research paper

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

Numerical Heat Transfer, Part B: Fundamentals

An International Journal of Computation and Methodology

ISSN: (Print) (Online) Journal homepage: www.tandfonline.com/journals/unhb20

Numerical computation of magnetic field with


melting heat and homogeneous-heterogeneous
chemical reaction effects on oblique stagnation
flow of variable viscosity micropolar Fe3O4
nanofluids

R. Mehmood, Rabil Tabassum, Noreen Sher Akbar & Taseer Muhammad

To cite this article: R. Mehmood, Rabil Tabassum, Noreen Sher Akbar & Taseer Muhammad
(29 Feb 2024): Numerical computation of magnetic field with melting heat and homogeneous-
heterogeneous chemical reaction effects on oblique stagnation flow of variable viscosity
micropolar Fe3O4 nanofluids, Numerical Heat Transfer, Part B: Fundamentals, DOI:
10.1080/10407790.2024.2321495

To link to this article: https://doi.org/10.1080/10407790.2024.2321495

Published online: 29 Feb 2024.

Submit your article to this journal

Article views: 75

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=unhb20
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS
https://doi.org/10.1080/10407790.2024.2321495

Numerical computation of magnetic field with melting heat


and homogeneous-heterogeneous chemical reaction effects
on oblique stagnation flow of variable viscosity micropolar
Fe3O4 nanofluids
R. Mehmooda, Rabil Tabassuma,b, Noreen Sher Akbarc, and Taseer Muhammadd
a
Department of Mathematics, Faculty of Natural Sciences, HITEC University, Taxila Cantt, Pakistan;
b
Department of Mathematics, Faculty of Basic and Applied Sciences, Air University, Islamabad, Pakistan;
c
DBS&H, CEME, National University of Sciences and Technology, Islamabad, Pakistan; dDepartment of
Mathematics, College of Science, King Khalid University, Abha, Saudi Arabia

ABSTRACT ARTICLE HISTORY


The complex micro-structural characteristics of electro-conductive sol gel Received 12 June 2023
materials require simultaneous consideration of magneto-hydrodynamics, Revised 30 September 2023
micro-rheology and also physico-chemical phenomena. In this editorial, a Accepted 12 December 2023
mathematical model is therefore developed to simulate the steady-state,
KEYWORDS
oblique (non-orthogonal) stagnation flow of electro-conductive micropolar ADM; boundary layers;
magneto-nano-liquid flow impacting on an extending horizontal plane homogeneous-heteroge­
under the impact of transverse magnetic field. To capture the sophisticated neous reactions; magnetic
physico-chemistry, the simultaneous presence of homogeneous and het­ solar nano-coatings;
erogeneous chemical reactions is considered. Viscosity depending upon MATLAB; Maxwell-Garnett
temperature is taken into consideration with Reynolds’ exponential model. model; micropolar
Tiwari-Das and Maxwell-Garnett nano-liquid models are deployed which nanofluid; oblique
modifies density, thermal conductivity and electrical conductivity with vol­ stagnation flow;
ume fraction of nano-sized particles. temperature-dependent vis­
cosity; thermo-solutal
magneto-hydrodynamics;
volume fraction

1. Introduction
There is a general consensus in the engineering community that traditional working fluids like
water, glycol and oil are poor heat conductors and are therefore less effective in processes which
involve heat transfer control. Although many techniques have been explored to enhance thermal
performance of working fluids, only in the 1990s did Choi et al. [1] successfully devise the con­
cept of doping working fluids with nano-particles, synthesizing a new generation of liquids which
they designated as “nanofluids”. These were shown to be astonishingly proficient heat transfer flu­
ent media. Subsequently many types of nano-particles have been tested including carbon nano­
tubes (CNTs), nano-shells, fullerenes, graphene and magnetic nano-particles. Ferrofluids are
magnetic nanofluids which are quite effective due to their facility for adjustable physical proper­
ties under the action of an applied magnetic field. They are used in electronic devices, intelligent
(“smart”) lubrication, loudspeakers (where they robustly reduce heat from voice coils), medical
applications such as MRI, drug delivery, optics, petro-chemical drilling fluids and smart solar
coatings. Nanoparticles in ferrofluids are Iron-based. Other solar coating materials may employ

CONTACT Noreen Sher Akbar noreen.sher@ceme.nust.edu.pk DBS&H, CEME, National University of Sciences and
Technology, Islamabad, Pakistan.
� 2024 Taylor & Francis Group, LLC
2 R. MEHMOOD ET AL.

copper and silver nano-particles. Although many base fluids exist including vegetable oils (e.g.,
coconut, canola), organic liquids (e.g., ethylene glycol, butanol), water-based magnetized poly­
meric solutions remain very popular for environmentally-free solar coating applications [2]. Such
smart coatings offer many advantages in commercial solar technology including surfactant stabil­
ity and resistance to thermal degradation, anti-abrasion, sustained performance at different scales
(volume size-effects) and excellent tribo-chemical properties. In the manufacturing of such mate­
rials, many complex flow phenomena are involved. These include stretching, blowing, extrusion,
enrobing and impingement. The flow toward a stagnation point with heat and mass transfer may
be normal or oblique and is fundamental to solar coating synthesis [3]. Stagnation flows consti­
tute a special application of boundary-layer theory [4] and were first examined by Hiemenz [5]
who explored Newtonian viscous stagnation flow approaching a non-deformable surface in two-
dimensions via similarity methods. With the advent of digital computing, many subsequent inves­
tigations have extended the classical Hiemenz model for different fluids and in the presence of
different body forces and under various physical constraints. An excellent perspective of many
works relating to thermal convection stagnation flows has been provided by Hoshizaki et al. [6].
A comprehensive discourse on numerous stagnation-point flows in thin film coating materials
fabrication fluid dynamics has been presented by Rossi [7]. Generally, in normal stagnation flows
the axial velocity is not dependent upon radial position. The flow field in neighborhood of stag­
nation point becomes one dimensional and the Navier-Stokes or equations representing bound­
ary-layer can be reduced to simpler ordinary differential ones. In recent years many authors have
generalized the Hiemenz model to nanofluids. Bachok et al. [8] examined the two-dimensional,
time-dependent stagnant flow of nanoliquid with various categories of nanoparticles showing the
dual results for negative values of the unsteadiness parameter. Ibrahim et al. [9] studied analytic­
ally the magnetic nanofluid stagnation point convection along an extending surface focusing on
thermophoretic and Brownian motion nanoscale effects, noting that increasing extending sheet
velocity ratio elevates Nusselt number, skin friction and Sherwood number, whereas greater mag­
netic field boosts the temperatures. B�eg et al. [10] engaged a homotopy method to investigate the
transient nanofluid stagnation flow from a rotating sphere with free stream velocity effects, noting
the considerable rise in concentration and thickness of thermal boundary layer with thermopho­
resis, Brownian motion and unsteadiness. Gireesha et al. [11] studied computationally the steady-
state magneto-convective stagnation flow of an electrically conducting nanoliquid from a melting
extending surface with magnetic induction and heat sink/source effects, observing that surface
heat transfer is elevated with stronger melting effect and nanoparticles Brownian motion, whereas
it is reduced with magnetic field. Rehman et al. [12] scrutinized the heat transport in viscous,
incompressible nonaligned stagnation point nanoliquid flow impacting on a shrinking surface,
identifying that flow obliqueness exerts a significant influence on heat, momentum and concen­
tration fields.
These studies generally neglected the rheological characteristics of nanofluids which are associ­
ated with their suspension structure. Newtonian models while they provide some insight into the
nanoscale thermal characteristics and volume fraction doping, however do not provide the neces­
sary sophistication to simulate non-Newtonian characteristics. In this regard many physically viable
rheological models have been combined with nanofluid models to develop more comprehensive
frameworks for predicting shear characteristics. These include Ostwald-deWaele (power-law pseudo­
plastic and dilatant) models [13], multi-phase particulate models [14], second and third order
Reiner-Rivlin viscoelastic differential models [15, 16] and short memory viscoelastic models [17].
Stagnation-point flows of rheological nanofluids have stimulated considerable attention in recent
years. Hayat et al. [18] used Runge-Kutta methods to study radiative stagnation point flow of cross-
nanofluids. Nadeem et al. [19] investigated the viscoplastic nanofluid non-orthogonal stagnation
flow from an elongating surface with the Buongiorno nanoscale model. Mehmood et al. [20] used
the Jeffery model to analyze viscoelastic characteristics of nanofluid oblique stagnation flow from a
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 3

stretching plane. These non-Newtonian models capture various features of nanofluids including
stress relaxation, retardation, memory, shear-thinning/thickening etc. However, they ignore micro-
structural properties of the liquids, i.e., suspended particle spin. The spin effect can apply a
profound effect on nanofluid behavior and heat transfer rates. Micro-fluids dynamics [21] were
introduced by Eringen to mimick the gyratory motions of particles suspended in fluids. The theory
presented initially was very complex, involving six times as many balance equations as in the clas­
sical Navier-Stokes theory. Eringen [22], therefore, introduced a particular kind of microliquids
termed as micropolar fluids.
Computational simulations of multi-physical micropolar flows in materials processing, medical
engineering, energy systems and aerospace/naval propulsion have been reviewed by B�eg et al.
[23]. Representative studies of micropolar convection flows include Ibrahim et al. [24] who also
considered magnetohydrodynamics and mass transfer, Lok et al. [25] who studied stagnation
flow, Othman and Zaki [26] who examined electrohydrodynamic viscoelastic micropolar flow sta­
bility. In recent years several models have been developed combining micropolar theory and
nanofluid mechanics. Iqbal et al. [27] considered viscoplastic hydromagnetic micropolar nanofluid
transport. Nadeem et al. [28] studied stagnation point boundary layer flow of micropolar nano­
fluids over a curved body. Hussain et al. [29] addressed Sakiadis flow of micro-nanofluids.
Further studies of micropolar nanofluid flows include Noor et al. [30] who also included hydro­
dynamic wall slip effects, Waqas et al. [31] who analyzed magnetic field effects, Hayat et al. [32]
who considered radiative heat flux, Prasad et al. [33] who analyzed axisymmetric enrobing flows,
Haq et al. [34] who reported on radiative-convective stagnation flow and Latiff et al. [35] who
examined transient gyrotactic bioconvection from extending/contracting sheets.
In the synthesis of solar coating rheological nanofluids [3], high temperature and chemical
reactions may occur. These may be both homogeneous and/or heterogeneous in nature as exem­
plified by catalysis, exothermics, endothermics etc. In fluid flow phenomena these reactions are
very complex in nature and they cause reactant species production with dissimilar rates. In add­
ition, with extending sheet dynamics (Sakiadis flows), throughout the developing procedure, the
stretching sheet also interacts with the ambient nanoliquid both mechanically and thermally.
Homogenous chemical reactions have been considered in detail for manufacturing micropolar
transport phenomena by Shamshuddin et al. [36] for non-conducting fluids and Shamshuddin
et al. [37] for hydromagnetic radiative fluids. Uddin et al. [38] analyzed homogenous chemical
reactions in magnetic nanofluid dynamics. Chaudhary and Merkin [39] developed a more com­
prehensive model (although steady state) for combined homogeneous and heterogeneous chemical
reactions in stagnant flow with identical diffusion coefficients of both reactants. They found sev­
eral solutions due to hysteresis bifurcations but only considered Newtonian viscous fluids. This
model is more appropriate for solar coating nanofluid materials [40]. Kiran et al. [41] utilized the
Chaudhary-Merkin homogeneous-heterogeneous reaction model and micropolar theory to eluci­
date the complex physio-chemical processes in intestinal transport. Sajid et al. [42] inspected the
Blasius flow of Fe3 O4 magnetic nanofluids with homogeneous-heterogeneous reactions and strong
radiative flux effects, showing that species concentration is significantly reduced with increasing
both homogeneous and heterogeneous reaction rates. Hayat et al. [43] studied with a homotopy
method the stagnation point dynamics of nanofluid containing both single (SWCNT) and multi-
walled carbon nanotubes (MWCNT) from an impervious elongating cylinder with chemical reac­
tions. It was noted that concentration distribution is suppressed by strengthening homogeneous
reaction parameter for both MWCNT and SWCNT cases, whereas the opposite trend is induced
with increasing heterogeneous reaction parameter.
Usman et al. [44] studied the analysis of Cu-O–water nanofluid over a moving wall with
oblique stagnation point under the influence of a heat source. CuO nanoparticles having shapes
of platelet, brick and spheres were used in this study. Huq et al. [45] studied mixed convection in
micropolar nanofluid near an oblique stagnation point in the presence of a magnetic field. Aly
4 R. MEHMOOD ET AL.

et al. [46] investigated the exact solutions of the classical Glauert’s laminar wall jet mass and heat
transfer under wall suction, wall contraction or dilation, and two thermal transport boundary
conditions. Soomro et al. [47] considered magneto-hydrodynamic hybrid nanofluid to study the
heat transfer performance due to stretching of inclined surface. The stretching surface is consid­
ered under the effects of magnetic field along the normal direction.
Review of the literature revealed that the important nano-materials manufacturing problem of
transverse stagnation-point micropolar magnetic nanoliquid flow impacting on an extending hori­
zontal plane influenced by transverse magnetic field with homogeneous-heterogeneous reactions
is unaddressed till now. The important characteristic of viscosity variation is also considered since
viscosity of industrial fluids shows noticeable variation change in temperature. Viscosity of genu­
ine liquids rises with decline in temperature and there exist a noteworthy association between the
viscous properties and thermal development of nanofluids. The current model employs the robust
Reynolds exponential viscosity model which has been utilized successfully in many nanofluid sim­
ulations- see Tabassum et al. [48], Nada [49], Akbar et al. [50] and Khan et al. [51]. Viscosity
depending upon temperature is taken into consideration with Reynolds’ exponential model. The
two-dimensional incompressible boundary layer conservation equations for mass, linear momen­
tum, angular momentum (micro-rotation), energy and dual reactive species (A, B) are normalized
via proper similarity transformations into a strongly coupled, nonlinear, 13th order system of
ordinary differential equations with associated boundary conditions. The examination of the
emerging multi-physical problem is managed by a numerical shooting scheme in MATLAB sym­
bolic software. Validation of the solutions is included with an alternative semi-numerical scheme
known as the Adomian decomposition method (ADM). Graphical results and tables are presented
for the evolution of normal and tangential velocity components, temperature, reactive species
concentration, micro-rotation, normal skin friction, tangential skin friction and wall heat transfer
gradient (Nusselt number function) with homogenous and heterogenous chemical reaction par­
ameter, magnetic body force parameter and nano-particle volume fraction. The simulations are
relevant to the fabrication of solar magnetic nano-coatings.

2. Reactive magnetic nanofluid oblique stagnation flow model


Motivated by elucidating the materials manufacturing of solar magnetic nanoliquids, we consider
the two-dimensional, time-independent, incompressible, magneto-hydrodynamic, transverse stag­
nation flow of a variable-viscosity micropolar nanofluid impacting a stretching horizontal plane
with homogeneous-heterogeneous reactions under the action of a transverse, static magnetic field.
Magnetic induction and Joule and viscous heating are neglected. The nanofluid is assumed to be
dilute with equally sized Fe3O4 nano-particles regardless of shape factor. The horizontal sheet of
solar coating nano-polymer is stretched with the help of two forces opposite in direction and
equal in magnitude along the X ^ � -axis with the fixed origin as depicted in Figure 1. Iso-thermal
and iso-solutal conditions are enforced at the sheet surface. The Chaudhary-Merkin homoge­
neous-heterogeneous reactions model [39] is assumed as below:
2
^ þ 2B
A ^ ! 3B, ^ c^a � ^b � ,
^ rate ¼ K (1)
The single iso-thermal reaction on the catalyst surface can be written as follows:
^ ! B,
A ^ rate ¼ K
^ s^a � : (2)

Here, concentrations of the reactive chemical species A, ^ B ^ are denoted by ^a and b
� ^ , and
^ s are constants. Cross diffusion effects are negated and the Tiwari-Das and Maxwell-
^K c , K
Garnett nanofluid model is adopted. Eringen’s [21] micropolar non-Newtonian model is imple­
mented. The bulk spin torque term is neglected, and furthermore micro-elements are able to spin
only in the ^x � −^y � plane without interacting with nano-particles which have independent motion
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 5

Figure 1. Physical model for oblique stagnation magnetohydrodynamic micropolar nanofluid flow.

[22, 23]. The vector equations for micropolar fluids are summarized in Appendix. With the
fusion of Lorentzian magnetohydrodynamic body force and viscosity varying with temperature,
the governing boundary-layer equations for the flow, i.e., mass, normal linear momentum, trans­
verse linear momentum, micro-rotation (angular momentum), energy and dual chemical species
conservation, may be shown to take the form:
u � @^v �
@^
þ ¼ 0, (3)
@^x � @^y �
! !
u� u� 1 @^ p� 2 �
^ u � @^ ^ �Þ @T
l nf ðT ^� 2 �
^ ^�
^ nf ^�@^ @^
v� � þ ^ �Þ @ u
^ nf ðT
@^ ^ @ u @N
−r ^ 2u �
q u � þ^ ¼l þ � þ k þ ^ nf B 0^ ,
@^x @^y ^ nf @^x �
q @^y �
2
@^y �
@T ^ @^y �
@^y �
2
@^y �
(4)
! � � �
!
@^v � @^v � p� 2 � � ^ Þ @T
l nf ðT ^ 2 � ^
^�
^ nf u
q v� � þ
1 @^
^ nf ðT^ � Þ @ ^v 2 þ @^v @^ ^ @ ^v 2 − @ N
k , (5)
�þ^ ¼l � þ
@^x @^y ^ nf @^y �
q @^y � @^y �
@T ^ @^y �
@^y � @^x �
� �
! � �
@ ^
N @ N^ � @^u� @2N^�
^ nf :j u
q ^ �
v � ^ ^
¼ −k 2N þ � þ ^c nf 2 , (6)
� þ^ �
@^x @^y @^y @^y �
^�
@T ^� 2^�
� @T � @ T
^�
u þ v
^ ¼ ^
a 2 , (7)
@^x � @^y � nf
@^y �
@^a � @^a � 2 �
^ A @ ^a 2 − K ^�2,
^ c^a � b
^�
u v� � ¼ D
�þ^ �
(8)
@^x @^y @^y
^�
@b ^� @2b^�
� @b ^ ^ a�b^�2:
^�
u � þ v
^ � ¼ D B � 2 þ K c^ (9)
@^x @^y @^y
Here, the following notation applies: Velocity components alongside ^x � and ^y � − axis are
ð^ � ^ Þ
^ � and ^v � , q
u ^ nf and p ^ � Þ ¼ l0 e−d T −2:5T 1 represents the
^ � denote the density and pressure, lnf ðT ð1−/Þ
dynamic nanofluid viscosity which is an exponential function of temperature [45]. This model is
modified form of Brinkman viscosity model and it demonstrates that rise in temperature causes
6 R. MEHMOOD ET AL.

decline in viscosity of nanofluid. Here, l0 indicates the reference viscosity, d signifies the vis­
^
cosity variation exponent, a^ �nf ¼ k nf
is effective thermal diffusivity, r
^ nf represents the electrical
ðq^ cp^ Þnf
conductivity ^0 ^ � , /,
� constant� and B denotes the magnetic field strength of the nanofluid. N
� ^
^ , ^c nf ¼ l
T ^ j ¼ �^ f and l
^ nf þ k2 j, k, ^ nf characterize the micro-rotation component, magnetite nano­
c
particles volume fraction, temperature, Eringen spin gradient viscosity, Eringen vortex viscosity,
micro-inertia density and dynamic viscosity of the magnetic nanoliquid, respectively. Also ^a � and b ^�
^ and B:
denote the species concentrations for the reactants A ^ D^ A and D
^ B are the diffusion coefficients
^ ^
of the chemical reactants A and B: The effective electric conductivity r ^ nf is presented by:
� �
r
^ nf 3 rr^^ fs − 1 /
¼1þ� � � � (10)
r
^f r
^s
þ 2 − r^ s − 1 /
r
^f r
^f

Here, r
^ s and r^ f represent the electrical conductivity of nano
� sized particles of magnetite and
water, respectively. Density q^ nf and thermal capacitance q ^ cp^ nf of the nanoliquid are defined as
[44]:
^ nf ¼ ð1 − /Þ^
q q f þ /^qs, (11)
� � �
^ cp^ nf ¼ ð1 − /Þ q
q ^ cp^ f þ / q^ cp^ s , (12)
where q ^ s and q ^ f signify the mass density of magnetite particles and carrier fluid (water). The
Maxwell-Garnet model [44] is engaged to approximate the effective thermal conductivity of nano­
fluid k^ nf as follows:

^k nf k^ s þ 2k^ f − 2/ k^ f − ^k s
¼ �: (13)
k^ f k^ s þ 2k^ f þ / k^ f − k^ s
Here, k^ f and k^ s denote the thermal conductivities of water as base fluid and magnetite
nanoparticles, respectively. Some thermal and physical characteristics of water and Fe3 O4 [44] are
summarized in Table 1, which approximate quite well the properties of aqueous iron nano-doped
magnetic solar coating (sheet) materials [3]:
At the widening sheet surface (wall) and in free stream region the following boundary condi­
tions are suggested (edge of the boundary layer):
3
� � � @^a �
@ ^�
b
� � �
^ ¼ c^x , ^v ¼ 0, N
u ^ ¼ −n � , T
u
@^ ^ ¼T
^ w, D
^A � ¼ K ^ s^a , D
� ^ B � ¼ −K
^ s^a at ^y ¼ 0, 7
� �
@^y @^y @^y 5 (14)
� ^� � �
u � � ^ ^ � ^
^ ¼ a^x þ b^y , N ¼ constant, T ¼ T 1 , ^a ! a0 , b ! 0 when ^y ! 1: �

^ 1 and T
In Eq. (14), a, a0 , b, c are constants, T ^ w signify the ambient temperature of the
magnetic nanofluid and wall temperature, respectively. Micro-element concentration is denoted
by n where n 2 ½0, 1�[21, 23, 25, 33]. The case corresponding to n ¼ 0 describes strong micro-
element concentration and n ¼ 0:5 characterizes the case when the concentration of micro-ele­
ments is weak and is the most appropriate for materials processing boundary layer flows [32–37].
The case n ¼ 1 is used to describe the boundary layer flows generated by turbulence and is there­
fore not relevant to the present laminar flow regime [35–37]. It is judicious to convert the

Table 1. Physical characteristics of magnetite and water [44] for water-based solar nano-coatings.
� � � � �
^ mkg3
q J
cp^ kg:K ^k W −1
Properties m:K ^ ðX:mÞ
r
H2O 997.1 4179 0.613 0.05
Fe3O4 5180 670 6 25000
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 7

primitive forms of the conservation equations to dimensionless form and this is achieved by
implementing the following transformations:
rffiffiffiffi rffiffiffiffiffiffi
c c 1 1 ^�
p
^x ¼ ^x �
, ^y ¼ ^y �
,u ^ � pffiffiffiffiffiffi , ^v ¼ ^v � pffiffiffiffiffiffi , p
^¼u ^¼ , (15)
�^ f �^ f �^ f c �^ f c ^ �^ Þf c
ðq
^� ^ ^�
^ ¼ T − T1 , N
T ^ ¼ N , ^a � ¼ a0 ið^y Þ, ^b � ¼ a0 qð^y Þ:
^w − T
T ^1 c
The above expressions defined respectively nondimensional ^x � (lateral) coordinate, nondimen­
sional ^y � (transverse) coordinate, nondimensional normal velocity, nondimensional tangential vel­
ocity, dimensionless pressure, dimensionless temperature, dimensionless micro-rotation and
dimensionless reactant species concentrations. �^ f is signifying the kinematic viscosity of base
liquid (water). The emerging Eqs. (3)–(9), (14) in nondimensional form are as follows:
u @^v
@^
þ ¼ 0, (16)
@^x @^y
0 ! !1
^
−mT 2 ^ 2 ^
B e @ u
^ u
@^ @ T @ u
^ @ N C
B ð1 − /Þ2:5 @^y 2 − m @^y @^y þ K @^y 2 þ @^y C
u
@^ u q
@^ ^ @^ p ^ B
q B 0 1 C
� � C
^
u þ ^v þ f ¼ f B r
^s
3 r^ f − 1 / C, (17)
@^x @^y q ^ nf B
^ nf @^x q B C C
B −M^ u @1 þ � � � � A C
@ r
^s r
^s A
r
^f þ 2 − r
^f − 1 /
0 ! !1
^
−mT 2 ^ ^ 2
@^v ^ @^
@^v q p ^
q e @ ^v @^v @ T @ ^v @ N A
u
^ þ ^v þ f ¼ f @ 2−m þK − , (18)
@^x @^y q ^ nf ð1 − /Þ2:5
^ nf @^y q @^y @^y @^y @^y 2 @^x
� � !
^
@N ^
@N ^
q u ^ 1 K @2N^
^
u þ ^v ¼ −K f 2N ^ þ @^ q
þ f , (19)
2:5 þ
@^x @^y ^ nf
q @^y ^ nf
q ð1 − / Þ 2 @^y 2

^
@T ^ ^a nf @ 2 T
@T ^
^
u þ ^v ¼ , (20)
@^x @^y �^ f @^y 2
1 00
^v i0 ð^y Þ ¼ i ð^y Þ − K1ið^y Þq2 ð^y Þ, (21)
Sc
^d
q00 ð^y Þ þ K1ið^y Þq2 ð^y Þ,
^v q0 ð^y Þ ¼ (22)
Sc
rffiffiffiffi rffiffiffiffi
u
@^
^ ¼ −n , ^T ¼ 1, D
^ A i ð^y Þ ¼ K
0 ^s �
^ f ^ s �^ f ið^y Þ at ^y ¼ 0,
^ B q ð^y Þ ¼ −K
0
u
^ ¼ ^x , ^v ¼ 0, N ið^y Þ, D (23)
@^y c c
a b ^ b ^
u
^ ¼ ^x þ ^y , N ¼ − ,T ¼ 0, ið^y Þ ! 1, qð^y Þ ! 0 when ^y ! 1: (24)
c c c
^
Here, K ¼ l^k symbolizes the Eringen material parameter (vortex viscosity parameter), m ¼
� f
^ 1 represents the Reynolds’ viscosity variation parameter, K1 ¼ K^ c ao defines the homo­
2
^w − T
d T c
geneous reaction strength (rate) parameter (Note – the heterogeneous reaction strength parameter
2
r ^0
^f B
is defined in due course), M ¼ ^f c
q denotes the magnetic body force parameter, Sc ¼ D^�^ f is the
A
^
Schmidt number and ^d ¼ DD^ A : is the reactant diffusivities ratio. Presenting the stream function
B
8 R. MEHMOOD ET AL.

relations as [48]:
@w @w
u
^¼ , ^v ¼ − : (25)
@^y @^x
Engaging stream function relation in Eqs. (16)–(24) and after eliminating the pressure term p ^
by utilizing the fact that p ^ ^x ^y ¼ p^ ^y ^x in Eqs. (17) and (18) we have:
� 2 !2
^ � ^ @ � ^ ^ ^
q^ f e−mT @ @ T @ 2
w @ T @3w @T @2w @2T
2:5 2 r2 w − m r2 w þ m 2 2 −m 3 −m 2 2
q
^ nf ð1 − /Þ @^y @^y @^y @^y @^y @^y @^y @^y @^y
!
@T^ @3w 2 ^ ^ @2w @2T ^ � q ^f @2 �
2 @ w @T @T 2 2^
−m þm −m þ K r w þr N
@^x @^y 2 @^x @^y @^x @^x @^y @^y @^x @^y @^x ^ nf
q @^y 2
0 � � 1
2 3 r
^s
− 1 / 2

q^f @ wB r
^f C @ w, r w
− M @1 þ � � � � Aþ ¼ 0, (26)
^ nf @^y 2
q r
^s r
^s @ ð^x , ^y Þ
^f þ 2 − r
r ^f − 1 /
! !
^ @w @ N
@w @ N ^ ^
q @ 2
w ^
q 1 K ^
@2N
− ¼ −K f ^þ
2N 2 þ f
2:5 þ , (27)
@^y @^x @^x , @^y ^ nf
q @^y ^ nf ð1 − /Þ
q 2 @^y 2
!
^ @w @ T
@w @ T ^ ^
^a nf @ 2 T
Pr − ¼ , (28)
@^y @^x @^x , @^y ^a f @^y 2

@w 0 1
i ð^y Þ ¼ − i00 ð^y Þ þ K1ið^y Þq2 ð^y Þ, (29)
@^x Sc
@w 0 d
q ð^y Þ ¼ − q00 ð^y Þ − K1ið^y Þq2 ð^y Þ, (30)
@^x Sc
where Pr ¼ �^^a ff is the Prandtl number. The transformed wall and free stream boundary conditions
take the form:
rffiffiffiffi rffiffiffiffi
2
@w ^ ¼ −n @ w , T ^ A @i ¼ K
^ ¼1 D ^s �^ f ^ @q ^ �^ f
w ¼ 0, ¼ ^x , N 2 ið^y Þ, D B ¼ −K s ið^y Þ at ^y ¼ 0, (31)
@^y @^y @^y c @^y c

a 1 ^ ¼ −b,T
^ ¼ 0, ið^y Þ ! 1, qð^y Þ ! 0 as ^y ! 1:
w ¼ ^x ^y þ c^y 2 , N (32)
c 2 c
Here, c ¼ bc is the stagnation flow obliqueness parameter. The stream function relation is
redefined following Nadeem et al. [48] as:

wð^x , ^y Þ ¼ ^x Fð^y Þ þ Gð^y Þ, N ^ ð^x , ^y Þ ¼ hð^y Þ,


^ ð^x , ^y Þ ¼ ^x Jð^y Þ þ Sð^y Þ, T (33)
where Gð^y Þ and Fð^y Þ are tangential and normal flow components and Sð^y Þ and Jð^y Þ are tangen­
tial and normal micro-rotation components. Utilizing Eq. (33) in Eqs. (26)–(30) and after integra­
tion, we arrive at the following nonlinear system of coupled, multi-degree, multi-order, ordinary
differential equations:
� 0 1 �
r
^s
e −mh − 1 / ^ nf
q 3 �
ð 000 0 00 Þ
ð 000 0Þ 0B � C
r
^f 2

2:5 F − mh F þ K F þ J − MF @1 þ r
� � Aþ FF00 − ðF0 Þ þ C1 ¼ 0,
ð1 − /Þ ^s r
^s ^f
q
^f þ 2 − r
r ^f − 1 /

(34)
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 9

�0 1 �
r
^s
e−mh − 1 / ^ nf
q 3
ð 000 0 00 Þ ð 000 0Þ 0B � � � � C
r
^f
ðFG00 − F0 G0 Þ þ C2 ¼ 0,
2:5 G − mh G þ K G þ S − MG @1 þ r Aþ
^f
ð1 − /Þ ^ s
þ 2 − r
^ s
− 1 / q
r
^f r
^f

(35)
!
1 K 00 ^
q
2:5 þ J − Kð2J þ F00 Þ þ nf ðFJ0 − F0 JÞ ¼ 0, (36)
ð1 − / Þ 2 ^f
q
!
1 K 00 q
^
2:5 þ S − Kð2S þ G00 Þ þ nf ðFS0 − G0 JÞ ¼ 0, (37)
ð1 − / Þ 2 ^f
q
� !
k^ nf 00 ^ cp^ s
q
h þ Pr 1 − / þ � / Fh0 ¼ 0, (38)
k^ f ^ cp^ f
q
^ and B
It is assumed that the dispersion constants of the organic reactant classes A ^ are of simi­
^ ^
lar extent, so let D A ¼ D B which implies d ¼ 1: It follows that [39]:

i00 ð^y Þ þ Sc Fi0 ð^y Þ − K1ið1 − ið^y ÞÞ2 ¼ 0, (39)
Here, constants C1 and C2 appeared due to integration. The rates of change with respect to ^y
are denoted by primes. The boundary conditions now take the form:
)
F ð0Þ ¼ 0, F0 ð0Þ ¼ 1, −nF00 ð0Þ ¼ Jð0Þ, − nG00 ð0Þ ¼ Sð0Þ, i0 ð0Þ ¼ K2ið0Þ, hð0Þ ¼ 1,
a b (40)
F0 ð1Þ ¼ , G00 ð1Þ ¼ c, J ð1Þ ¼ 0, Sð1Þ ¼ − , ið1Þ ¼ 1, hð1Þ ¼ 0:
cqffiffiffi c
^
Here, K2 ¼ D^K s �^cf represents the heterogeneous reaction strength (rate) parameter. After
A
setting the limit when ^y ! 1 on Eq. 0 (34) and making use of the condition at boundary
r
^s
� 1
3 −1 /
r
^
B1þ r^ s þ2�−f r^ s −1�/C
0 a ^ nf
q 2 @ r^ f r
^f A
F ð1Þ ¼ c ¼ B we get C1 ¼ q^ ðBÞ þ q
^s MðBÞ: Boundary layer analysis of
f 1−/þq^ /
f

Eq. (34) indicates that Fð^y Þ ¼ ðBÞ^y þ A when ^y approaches 1 where boundary layer displace­
0 � 1
r
^s
3 −1 /
r
^f
B1þ r^ s þ2�− r^ s −1�/C
@ r^ f A
ment is denoted by A: Similarly C2 ¼ −A q^q^nf c þ
r
^f
q
^s Mc^y is attained by taking the
f 1−/þq^ /
f
00
limit ^y toward 1 on Eq. (35) and employing G ð1Þ ¼ c: Thus, Eqs. (34) and (35) take the
form:
0 � 1
r
^s
3 ^ f −1 /
r
B1 þ � � C � �
e−mh
r
^s
B /C
þ2 − rr^^ s −1
^ nf
q
ðF000 − mh0 F00 Þ þ KðF000 þ J0 Þ − MðF0 − BÞ@ r
^f
A f 00 ð 0 Þ2 ð Þ 2
þ FF − F þ B ¼ 0,
ð1 − /Þ2:5 1 − / þ qq^^ s / ^f
q
f

(41)
0 � 1
r
^s
r
^ 3 −1 /
−mh B1 þ r^ s þ2�−f r^ s −1�/C
e �B r
^ r
^f C q^
ð 000 0 00 Þ ð 000 0Þ
y − G0 @
2:5 G − mh G þ K G þ S þ M c^
f
A þ nf ðFG00 − F0 G0 − AcÞ ¼ 0,
^s
q ^f
q
ð1 − /Þ 1 − / þ q^ /
f

(42)
a b
where B ¼ represents stretching ratio parameter and c ¼ indicates flow obliqueness parameter.
c c
Now, presenting the following relation:
10 R. MEHMOOD ET AL.

G0 ð^y Þ ¼ cHð^y Þ, (43)


Using Eq. (43) in Eqs. (42) and (37) respectively, then, the final form of the tangential
momentum and secondary micro-rotation (angular momentum) equations become:
0 � 1
r
^s
3 −1 /
� � B1 þ r^ s þ2 − r^ s −1�/C �r^ f
e−mh S0 B r
^f r
^f C q^
ð 00 0 0 Þ 00
2:5 H − mh H þ K H þ þ Mð^y − HÞ@ A þ nf ðFH0 − F0 H − AÞ ¼ 0,
c ^
q ^f
q
ð1 − /Þ 1 − / þ q^ /s
f

(44)
!
1 K 00 ^
q
þ S − Kð2S þ cH0 Þ þ nf ðFS0 − cHJÞ ¼ 0, (45)
ð1 − /Þ2:5 2 ^f
q

With the modified tangential velocity boundary conditions as follows:


Hð0Þ ¼ 0, H0 ð1Þ ¼ 1: (46)

3. MATLAB shooting quadrature numerical outcomes


The finalized boundary value problem is defined by six ordinary differential equations of collect­
ive order thirteen, i.e., Eq. (41) (third order normal momentum), Eq. (44) (reduced second order
tangential momentum), Eq. (36) (second order primary micro-rotation, i.e., angular momentum),
Eq. (45) (second order secondary micro-rotation), Eq. (38) (second order temperature) and
Eq. (39) (second order species concentration) with the boundary conditions in Eq. (40) modified
with the replacement conditions in Eq. (46), i.e., 13 boundary conditions are required and pro­
vided. Evidently this system is very difficult to solve analytically if not completely intractable. A
computational approach is therefore chosen in which numerical quadrature is executed (i.e., a
shooting algorithm) composed of the robust Runge-Kutta Fehlberg algorithm with an approxi­
mate executed CPU time of 10−1 Sec. This approach can effortlessly handle multi-order ordinary
differential BVPs and has been implemented by means of different symbolic codes in numerous
studies. Its greater advantage is its stability when applied to stiff equations. This method [49, 50]
employs an efficient procedure to determine a suitable step size h for stepping. Two different
approximations are made for the solution and compared. The step size is assumed to be accept­
able if the obtained results are in excellent agreement. Otherwise the process is repeated with a
smaller step size. Following six values are required in each step:
l1 ¼ hf ðtk , yk Þ,
� �
1 1
l2 ¼ hf tk þ h, yk þ l1 ,
4 4
� �
3 3 9
l3 ¼ hf tk þ h, yk þ l1 þ l2 ,
8 32 32
� �
12 1932 7200 7296
l4 ¼ hf tk þ h, yk þ l1 − l2 þ l3 ,
� 13 2197 2197 2197�
439 3680 845
l5 ¼ hf tk þ h, yk þ l1 − 8l2 þ l3 − l4 , (47)
� 216 513 4104 �
1 8 3544 1859 11
l6 ¼ hf tk þ h, yk − l1 þ 2l2 − l3 þ l4 − l5 :
2 27 2565 4104 40
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 11

Then I.V.P is approximated by using a Runge-Kutta method of order 4:

25 1408 2197 1
ykþ1 ¼ yk þ l1 þ l3 þ l4 − l5 , (48)
216 2565 4104 5

By utilizing following substitutions respectively in Eqs. (41), (44), (36), (45), (38) and (39), i.e.,
the finalized coupled six ordinary differential equations of collective order thirteen and the modi­
fied boundary conditions (40) and (46):
0 1 0 1 9
0 1 0 1 H y4 >
F y1 >
B C B 0 0 C >>
@ F 0 A ¼ @ y0 ¼ y2 A, B H C ¼ B y0 4 ¼ y5 C, > >
1 @ J A @ y ¼ y6 A > >
>
F 00 0
y 2 ¼ y3 0 0
5 >
=
0 1 0 1 J y 6 ¼ y 7
(49)
S y8 � � � � >>
B S0 C B y0 ¼ y9 C I >
>
B C¼ B 08 C, 0 ¼ y12 >
>
@ h A @ y ¼ y10 A I 0 : >
>
9 y 12 ¼ y13 >
>
0 0 ;
h y10 ¼ y11

Then, the following numerical differential equation system defining the transformed six flow
variables (both linear velocity components, both micro-rotation components, temperature and
reactant concentration) is obtained:
0 1
� � � � 2 !!
B 1 −my10 a ^ nf
q 2 a C
B Cð/Þe my11 y3− Ky7 þ M y2 − K1 − y1 y 3 − y 2 þ , C
B Cð/Þe−my10 þ K c q^f c C
B � � C
B −1 K ^
q C
0 0 1 B −my
Cð/Þe−my10 my11 y5− y9 − Mð^y − y4 ÞK1 − nf ðy1 y5 − y2 y4 − AÞ , C
y3 B Cð/Þe 10 þK � c ^ f�
q C
B C
B y0 5 C B 1 ^ nf
q C
B 0 C B Kð2y6 þ y3 Þ − y y − y2 y 6 Þ , C
B y7 C B
B 0 C¼B Cð/Þ þ K2� q^f ð 1 7 C
C
By C B � C
B 09 C B 1 ^ nf
q C
K 2y
ð 8 þ cy 5Þ − y y − cy y
4 6Þ ,
@y A B
11 B Cð/Þ þ K2 ^f ð 1 9
q C
C
y0 13 B �! C
B
B k^ f ^ cp^ s
q C
C
B −Pr 1−/þ/ � ðy1 y11 Þ, C
B ^
k nf ^
q cp^ f C
B � C
@ Sc K1y12 ð1 − y12 Þ2 − y1 y13 , A

(50)
0 � 1
r
^s
3 −1 /
r
^f
B1þ r
^s
� r
^s
�C
@ r
^f
þ2 −
r
^f A
−1 /
Here, Cð/Þ ¼ ð1−/1Þ2:5 and K1 ¼ q
^s : The finalized numerical boundary conditions
1−/þq^ /
become: f

9
b1 >
y1 ð0Þ ¼ 0, y2 ð0Þ ¼ 1, y3 ð0Þ ¼ − , y4 ð0Þ ¼ 0, >
>
n =
b2 : (51)
y5 ð0Þ ¼ − , y6 ð0Þ ¼ b1 , y7 ð0Þ ¼ b3 , y8 ð0Þ ¼ b2 , >
>
nc >
;
y9 ð0Þ ¼ b4 , y10 ð0Þ ¼ 1, y11 ð0Þ ¼ b5 , y12 ð0Þ ¼ b6 , y13 ð0Þ ¼ K2b6 :

The subsequent IVP is solved numerically by executing in MATLAB the fourth order Runge-
Kutta-Fehlberg method with an associated shooting scheme [47, 49]. The numerical shooting
parameters b1 , b2 , b3 , b4 , b5 and b6 are optimized by the Newton-Raphson iterative scheme.
Further details are provided in Keller [50] and specific applications to nonlinear ODE systems in
magneto-hydrodynamic flows are addressed by B�eg [51].
12 R. MEHMOOD ET AL.

4. Validation with Adomian decomposition method code (ADSIMNAN)


To validate the accuracy of the generalized magneto-hydrodynamic micropolar nanofluid oblique
reactive flow model, an alternative numerical scheme is required. We employed an efficient
Adomian decomposition method (ADM) [54]. This Scheme has been implemented in numerous
multi-physical and multi-scale fluid dynamics problems in recent year, including periodic (pulsa­
tile) as considered by Liu [55], Sisko non-Newtonian thin film flows as examined by Siddiqui
et al. [56], bio-magnetic orthopedic lubrication flows as simulated by B�eg et al. [57], micropolar
channel hydrodynamics in channels as investigated by Aski et al. [58], pulsatile micropolar flow
as studied by Adanhounm�e et al. [59], and rotating nanofluid bio-convection systems as simu­
lated by B�eg et al. [58]. ADM [60] deploys an infinite series solution for the unknown functions
(F,H, J, S, h, i) and utilizes recursive relations. The present ordinary differential nonlinear bound­
ary value problem (BVP) is re-written using the standard operator, following B�eg et al. [57]:
LðWÞ þ RðWÞ þ NðWÞ ¼ Q (52)
where W is the function to be determine, L symbolizes the highest order linear derivative, R is a
linear differential operator of order less than L, nonlinear terms are represented by N, and Q is
the source term. Applying the inverse operator L−1 to both sides of Eq. (52) and using the given
conditions we obtain:
W ¼ v − L−1 ðRWÞ − L−1 ðNÞ (53)
where � denotes the terms appearing after integrating the source term Q and from the auxiliary
conditions. ADM gives the following series solution W :
X
1
W¼ Wn (54)
n¼0

The solution for the nonlinear terms is:


X
1
N¼ Bn (55)
n¼0

Here, Adomian polynomials, denoted by Bn , can be evaluated via the following relation
[52–58]:
( !)
1 dk X 1
k
Bk ¼ N k Wk (56)
k! dkk k¼0 k¼0

If a nonlinear function fðWÞ is used to express the nonlinear term, then Adomian polynomials
take the form:
B0 ¼ fðW0 Þ (57)

B1 ¼ W1 f ð1Þ ðW0 Þ (58)


1
B2 ¼ W2 f ð1Þ ðW0 Þ þ W1 2 f ð2Þ ðW0 Þ (59)
2!
1
B3 ¼ W3 f ð1Þ ðW0 Þ þ W1 W2 f ð2Þ ðW0 Þ þ W1 3 f ð3Þ ðW0 Þ (60)
3!
The components W0 , W1 , W2 , ::::::::: are then evaluated recursively by employing the rela­
tion:
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 13

W0 ¼ v
(61)
Wkþ1 ¼ −L−1 RWk − L−1 Bk , k � 0
Here, W0 is denoting the zeroth component. Finally obtained truncated series solution is
given as:
X
1
Sn ¼ Wn (62)
n¼0

The individual Adomian polynomial series solutions for the present 13th order nonlinear
ordinary differential oblique stagnation flow boundary value problem, i.e., the unknown func­
tions, linear normal and tangential velocity components, primary and secondary micro-rotation
(angular velocity components), temperature and reactant species functions, which are designated
as FðyÞ, HðyÞ, JðyÞ, SðyÞ, hðyÞ and iðyÞ are defined by:
X
1
FðyÞ ¼ fn ðyÞ (63)
n¼0

X
1
HðyÞ ¼ hn ðyÞ (64)
n¼0

X
1
JðyÞ ¼ jn ðyÞ (65)
n¼0

X
1
SðyÞ ¼ sn ðyÞ (66)
n¼0

X
1
hðgÞ ¼ hn ðgÞ (67)
n¼0

X
1
iðyÞ ¼ In ðyÞ (68)
n¼0

As elaborated earlier the components f0 , f1 , f2 ::::, h0 , h1 , h2 ::::, j0 , j1 , j2 ::::, s0 , s1 , s2 ::::, h0 , h1 , h2 ::::


and finally I0 , I1 , I2 :::: are determined recursively. Computations are performed in the
MATHEMATICA symbolic software on an SGI Octane Desk workstation and compute in sec­
onds. Comparison between the MATLAB shooting and ADM solutions for selected values of key
parameters are shown in Tables 2–4. In all cases very close correlation is achieved. Confidence in
the MATLAB shooting solutions for the general magnetic micropolar nanofluid model is there­
fore justifiably high.
Table 2 shows that surface concentration is decreased with growing homogenous reaction
strength parameter K1 and also decreases with increasing magnetic parameter M. However, the
impact of K1 is more dramatic. Table 3 shows that surface concentration is depleted with increas­
ing heterogeneous reaction strength parameter K2 to a much greater extent than homogenous
reaction strength parameter K1: Similarly surface concentration is also reduced substantially with
increasing magnetic parameter M. Table 3 demonstrates that the normal skin friction component
(F00 ð0Þ) is decreased with increasing nano-particle volume fraction (/), whereas there is a slight
elevation in tangential skin friction (H0 ð0Þ). Heat transfer gradient at the wall (−h0 ð0Þ) is mark­
edly reduced with higherr nano-particle volume fraction indicating that there is an increase in
thickness of thermal boundary layer and boost in temperatures in the body of the micropolar
nanofluid.
14 R. MEHMOOD ET AL.

Table 2. Concentration at the surface with magnetic parameter M for three different values of homogeneous reaction strength
parameter K1 and B ¼ 0:1, m ¼ 0:1, K ¼ 0:1, n ¼ 0:5, c ¼ 0:1, Sc ¼ 0:1, K2 ¼ 0:1, / ¼ 0:1:
K1 ¼ 0.1 K1 ¼ 0.4 K1 ¼ 0.7
MATLAB K1 ¼ 0.1 MATLAB K1 ¼ 0.4 MATLAB K1 ¼ 0.7
M R-K-F ADM R-K-F ADM R-K-F ADM
0.05 0.595345 0.595364 0.558176 0.558171 0.50721 0.507241
0.1 0.594414 0.594420 0.557064 0.557058 0.505594 0.505589
0.15 0.593518 0.593514 0.55599 0.556041 0.504034 0.504029
0.2 0.592657 0.592643 0.554954 0.554929 0.502527 0.502519
0.25 0.591828 0.591809 0.553953 0.553947 0.50107 0.501068
0.3 0.591028 0.591034 0.552986 0.552302 0.499661 0.499658
0.35 0.590257 0.590249 0.552049 0.552071 0.498297 0.498291
0.4 0.589513 0.589519 0.551143 0.551101 0.496976 0.496972
0.45 0.588794 0.588788 0.550265 0.550228 0.495695 0.495675
0.5 0.588098 0.588091 0.549414 0.549417 0.494454 0.494444
0.55 0.587426 0.587430 0.548589 0.548592 0.493249 0.493237
0.6 0.586775 0.586768 0.547788 0.547784 0.49208 0.492082
0.65 0.586144 0.586139 0.547011 0.547018 0.490944 0.490923
0.7 0.585532 0.585529 0.546255 0.546248 0.48984 0.489836
0.75 0.584939 0.584932 0.558176 0.558174 0.50721 0.507196
0.8 0.584364 0.584361 0.557064 0.557059 0.505594 0.505585
0.85 0.583805 0.583811 0.55599 0.555998 0.504034 0.504031

Table 3. Concentration at the surface against magnetic parameter M for three different values of heterogeneous reaction
strength parameter K2 with B ¼ 0:1, m ¼ 0:1, K ¼ 0:1, n ¼ 0:5, c ¼ 0:1, Sc ¼ 0:1, K1 ¼ 0:1, / ¼ 0:1:
K2 ¼ 0.1 K2 ¼ 0.2 K2 ¼ 0.4
MATLAB K2 ¼ 0.1 MATLAB K2 ¼ 0.2 MATLAB K2 ¼ 0.4
M R-K-F ADM R-K-F ADM R-K-F ADM
0.05 0.595345 0.595341 0.422349 0.422342 0.26769 0.26765
0.1 0.594414 0.594413 0.421407 0.421406 0.266941 0.266943
0.15 0.593518 0.593520 0.420503 0.420504 0.266223 0.266227
0.2 0.592657 0.592646 0.419635 0.419632 0.265534 0.265529
0.25 0.591828 0.591824 0.418799 0.418797 0.264872 0.264868
0.3 0.591028 0.591025 0.417995 0.417993 0.264235 0.264238
0.35 0.590257 0.590251 0.41722 0.417223 0.263623 0.263626
0.4 0.589513 0.589516 0.416473 0.416477 0.263033 0.263036
0.45 0.588794 0.588793 0.415752 0.415754 0.262464 0.262467
0.5 0.588098 0.588092 0.415056 0.415059 0.261916 0.261918
0.55 0.587426 0.587424 0.414384 0.414382 0.261386 0.261381
0.6 0.586775 0.586773 0.413733 0.413735 0.260874 0.260867
0.65 0.586144 0.586141 0.413104 0.413105 0.260379 0.260374
0.7 0.585532 0.585534 0.412494 0.412493 0.2599 0.259931
0.75 0.584939 0.584933 0.411903 0.411908 0.259437 0.259433
0.8 0.584364 0.584362 0.41133 0.411334 0.258988 0.258984
0.85 0.583805 0.583806 0.410775 0.410771 0.258552 0.258549

00 0 0
Table 4. Values of F ð0Þ, H ð0Þ and −h ð0Þ against / with B ¼ 0:1, m ¼ 0:1, K ¼ 0:1, n ¼ 0:5, c ¼ 0:1, Sc ¼ 0:1, K1 ¼
0:1, K2 ¼ 0:1, M ¼ 0:1:
00 0 0
F ð0Þ 00
H ð0Þ 0
−h ð0Þ 0
MATLAB F ð0Þ MATLAB H ð0Þ MATLAB −h ð0Þ
/ R-K-F ADM R-K-F ADM R-K-F ADM
0 −1.05374567 −1.05374562 0.34286576 0.34286571 1.76795557 1.76795551
0.02 −1.06969370 −1.06969371 0.34272998 0.34272992 1.71096957 1.71096959
0.04 −1.08229063 −1.08229059 0.34282208 0.34282210 1.65661679 1.65661672
0.06 −1.09180156 −1.09180152 0.34310638 0.34310644 1.60471141 1.60471143
0.08 −1.098453853 −1.09845385 0.34355583 0.34355577 1.55508720 1.55508717
0.1 −1.102444846 −1.10244485 0.34414932 0.34414928 1.50759460 1.50759457
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 15

5. Results and discussion


Extensive solutions are presented for the MATLAB R-K-F code in Figures 2–12.
Note that the notation (^) is dropped in the graphs for clarity. Furthermore, in some cases the
MATLAB R-F-K numerical terminology is used, i.e., N1 for J (primary micro-rotation Using
the data in Table 1, a nominal value of 0.1 is computed which is representative of aqueous
nano-polymer solar coatings [2, 3]. Ten independent parameters arise in the final boundary value
problem: B (stretching ratio parameter), m (Reynolds’ viscosity variation parameter), / (nano-par­
ticle volume fraction), K (Eringen vortex viscosity parameter), n (Eringen micro-gyration constant, c
(stagnation flow obliqueness parameter), Sc (Schmidt number), K1 (homogeneous reaction param­
eter), K2 (heterogeneous reaction parameter), M (magnetic parameter). In the present computations
we prescribe n ¼ 0.5 which is appropriate for nano-polymer micropolar suspensions. In the bound­
ary conditions, N^ � ¼ −n @^u �� which defines the relationship between micro-rotation variable and the
@^y

0
Figure 2. Normal velocity profile F ðyÞ for different values of M:

0
Figure 3. Tangential velocity profile H ðyÞ for different values of M:
16 R. MEHMOOD ET AL.

Figure 4. Temperature profile hðyÞ for different values of M:

Figure 5. Concentration profile iðyÞ for different values of M:

normal velocity surface shear stress, i.e., it allows the simulation of the inter-relationship between
micro-rotation vector and shear stress. The state for which n ¼ 0 relates to the situation where the
particle density is sufficiently large, hence microelements nearby the wall are not able to rotate and
therefore this case not applicable. Many studies have confirmed that n ¼ 0.5 is physically viable for
engineering microstructural fluid analysis Figure 1. The default values implemented in the computa­
tions are as follows with the physical explanation in brackets B ¼ 0.1 (strong nano-particle doping
and high electrical conductivity of micropolar nanofluid), m ¼ 0.1 (high viscosity dependence on
temperature), / ¼0.1 (high nano-particle doping), K ¼ 0.1 (weak vortex viscosity), n ¼ 0.5 (weak
concentration of micro-elements), c ¼ 0.1 (weak obliqueness), Sc ¼ 0.1 (reactant diffusivity is one
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 17

Figure 6. Primary micro-rotation profile Nð1, yÞ for different values of M:

Figure 7. Concentration profile iðyÞ for different homogeneous reaction strength parameters K1:

tenth of the momentum diffusivity), K1 ¼ 0.1 (weak homogeneous reaction strength parameter),
K2 ¼ 0.1 (weak heterogeneous reaction strength parameter), M ¼ 0.1 (weak magnetic field).
Figures 2 and 3 illustrate the impact of magnetic
� (body force) parameter on respectively the nor­
mal and tangential velocity profiles F0 ðyÞ, H0 ðyÞ : Evidently there is a consistent decrease in normal
flow velocity with elevation in magnetic parameter M: This response is sustained throughout the
boundary layer transverse to the stretching sheet (Figure 2). The secondary (tangential) velocity
response is however very different – there is an initial boost in tangential velocity values close to the
sheet with increasing magnetic parameter, whereas some distance from the sheet the trend is then
reversed with subsequent deceleration in tangential flow with increasing magnetic field effect. The
modified Lorentzian (magnetic drag) term in the normal momentum Eq. (41) takes the form
18 R. MEHMOOD ET AL.

Figure 8. Concentration profile iðyÞ for different heterogeneous reaction strength parameters K2:

Figure 9. Concentration at the surface ið0Þ for different homogeneous reaction strength parameters K1:

0 � 1
r
^s
3 −1 /
r
^f
� �C
�B
1þ r
^s r
^s
@ r
^f
þ2 −
r
^f A
−1 /
−M F0 − ac , on the other hand in the normal momentum Eq. (44) corresponding
q
^s
1−/þq^ /
f
0 � 1
r
^s
3 −1 /
r
^f
B1þ r^ s þ2 − r^ s −1�/C

@ r^ f r
^f A
Lorentzian (magnetic drag) term is þMð^y − HÞ q
^ : In normal momentum a retard­
1−/þ s / q
^f

ing motion is caused by magnetic field. However, in tangential case the field tend to dragged with
the free stream velocity and this initially serves to accelerate the flow for slight distances from the
stretching sheet surface. But, with increase in transverse coordinate the constraining effect leads and
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 19

Figure 10. Concentration at the surface ið0Þ for different heterogeneous reaction strength parameters K2:

Figure 11. Concentration at the surface for homogeneous reaction strength parameter K1:

Figure 12. Concentration at the surface for heterogeneous reaction strength parameter K2:

consequently causes a deceleration. A similar effect has been reported in other magneto-hydro­
dynamic flows by for example Chamkha and Camille [61] and Zueco et al. [62]. The implication is
that the tangential flow causes the magnetic field to move with the free stream which initially
20 R. MEHMOOD ET AL.

enhances tangential velocities but eventually leads to a marked deceleration. Similar observations
have been made for magneto-hydrodynamic stagnation flows by Cramer and Pai [63], although
only for orthogonal impingement. In all the cases computed magnetic drag is relatively weak and
less than the viscous hydrodynamic force. Reverse (back) flow is never generated and consistently
positive normal and tangential velocities are computed. Normal hydro-dynamic (velocity) boundary
layer thickness is therefore increased, whereas there is an initial thinning in the tangential boundary
layer thickness followed by a subsequent thickening.
Figure 4 depicts the evolution in temperature distribution, hðyÞ with magnetic parameter M: The
case of M ¼ 1 corresponds to an equivalence in inertial forces and magnetic drag in the regime. For
M > 1 the magnetic force leads the inertial force. This causes elevation in temperatures and increases
thickness of thermal boundary layer. In consistency with this, temperature is minimum for M ¼ 1
(weak magnetic field case), whereas the maximum temperature is achieved with M ¼ 7 (strongest
magnetic field case). Temperature decayed uniformly from the stretching sheet toward the free stream
and attained a smooth convergent profile. Strong magnetic fields can therefore be employed to
strengthen the boundary layer flow and to decline rate of heat transfer at the boundary (sheet sur­
face), which is an important aspect in thermal control involved in materials manufacturing [3].
Figure 5 presents the distribution for concentration of chemical species versus transverse normal­
ized coordinate with various magnetic parameter values M: The profiles are the opposite in nature to
the temperature profiles. The former exhibit monotonic growths, whereas the latter are clearly mono­
tonic decaying profiles (Figure 4). In other words, the maximum temperature is imposed at the sheet
surface and decays to zero in the free stream, whereas the minimum reactant concentration is pre­
scribed at the wall and ascends to a maximum in the free stream. There is a strong reduction in
reactant concentration (ið^y Þ) with greater transverse magnetic field. Clearly the Lorentzian magnetic
body force term arising in the normal momentum equation exerts a non-trivial influence on the
reactant species concentration distribution via the coupling term, Fi0 ð^y Þ: in Eq. (39). The deceleration
in the normal flow (Figure 2) reduces species diffusion in the boundary layer which decreases the
concentration magnitudes. The mass flux to the wall (species gradient at the wall) is therefore
enhanced. The imposition of a magnetic field therefore successfully alters the distribution of reactant
species in the regime and is a useful mechanism therefore for controlling the organization of species
in the solar magnetic micropolar nanofluid coating during manufacturing.
Figure 6 illustrates the response in the primary micro-rotation component N(1,y) (i.e., J) in the
boundary layer regime to a variation in magnetic parameter values (MÞ: With a relatively small
increment in M from 0.1 to 0.7, there is a substantial elevation in micro-rotation. As with the tem­
perature and reactant species concentration fields, the influence of Lorentzian magnetic body force
is experienced indirectly by the micro-rotation field. Magnetic body forces do not arise in either
Eq. (36) (second order primary micro-rotation, i.e., angular momentum) or Eq. (45) (second order
secondary micro-rotation). The coupling terms between the primary micro-rotation Eq. (36) and
the normal momentum Eq. (41), i.e., −Kð2J þ F00 Þ þ q^q^nf ðFJ0 − F0 JÞ and KðF000 þ J0 Þ, imply that there
f
is a strong dependence of the micro-rotation on the normal velocity. The deceleration in the nor­
mal flow allows for faster rotation of micro-elements at low micropolarity (K ¼ 0.1) since the flow
is less dominated by linear (translational) momentum and micro-elements are permitted to perform
gyratory motions more easily. This has also been identified by Ahmadi [65]. Furthermore, the decay
in micro-rotation profiles exhibits an upward trend, i.e., micro-rotation is not completely eliminated
in the free stream, a feature which has also been corroborated in two-dimensional flows by
Hudimoto and Tokuoka [64]. The prescription of m ¼ 0.1 and K ¼ 0.1 indicates a higher dynamic
viscosity and relatively lower micropolar vortex viscosity. The resistance to the flow is therefore
increased which induces translational momentum destruction and increased angular momentum,
manifesting in gyratory motion acceleration (increase in angular velocity, i.e., micro-rotation).
Figures 7 and 8 demonstrate the influence of homogeneous-heterogeneous reaction rate K1 and K2
on concentration of chemical species iðyÞ: The parameters K1 and K2 induce a significant reduction
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 21

in concentration values. K1 arises�in the species conservation (concentration boundary layer) equation
in the term, Sc −K1ið1 − ið^y ÞÞ2 : However, K2 features as a boundary condition, i0 ð0Þ ¼ K2ið0Þ:
Both parameters generate the same modification in profiles. The monotonic growth from the surface
to the maximum concentration in free stream is however more severely altered by homogenous reac­
tions (K1) rather than heterogeneous reaction rate (K2). The intensity of either chemical reaction
results in the conversion of the original reactant to a new species. There is therefore logically a
decrease in concentrations as a direct result of the “destructive” nature of the reactions. Asymptotically
smooth convergence of concentration profiles is sustained into the free stream, again confirming that
in the MATLAB computations we have deployed a sufficiently large infinity boundary condition. It is
important to note that multiple species may be present in organo-metallic solar nano-polymer coatings.
In the thermal manufacture of these materials both homogeneous and heterogeneous reactions are
known to arise. The homogeneous reaction is confined to the bulk of the fluid, whereas heterogeneous
reactions specifically flourish on a catalytic surface (the sheet, i.e., wall) [65, 66]. There is an inversion
in the curvature of profiles for the homogenous reaction case (Figure 7), whereas the growths from
the wall are essentially linear for the heterogeneous reaction case (Figure 8). Overall however both
reaction types impart a non-trivial influence on the concentration distributions.
Figures 9 and 10 are presented to inspect how the homogeneous-heterogeneous reactions rates, K1
and K2, influence the concentration at the surface, ið0Þ, of the micropolar nanofluid sheet with a
simultaneous variation in magnetic parameter M: Figure 9 shows that concentration at the surface
decreases weakly with an elevation in homogeneous reaction rate, K1, whereas it is reduced dramatic­
ally with greater magnetic parameter, M (an inverse relationship), with the strength of heterogeneous
chemical reaction fixed (K2 ¼ 0:1Þ:Figure10 however, indicates that the depletion in surface concen­
tration is much more profound with increasing heterogeneous reaction rate, K2, whereas only very
weak decrease is induced with an increase in magnetic parameter, M, this time with the homogeneous
reaction fixed (K1 ¼ 0:1Þ: This is related to the presence of K2 only in the wall solutal boundary con­
dition, as elaborated earlier and the presence of K1 in the concentration boundary layer equation.
The former therefore exerts a substantially greater impact on surface (wall) concentration (since auto-
catalysis is present only here) while the latter principally influences the concentration within the body
of the nanofluid. Similar impacts of surface (heterogeneous) reaction rate K2 and magnetic param­
eter M are observed on concentration profile at the surface in Figure 10.
Figures 11 and 12 present bar graphs to describe how the reactions parameters K1 and K2 influ­
ence the concentration profile at the surface with different magnetic field parameters (these graphs
basically visualize the results in Figures 9 and 10 in a different fashion but with much stronger mag­
netic parameters ranging from M ¼ 1, through 4 to 8). It is apparent, as elaborated earlier, that sur­
face concentration of reactant species declines with both increasing magnetic field strength as well as
with more intense homogeneous-heterogeneous reactions rates K1 and K2: The behavior computed
in Figures 9 and 10 is therefore sustained at even extremely high magnetic fields indicating that the
facility of deploying a transverse magnetic static field can consistently manipulate surface concentra­
tions of reactant species whether at low magnetic field strength (M < 1) or at much higher field inten­
sities (M > 1). A comparison of present and previously published results by Mahaptra [52] and Pop
[53] is shown in Table 5. An admirable agreement is noted between the both outcomes.

Table 5. Comparison table for several values of stretching ratio B ¼ a=c when / ¼ m ¼ M ¼ K ¼ 0:
F00 ð0Þ H0 ð0Þ −h0 ð0Þ
B ¼ a=c Present Mahapatra [52] Pop et al [53] Present Pop et al [53] Present Pop et al [53]
0:1 −0:96938 −0:9694 −0:96938 0:26341 0:26278 0:60215 0:60281
0:3 −0:84937 −0:84942 0:60631 0:60573 0:64728 0:64732
0:8 −0:29937 −0:29938 0:93472 0:93430 0:75709 0:75709
1:0 0:0 0:0 1:0 1:0 0:79788 0:79788
2:0 2:01750 2:01750 2:01750 1:16522 1:16489 0:97872 0:97872
3:0 4:72928 4:72930 4:72928 1:23465 1:23438 1:13209 1:13209
22 R. MEHMOOD ET AL.

6. Conclusions
Inspired by simulating multi-physio-chemical transport phenomena in non-Newtonian nanofluid
magneto-hydrodynamic materials manufacturing, a mathematical model is described for the time-
independent, nonaligned stagnation flow, heat and mass transfer in an electro-conductive magnetite
micropolar nanoliquid polymer impinging on an extending surface accompanied with chemical
reactions and temperature-dependent viscosity. Eringen’s robust micropolar model is employed to
simulate micro-rheological characteristics. Presented computational outcomes have revealed that:

I. Normal velocity component and species concentration are suppressed with larger magnetic
parameter, whereas micro-rotation and temperature are enhanced.
II. Concentration profile decreases with an increase in both homogenous and heterogeneous
reaction rate parameters.
III. Surface concentration demonstrates an inverse relationship with magnetic parameter and is
also significantly depressed with homogenous reaction rate parameter.
IV. Surface concentration is also strongly decreased with heterogeneous reaction rate parameter –
however the depletion with magnetic parameter is much weaker. This trend is sustained at
low magnetic fields or high magnetic fields.
V. Normal skin friction component is decreased with growing volume fraction, whereas there
is a slight initial elevation in tangential skin friction which is ensued with subsequent tan­
gential deceleration.
VI. Heat transfer gradient at the elongating surface (Nusselt number) is strongly lowered with
greater nano-particle volume fraction indicating an increase in thickness of thermal bound­
ary layer and a concomitant boost in temperatures in the body of the micropolar nanofluid.
VII. Both MATLAB R-K-F quadrature and Adomian Simulation Methodology (ADM) perform
very efficiently in the numerical solution of high order ordinary differential coupled nonlin­
ear boundayr value problems and offer significant promise for simulating multi-physico-
chemical magnetic non-Newtonian nanofluid manufacturing flows.

The current study has considered only magnetic fields. Future studies will also address the
dual action of mutually perpendicular electrical and magnetic fields on stagnation flows of micro­
polar nanofluids and will be presented imminently [67].

Disclosure statement
We declare that we have no conflict of interest regarding publishing this research work with any individual or
organization.

Funding
The authors extend their appreciation to the Deanship of Scientific Research at King Khalid University, Abha,
Saudi Arabia for funding this work through Large Groups Project under grant number RGP.2/536/44.

ORCID
Taseer Muhammad http://orcid.org/0000-0002-0838-2731

Data availability statement


Our article has no associated data.
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 23

References
0[1] E. A. Eastman and S. U. S. Choi, “Enhancing thermal conductivity of fluids with nanoparticles,” ASME Int.
Mech. Eng. Cong. Expos. vol. 11, pp. 1–8, 1995.
0[2] G. Xu, S. Zhao, X. Zhang, and X. Zhou, “Experimental thermal evaluation of a novel solar collector using
magnetic nano-particles,” Energy Conver. Manag., vol. 130, pp. 252–259, 2016. DOI: 10.1016/j.enconman.
2016.10.036.
0[3] P. Krajnik, F. Pusavec, and A. Rashid, “Nanofluids: properties, applications and sustainability aspects in mate­
rials processing technologies,” in Advances in Sustainable Manufacturing. Berlin, Heidelberg: Springer, 2011.
0[4] H. Schlichting, Boundary-Layer Theory. New York: MacGraw-Hill, 1979.
0[5] K. Hiemenz, “Die Grenzschict neinem in den gleichformigen flussigkeitsstrom eingetauchten geraden
Kreiszylinder,” Dingler’s Polytechnic J., vol. 326, pp. 321–410, 1911.
0[6] H. Hoshizaki, Y. S. Chou, N. G. Kulgein, and J. W. Meyer, Critical Review of Stagnation Point Heat
Transfer Theory, Lockheed Palo Alto Research Laboratory, Technical Report Tr-AFFDL-75-85, Palo Alto,
California 94304, July, 1975.
0[7] R. C. Rossi, Handbook of Thin Film Deposition Processes and Techniques, K.K. Schuegraf, Ed. Park Ridge,
New Jersey, USA: Noyes Publications, 1988.
0[8] N. Bachok, A. Ishak, and I. Pop, “The boundary layers of an unsteady stagnation-point flow in a nano­
fluid,” Int. J. Heat Mass Trans., vol. 55, no. 23–24, pp. 6499–6505, 2012. DOI: 10.1016/j.ijheatmasstransfer.
2012.06.050.
0[9] W. Ibrahim, B. Shankar, and M. M. Nandeppanavar, “MHD stagnation point flow and heat transfer due to
nanofluid towards a stretching sheet,” Int. J. Heat Mass Transf., vol. 56, no. 1–2, pp. 1–9, 2013. DOI: 10.
1016/j.ijheatmasstransfer.2012.08.034.
[10] O. A. B�eg, F. Mabood, and M. Nazrul Islam, “Homotopy simulation of nonlinear unsteady rotating nano­
fluid flow from a spinning body,” Int. J. Eng. Math., vol. 2015, pp. 1–15, 2015. DOI: 10.1155/2015/272079.
[11] B. J. Gireesha, B. Mahanthesh, I. S. Shivakumara, and K. M. Eshwarappa, “Melting heat transfer in bound­
ary layer stagnation-point flow of nanofluid toward a stretching sheet with induced magnetic field,” Eng.
Sci. Tech. Int. J., vol. 19, no. 1, pp. 313–321, 2016. DOI: 10.1016/j.jestch.2015.07.012.
[12] M. M. Rehman and T. Grosser, “Oblique stagnation-point flow of a nanofluid past a shrinking sheet,” Int.
J. Numerical Methods Heat Fluid Flow, vol. 26, pp. 189–213, 2016.
[13] M. Jashim Uddin, N. H. Yusoff, O. Anwar B�eg, and A. Ismail, “Lie group analysis and numerical solutions
for non-Newtonian nanofluid flow in a porous medium with internal heat generation,” Phys. Scr., vol. 87,
no. 2, pp. 025401, 2013. DOI: 10.1088/0031-8949/87/02/025401.
[14] O. Anwar B�eg, M. M. Rashidi, M. Akbari, and A. Hosseini, “Comparative numerical study of single-phase
and two-phase models for bio-nanofluid transport phenomena,” J. Mech. Med. Biol., vol. 14, pp. 1–31, 2014.
[15] P. Rana, R. Bhargava, O. A. B�eg, and A. Kadir, “Finite element analysis of viscoelastic nanofluid flow with
energy dissipation and internal heat source/sink effects,” Int. J. Appl. Comput. Math., vol. 3, no. 2,
pp. 1421–1447, 2016. DOI: 10.1007/s40819-016-0184-5.
[16] M. J. Uddin, O. A. B�eg, P. K. Ghose, and A. I. M. Ismael, “Numerical study of non-Newtonian nanofluid
transport in a porous medium with multiple convective boundary conditions and nonlinear thermal radi­
ation effects,” Int. J. Numer. Methods Heat Fluid Flow, vol. 26, no. 5, pp. 1526–1547, 2016. DOI: 10.1108/
HFF-03-2015-0123.
[17] S. Noreen, “Mixed convection peristaltic flow of third order nanofluid with an induced magnetic field,”
Plos One, vol. 8, no. 12, pp. 1–10, 2013. DOI: 10.1371/annotation/c18260c7-d0ff-4f79-82f7-55e9cf254f04.
[18] M. I. Khan et al., “Activation energy impact in nonlinear radiative stagnation point flow of Cross nanofluid,”
Int. Commun. Heat Mass Transf., vol. 91, pp. 216–224, 2018. DOI: 10.1016/j.icheatmasstransfer.2017.11.001.
[19] S. Nadeem, R. Mehmood, and N. S. Akbar, “Oblique stagnation point flow of a Casson-nano fluid towards
a stretching surface with heat transfer,” J. Comp. Theo. Nano., vol. 11, no. 6, pp. 1422–1432, 2014. DOI: 10.
1166/jctn.2014.3513.
[20] R. Mehmood, S. Saleem, S. Nadeem, and N. S. Akbar, “Flow and heat transfer analysis of Jeffery nanofluid
impinging obliquely over a stretched plate,” J. Taiwan Inst. Chem. Eng., vol. 74, pp. 49–58, 2017. DOI: 10.
1016/j.jtice.2017.02.001.
[21] A. C. Eringen, “Theory of micropolar fluids,” Indiana Univ. Math. J., vol. 16, no. 1, pp. 1–18, 1966. DOI:
10.1512/iumj.1966.16.16001.
[22] A. C. Eringen, Micro-Continuum Field Theories: II – Fluent Media. New York: Springer, 2001.
[23] O. A. B�eg, Numerical Simulation in Micropolar Fluid Dynamics. Germany: Lambert, 2011, pp. 280.
[24] F. S. Ibrahim, I. A. Hassanien, and A. A. Bakr, “Unsteady magneto-hydrodynamic micropolar fluid flow
and heat transfer over a vertical porous plate through a porous medium in the presence of thermal and
mass diffusion with a constant heat source,” Can. J. Phys., vol. 82, no. 10, pp. 775–790, 2004. DOI: 10.
1139/p04-021.
24 R. MEHMOOD ET AL.

[25] Y. Y. Lok, N. Amin, D. Campean, and I. Pop, “Steady mixed convection flow of a micropolar fluid near the
stagnation point on a vertical surface,” Int. J. Numer. Methods Heat Fluid Flow, vol. 15, no. 7, pp. 654–670,
2005. DOI: 10.1108/09615530510613861.
[26] M. I. A. Othman and S. A. Zaki, “Thermal instability in a rotating micropolar viscoelastic fluid layer under
the effect of electric field,” Mech. Mech. Eng., vol. 12, pp. 171–184, 2008.
[27] Z. Iqbal, R. Mehmood, and Z. Mehmood, “Thermal deposition on magneto-hydrodynamic nanofluidic
transport of viscoplastic fluid with microrotations,” J. Mol. Liquids, vol. 243, pp. 341–347, 2017. DOI: 10.
1016/j.molliq.2017.08.041.
[28] S. Nadeem, A. Rehman, K. Vajravelu, J. Lee and C. Lee, “Axisymmetric stagnation flow of a micropolar
nanofluid in a moving cylinder,” Math. Prob. Eng., vol. 2012, pp. 1–12, 2012.
[29] S. T. Hussain, S. Nadeem, and R. U. Haq, “Model-based analysis of micropolar nanofluid flow over a
stretching surface,” Euro. J. Phys. Plus, vol. 129, pp. 161–171, 2014.
[30] N. F. M. Noor, R. U. Haq, S. Nadeem, and I. Hashim, “Mixed convection stagnation flow of a micropolar
nanofluid along a vertically stretching surface with slip effects,” Meccanica, vol. 50, no. 8, pp. 2007–2022,
2015. DOI: 10.1007/s11012-015-0145-9.
[31] M. Waqas, M. Farooq, M. I. Khan, A. Alsaedi, T. Hayat, and T. Yasmeen, “Magneto-hydrodynamic (MHD)
mixed convection flow of micropolar liquid due to nonlinear stretched sheet with convective condition,”
Int. J. Heat Mass Transf., vol. 102, pp. 766–772, 2016. DOI: 10.1016/j.ijheatmasstransfer.2016.05.142.
[32] T. Hayat, I. Khan, M. Waqas, A. Alsaedi, and M. I. Khan, “Radiative flow of micropolar nanofluid account­
ing thermophoresis and Brownian moment,” Int. J. Hydrogen Energy, vol. 42, no. 26, pp. 16821–16833,
2017. DOI: 10.1016/j.ijhydene.2017.05.006.
[33] V. R. Prasad, S. A. Gaffar, and O. A. B�eg, “Heat and mass transfer of a nanofluid from a horizontal cylinder to
a micropolar fluid,” J. Thermophysics Heat Transfer, vol. 29, no. 1, pp. 127–139, 2015. DOI: 10.2514/1.T4396.
[34] R. U. Haq, S. Nadeem, N. S. Akbar, and Z. H. Khan, “Buoyancy and radiation effect on Stagnation point
flow of micropolar nanofluid along a vertically convective stretching surface,” IEEE Trans. Nanotech.,
vol. 14, no. 1, pp. 42–50, 2015. DOI: 10.1109/TNANO.2014.2363684.
[35] A. Latiff, M. J. Uddin, O. A. B�eg, and A. Izani Md. Ismail, “Unsteady forced bio-convection slip flow of a
micropolar nanofluid from a stretching/shrinking sheet,” Proc. IMECHE – Part N: J. Nanoeng. Nanosyst.,
vol. 230, no. 4, pp. 177–187, 2016. DOI: 10.1177/1740349915613817.
[36] S. M. D. Siva Reddy Sheri and O. A. B�eg, “Oscillatory dissipative conjugate heat and mass transfer in
chemically reacting micropolar flow with wall couple stress: a finite element numerical study,” J. Process
Mech. Eng., vol. 233, pp. 1–10, 2017.
[37] M. D. Shamshuddin, S. R. Mishra, O. Anwar B�eg, and A. Kadir, “Unsteady reactive magnetic radiative
micropolar flow, heat and mass transfer from an inclined plate with Joule heating: a model for magnetic
polymer processing,” J. Process Mech. Eng. Sci., vol. 23, pp. 1–10, 2018.
[38] M. J. Uddin, O. Anwar B�eg, A. Aziz, and A. I. M. Ismail, “Group analysis of free convection flow of a mag­
netic nanofluid with chemical reaction,” Math. Problems Eng., vol. 2015, pp. 1–11, 2015. DOI: 10.1155/
2015/621503.
[39] M. A. Chaudhary and J. H. Merkin, “A simple isothermal model for homogeneous-heterogeneous reactions in
boundary layer flow,” Fluid Dyn. Res., vol. 16, no. 6, pp. 311–333, 1995. DOI: 10.1016/0169-5983(95)00015-6.
[40] X. Pu and Y. Su, “Heterogeneous catalysis in microreactors with nanofluids for fine chemicals syntheses:
benzylation of toluene with benzyl chloride over silica-immobilized FeCl3 catalyst,” Chem. Eng. Sci.,
vol. 184, pp. 200–208, 2018. DOI: 10.1016/j.ces.2018.03.049.
[41] G. Ravi Kiran, G. Radhakrishnamacharya, and O. A. B�eg, “Peristaltic flow and hydrodynamic dispersion
of a reactive micropolar fluid: simulation of chemical effects in the digestive process,” J. Mech. Med. Biol.,
vol. 32, pp. 12–22, 2016.
[42] M. Sajid, S. A. Iqbal, M. Naveed, and Z. Abbas, ““Effect of homogeneous-heterogeneous reactions and mag­
netohydrodynamics on Fe3 O4 nanofluid for the Blasius flow with thermal radiations” J. Mol. Liquids,
vol. 23, pp. 1–12, 2017.
[43] T. Hayat, M. Farooq, and A. Alsaedi, “Homogeneous-heterogeneous reactions in the stagnation point flow of
carbon nanotubes with Newtonian heating,” AIP Adv., vol. 5, no. 2, pp. 1–12, 2015. DOI: 10.1063/1.4908602.
[44] T. Sajjad, R. U. Haq, and M. Usman, “Entropy generation and mixed convection of CuO–water near an
oblique stagnation point: modified Chebyshev wavelets approach,” Waves Random Complex Media, pp. 1–24,
2022. DOI: 10.1080/17455030.2022.2048121.
[45] R. U. Haq, T. Sajjad, M. Usman, and A. Naseem, “Oblique stagnation point flow of micropolar nanofluid
impinge along a vertical surface via modified Chebyshev collocation method,” Phys. Fluids, vol. 34, no. 10,
pp. 1–12, 2022. DOI: 10.1063/5.0099251.
[46] W. K. Usafzai, R. U. Haq, and E. H. Aly, “Wall laminar nanofluid jet flow and heat transfer,” HFF, vol. 33,
no. 5, pp. 1818–1836, 2023. DOI: 10.1108/HFF-09-2022-0528.
NUMERICAL HEAT TRANSFER, PART B: FUNDAMENTALS 25

[47] F. A. Soomro, M. Usman, S. El-Sapa, M. Hamid, and R. U. Haq, “Numerical study of heat transfer per­
formance of MHD Al2O3-Cu/water hybrid nanofluid flow over inclined surface,” Arch. Appl. Mech.,
vol. 92, no. 9, pp. 2757–2765, 2022. DOI: 10.1007/s00419-022-02214-1.
[48] R. Tabassum, R. Mehmood, and S. Nadeem, “Impact of viscosity variation and micro rotation on oblique
transport of Cu-water fluid,” J. Colloid Interf. Sci., vol. 501, pp. 304–310, 2017. DOI: 10.1016/j.jcis.2017.
04.060.
[49] E. A. Nada, “Effects of variable viscosity and thermal conductivity of CuO-water nanofluid on heat transfer
enhancement in natural convection: mathematical model and simulation,” ASME J. Heat Mass Transf.,
vol. 132, pp. 1–30, 2010.
[50] N. S. Akbar, D. Tripathi, Z. Khan, and O. A. B�eg, “A numerical study of magneto-hydrodynamic transport
of nanofluids from a vertical stretching sheet with exponential temperature-dependent viscosity and buoy­
ancy effects,” Chem. Phys. Lett., vol. 661, pp. 20–30, 2016. DOI: 10.1016/j.cplett.2016.08.043.
[51] W. A. Khan, O. D. Makinde, and Z. H. Khan, “Non-aligned MHD stagnation point flow of variable viscos­
ity nanofluids past a stretching sheet with radiative heat,” Int. J. Heat Mass Transf., vol. 96, pp. 525–534,
2016. DOI: 10.1016/j.ijheatmasstransfer.2016.01.052.
[52] T. R. Mahapatra and A. S. Gupta, “Heat transfer in stagnation-point flow towards a stretching sheet,” Heat
Mass Transf., vol. 38, no. 6, pp. 517–521, 2002. DOI: 10.1007/s002310100215.
[53] F. Labropulu, D. Li, and I. Pop, “Heat transfer in stagnation-point flow towards a stretching sheet,” Int. J.
Thermal Sci., vol. 49, no. 6, pp. 1042–1050, 2010. DOI: 10.1016/j.ijthermalsci.2009.12.005.
[54] G. Adomian, Solving Frontier Problems in Physics: The Decomposition Method. Dordrecht, USA: Kluwer, 1994.
[55] C. M. Liu, “Application of the Adomian decomposition method to oscillating viscous flows,” ACM, vol. 5,
no. 3, pp. 121–132, 2016. DOI: 10.11648/j.acm.20160503.15.
[56] A. M. SiddiquiH. Ashraf, T. Haroon, and A. Walait, “Analytic solution for the drainage of Sisko fluid film
down a vertical belt,” Appl. Appl. Math., vol. 8, pp. 465–470, 2013.
[57] O. Anwar B�eg, D. Tripathi, T. Sochi, and P. K. Gupta, “Adomian decomposition method (ADM) simula­
tion of magneto-bio-tribological squeeze film with magnetic induction effects,” J. Mech. Med. Biol., vol. 15,
pp. 1–20, 2015.
[58] F. S. Aski, “Application of Adomian decomposition method for micropolar flow in a porous channel,”
Propulsion Power Res., vol. 3, pp. 15–21, 2014.
[59] V. F. Adanhounm�e, F. de Paule Codo, and A. Adomou, “Solving the Navier Stokes flow equations of
micro-polar fluids by Adomian decomposition method,” BMSA, vol. 2, pp. 30–37, 2012. DOI: 10.18052/
www.scipress.com/BMSA.2.30.
[60] O. A. B�eg, “Multi-physical computational modelling of nanofluid bioconvection flows in spacecraft bio­
reactors”, Computational Approaches in Biomedical Nano-Engineering, Wiley-CVH, China, Chapter 5,
pp. 100–150, 2019.
[61] A. J. Chamkha and I. Camille, “Effects of heat generation/absorption and thermophoresis on hydromagnetic
flow with heat and mass transfer over a flat surface,” Int. J. Numerical Methods Heat Fluid Flow, vol. 10,
no. 4, pp. 432–449, 2000. DOI: 10.1108/09615530010327404.
[62] J. Zueco, O. A. B�eg, H. S. Takhar, and V. R. Prasad, “Thermophoretic hydromagnetic dissipative heat and
mass transfer with lateral mass flux, heat source, Ohmic heating and thermal conductivity effects: network
simulation numerical study,” Appl. Ther. Eng., vol. 29, no. 14–15, pp. 2808–2815, 2009. DOI: 10.1016/
j.applthermaleng.2009.01.015.
[63] K. C. Cramer and S. I. Pai, Applied Magneto-Fluid Dynamics for Engineers and Applied Physicists. New
York: MacGraw-Hill, 1973.
[64] B. Hudimoto and T. Tokuoka, “Two-dimensional shear flows of linear micropolar fluids,” Int. J. Eng. Sci.,
vol. 7, pp. 515–522, 1969.
[65] P. K. Kameswaran, S. Shaw, P. Sibanda, and P. V. S. N. Murthy, “Homogeneous-heterogeneous reactions in
a nanofluid flow due to a porous stretching sheet,” Int. J. Heat Mass Transf., vol. 57, no. 2, pp. 465–472,
2013. DOI: 10.1016/j.ijheatmasstransfer.2012.10.047.
[66] P. L. Chambr�e and A. Acrivos, “On chemical surfaces reactions in laminar boundary layer flows,” J. Appl.
Phys., vol. 27, no. 11, pp. 1322–1328, 1956. DOI: 10.1063/1.1722258.
[67] F. T. Zohra, M. J. Uddin, A. I. Ismail, and O. A. B�eg, “Bioconvective electromagnetic nanofluid transport
from a wedge geometry: simulation of smart electro-conductive bio-nano-polymer processing,” Heat
Transf. Asian Res., vol. 34, pp. 1–30, 2017.
26 R. MEHMOOD ET AL.

Appendix

The governing equations for the flow of an incompressible micropolar fluid in the absence of
body force and body couple, in a wave frame of reference, following Eringen [21, 22] are as
follows:
!
r � V ¼ 0, (A1)
! ! � �! !
q V � r V ¼ −rp þ jr � W þ ðl þ jÞr2 V , (A2)
! �! � �! ! �! � � �! �
qj V � r W ¼ −2j W þ jr � V − c r � r � W þ a0 þ b0 þ c r r � W : (A3)
! �!
where V ¼ ðv, 0, uÞ and W ¼ ð0, x, 0Þ are the velocity vector and the micro-rotation vector,
respectively, p is the fluid pressure, q is the micropolar fluid density j is the micro-gyration
(Eringen’s micro-inertia density parameter). The constants l, j, a0 , b0 and c are material constants
which satisfy the following inequalities: 2l þ j � 0, j � 0 3a0 þ b0 þ c � 0, c � jb0 j: Since the
�!
micro-rotation vector W is solenoidal, a0 , b0 do not appear in the governing equations.

You might also like