5em C=0\kern1.75em \mathrm{at}\kern0.75em \xi =0\\ {}\kern3em u\to 0,\theta \to 0,\kern0.5em C\to 0\kern0.5em \mathrm{at}\kern0.75em \eta \to 0\end{array}\right\} $$ (14)
The thermos-physical properties of nanofluids, including density, heat capacitance, dynamic viscosity, thermal conductivity, electrical conductivity and thermal expansion coefficient, significantly influence their heat transfer efficiency. These properties depend on nanoparticle type, concentration, shape, size, and base fluid characteristics. High-conductivity nanoparticles like Al 2 O 3 $$ {\mathrm{Al}}_2{\mathrm{O}}_3 $$ , ethylene glycol or Ag $$ \mathrm{Ag} $$ (Table 2) enhance thermal performance but may alter viscosity and flow behavior. Specific heat and density affect energy storage and thermal response. Understanding these properties is essential for optimizing industrial, energy, and cooling nanofluids. These thermo-physical relations of nanofluids are mathematically presented as:
  1. Density
ρ nf = ( 1 φ ) ρ f + φ ρ s $$ {\rho}_{nf}=\left(1-\varphi \right){\rho}_f+{\varphi \rho}_s $$
TABLE 2. Thermal-physical properties of nanofluids.
H 2 O $$ {\mathrm{H}}_2\mathrm{O} $$ Al 2 O 3 $$ {\mathrm{Al}}_2{\mathrm{O}}_3 $$ Ethylene glycol Ag $$ \mathrm{Ag} $$
c p kg 1 K 1 $$ {c}_p\left({kg}^{-1}{K}^{-1}\right) $$ 4179 765 2400.00 235.00
ρ kgm 3 $$ \rho \left({kgm}^{-3}\right) $$ 997.1 3970 1109.00 10,500
k Wm 1 K 1 $$ k\left({Wm}^{-1}{K}^{-1}\right) $$ 0.613 40 0.2580 429.00
β × 10 5 K 1 $$ \beta \times {10}^{-5}\left({K}^{-1}\right) $$ 21 0.850 1.8900
  1. Heat capacitance
ρ c p nf = ( 1 φ ) ρ c p f + φ ρ c p s $$ {\left(\rho {c}_p\right)}_{nf}=\left(1-\varphi \right){\left(\rho {c}_p\right)}_f+\varphi {\left(\rho {c}_p\right)}_s $$
  1. Dynamic viscosity
μ nf = μ f ( 1 φ ) 2.5 $$ {\mu}_{nf}=\frac{\mu_f}{{\left(1-\varphi \right)}^{2.5}} $$
  1. Thermal conductivity
λ 1 = k nf k f = k s + 2 k f 2 φ k f k s k s + 2 k f + φ k f k s $$ {\lambda}_1=\frac{k_{nf}}{k_f}=\frac{k_s+2{k}_f-2\varphi \left({k}_f-{k}_s\right)}{k_s+2{k}_f+\varphi \left({k}_f-{k}_s\right)} $$
  1. Electrical conductivity
σ nf = σ f 1 + 3 ( σ 1 ) φ ( σ + 2 ) ( σ 1 ) φ where σ = σ f σ s $$ {\sigma}_{nf}={\sigma}_f\left(1+\frac{3\left(\sigma -1\right)\varphi }{\left(\sigma +2\right)-\left(\sigma -1\right)\varphi}\right)\kern0.75em \mathrm{where}\ \sigma =\frac{\sigma_f}{\sigma_s} $$
  1. Thermal expansion coefficient
ρ β T nf = ( 1 φ ) ρ β T f + φ ρ β T s $$ {\left({\rho \beta}_T\right)}_{nf}=\left(1-\varphi \right){\left({\rho \beta}_T\right)}_f+\varphi {\left({\rho \beta}_T\right)}_s $$
where ϕ , ρ f , ρ s , σ f , σ s , μ f , ρ c p f , ρ c p s , k f , k s $$ \phi, {\rho}_f,{\rho}_s,{\sigma}_f,{\sigma}_s,{\mu}_f,{\left(\rho {c}_p\right)}_f,{\left(\rho {c}_p\right)}_s,{k}_f,{k}_s $$ is the nanoparticle volume fraction, density of the base fluid, density of the nanoparticles, electrical conductivity of the base fluid, electrical conductivity of the nanoparticles, viscosity of the base fluid, heat capacitance of the base fluid, heat capacitance of the nanoparticles, thermal conductivity of the base fluid, and thermal conductivity of the nanoparticles respectively.

With considerable dimensionless engineering curiosity analogous to heat and mass flow potentials encompass the wall shear stress C f $$ {C}_f $$ , heat and mass transfer rates in terms of Nusselt number N u $$ {N}_u $$ , and Sherwood number S h $$ {S}_h $$ , which are defined as:

C f = 1 + 1 α u η η = 0 , N u = θ η η = 0 and S h = C η η = 0 . $$ {C}_f=-\left(1+\frac{1}{\alpha}\right){\left(\frac{\partial u}{\partial \eta}\right)}_{\eta =0},{N}_u=-{\left(\frac{\partial \theta }{\partial \eta}\right)}_{\eta =0}\ \mathrm{and}\ {S}_h=-{\left(\frac{\partial C}{\partial \eta}\right)}_{\eta =0}. $$ (15)

3 Finite Element Computation Analysis

The dimensionless leading structure of equations with included values of parameter (11-13) imposed to initial and boundary constraints (14) is cracked numerically employing the GFEM since the analytical approaches cannot be applied to these highly coupled nonlinear equations. In diverse technical and engineering implications, the FEM is a potent computer tool that may be employed to execute ODEs and PDEs. Bathe [65] and Rao [66] provided a thorough analysis of the method. The FEM offers a profound advantage in the simulation of fluid flow phenomena, owing to its unparalleled versatility in managing intricate geometries and boundary conditions, coupled with its adeptness at dynamically refining the mesh in regions of high gradient. Furthermore, FEM excels in modeling nonlinearities inherent in complex fluid behaviors, such as turbulence and viscosity variations, while also proving indispensable for multiphysics simulations that necessitate the seamless integration of fluid dynamics with heat transfer, chemical reactions, and MHD effects. This method guarantees precise and convergent solutions, accommodates parallel computing for large-scale simulations, and enhances post-processing efficiency for in-depth analysis. In essence, FEM presents a robust, sophisticated, and highly accurate framework for the simulation of fluid dynamics across a vast array of complex and challenging scenarios. The scheme of the computational method is delineated as follows Bathe [65]:
  • Discretization of the finite element
  • Element equations formulation
  • Compiling of element equations
  • Introducing frontier settings
  • Solving the compiled equation

To use the GFEM for structures of coupled equations, the leading equations are multiplied by weight functions, and the integral over the element field is performed, resulting in the setting of the results to zero. This process leads to the following variational equations:

ω e W i u t + u u ξ + v u η 1 J 0 1 + 1 α 2 u y 2 + J 2 M + 1 K A u J 1 θ ( cos β ) J 2 N 1 C ( cos β ) dξdη = 0 $$ {\iint}_{\omega_e}{W}_i\left[\begin{array}{l}\frac{\partial u}{\partial t}+u\frac{\partial u}{\partial \xi }+v\frac{\partial u}{\partial \eta }-\frac{1}{J_0}\left(1+\frac{1}{\alpha}\right)\frac{\partial^2u}{\partial {y}^2}+{J}_2\left(M+\frac{1}{K_A}\right)u-{J}_1\theta \left(\cos \beta \right)\\ {}-{J}_2{N}_1C\left(\cos \beta \right)\end{array}\right] d\xi d\eta =0 $$ (16)
ω e W i θ t + u θ ξ + v θ η J 3 λ 1 1 P r 2 θ η 2 1 P r A * u + B * θ D u 2 C η 2 J 3 P r 1 + 4 3 N r θ J 4 E c 1 + 1 α u η 2 dξdη = 0 $$ {\iint}_{\omega_e}{W}_i\left[\begin{array}{l}\frac{\partial \theta }{\partial t}+u\frac{\partial \theta }{\partial \xi }+v\frac{\partial \theta }{\partial \eta }-{J}_3\left[{\lambda}_1\frac{1}{P_r}\frac{\partial^2\theta }{\partial {\eta}^2}-\frac{1}{P_r}\left({A}^{\ast }u+{B}^{\ast}\theta \right)\right]-{D}_u\frac{\partial^2C}{\partial {\eta}^2}\\ {}-\frac{J_3}{P_r}\left(1+\frac{4}{3{N}_r}\right)\theta -{J}_4{E}_c\left(1+\frac{1}{\alpha}\right){\left(\frac{\partial u}{\partial \eta}\right)}^2\end{array}\right] d\xi d\eta =0 $$ (17)
ω e W i C t + u C ξ + v C η 1 S c 2 C η 2 + k 1 C n S r 2 θ η 2 dξdη = 0 $$ {\iint}_{\omega_e}{W}_i\left[\frac{\partial C}{\partial t}+u\frac{\partial C}{\partial \xi }+v\frac{\partial C}{\partial \eta }-\frac{1}{S_c}\left(\frac{\partial^2C}{\partial {\eta}^2}\right)+{k}_1{C}^n-{S}_r\left(\frac{\partial^2\theta }{\partial {\eta}^2}\right)\right] d\xi d\eta =0 $$ (18)
The Galerkin finite element formulation may be attained from Equations (16-18) by replacing the finite element estimates of the arrangement:
u = j = 1 3 W j u j , θ = j = 1 3 W j θ j , C = j = 1 3 W j C j $$ u=\sum \limits_{j=1}^3{W}_j{u}_j,\kern0.5em \theta =\sum \limits_{j=1}^3{W}_j{\theta}_j,\kern0.5em C=\sum \limits_{j=1}^3{W}_j{C}_j $$ (19)
where W j ' s $$ {W_j}^{\hbox{'}}s $$ are the linear interpolation functions for a triangular element ω e $$ \left({\omega}_e\right) $$ .

It must be emphasized that the demand for continuity in field variables is significantly stricter in its current robust manifestations as depicted in Equations (16-18). To address these challenges, there is a preference for weak formulations. Undoubtedly, these weak formulations assist in lowering the continuity requirements for selected elements, thereby simplifying the approximation functions and allowing the utilization of easily constructed and implemented polynomials. Additionally, weak formulations inherently enforce natural frontier settings. Consequently, the chosen weak variations of the variational formulations are derived by substituting the unknowns in the weighted residual approach in Equations (16-18) with approximate trial solutions, represented by polynomial relationships in Equation (19), and performing integrations by parts over the element domain. The incorporation of borderline settings directly into the weak formulations follows the integrations by parts over the element space. The finite element model of the equations in matrix form is presented as:

K 11 K 12 K 13 K 21 K 22 K 23 K 31 K 32 K 33 u θ C + N 11 N 12 N 13 N 21 N 22 N 23 N 31 N 32 N 33 { u } { θ } { C } = F ( 1 ) F ( 2 ) F ( 3 ) $$ \left[\begin{array}{l}\left[{K}^{11}\right]\kern0.75em \left[{K}^{12}\right]\kern0.75em \left[{K}^{13}\right]\\ {}\left[{K}^{21}\right]\kern0.75em \left[{K}^{22}\right]\kern0.75em \left[{K}^{23}\right]\\ {}\left[{K}^{31}\right]\kern0.75em \left[{K}^{32}\right]\kern0.75em \left[{K}^{33}\right]\end{array}\right]\ \left[\begin{array}{l}\left\{\ {u}^{\prime}\right\}\\ {}\left\{\ {\theta}^{\prime}\right\}\\ {}\left\{\ {C}^{\prime}\right\}\end{array}\right]+\left[\begin{array}{l}\left[{N}^{11}\right]\kern0.75em \left[{N}^{12}\right]\kern0.75em \left[{N}^{13}\right]\\ {}\left[{N}^{21}\right]\kern0.75em \left[{N}^{22}\right]\kern0.75em \left[{N}^{23}\right]\\ {}\left[{N}^{31}\right]\kern0.75em \left[{N}^{32}\right]\kern0.75em \left[{N}^{33}\right]\end{array}\right]\ \left[\begin{array}{l}\left\{u\right\}\\ {}\left\{\theta \right\}\\ {}\left\{C\right\}\end{array}\right]=\left[\begin{array}{l}\left\{\ {F}^{(1)}\right\}\\ {}\left\{\ {F}^{(2)}\right\}\\ {}\left\{\ {F}^{(3)}\right\}\end{array}\right] $$
where the K th , N th , and F th $$ {K}^{th},{N}^{th},\mathrm{and}\ {F}^{th} $$ elements are defined as follows, while u , θ and C $$ {u}^{\prime },{\theta}^{\prime }\ \mathrm{and}\ {C}^{\prime } $$ represents the derivatives of u , θ and C $$ u,\theta\ \mathrm{and}\ C $$ with respect to time respectively.
K ij 11 = ω e W i W j dξdη , K ij 12 = 0 , K ij 13 = 0 $$ {K}_{ij}^{11}={\iint}_{\omega_e}{W}_i{W}_j d\xi d\eta, \kern0.5em {K}_{ij}^{12}=0,\kern0.5em {K}_{ij}^{13}=0 $$
K ij 21 = 0 , K ij 22 = ω e W i W j dξdη , K ij 23 = 0 $$ {K}_{ij}^{21}=0,\kern0.5em {K}_{ij}^{22}={\iint}_{\omega_e}{W}_i{W}_j d\xi d\eta, \kern0.5em {K}_{ij}^{23}=0 $$
K ij 31 = 0 , K ij 32 = 0 , K ij 33 = ω e W i W j dξdη $$ {K}_{ij}^{31}=0,\kern0.5em {K}_{ij}^{32}=0,\kern0.5em {K}_{ij}^{33}={\iint}_{\omega_e}{W}_i{W}_j d\xi d\eta $$
N ij 11 = ω e W i u W j ξ dξdη + ω e W i v W j η dxdξ 1 J 0 1 + 1 α ω e W i ξ W j ξ dξdη + J 2 M + 1 K A ω e W i W j dξdη $$ {\displaystyle \begin{array}{ll}{N}_{ij}^{11}& ={\iint}_{\omega_e}{W}_i{u}^{\bullet}\left(\frac{\partial {W}_j}{\partial \xi}\right) d\xi d\eta +{\iint}_{\omega_e}{W}_i{v}^{\bullet}\left(\frac{\partial {W}_j}{\partial \eta}\right) d xd\xi \\ {}& \kern1em -\frac{1}{J_0}\left(1+\frac{1}{\alpha}\right){\iint}_{\omega_e}\frac{\partial {W}_i}{\partial \xi}\frac{\partial {W}_j}{\partial \xi } d\xi d\eta \\ {}& \kern1em +{J}_2\left(M+\frac{1}{K_A}\right){\iint}_{\omega_e}{W}_i{W}_j d\xi d\eta \end{array}} $$
N ij 12 = J 1 ( cos β ) ω e W i W j dξdη $$ {N}_{ij}^{12}=-{J}_1\left(\cos \beta \right){\iint}_{\omega_e}{W}_i{W}_j d\xi d\eta $$
N ij 13 = J 2 N 1 C ( cos β ) ω e W i W j dξdη $$ {N}_{ij}^{13}={J}_2{N}_1C\left(\cos \beta \right){\iint}_{\omega_e}{W}_i{W}_j d\xi d\eta $$
N ij 21 = 0 , N ij 23 = 0 $$ {N}_{ij}^{21}=0,\kern0.5em {N}_{ij}^{23}=0 $$
N ij 22 = ω e W i u W j ξ dξdη + ω e W i v W j η dξdη J 3 λ 1 P r ω e W i η W j η dξdη 1 P r A * W j + B * W j ω e W i W j dξdη D u ω e W i η W j η dξdη J 4 E c 1 + 1 α ω e W j η 2 dξdη J 3 P r 1 + 4 3 N r ω e W i η W j η dξdη N ij 31 = 0 , N ij 32 = S r ω e W i ξ W j ξ dξdη $$ {\displaystyle \begin{array}{ll}{N}_{ij}^{22}& \kern1em ={\iint}_{\omega_e}{W}_i{u}^{\bullet}\left(\frac{\partial {W}_j}{\partial \xi}\right) d\xi d\eta +{\iint}_{\omega_e}{W}_i{v}^{\bullet}\left(\frac{\partial {W}_j}{\partial \eta}\right) d\xi d\eta \\ {}& \kern1em -{J}_3\left(\frac{\lambda_1}{P_r}{\iint}_{\omega_e}\frac{\partial {W}_i}{\partial \eta}\frac{\partial {W}_j}{\partial \eta } d\xi d\eta -\frac{1}{P_r}\left({A}^{\ast }{W}_j+{B}^{\ast }{W}_j\right){\iint}_{\omega_e}{W}_i{W}_j d\xi d\eta \right)\\ {}& \kern1em -{D}_u{\iint}_{\omega_e}\frac{\partial {W}_i}{\partial \eta}\frac{\partial {W}_j}{\partial \eta } d\xi d\eta -{J}_4{E}_c\left(1+\frac{1}{\alpha}\right){\iint}_{\omega_e}{\left(\frac{\partial {W}_j}{\partial \eta}\right)}^2 d\xi d\eta \\ {}& \kern1em -\frac{J_3}{P_r}\left(1+\frac{4}{3{N}_r}\right){\iint}_{\omega_e}\frac{\partial {W}_i}{\partial \eta}\frac{\partial {W}_j}{\partial \eta } d\xi d\eta \\ {}{N}_{ij}^{31}& =0,\kern0.5em {N}_{ij}^{32}={S}_r{\iint}_{\omega_e}\frac{\partial {W}_i}{\partial \xi}\frac{\partial {W}_j}{\partial \xi } d\xi d\eta \end{array}} $$
N ij 33 = w e W i u W j ξ dξdη + w e W i v W j η dxdξ + 1 S c w e W i ξ W j ξ dξdη $$ {N}_{ij}^{33}={\iint}_{w_e}{W}_i{u}^{\bullet}\left(\frac{\partial {W}_j}{\partial \xi}\right) d\xi d\eta +{\iint}_{w_e}{W}_i{v}^{\bullet}\left(\frac{\partial {W}_j}{\partial \eta}\right) d xd\xi +\frac{1}{S_c}{\iint}_{w_e}\frac{\partial {W}_i}{\partial \xi}\frac{\partial {W}_j}{\partial \xi } d\xi d\eta $$
F ( 1 ) = 0 , F ( 2 ) = 0 , F ( 3 ) = 0 . $$ {F}^{(1)}=0,\kern0.75em {F}^{(2)}=0,\kern0.75em {F}^{(3)}=0. $$

Here, u = j = 1 3 W j u j $$ {u}^{\bullet }=\sum \limits_{j=1}^3{W}_j{u^{\bullet}}_j $$ and v = j = 1 3 W j v j $$ {v}^{\bullet }=\sum \limits_{j=1}^3{W}_j{v^{\bullet}}_j $$ are the presumed relations that are introduced to simplify the structure of the equations. The Crank–Nicolson technique is used to approximate the time derivatives. Subsequently, the computational domain is split into 448 triangular components, each consisting of three nodes. As a result, there are 285 nodes in the entire domain. Meanwhile, 3 functions are computed at their respective node, then, the order of apiece component matrix is 9 × 9 . $$ 9\times 9. $$ Upon assembling the fundamental equations, an initial arrangement of 855 nonlinear coupled equations is formed. As mentioned before, and are used to linearize the structure of nonlinear coupled equations. An arrangement of 660 linear coupled equations of order is formed after the frontier settings are executed on the structure of linearized coupled equations. The Gauss–Seidel iteration scheme is employed to compute the linearized structure of equations. When the relative error for each variable amid successive iterations is calculated under the convergence requirement in a way that indicates ζ i n ζ i n 1 10 6 , $$ \sum \mid {\zeta}_i^n-{\zeta}_i^{n-1}\mid \le {10}^{-6}, $$ where ζ $$ \zeta $$ is the overall reliant variable u , θ and ϕ $$ u,\theta\ \mathrm{and}\ \phi $$ and n $$ n $$ is the number of successive iterations.

4 Validation of the Computation Procedures

To display the authentication and precision of the current arithmetical scheme, a comparative analysis of the rate of wall shear stress and mass transfer was computed by removing the nanofluid characteristics and letting M = 0 and E c = 0 $$ M=0\ \mathrm{and}\kern0.5em {E}_c=0 $$ in this present study with the preceding work of Mythili et al. [56] attained by FDM is depicted in Table 3. It is professed that there is an excellent pact between the outcomes. This describes the precision of the current numerical method.

TABLE 3. Numerical values of C f and S h $$ {C}_f\ \mathrm{and}\ {S}_h $$ when M = 0 , E c = 0 $$ M=0,{E}_c=0 $$ considering Casson fluid.
C f $$ {C}_f $$ S h $$ {S}_h $$
α $$ \alpha $$ K A $$ {K}_A $$ D u $$ {D}_u $$ S r $$ {S}_r $$ Mythili et al. [56] FEM results Mythili et al. [56] FEM results
0.5 1.0 0.02 0.3 0.19780 0.19781 0.47761 0.47762
1.0 0.24137 0.24139 0.47090 0.47093
1.5 0.26442 0.26445 0.46836 0.46839
2.0 0.24137 0.24136 0.47090 0.47092
3.0 0.25658 0.25659 0.46709 0.46706
4.0 0.26307 0.26309 0.46607 0.46608
0.03 0.18238 0.18238 0.15175 0.15170
0.05 0.18314 0.18317 0.14883 0.14887
0.15 0.18642 0.18641 0.15073 0.15076
0.4 0.18642 0.18644 0.15073 0.15079
1.2 0.18314 0.18315 0.14883 0.14884
2 0.18238 0.18239 0.15175 0.15176

5 Results and Discussion

To comprehend the insight of the problem at hand more thoroughly, we performed numerical computations for non-dimensional embedded flow profiles and the engineering quantities, using pertinent parameter values for two distinct Casson nanofluids. We presented twofold solutions for Ag WEG $$ Ag- WEG $$ and Al 2 O 3 WEG $$ {Al}_2{O}_3- WEG $$ nanofluids to achieve this. Shear stress, thermal, and solutal numbers are calculated numerically and exhibited in table format. To accomplish the numerical results of the current model problem, we adopted the following default values throughout the analysis: E c = 0.5 , K A = 1 , M = 2 , S r = 0.2 , α = 0.5 , A * = 2 , B * = 2.1 , P r = 0.71 , D u = 0.3 , t = 1 , β = π / 4 , N 1 = 1 , ϕ = 0.01 , n = 2 , S c = 0.2 , and k 1 = 0.2 . $$ {E}_c=0.5,{K}_A=1,M=2,{S}_r=0.2,\alpha =0.5,{A}^{\ast }=-2,{B}^{\ast }=-2.1,{P}_r=0.71,{D}_u=0.3,t=1,\beta =\pi /4,{N}_1=1,\phi =0.01,n=2,{S}_c=0.2,\mathrm{and}\ {k}_1=0.2. $$

Figure 2 exemplifies the impact of magnetic field intensity M $$ M $$ on the velocity of the nanofluid. It is observed that the velocity profile of both nanofluids exhibits a declining trend with increasing M $$ M $$ . This reduction arises from the imposition of a transverse magnetic field, which persuades a Lorentz force acting on the electrically conductive fluid. The resultant retarding force impedes the fluid's motion within the boundary layer, thereby decreasing its velocity. This scenario holds significant importance in the domain of motor rewinding technology, particularly in the precise regulation and control of machine processes across a diverse range of applications. Figure 3 showcases the impact of the absorbent parameter K A $$ {K}_A $$ on nanofluid velocity. It is perceived that nanofluid velocity increases as K A $$ {K}_A $$ is elevated. The underlying physics of this observation lies in the fact that the porosity parameter defines the proportion of the medium's volume composed of void spaces, devoid of solid material. As K A $$ {K}_A $$ increases, the impediment imposed by the absorbent matrix diminishes, owing to the presence of more void spaces through which the nanofluid can traverse. This reduction in resistance facilitates a freer flow of the nanofluid, thereby increasing fluid velocity. In manufacturing, controlled porosity enhances filtration efficiency, thermal insulation, and catalytic reactions, making it essential in sectors such as aerospace, biomedical engineering, and metallurgy. Figure 4 elucidates the variation in fluid velocity for diverse parameters of the Soret number S r $$ {S}_r $$ . It is seen that, as S r $$ {S}_r $$ increases, fluid velocity correspondingly rises. Physically, this behavior is attributed to enhanced solute diffusion within the boundary layer, driven by an intensified thermal gradient. The resulting increase in diffusion accelerates the fluid flow, thereby elevating its velocity. Figure 5 illustrates the variation in velocity distribution as a function of the viscous dissipation parameter E c $$ {E}_c $$ . It is detected that the fluid's momentum upsurges with the rise in E c $$ {E}_c $$ . Physically, the enhancement of E c $$ {E}_c $$ converts kinetic energy into internal energy, which, in turn, augments the fluid's velocity. Considerating and enhancing viscous dissipation is crucial for advancing energy efficiency, thermal regulation, and material processing in industrial systems. Figure 6 presents the influence of the Casson parameter α $$ \alpha $$ on fluid velocity. It is evident that increasing α $$ \alpha $$ reduces the fluid's velocity. This effect is attributed to the rise in the fluid's yield stress, which introduces greater opposition to the fluid movement. As α $$ \alpha $$ increases, the fluid requires a higher force to initiate movement, resulting to a thickening of the borderline stratum. This, in turn, diminishes the shear rate, ultimately resulting in a decrease in velocity. The stimulus of space and temperature-dependent thermal consumption A * < 0 ; B * < 0 $$ \left({A}^{\ast }<0;{B}^{\ast }<0\right) $$ and heat generation A * > 0 ; B * > 0 $$ \left({A}^{\ast }>0;{B}^{\ast }>0\right) $$ on the velocity curves are exhibited in Figures 7 and 8, respectively. It is evident from these depictions that snowballing the parameters of heat generation augment the velocity curves by generating heat within the fluid while lessening the values of heat absorption diminishes the fluid pace by making the fluid cooler. Figure 9 depicts the stimulus of the Prandtl number P r $$ {P}_r $$ on the velocity of the nanofluid. It is noted that as the P r $$ {P}_r $$ surges, the fluid's pace shrinks. This behavior can be explained by the reason that a higher P r $$ {P}_r $$ corresponds to greater fluid viscosity, which effectively thickens the fluid. The augmented viscosity dissipates kinetic energy, thereby reducing the fluid's speed. The Prandtl number balances momentum and thermal conductivity to regulate heat distribution in the MHD nanofluid flow. It is important for energy systems and cooling technologies. Figure 10 demonstrates the upshot of the semi-apex angle β $$ \beta $$ on fluid velocity. It is evident that increasing β $$ \beta $$ results in a reduction in fluid momentum. This behavior arises because the inclination angle modifies the influence of gravitational forces, increasing frictional resistance and reducing buoyancy-driven convection. These combined mechanisms slow down the fluid, yielding to a noticeable decrease in velocity as the inclination angle grows. The fluctuation of the ratio mass Grashoff number to the thermal Grashoff number on fluid velocity is illustrated in Figure 11. It is well concluded that fluid speed ascends when N 1 $$ {N}_1 $$ is raised. Physically, N 1 $$ {N}_1 $$ impacts fluid velocity by controlling the relative intensity of buoyancy forces due to concentration and temperature gradients. These buoyancy forces drive convection currents within the fluid, hence augmenting fluid velocity. Figure 12 explicates the fluid velocity variability for the sprouting radiation parameter N r . $$ {N}_r. $$ There is an enlargement noted in the distributions of the velocity when N r . $$ {N}_r. $$ values are progressed. This is due to kinetic energy within the fluid being enhanced causing the fluid to flow quickly.

Details are in the caption following the image
Velocity dynamics in response to M $$ M $$ .
Details are in the caption following the image
Velocity dynamics in response to K A $$ {K}_A $$ .
Details are in the caption following the image
Velocity dynamics in response to S r $$ {S}_r $$ .
Details are in the caption following the image
Velocity dynamics in response to E c $$ {E}_c $$ .
Details are in the caption following the image
Velocity dynamics in response to α $$ \alpha $$ .
Details are in the caption following the image
Velocity dynamics in response to A * $$ {A}^{\ast } $$ .
Details are in the caption following the image
Velocity dynamics in response to B * $$ {B}^{\ast } $$ .
Details are in the caption following the image
Velocity dynamics in response to P r $$ {P}_r $$ .
Details are in the caption following the image
Velocity dynamics in response to β $$ \beta $$ .
Details are in the caption following the image
Velocity dynamics in response to N 1 $$ {N}_1 $$ .
Details are in the caption following the image
Velocity dynamics in response to N r $$ {N}_r $$ .

Figure 13 depicts the temperature behavior in response to the Eckert number E c $$ {E}_c $$ . A growing E c $$ {E}_c $$ contributes to the heightened activity of the material particles as a result of the substantial energy stored in the boundary stratum, resulting in an elevation of the fluid temperature. Figure 14 elucidates the impact of the Prandtl number P r $$ {P}_r $$ on fluid temperature, showing that the fluid temperature shrinks with mounting P r $$ {P}_r $$ . The underlying reason for this behavior is that higher P r $$ {P}_r $$ indicate that thermal diffusion is comparatively lower likened to momentum diffusion, meaning that thermal energy is transmitted more slowly through the fluid. As a result, in fluids with elevated P r $$ {P}_r $$ , heat propagation occurs at a slower rate, leading to a more gradual change in temperature. This phenomenon is of paramount importance in maintaining thermal equilibrium within cooling systems, particularly in engines and high-performance machinery operating under extreme conditions. Figure 15 showcases the impact of nanoparticle volume fraction ϕ $$ \phi $$ on fluid temperature. It is detected that, the growth of ϕ $$ \phi $$ demotes the fluid temperature. This fallout is the result of the reduction of the width of the thermal boundary stratum in the flow domain. The distinction of Soret number S r $$ {S}_r $$ on fluid temperature is exposed in Figure 16. It is distinguished that the escalation of S r $$ {S}_r $$ enhances fluid temperature. Physically, an increase in the Soret parameter signifies that temperature variations exert a significant influence on solute diffusion. When a thermal gradient is established, solutes exhibit a propensity to thrust from areas of lower temperature to those of higher temperature due to this coupling effect. This propulsion can result in localized heating, as the solute particles transport thermal energy with them, thereby contributing to an elevation in the overall fluid temperature. Figures 17 and 18 explain that, the heat boundary stratum produces the energy for snowballing the heat generating values, that is, A > 0 and B > 0 which improve the thermal transport, and the energy is engrossed by lessening the heat consumption values, that is, A < 0 and B < 0 yielding to temperature decline pointedly adjacent the frontier layer. It is to be distinguished that the extent of thermal transference for temperature-dependent heat generation is more noticeably associated with surface-dependent heat generation. The performance of Dufour number D u $$ {D}_u $$ on fluid temperature is illustrated in Figure 19. It is perceived that fluid temperature grows when is D u $$ {D}_u $$ enhanced. The physics behind this process is, that in mass diffusion, kinetic energy is transported toward the fluid due to molecular exchanges, resulting in energy imbalance. The energy transmission enhances heat, which boosts the local temperature. The higher the concentration gradient, the greater the energy flow, resulting in a perceptible temperature intensification in improved diffusion processes. The upshot of radiative property N r $$ {N}_r $$ is exemplified in Figure 20. It is detected that the fluid temperature rises with a higher value of N r $$ {N}_r $$ . Practically, when fluid is subjected to thermal radiation it engrosses radiant energy which is rehabilitated into heat energy inside the fluid yielding to an escalation of fluid temperature.

Details are in the caption following the image
Temperature dynamics in response to E c $$ {E}_c $$ .
Details are in the caption following the image
Temperature dynamics in response to P r $$ {P}_r $$ .
Details are in the caption following the image
Temperature dynamics in response to ϕ $$ \phi $$ .
Details are in the caption following the image
Temperature dynamics in response to S r $$ {S}_r $$ .
Details are in the caption following the image
Temperature dynamics in response to A * $$ {A}^{\ast } $$ .
Details are in the caption following the image
Temperature dynamics in response to B * $$ {B}^{\ast } $$ .
Details are in the caption following the image
Temperature dynamics in response to D u $$ {D}_u $$ .
Details are in the caption following the image
Temperature dynamics in response to N r $$ {N}_r $$ .

The performance of order of chemical reaction n $$ n $$ toward fluid concentration is professed in Figure 21. It highlighted that n $$ n $$ is proportional to the concentration of the fluid. The physics behind this scenario is higher n $$ n $$ values tend to intensify in fluid concentration due to improved reaction rates, more operative exchanges among reactants, and response processes that endorse further reaction, eventually contributing to an advanced concentration of reactants or products in the fluid. Figure 22 elucidates the impact of the Schmidt number S c $$ {S}_c $$ on concentration. species. It is detected that concentration of the fluid diminutions progressively with the elevation of S c $$ {S}_c $$ From a physical standpoint, an expansion in the S c $$ {S}_c $$ signifies a lessening in molecular diffusion. Consequently, species concentration is augmented at lower Schmidt number, while it experiences a marked decline at higher values. Figure 23 reveals the influence of the thermo-diffusion S r $$ {S}_r $$ on fluid concentration. It is noted that snowballing the Soret number enhances fluid concentration. Physically, the S r $$ {S}_r $$ quantifies the relative strength of thermal diffusion compared to molecular diffusion. Optimizing the Soret consequence can improve efficiency in energy systems, chemical processing, biomedical applications, and advanced material technologies. An increase in this parameter amplifies the effect of temperature gradients on concentration distribution, leading to more pronounced variations in fluid reactions. The impression of the chemical reaction parameter k 1 $$ {k}_1 $$ on fluid concentration is portrayed in Figure 24. A pronounced decline in species concentration is observed with the intensification of k 1 $$ {k}_1 $$ . This reduction is attributable to the chemical reaction's repressive effect, which diminishes the molecular propulsion within the fluid, consequently leading to a significant attenuation in fluid concentration. Figure 25 describes the concentration behavior with time t $$ t $$ variation. It is perceived that fluid concentration enlarges as time progresses.

Details are in the caption following the image
Concentration dynamics in response to n $$ n $$ .
Details are in the caption following the image
Concentration dynamics in response to S c $$ {S}_c $$ .
Details are in the caption following the image
Concentration dynamics in response to S r $$ {S}_r $$ .
Details are in the caption following the image
Concentration dynamics in response to k 1 $$ {k}_1 $$ .
Details are in the caption following the image
Concentration dynamics in response to t.

5.1 Wall Shear Stress, Rate of Thermal Flow, and Rate of Solutal Flow

The distinction of the wall shear stress for both nanofluids with dissimilar parameters is visible in Table 4. As indicated in this table, the wall shear stress is elevated with the upsurge in M , K A , α , D u , and S r $$ M,{K}_A,\alpha, {D}_u,\mathrm{and}\ {S}_r $$ but it lessens with an enhancement in ϕ $$ \phi $$ . The rate of thermal transport (Nusselt number) variation with several physical flow parameters is disclosed in Table 5. It is stated from this table that the Nusselt number for both nanofluids upsurges with an increment in ϕ $$ \phi $$ but a contrary trend is perceived with a surge of M , A * , B * , and P r . $$ M,{A}^{\ast },{B}^{\ast },\mathrm{and}\ {P}_r. $$ The deviation of the Sherwood number for both nanofluids is accessible in Table 6. It is perceived that the Sherwood number heightens by intensifying S c and S r $$ {S}_c\ \mathrm{and}\ {S}_r $$ but exhibits a converse trend for a surge in k 1 , n , and ϕ $$ {k}_1,n,\mathrm{and}\ \phi $$ values.

TABLE 4. Numerical solutions of the wall shear stress C f $$ {C}_f $$ for distinct parameters.
C f $$ {C}_f $$
M $$ M $$ K A $$ {K}_A $$ α $$ \alpha $$ D u $$ {D}_u $$ S r $$ {S}_r $$ ϕ $$ \phi $$ Al 2 O 3 WEG $$ {Al}_2{O}_3- WEG $$ Ag WEG $$ Ag- WEG $$
2.0 1.0 0.5 0.3 0.2 0.01 1.345245 1.923434
4.0 1.357643 1.932111
6.0 1.363215 1.945223
2.0 1.312364 2.082168
3.0 1.321345 2.125452
4.0 1.334542 2.172371
1.0 1.439871 2.345757
1.5 1.442152 2.412359
2.0 1.447654 2.490121
0.4 1.213476 1.823189
0.6 1.217652 1.829874
0.9 1.345615 1.835528
0.4 1.233476 1.863189
0.6 1.247652 1.899874
0.8 1.325615 1.905528
0.03 1.453216 1.976525
0.05 1.432672 1.964326
0.07 1.423447 1.936540
TABLE 5. Numerical calculations of the thermal transport rate N u $$ {N}_u $$ for distinct parameters.
N u $$ {N}_u $$
M $$ M $$ A * $$ {A}^{\ast } $$ B * $$ {B}^{\ast } $$ P r $$ {P}_r $$ ϕ $$ \phi $$ Al 2 O 3 WEG $$ {Al}_2{O}_3- WEG $$ Ag WEG $$ Ag- WEG $$
2.0 −2.0 −2.1 0.71 0.01 1.295258 1.723471
4.0 1.257643 1.632745
6.0 1.163215 1.545639
−1.0 −1.682187 −1.212364
0.0 −1.725438 −1.425438
2.0 −1.972376 −1.572376
−1.5 1.239871 1.845761
2.0 1.042152 1.412352
3.0 0.747654 1.090125
1.0 1.313476 1.923188
1.5 1.217652 1.829871
2.0 1.145615 1.735523
0.03 1.233476 1.863189
0.05 1.347652 1.999864
0.07 1.425615 2.105628
TABLE 6. Numerical solutions of the mass transfer rate N u $$ {N}_u $$ for distinct parameters.
S h $$ {S}_h $$
S c $$ {S}_c $$ k 1 $$ {k}_1 $$ n $$ n $$ S r $$ {S}_r $$ ϕ $$ \phi $$ Al 2 O 3 WEG $$ {Al}_2{O}_3- WEG $$ Cu WEG $$ Cu- WEG $$
0.2 0.2 2.0 0.71 0.01 1.395258 1.823463
0.4 1.457643 1.932741
0.6 1.563215 2.145656
0.4 1.382181 1.712361
0.6 1.225433 1.625432
0.8 1.172375 1.572371
3.0 1.539871 1.715761
4.0 1.442152 1.642352
5.0 0.247654 1.500125
1.0 1.313465 1.823188
1.5 1.417651 1.929851
2.0 1.545613 2.135522
0.03 1.433476 1.863145
0.05 1.347652 1.799821
0.07 1.225615 1.505676

Figures 26-31 present 3D plots that elucidate the response of various parameters to engineering quantities. These visualizations offer a comprehensive insight into the multifaceted factors affecting frictional drag, as well as thermal and solutal transport rates. By revealing the complex interdependencies between these parameters, the diagrams serve as essential tools for rigorous engineering analysis, design refinement, and process optimization. Their capacity to highlight intricate relationships makes them invaluable assets in advancing the understanding of these critical transport phenomena. Figures 26 and 27 exemplify the impression of parameters K A $$ {K}_A $$ and ϕ $$ \phi $$ on the wall shear stress on both nanofluids, respectively. It is detected that a rise in both K A $$ {K}_A $$ and ϕ $$ \phi $$ results in an elevation in drug friction. Figures 28 and 29 demonstrate the consequence of parameters B * $$ {B}^{\ast } $$ and n $$ n $$ on the rate of thermal transport N u $$ {N}_u $$ . It is detected that an expansion in B * $$ {B}^{\ast } $$ results in a lessening in N u $$ {N}_u $$ . Conversely, when n $$ n $$ values are amplified, N u $$ {N}_u $$ exhibits an upward trend. Figures 30 and 31 portray the performance of α $$ \alpha $$ and A * $$ {A}^{\ast } $$ on the Sherwood number S h $$ {S}_h $$ . It is perceived that the upgrading α $$ \alpha $$ lessens the Sherwood number and enlarges when A * $$ {A}^{\ast } $$ values are augmented.

Details are in the caption following the image
A 3D plot of C f $$ {C}_f $$ behavior with K A and ϕ $$ {K}_A\ \mathrm{and}\ \phi $$ for Al 2 O 3 WEG $$ {Al}_2{O}_3- WEG $$ .
Details are in the caption following the image
A 3D plot of C f $$ {C}_f $$ behavior with K A and ϕ $$ {K}_A\ \mathrm{and}\ \phi $$ for Ag WEG $$ Ag- WEG $$ .
Details are in the caption following the image
A 3D plot of N u $$ {N}_u $$ behavior with B * and n $$ {B}^{\ast }\ \mathrm{and}\ n $$ for Ag WEG $$ Ag- WEG $$ .
Details are in the caption following the image
A 3D plot of N u $$ {N}_u $$ behavior with B * and n $$ {B}^{\ast }\ \mathrm{and}\ n $$ for Ag WEG $$ Ag- WEG $$ .
Details are in the caption following the image
A 3D plot of S h $$ {S}_h $$ behavior with α and A * $$ \alpha\ \mathrm{and}\ {A}^{\ast } $$ for Al 2 O 3 WEG $$ {Al}_2{O}_3- WEG $$ .
Details are in the caption following the image
A 3D plot of S h $$ {S}_h $$ behavior with α and A * $$ \alpha\ \mathrm{and}\ {A}^{\ast } $$ for Ag WEG $$ Ag- WEG $$ .

6 Conclusions

In this investigation, Al 2 O 3 and Ag $$ {Al}_2{O}_3\ \mathrm{and}\ Ag $$ nanoparticles were independently analyzed, utilizing a water-ethylene glycol mixture ( WEG ) $$ (WEG) $$ as the Casson base fluid. The dimensional governing equations, consisting of partial derivatives, were transformed into dimensionless PDEs by bringing forth suitable dimensionless parameters. The transmuted equations were then numerically cracked using the GFEM to examine the unsteady dissipative effects on hydromagnetic Casson nanofluid transport past a cone, subject to a non-uniform heat generation/absorption and a higher-order reacting species. Applications for this research findings comprise improving heat exchangers, optimizing cooling systems, and progressing nanotechnology in energy systems, electronics, and biomedicine. It helps strategy effective thermal management systems for sectors such as geothermal, solar and aerospace. The critical findings of this study can be concisely recapitulated as follows:
  1. The velocity profile is escalated with expansion in K A , E c , S r , A * , B * , N 1 , and N r $$ {K}_A,{E}_c,{S}_r,{A}^{\ast },{B}^{\ast },{N}_1,\mathrm{and}\ {N}_r $$ while a converse trend is detected in enhancing M , α , P r and β $$ M,\alpha, {P}_r\ \mathrm{and}\ \beta $$ .
  2. The fluid temperature exhibited an upsurge with growing values of E c , S r , D u , and N r $$ {E}_c,{S}_r,{D}_u,\mathrm{and}\ {N}_r $$ , and the contrary outcome is perceived when improving P r and ϕ $$ {P}_r\ \mathrm{and}\ \phi $$ .
  3. The concentration trend of nanofluid upsurged with an increment in n and S r $$ n\ \mathrm{and}\ {S}_r $$ while S c , and k 1 $$ {S}_c,\mathrm{and}\ {k}_1 $$ .
  4. The wall friction expands with the progress in M , K A , α , D u , and S r $$ M,{K}_A,\alpha, {D}_u,\mathrm{and}\ {S}_r $$ but it lessens with an enhancement of ϕ $$ \phi $$ .
  5. The Nusselt number expands with an augmentation in ϕ $$ \phi $$ but a converse tendency is detected with an escalation of M , A * , B * , and P r . $$ M,{A}^{\ast },{B}^{\ast },\mathrm{and}\ {P}_r. $$
  6. Sherwood number heightens by intensifying S c and S r $$ {S}_c\ \mathrm{and}\ {S}_r $$ but exhibits a converse trend for the surge in k 1 , n , and ϕ $$ {k}_1,n,\mathrm{and}\ \phi $$ values.

Nomenclature

  • u $$ u $$
  • velocity along x $$ x- $$ axis ( m / s ) $$ \left(m/s\right) $$
  • k $$ k $$
  • thermal conductivity ( W / mK ) $$ \left(W/ mK\right) $$
  • T $$ {T}^{\prime } $$
  • fluid temperature ( K ) $$ (K) $$
  • σ $$ \sigma $$
  • electrical conductivity ( S / m ) ( K ) $$ \left(S/m\right)(K) $$
  • C $$ C $$
  • fluid concentration mol / m 3 $$ \left( mol/{m}^3\right) $$
  • q r $$ {q}_r $$
  • radiative heat flux W / m 2 $$ \left(W/{m}^2\right) $$
  • B * $$ {B}^{\ast } $$
  • temperature dependent thermal consumption
  • θ $$ \theta $$
  • dimensionless temperature ( K ) $$ (K) $$
  • C $$ {C}_{\infty } $$
  • free stream concentration mol / m 3 $$ \left( mol/{m}^3\right) $$
  • M $$ M $$
  • magnetic field parameter
  • β nf θ $$ {\beta}_{nf}^{\theta } $$
  • heat expansivity coefficient ( K ) $$ (K) $$
  • v $$ v $$
  • velocity along y $$ y- $$ axis ( m / s ) $$ \left(m/s\right) $$
  • β θ $$ {\beta}^{\theta } $$
  • thermal expansion coefficient ( K ) $$ (K) $$
  • P r $$ {P}_r $$
  • Prandtl number
  • σ * $$ {\sigma}^{\ast } $$
  • effective thermal diffusivity of the fluid m 2 / s $$ \left({m}^2/s\right) $$
  • S c $$ {S}_c $$
  • Schmidt number
  • q $$ {q}^{{\prime\prime\prime} } $$
  • non-uniform heat flux W / m 2 $$ \left(W/{m}^2\right) $$
  • K A $$ {K}_A $$
  • porosity constant
  • β C $$ {\beta}^C $$
  • concentration expansion coefficient 1 / Kg / m 3 $$ \left(1/ Kg/{m}^3\right) $$
  • T w $$ {T}_w $$
  • temperature at the wall ( K ) $$ (K) $$
  • β nf C $$ {\beta}_{nf}^C $$
  • coefficient of volumetric expansivity ( K ) $$ (K) $$
  • T $$ {T}_{\infty } $$
  • free stream temperature ( K ) $$ (K) $$
  • g $$ g $$
  • acceleration on account of gravity m / s 2 $$ \left(m/{s}^2\right) $$
  • k 1 $$ {k}_1 $$
  • chemical reaction
  • C $$ C $$
  • dimensionless concentration
  • D u $$ {D}_u $$
  • Dufour parameter
  • C w $$ {C}_w $$
  • concentration at the wall molm 3 $$ \left({molm}^{-3}\right) $$
  • k T $$ {k}_T $$
  • thermal diffusion rate
  • ρ f $$ {\rho}_f $$
  • density of the fluid Kg / m 3 $$ \left( Kg/{m}^3\right) $$
  • S r $$ {S}_r $$
  • Soret parameter
  • ρ C p p $$ {\left(\rho {C}_p\right)}_p $$
  • base fluid heat capacity Jkg 1 k $$ \left({Jkg}^{-1}k\right) $$
  • C f $$ {C}_f $$
  • skin-friction N / m 2 $$ \left(N/{m}^2\right) $$
  • B 0 $$ {B}_0 $$
  • uniform magnetic field strength ( A / m ) $$ \left(A/m\right) $$
  • β $$ \beta $$
  • half angle ( rad ) $$ \left(\mathrm{rad}\right) $$
  • ρ C p p $$ {\left(\rho {C}_p\right)}_p $$
  • ratio of the nanoparticle heat capacity
  • N r $$ {N}_r $$
  • radiation parameter
  • D B $$ {D}_B $$
  • Brownian diffusion coefficient m 2 / s $$ \left({m}^2/s\right) $$
  • S h $$ {S}_h $$
  • Sherwood number
  • D m $$ {D}_m $$
  • chemical molecular diffusion
  • N u $$ {N}_u $$
  • Nusselt number
  • A * $$ {A}^{\ast } $$
  • space dependent thermal consumption
  • α $$ \alpha $$
  • Casson fluid parameter
  • Author Contributions

    L. Joseph Sademaki: conceptualization, writing – original draft, methodology, software, data curation, writing – review and editing, investigation, formal analysis, validation. B. Prabhakar Reddy: investigation, writing – original draft, writing – review and editing, visualization, methodology, supervision, resources. P. M. Matao: conceptualization, investigation, writing – original draft, writing – review and editing, methodology, formal analysis, supervision, project administration, visualization, resources.

    Conflicts of Interest

    The authors declare no conflicts of interest.

    Data Availability Statement

    Data sharing not applicable to this article as no datasets were generated or analysed during the current study.

      The full text of this article hosted at iucr.org is unavailable due to technical difficulties.