Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Study of incompressible MHD flow in a circular pipe

with transverse magnetic field using a spectral/finite


element solver.
X. Dechamps ∗ and G. Degrez †

Université Libre de Bruxelles, Av. F-D. Roosevelt 50 (CP 165/41), Brussels, Belgium

M. Rasquin
Argonne Leadership Computing Facility, Argonne National Laboratory, Argonne (IL), USA
§
K.E. Jansen
Department of Aerospace Engineering, University of Colorado Boulder, Boulder (CO), USA

This work is concerned with the numerical study of an unsteady and incompressible
flow of an electrically conducting liquid metal inside a pipe of circular cross-section at low
Reynolds number. In particular, the aim of this work is to understand better the influence
of an external transverse magnetic field on the flow field and the transition between a
fully turbulent regime to a fully laminar regime at moderate Reynolds number as the
Hartmann increases. For that purpose, direct numerical simulations have been carried out
and provide mean velocity profiles, turbulence intensity profiles, physical and numerical
dissipation profiles and skin friction data. The obtained results show the damping of
turbulence due to the action of the magnetic field and the development of anisotropy in
the flow.

Nomenclature

µ0 magnetic permeability of vacuum


ν kinematic viscosity
φ electric potential
ρ density of the fluid
σ electrical conductivity of the fluid
B external magnetic field
J current density
u velocity field = (uz , ur , uθ )
Ha Hartmann number
p kinematic pressure P/ρ
Re hydrodynamic Reynolds number
Rem magnetic Reynolds number
ub mean streamwise velocity

∗ Graduate student, department of Aero-Thermo-Mechanics, Université Libre de Bruxelles, Av. F-D. Roosevelt 50 (CP

165/41), Brussels, Belgium, AIAA member.


† Professor, department of Aero-Thermo-Mechanics, Université Libre de Bruxelles, Av. F-D. Roosevelt 50 (CP 165/41),

Brussels, Belgium, AIAA member.


‡ Research associate, Argonne Leadership Computing Facility, Argonne National Laboratory, Argonne (IL), USA.
§ Professor, department of Aerospace Engineering, University of Colorado Boulder, Boulder (CO), USA.

1 of 13

American Institute of Aeronautics and Astronautics


I. Introduction
he study of duct flow for electrically conducting liquid metal fluids subject to an externally applied
T magnetic field is acknowledged to be a good approach for a better understanding of the fundamental
properties of magnetohydrodynamic (MHD) turbulence. This fundamental type of flow is encountered fre-
quently in industrial processes such as the casting/stirring of steel or the Czochralski process used to obtain
single crystals of semiconductors1, 2 .
The experimental study of the MHD pipe flow began in 1937 with Hartmann3. In the 60’s-70’s, numerous
experiments4, 5, 6, 7, 8 were conducted on several configurations (non conductive wall or with finite conductivity
- transverse or longitudinal magnetic field). In the late 90’s and early 2000’s, numerical results began to
appear9, 10, 11. Finally, in 2008, additional experimental results were obtained12 for a Reynolds number
Reb = 11, 300. Overall, little information is available concerning the properties of turbulence in the case of a
non-conducting pipe flow under a transverse magnetic field. The available results are generally limited to the
velocity profiles, velocity fluctuations and global skin friction coefficient. Up to now, no relevant data about
the physical characteristics of turbulent structures are available and the few available numerical results are
restricted to low Reynolds numbers.
In contrast, many more results were obtained in the case of a flow inside a channel13, 14, 15, 16, 17, 18 or
a duct with rectangular/square cross-section19, 20, 21, 22, 23, 24 . This is linked with the greater choice of the
electrical conductivity of the walls and the different resulting physical phenomena. Indeed, through a given
combination of electrically conducting/non-conducting walls, jets can for instance appear along sidewalls.
These jets are thus source of flow instabilities that do not occur (or not so easily) in the case of the pipe
flow. Due to this increasing interest of the scientific community to duct flows, numerical simulations at
higher Reynolds numbers were performed. Ref. 24 for instance considered a Reynolds number equal to 105
and showed a quasi-laminar core in the center of the square duct and in the Hartmann layers, with still a
turbulent flow near the walls parallel to the magnetic field (Shercliff layers), a flow pattern that still remains
to be observed in the case of pipe flows.
One main objective of this study is to improve our understanding of turbulent MHD pipe flows subject
to a uniform transverse magnetic field and the transition between a fully turbulent regime to a fully laminar
regime at moderate Reynolds numbers as the Hartman number increases. The bulk Reynolds number Re is
chosen equal to 8000 and the Hartmann number Ha ranges between 0 and 100. This range allows to cover
the turbulent and laminar regimes with the laminarization process being linked with the Joule dissipation.
The physical model is discussed in section II, whereas the numerical method is presented in section III. The
numerical results obtained from direct numerical simulations are discussed in section IV and include the
velocity profiles, velocity fluctuations, skin friction coefficient and physical/numerical dissipations. Finally,
the conclusions are presented in section V.

II. Physical model


This work is concerned with the unsteady, incompressible and
isothermal flow of an electrically conducting and Newtonian fluid
(i.e. a liquid metal) inside a pipe of circular cross-section and di- Lz
ameter d = 2R (see figure 1). An external, constant and uniform r
magnetic field B = B ex is applied along the transverse direction x d = 2R
(θ = 0). The region near the wall in the direction parallel to B is ! z
y
referred to as the Hartmann layer and the one in the direction per-
pendicular to the magnetic field is referred to as the Roberts layer. x z
The wall of the cylinder is solid, smooth and electrically insulated.
Periodic conditions are imposed at the inlet (z = 0) and the outlet B
(z = Lz ) of the domain. Under the flow conditions specified later,
we observe that the magnetic Reynolds number Rem = µ0 σub d  1. Figure 1. Geometry of the domain and
Under this condition, the inductionless form of the MHD equations definition of the axes.
described by the set of equations Eqs. (1) can be used. As a conse-
quence of the assumption Rem  1, the contribution of the induced magnetic field is negligible in comparison
with the external magnetic field B. Another way to interpret this assumption is that the velocity field is
influenced by the magnetic field but the opposite is not true. Eqs. (1) are respectively the continuity equa-

2 of 13

American Institute of Aeronautics and Astronautics


tion, the momentum equations, the conservation of charge ∇ · J = 0 and the definition of the current density
through Ohm’s law. The validity of this model has been confirmed in numerous works (see e.g. Refs. 2, 25).
Additionally, numerical results of Knaepen26 show that no significant deviations are observed for Rem ≤ 1
between the inductionless model and the full set of MHD equations. The unknowns of the problem are
the components of the velocity field u, kinematic pressure p = P/ρ and the electric potential φ. In the
momentum equations, the Lorentz force fL = J × B creates a link between the flow and the magnetic field.
This force is generally opposed to the flow and tends to flatten the velocity profile.

Rcont = ∇ · u = 0 (1a)
∂u J×B
Rmom = + (u · ∇) u + ∇p − ν∇2 u − =0 (1b)
∂t ρ
Rcharge = ∇2 φ − ∇ · (u × B) = 0 (1c)
J = σ (−∇φ + u × B) (1d)

The set of boundary conditions specified in Eqs.(2) are applied along the wall in order to close the
problem (n stands for the normal to the wall). Regularity conditions are imposed on the axis (r = 0) for
velocity, pressure and electric potential (see section III.A).

~u(z, r = R, θ) = 0 (2a)
∂φ
(z, r = R, θ) = 0 (2b)
∂n
The non-dimensional parameters that characterize the problem are
Inertial forces ub d
Reb = = Reynolds number (3a)
Viscous forces ν r
Lorentz force σ
Ha = = BR Hartmann number (3b)
Viscous forces ρν

III. Numerical method


The unsteady equations (1) are solved by the in-house hybrid Spectral/Finite Element LES algorithm
(SFELES). The spatial discretization in SFELES is based on a stabilized finite element method in 2D planes,
whereas a collocated spectral method describes the problem in the out-of-plane direction. The stabilization
of the 2D finite element method is ensured by the traditional Streamline Upwind Petrov-Galerkin (SUPG)
and Pressure-Stabilized Petrov-Galerkin (PSPG) formulation27, 28 . The main assumption in SFELES relies
therefore on a direction of periodicity in the flow and can simulate 3D flows associated with complex planar
of axisymmetric geometries. Several benefits result from this combination of discretizations:
• The in-plane finite element method is well suited for unstructured meshes of complex 2D geometries;
• Memory requirements are reduced because only a single 2D mesh needs to be stored (connectivity,
coordinates, etc);
• As it will be demonstrated below, a pseudo-spectral treatment of the non-linear terms leads to a
decoupling of the discretized equations for each Fourier mode. For each time step, a set of independent
2D linear systems are solved in Fourier space, which is a natural way to introduce a first level of
parallelism.
This pseudo-spectral approach implies the evaluation of the non-linear terms in the physical domain and their
transfer back to Fourier space with a Direct Fourier Transform. In order to avoid aliasing problems during
the Direct Fourier Transforms, only Ncomp = 2/3N modes are kept and effectively computed in Fourier
space. Finally, the full 3D solution in physical space is reconstructed from an Inverse Fourier Transform
applied to the Fourier modes computed in Fourier space. Although 1/3 of the modes are thrown out during
the Direct Fourier Transform, the solution in physical space is evaluated at N locations in the periodic
direction (i.e. N 2D planes).

3 of 13

American Institute of Aeronautics and Astronautics


III.A. Spatial discretization
Considering the axisymmetric version of SFELES, the unknowns q = {uz , ur , uθ , p, φ} in the planes θ = cst
PNnodes
are represented by linear, triangular elements q(z, r, θ, t) = j=1 qj (θ, t)Nj (z, r) where the index j runs
over all Nnodes nodes in the 2D in-plane finite element mesh. The second-order accurate discretization
in space of Eq. (1a)-(1c) is performed in the 2D finite element planes according to Eqs. (4) where ∇ ˜ =
( ∂/ ∂z, ∂/ ∂r, 0), ũ = (uz , ur , 0) and dΩ = dr dz:
Z  
˜ i · Rmom dΩ = 0
r Ni Rcont + τpspg ∇N (4a)
ZΩ  
˜ i Rmom dΩ = 0
r Ni + τsupg ũ · ∇N (4b)
ZΩ
rNi Rcharge dΩ = 0 (4c)

The definitions of the stabilization parameters are constant over each triangular element and are inspired
from Ref. 27, 28:
 −1/2  −1/2
1 1 1 1 1 1 1 1
τpspg = + + + τsupg = + + + (5a)
τt2 2
τc,p 2
τν,p τJ2 τt2 2
τc,u 2
τν,u τJ2
where
∆t ρ
τt = τJ = (5b)
2 σB 2 r
he,p h2e,p 4Ωe
τc,p = τν,p = he,p = (5c)
2U 4ν π
he,u h2e,u 2|Uc |
τc,u = τν,u = he,u = P3 (5d)
2|Uc | 4ν ˜ j|
|Uc · ∇N
j=1

with Ωe the area of the element and Uc the mean planar component of the velocity inside the element. The
discretized Eq. (4) can be written in the more compact form described in Eq. (6). For the sake of brevity, the
matrices M1 , M2 , M3 , L1 , L2 and L3 are not detailled in this paper (see Ref. 29 for more details about their
hydrodynamic components). The mass matrices M1 , M2 and M3 are linked to the time-derivative dependent
terms. The matrices L1 , L2 and L3 contain all the linear terms of the Navier-Stokes equations. The right-
hand side terms Cu , Cp and Cφ contain the non-linear terms that are treated through the pseudo-spectral
approach.
     
u u Cu
∂2 ∂  ∂2 
   
∂ ∂
M1 + M2 + M3 2  p  + L1 + L2 + L3 2  p  = −  Cp  (6)
   
∂θ ∂θ ∂t ∂θ ∂θ
φ φ Cφ
The discretization in the θ-direction is introduced by means of a discrete sum of N Fourier modes
N
−1
2
1 X 2π
q(z, r, θ, t) = q̂k (z, r, t) eIk θmax θ (7)
N
k=− N
2 +1

where θmax is the maximal azimuthal coordinate of the domain in the direction of periodicity (θmax = 2π
for domain with a symmetry of revolution such as in this study). The last step of the spatial discretization
consists in applying a Direct Fourier Transform to Eq. (6), which leads to a set of independent systems of
equations in Fourier space for the unknowns q̂k :
 
"    2 # ûk
2π 2π ∂ 
M1 + M2 Ik + M3 Ik  p̂k  +

θmax θmax ∂t
φˆk
   
"    2 # ûk Ĉu
2π 2π
L1 + L2 Ik + L3 Ik  p̂k = − Ĉp  (8)
   
θmax θmax
φ̂k Ĉφ

4 of 13

American Institute of Aeronautics and Astronautics


or in a more compact form:
∂ q̂k  
Mk + Lk q̂k = −Ĉ q− N +1 , ..., q N (9)
∂t 2 2

where qp is the unknown vector in physical space associated with the p-th finite element plane. We clearly
observe here the decoupling of each matrix system for the Fourier modes. The right-hand side, composed
of the non-linear terms, couples all the Fourier modes together. For each mode, the matrices Mk and Lk
are constant in time but differ for each Fourier mode. Once the linear systems presented in Eq. (9) are
assembled, both the real and complex components of the Fourier coefficients q̂k need to be solved. As the
physical solution is purely real, the following symmetry appears in Fourier space:

q̂−k = [q]∗k (10)

so that only the first N/2 + 1 modes need to be solved, the other ones being evaluated through the symmetry
relation of Eq. (10). Furthermore, modes 0 and N/2 are both purely real. They are combined in a single
complex-valued linear system so that the size of every systems is always identical for a load balance point of
view and only N/2 linear systems need to be solved in practice.
To impose boundary conditions on the axis (r = 0) is mathematically not compulsory because this edge
is of artificial nature. However, the regularity conditions summarized in table 1 are applied in order to
obtain a better quality results29 . The terms between parentheses are not treated explicitly as they are
treated implicitly through an integration by parts with the finite element method described in the previous
paragraph. Because of the form of the regularity condition for the radial and azimuthal components of
velocity of mode k = 1, the r- and θ- momentum equations are replaced by the linear combination [r-
momentum]±I[θ-momentum]. The regularity condition ûr,1 + I ûθ,1 = 0 replaces the first linear combination
for all the nodes lying on the axis.

Mode ûz,k ûr,k ûθ,k p̂k φ̂k


0 (∂ ûz,0 /∂r = 0) ûr,0 = 0 ûθ,0 = 0 (∂ p̂0 /∂r = 0) (∂ φ̂0 /∂r = 0)
1 ûz,1 = 0 ûr,1 + I ûθ,1 = 0 p̂1 = 0 φ̂1 = 0
k 6= 0, 1 ûz,k = 0 ûr,k = 0 ûθ,k = 0 p̂k = 0 φ̂k = 0
Table 1. Axis boundary conditions.

III.B. Temporal integration


At this point, we still need to perform a temporal integration on time-derivative dependent terms in Eq. (9).
An implicit second-order accurate Crank-Nicolson scheme is used for the linear terms whereas explicit
schemes (either Adams-Bashforth or Runge-Kutta with 1 up to 4 stages) are chosen for the non-linear
convective and stabilization terms. The family of temporal integrators available in SFELES allows us to
keep the numerical dissipation at a low value while the numerical stability of the convective terms is con-
trolled. Finally, it is found more convenient to solve the variations in time of the unknowns δ q̂k = q̂kn+1 − q̂kn .
The combination of the spatial discretization and temporal integration boils down to the form:
  S−1
Mk Lk X
+ δ q̂k = −Lk q̂kn − βs Ĉ s (11)
∆t 2 s=1
| {z } | {z }
Ak bk

where βs and S depends on the chosen time integrator for the non-linear terms.

III.C. Parallel implementation


To exploit the advantages of the combined spectral/finite element discretization, three levels of parallelisms
coexist in SFELES.

• First, the physical domain is partitioned by METIS30 in as many parts as the total number of processes
launched for the computation. Each process assembles its own local matrices Ak and right-hand side

5 of 13

American Institute of Aeronautics and Astronautics


bk for all the required modes. Doing so, the direct and inverse Fourier transforms are performed locally
and do not require any communication between processes.
• As the matrices Ak in Eq. (11) are constant in time, one can factorize each matrix Ak in LU factors
independently from the other matrices. Similarly, as the linear systems associated with each Fourier
mode in Eq. (11) are independent from each other within each time step, the substitution phase for
one linear system does not require any information from the other systems and can be performed
independently.
• The last and most recent level of parallelism is related to the ability to perform the factorization and
substitution phases for each mode on several processes thanks to the parallel direct solver MUMPS31, 32.

IV. Numerical results


The numerical results obtained for a hydrodynamic Reynolds number Re pb = ub d/ν = 8000 based on
the bulk velocity and pipe diameter and a Hartmann number 0 ≤ Ha = BR σ/ρν ≤ 100 are presented in
this section. A time-varying forcing term in the z-momentum equation ensures a constant Reb during the
whole computation. The length of the pipe is set to Lz = 5d. The present computations are performed
with 250 × 100 × 128 nodes/modes in respectively the axial, radial and azimuthal directions for a total of 3.2
million nodes. The nodes of the bidimensional structured finite element mesh are uniformly spaced in the
axial direction, whereas the mesh is refined near the wall of the pipe in the radial direction. In viscous wall
units (ν/uτ ), the grid spacing is ∆r+ ≈ 0.09 at the pipe wall, ∆r+ ≈ 6.6 near the centreline of the pipe and
∆z + ≈ 5.3. The circumferential gridspacing is (d∆θ/2)+ ≈ 8.5 near the wall. These characteristics of the
mesh are quite similar to those of Ref. 33, 34. The time integration step is equal to ∆t = 10−4 ≈ 1.3 10−5 t∗ ,
where t∗ is the dimensionless time scale defined as t∗ = d/uτ . The flow is computed until it reaches a fully
developed state (until t = 10t∗ ). From t = 10t∗ , the statistics were gathered until t = 14t∗ and result from a
spatial average along the axial direction and a time average over each time step. Each computation required
a total CPU-time of 50.000 CPU-hours. Table 2 provides a summary of the flow conditions.
Table 2. Mean flow configurations for the numerical simulations.
p
Ha= BR σ/ρν * 0 5 10 15 20 30 100

Reb = ub d/ν 8000 8000 8000 8000 8000 8000 8000
Reτ = uτ d/ν † 517 510.5 494.8 421 477.2 572.7 1015.7
* Hartmann number with B the applied magnetic field, R the radius of the pipe, σ
the electrical conductivity of the fluid, ρ the density of the fluid and ν the kinematic
viscosity of the fluid.
† Hydrodynamic Reynolds numbers with d the diameter of the pipe, u the average
b
velocity and uτ the wall shear velocity defined by the wall shear stress τw = ρu2τ .

IV.A. Instantaneous solution


The general pattern of the solution is presented in figure 2 with the instantaneous streamwise velocity
component uz in a cross-section at z = Lz /2 and in the meridional x − z plane. We clearly see that for
Ha = 0, 5 and 10, the regime is turbulent and from Ha = 15 the flow is entirely laminarized. For the
turbulent cases, the core is flattened when the Hartmann number is increased. This laminarization process
first occurs in the core and the Hartmann layer as clearly shown in figure 2(h). This laminarization process
in the core of the pipe and the Hartmann layers is also highlighted in figure 3, where turbulent structures
vanish in the plane parallel to B as Ha increases. The Roberts layers are less affected by the increasing
Hartmann number than the Hartmann layers. For the considered Reynolds number Reb = 8000, we do
not obtain the quasi-laminar state observed with a Reynolds number Reb = 105 in a square duct24 . From
Ha = 15, the turbulent structures completely disappear and we retrieve the usual laminar distribution of
velocity predicted for low Hartmann numbers and by an asymptotic approach35, 36 . It is interesting to note
that the transition between a partially turbulent and a fully laminar flow is brutal, as a slight increase of
the magnetic field between the case Ha = 10 and Ha = 15 is sufficient to completely laminarize the flow.

6 of 13

American Institute of Aeronautics and Astronautics


Y

(a) Ha = 0 (b) Ha = 5 (c) Ha = 10

(d) Ha = 15 (e) Ha = 100

(f) Ha = 0 (g) Ha = 5

(h) Ha = 10 (i) Ha = 15

Figure 2. Instantaneous velocity field (a) to (e) in cross-section x − y located at z = Lz /2 and (f ) to (i) in the meridional
x-z plane (parallel to B). All the plots have the same contour level ranging from 0 (blue) to 5.03 (red).

7 of 13

American Institute of Aeronautics and Astronautics


(a) (b)

Figure 3. Instantaneous isosurfaces of Q-criterion for (a) Ha = 0 and (b) Ha = 10. These isosurfaces are coloured by
the norm of velocity.

IV.B. Mean flow and turbulent fluctuations


The time and streamwise-averaged uz velocity profile is shown in figure 4 in the direction parallel and
perpendicular to the external magnetic field B. An obvious consequence of an increasing Hartmann number
is the flattening of the velocity profile in the core of the pipe in the direction parallel to B. In this direction,
the flow becomes nearly r-independent outside the boundary layers. The MHD turbulent profiles (Ha = 5
and Ha = 10) are slightly less rounded than their hydrodynamic counterpart. This alteration is more visible
in the direction parallel to B because of the so-called ‘Hartmann effect’2 . In the direction perpendicular
to the magnetic field, the core velocity slightly increases between Ha = 0 and Ha = 5, then decreases for
Ha = 10. As explained in Ref. 24, this behaviour is attributed to the reduction of the wall-normal turbulent
transport of momentum. The decreasing core velocity is further observed with the laminar results. This is
linked with the flattening of the velocity profile.

1.2 1.2

1 1

0.8 0.8
uz / ub

uz / ub

0.6 0.6
Ha = 0 Ha = 0
Ha = 5 Ha = 5
0.4 Ha = 10 0.4 Ha = 10
Ha = 15 Ha = 15
Ha = 20 Ha = 20
0.2 0.2
Ha = 30 Ha = 30
Ha = 100 Ha = 100
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Radius r / R Radius r / R

(a) (b)

Figure 4. Mean uz streamwise component of velocity in a direction parallel (a) and perpendicular (b) to B. The
averages are performed over time and along the streamwise coordinate z. The velocity profiles are normalized by the
bulk velocity ub .

The velocity fluctuations are represented by the root-mean-square values u2ij,rms = ui uj − ui uj where
(. . .) is the temporal average operator. Figure 5 shows the distribution of the fluctuations of each velocity
component as well as the Reynolds-stress components uzr,rms . The results are non-dimensionalized by the
friction velocity uτ . In these cuts, we represent only the turbulent cases because the fluctuations are equal
to zero for laminar cases. An obvious observation is the damping of turbulence with an increasing Hartmann
number. We can see that this damping is slightly more pronounced in the direction parallel to B than in
the direction perpendicular to B.

8 of 13

American Institute of Aeronautics and Astronautics


3 3
Ha = 0 Ha = 0
Ha = 5 Ha = 5
2.5 Ha = 10 2.5 Ha = 10

2 2
u+z,rms

u+z,rms
1.5 1.5

1 1

0.5 0.5

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Radius r / R Radius r / R
1 1

0.9 0.9

0.8 0.8

0.7 0.7

0.6 0.6
u+r,rms

u+r,rms
0.5 0.5
Ha = 0 Ha = 0
0.4 Ha = 5 0.4 Ha = 5
0.3
Ha = 10 0.3
Ha = 10

0.2 0.2

0.1 0.1

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Radius r / R Radius r / R

1 1

0.8 0.8
u+θ,rms

u+θ,rms

0.6 0.6
Ha = 0
Ha = 5
0.4 Ha = 10 0.4

Ha = 0
0.2 0.2
Ha = 5
Ha = 10
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Radius r / R Radius r / R
0.8 0.8
Ha = 0 Ha = 0
0.7
Ha = 5 0.7
Ha = 5
Ha = 10 Ha = 10
0.6 0.6

0.5 0.5
u+zr,rms

u+zr,rms

0.4 0.4

0.3 0.3

0.2 0.2

0.1 0.1

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Radius r / R Radius r / R
Figure 5. Mean velocity fluctuations in a direction parallel (left) and perpendicular (right) to B. The averages are
performed over time and along the streamwise coordinate z.

9 of 13

American Institute of Aeronautics and Astronautics


IV.C. Friction coefficient
The wall shear stress τw = ρu2τ defined in table 2
is computed using the global wall-normal derivative 35 * 10−3

of the mean axial velocity. In figure 6, the global Satake (2002) − Reb=5300
skin friction coefficient Cf = τw /ρu2b /2 is repre-

Friction coefficient Cf = τw/ρ u2b/2


30
Gardner & Lykoudis (1971) − Reb=25000
sented as a function of the ratio Ha/Reb . For the
Loeffler & co. (1969) − Reb=40000
sake of comparison, we also provide numerical re- 25
Present work − Reb=8000
sults of Satake11 and experimental values of Gardner
7 6 20
& Lykoudis and Loeffler et al. The skin friction
coefficient tends to decrease in the turbulent regime
(2Ha/Re 104 ≤ 25) and then to increase linearly 15

4
in the laminar regime (2Ha/Re 10 ≥ 37.5). This
10
evolution of the skin friction coefficient is a combi-
nation of two effects: the steeper velocity profile in 1,271 (Ha/Reb)+0,000242
5
the Hartmann layer in the laminar regime and the
damping of turbulence in the turbulent regime. The
0
fact that the flow at 2Ha/Re 104 = 37.5 in Ref. 11 0 50 100 150 200 250
2 Ha / Reb x 104
is still in the turbulent regime is maybe an indicator
that our result at Ha = 15 is not fully reliable. In- Figure 6. Skin friction coefficient Cf = ρuτ2w/2 as a function
b
deed, as underlined in Ref. 37, the transition point of the ratio Re Ha
b
.
between turbulent and laminar case covers a rela-
tively wide range of Ha/Re. Several factors can
influence the transition: the periodic length along the streamwise direction, the regime of the initial con-
dition, etc. In the purely hydrodynamic case (Ha = 0), the skin friction coefficient was found to be very
−1/4
close to the analytical value obtained by Blasius’ law Cf = 0.079Reb . In the laminar regime, the slope of
the line is slightly larger than the analytical slope of 3π/8 obtained by Shercliff.38 This discrepancy is also
observed for the other numerical/experimental values but no satisfactory explanation has been found so far.

IV.D. Physical and numerical dissipations


One main advantage of the finite element method is the ability to compute a posteriori the energy budget
given by Eq. (12a). The variation of the kinetic energy e is compensated by the viscous and Joule dissipations
(ν and J ), by the introduction of energy through the source term f that maintains a constant bulk Reynolds
number Reb and the contribution of the PSPG and SUPG stabilization terms stab .
∂e
= −ν − J + f − stab (12a)
∂t
where
u·u
Z
1
e= dV (12b)
V V 2
Z
1
ν = ν∇u : ∇u dV ≥ 0 (12c)
V V
J·J
Z
1
J = dV ≥ 0 (12d)
V V ρσ
Z
1
f = u · f dV (12e)
V V

All these contribution are shown in figure 7. Concerning the viscous and Joule dissipations, we notice
the appearance of peaks of dissipation in the Hartmann layers. These peaks are more important for high
Hartmann numbers because of the narrowing Hartmann layer. The narrowing Hartmann layer causes steeper
velocity gradients and narrower passages for the electric current. In contrast, the evolution of the dissipations
is smooth in the Roberts layers. Concerning the contributions of the stabilization terms in the energy
budget, we see that their amplitude is 2 up to 3 orders of magnitude lower than the physical dissipations.
The additional numerical stabilization terms do not alter the physics of the solution and allow a smooth
evolution of the solution without any spurious oscillations.

10 of 13

American Institute of Aeronautics and Astronautics


−7
x 10 −7
8 x 10
4
Ha = 0
7
3.5 Ha = 0
Ha = 5
Ha = 5
6 Ha = 10
3 Ha = 10
Ha = 15
5
Ha = 15
Ha = 20 2.5
Ha = 20
εviscous

εviscous
Ha = 30
4 2 Ha = 30
Ha = 100
Ha = 100
3 1.5

2 1

1 0.5

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Radius r / R Radius r / R
−7 −8
1
x 10
4
x 10 Ha = 0
Ha = 0 Ha = 5
0.9
Ha = 5 3.5 Ha = 10
0.8
Ha = 10 Ha = 15
3
0.7 Ha = 15 Ha = 20
Ha = 20 2.5
Ha = 30
0.6
Ha = 30 Ha = 100
εJoule

εJoule

0.5 2
Ha = 100
0.4
1.5
0.3
1
0.2
0.5
0.1

0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Radius r / R Radius r / R
−10 −10
x 10 x 10
14 18
Ha = 0 Ha = 0
16 Ha = 5
12 Ha = 5
Ha = 10 14 Ha = 10
10 Ha = 15
Ha = 15 12
Ha = 20 Ha = 20
8
10 Ha = 30
Ha = 30
εSUPG
εPSPG

6 Ha = 100 8 Ha = 100
6
4
4
2
2
0
0

−2 −2
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Radius r / R Radius r / R
Figure 7. Mean physical and numerical dissipations in a direction parallel (left) and perpendicular (right) to B. The
averages are performed over time and along the streamwise coordinate z. The numerical dissipations are shown only
in the direction parallel to B.

11 of 13

American Institute of Aeronautics and Astronautics


V. Conclusion
The present work consisted of a numerical study of isothermal, turbulent MHD pipe flow in a transverse,
uniform and constant magnetic field. A numerical algorithm was developed for this purpose. At low Reynolds
number, we did not observe the particular pattern of quasi-laminar flow observed at higher Reynolds number
in a square duct.24 The transition between turbulent and laminar regimes was found to occur around
Re/2Ha ∼ 250, which is in good accordance with previous works7, 11 .
Velocity profiles in the direction parallel to the magnetic field flatten as the Hartmann number increases.
Along the direction perpendicular to the magnetic field, the profiles became more rounded than the zero-field
profile. We observed that the laminarization process in the turbulent regime starts first in the core and the
Hartmann layers. This observation was made on the basis of the velocity fluctuations.
The skin friction coefficient has been found to be in good agreement with the literature. As the Hartmann
number increases, the friction tends to decrease in the turbulent regime and then to increase linearly in the
laminar regime.
The viscous and Joule dissipations show why the laminarization process occurs first in the Hartmann
layers and the core of the pipe. The combination of the two energy-dissipating processes is more pronounced
in these zones. It was also proved that the contribution of the stabilization terms in the numerical scheme
is negligible with respect to the viscous and Joule dissipations and does not alter the physics of the flow.

Acknowledgments
This work was supported by a grant from the Belgian Fonds de la Recherche Scientifique de Belgique
(FRS-FNRS). This work utilized the Janus supercomputer, which is supported by the National Science Foun-
dation (award number CNS-0821794) and the University of Colorado Boulder. Computational resources have
also been provided by the supercomputing facilities of the Université Catholique de Louvain (CISM/UCL)
and the Consortium des Équipements de Calcul Intensif en Fédération Wallonie Bruxelles (CÉCI) funded
by the Fonds de la Recherche Scientifique de Belgique (FRS-FNRS).

References
1 Davidson, P., “Magnetohydrodynamics in materials processing,” Annual Review of Fluid Mechanics, Vol. 31, No. 1, 1999,

pp. 273–300.
2 Davidson, P., An introduction to Magnetohydrodynamics, Cambridge University Press, Cambridge, 2001.
3 Hartmann, J., “Hg dynamics. I - Theory of the laminar flow of an electrically conductive liquid in a homogeneous magnetic

field,” Det Kgl. Danske Videnskabernes Selskab. Mathematisk-Fysiske Meddelelser , Vol. XV, No. 6, 1937, pp. 1–28.
4 Ihara, S., Tajima, K., and Matsushima, A., “The flow of conductive fluids in circular pipes with finite conductivity under

uniform transverse magnetic fields,” Journal of Applied Mechanics, Vol. 34, No. 1, 1967, pp. 29–36.
5 Fraim, F. and Heiser, W., “The effect of a strong longitudinal magnetic field on the flow of mercury in a circular tube.”

Journal of Fluid Mechanics, Vol. 33, 1968, pp. 397–413.


6 Loeffler, A., Maciulaitis, A., and Hoff, M., “Circular channel magnetohydrodynamic experiments in a transverse magnetic

field,” Physics of Fluids, Vol. 12, 1969, pp. 2445.


7 Gardner, R. and Lykoudis, P., “Magneto-fluid-mechanic pipe flow in a transverse magnetic field. Part 1. Isothermal flow,”

Journal of Fluid Mechanics, Vol. 47, No. 1, 1971, pp. 737–764.


8 Hochreiter, L. and Sesonske, A., “Turbulent structure of isothermal and nonisothermal liquid metal pipe flow,” Interna-

tional Journal of Heat and Mass Transfer , Vol. 17, No. 1, 1974, pp. 113–123.
9 Ji, H. and Gardner, R., “Numerical analysis of turbulent pipe flow in a transverse magnetic field,” International Journal

of Heat and Mass Transfer , Vol. 40, No. 8, 1997, pp. 1839–1851.
10 Barrett, K., “Duct flow with a transverse magnetic field at high Hartmann numbers,” International Journal for Numerical

Methods in Engineering, Vol. 50, No. 8, 2001, pp. 1893–1906.


11 Satake, S., Kunugi, T., and Smolentsev, S., “Direct numerical simulations of turbulent pipe flow in a transverse magnetic

field,” Journal of Turbulence, Vol. 3, No. 1, 2002, pp. 020.


12 Takeuchi, J., Satake, S., Morley, N., Kunugi, T., Yokomine, T., and Abdou, M., “Experimental study of MHD effects on

turbulent flow of Flibe simulant fluid in circular pipe,” Fusion Engineering and Design, Vol. 83, No. 7-9, 2008, pp. 1082–1086.
13 Lee, D. and Choi, H., “Magnetohydrodynamic turbulent flow in a channel at low magnetic Reynolds number,” Journal

of Fluid Mechanics, Vol. 439, 2001, pp. 367–394.


14 Kobayashi, H., “Large eddy simulation of magnetohydrodynamic turbulent channel flows with local subgrid-scale model

based on coherent structures,” Physics of Fluids, Vol. 18, 2006, pp. 045107.
15 Satake, S., Kunugi, T., Takase, K., and Ose, Y., “Direct numerical simulation of turbulent channel flow under a uniform

magnetic field for large-scale structures at high Reynolds number,” Physics of Fluids, Vol. 18, 2006, pp. 125106.

12 of 13

American Institute of Aeronautics and Astronautics


16 Boeck, T., Krasnov, D., and Zienicke, E., “Numerical study of turbulent magnetohydrodynamic channel flow,” Journal

of Fluid Mechanics, Vol. 572, No. 1, 2007, pp. 179–188.


17 Sarris, I., Kassinos, S., and Carati, D., “Large-eddy simulations of the turbulent Hartmann flow close to the transitional

regime,” Physics of Fluids, Vol. 19, No. 8, 2007, pp. 085109–085109–9.


18 Krasnov, D., Zikanov, O., Schumacher, J., and Boeck, T., “Magnetohydrodynamic turbulence in a channel with spanwise

magnetic field,” Physics of Fluids, Vol. 20, No. 9, 2008, pp. 095105.
19 Burr, U., Barleon, L., Müller, U., and Tsinober, A., “Turbulent transport of momentum and heat in magnetohydrody-

namic rectangular duct flow with strong sidewall jets,” Journal of Fluid Mechanics, Vol. 406, No. 1, 2000, pp. 247–279.
20 Kobayashi, H., “Large eddy simulation of magnetohydrodynamic turbulent duct flows,” Physics of Fluids, Vol. 20, 2008,

pp. 015102.
21 Chaudhary, R., Vanka, S., and Thomas, B., “Direct numerical simulations of magnetic field effects on turbulent flow in

a square duct,” Physics of Fluids, Vol. 22, 2010, pp. 075102.


22 Shatrov, V. and Gerbeth, G., “Marginal turbulent magnetohydrodynamic flow in a square duct,” Physics of Fluids,

Vol. 22, 2010, pp. 084101.


23 Chaudhary, R., Shinn, A., Vanka, S., and Thomas, B., “Direct numerical simulations of transverse and spanwise magnetic

field effects on turbulent flow in a 2:1 aspect ratio rectangular duct,” Computers and Fluids, Vol. 51, 2011, pp. 100–114.
24 Krasnov, D., Zikanov, O., and Boeck, T., “Numerical study of magnetohydrodynamic duct flow at high Reynolds and

Hartmann numbers,” Journal of Fluid Mechanics, Vol. 704, 2012, pp. 421–446.
25 Müller, U. and Bühler, L., Magnetofluiddynamics in Channels and Containers, Springer Berlin, 2001.
26 Knaepen, B., Kassinos, S., and Carati, D., “Magnetohydrodynamic turbulence at moderate magnetic Reynolds number,”

Journal of Fluid Mechanics, Vol. 513, 2004, pp. 199–220.


27 Tezduyar, T., Mittal, S., Ray, S., and Shih, R., “Incompressible flow computations with stabilized bilinear and linear

equal-order-interpolation velocity-pressure elements,” Computer Methods in Applied Mechanics and Engineering, Vol. 95, No. 2,
1992, pp. 221–242.
28 Tezduyar, T. and Osawa, Y., “Finite element stabilization parameters computed from element matrices and vectors,”

Computer Methods in Applied Mechanics and Engineering, Vol. 190, No. 3-4, 2000, pp. 411–430.
29 Detandt, Y., Numerical Simulation of Aerodynamic Noise in Low Mach Number Flows, Ph.D. thesis, Université Libre

de Bruxelles, 2007.
30 George, K. and Vipin, K., “A Fast and Highly Quality Multilevel Scheme for Partitioning Irregular Graphs,” SIAM

Journal on Scientific Computing, Vol. 20, No. 1, 1999, pp. 359–392.


31 Amestoy, P., Duff, I., Koster, J., and L’Excellent, J.-Y., “A Fully Asynchronous Multifrontal Solver Using Distributed

Dynamic Scheduling,” SIAM Journal on Matrix Analysis and Applications, Vol. 23, No. 1, 2001, pp. 15–41.
32 Amestoy, P., Guermouche, A., L’Excellent, J.-Y., and Pralet, S., “Hybrid scheduling for the parallel solution of linear

systems,” Parallel Computing, Vol. 32(2), 2006, pp. 136–156.


33 Eggels, J., Unger, F., Weiss, M., Westerweel, J., Adrian, R., Friedrich, R., and Nieuwstadt, F., “Fully developed turbulent

pipe flow: a comparison between direct numerical simulation and experiment,” Journal of Fluid Mechanics, Vol. 268, No. 1,
1994, pp. 175–209.
34 Loulou, P., Moser, R., Mansour, N., and Cantwell, B., Direct numerical simulation of incompressible pipe flow using a

B-spline spectral method, NASA technical memorandum 110436, 1997.


35 Gold, R., “Magnetohydrodynamic pipe flow. Part 1,” Journal of Fluid Mechanics, Vol. 13, No. 04, 1962, pp. 505–512.
36 Shercliff, J., “Magnetohydrodynamic pipe flow. Part 2. High Hartmann number,” Journal of Fluid Mechanics, Vol. 13,

No. 04, 1962, pp. 513–518.


37 Krasnov, D., Zikanov, O., Thess, A., and Boeck, T., “Direct numerical simulation of transition in MHD duct flow,” Appl.

Math. Mech., Vol. 11, 2011, pp. 659–660.


38 Shercliff, J., “The flow of conducting fluids in circular pipes under transverse magnetic fields,” Journal of Fluid Mechanics,

Vol. 1, No. 6, 1956, pp. 644–666.

13 of 13

American Institute of Aeronautics and Astronautics

You might also like