large eddy simulation of turbulent flow past a bluff body ...336868/... ·...

27
Large Eddy Simulation of Turbulent Flow Past a Bluff Body using OpenFOAM A Thesis Presented By David Joseph Hensel To The Department of Mechanical and Industrial Engineering in partial fulfillment of the requirements for the degree of Master of Science in Mechanical Engineering Northeastern University Boston, Massachusetts August 2014

Upload: lynhi

Post on 17-Mar-2018

218 views

Category:

Documents


0 download

TRANSCRIPT

Large  Eddy  Simulation  of  Turbulent  Flow  Past  a  Bluff  Body  using  OpenFOAM  

 

A  Thesis  Presented  

By  

David  Joseph  Hensel  

To  

The  Department  of  Mechanical  and  Industrial  Engineering    

in  partial  fulfillment  of  the  requirements    for  the  degree  of  

 

Master  of  Science  

in  

Mechanical  Engineering  

Northeastern  University  Boston,  Massachusetts  

 

August  2014  

   

 

 

Large  Eddy  Simulation  of  Turbulent  Flow  Past  a  Bluff  Body  using  OpenFOAM  

M.S.    Defense  by  

David  Hensel  

Tuesday,  July  29th  2014,  12:00pm  –  1:00pm,  09  Forsythe  Northeastern  University,  2014  

Abstract    Numerical  simulation  of  a  turbulent  bluff-­‐body  flow  is  conducted  using  large  eddy  simulation  (LES).    The  open  source  CFD  software  package,  OpenFOAM,  is  employed  to  solve  the  LES  filtered  transport  equations  governing  the  three-­‐dimensional  incompressible  flow  in  the  wake  of  the  body.    This  work  is  motivated  by  the  importance  of  bluff-­‐body  flows,  for  example,  in  flame  stabilization  in  industrial  combustors  and  burners,  as  well  as  in  aerodynamics  applications.    The  focus  of  the  study  is  on  the  proper  generation  of  turbulence  at  the  inlet  boundary.    Standard  boundary  conditions  available  in  OpenFOAM   are   not   sufficient   for   providing   an   accurate   turbulent   inlet   condition   without  modification  of  the  bluff  geometry.    An  improved  method  describing  the  inflow  boundary  condition  is  developed  based  on  the  existing  OpenFOAM  mapping-­‐type  boundary  condition.    In  this  method,  the  boundary  condition  scales  the  standard  deviation  and  mean  value  of  the  velocity  field  onto  the  prescribed  values  provided  by  the  experimental  data.    The  method   is   implemented   in  OpenFOAM  and  employed   in  LES  prediction  of  a   turbulent  bluff-­‐body   flow,   studied   in   the  experiments  of   the  Clean   Combustion   Research   Group   at   the   University   of   Sydney.     The   LES   results   show   favorable  agreements  with  the  experimental  data.      

Thesis  Committee  Members  

Prof.    Reza  Sheikhi  

 

 

 

 

Contents  

Table  of  Figures  ..........................................................................................................................................................................  1  

1.   Introduction  ........................................................................................................................................................................  2  

2.   Formulation  ........................................................................................................................................................................  3  

3.   Simulation  ............................................................................................................................................................................  6  

3.1.   OpenFOAM  .................................................................................................................................................................  6  

3.2.   Numerical  Specification  .......................................................................................................................................  6  

3.3.   Grid  ...............................................................................................................................................................................  8  

3.4.   Turbulent  Inlet  Boundary  Condition  ...........................................................................................................  10  

3.4.1.   Scaling  Method  .................................................................................................................................................  11  

4.   Results  ................................................................................................................................................................................  13  

5.   Summary  and  Concluding  Remarks  ......................................................................................................................  22  

References  ..................................................................................................................................................................................  24  

     

 1  

Table  of  Figures  

Figure  1:  Bluff  Body  Schematic  ............................................................................................................................................................  7  

Figure  2:  Computational  Domain  Representation  using  ParaView  (dimensions  in  millimeters)  ..........................  9  

Figure  3:  Resolution  of  Circular  Jet  (37  cells)  ................................................................................................................................  9  

Figure  4:  Instantaneous  Streamwise  Filtered  Velocity  Iso-­‐surfaces  (clipped  by  X-­‐Y  plane)  .................................  14  

Figure  5:  Magnitude  of  Instantaneous  Vorticity  Iso-­‐surfaces  (clipped  by  X-­‐Y  plane)  ..............................................  15  

Figure  6:  Line  Integral  Convolution  of  Streamwise  Filtered  Velocity  in  Bluff  Region,  X-­‐Y  Plane  ........................  15  

Figure  7:  LES  filtered  Streamwise  Velocity  Contours  Predicted  by  the  Smagorinsky  Model.  ...............................  16  

Figure  8:  LES  filtered  Streamwise  Velocity  Contours  Predicted  by  the  Dynamic  One  Equation  Model.  ..........  17  

Figure  9:  Radial  Profiles  of  the  Mean  and  Resolved  RMS  Streamwise  Velocity  (x=0.003,  0.01,  0.02  [m]).  .....  18  

Figure  10:  Radial  Profiles  of  the  Mean  and  Resolved  RMS  Streamwise  Velocity  (x=0.03,  0.04,  0.05  [m]).  .....  19  

Figure  11:  Radial  Profiles  of  the  Mean  and  Resolved  RMS  Streamwise  Velocity  (x=0.06  [m]).  ...........................  20  

Figure  12:  Radial  Profiles  of  the  Mean  Radial  Velocity  (x=0.003,  0.01  [m]).  ................................................................  20  

Figure  13:  Radial  Profiles  of  the  Mean  Radial  Velocity  (x  =  0.02,  0.03,  0.04[m]).  .......................................................  21  

Figure  14:  Radial  Profiles  of  the  Mean  Radial  Velocity  (x  =  0.05,  0.06[m]).  ..................................................................  22  

   

 2  

1. Introduction    

Approaches  for  simulations  of  turbulent  reacting  flows  can  be  divided  into  three  categories:  direct  

numerical   simulation   (DNS),   large   eddy   simulation   (LES),   and   Reynolds   averaged   Navier-­‐Stokes  

simulation  (RANS).    DNS  provides  the  most  detailed  predictions  and  is  a  useful  tool  for  the  studying  

the  physics   of   turbulent   flows;   however,   the   large  number   of   grid  points   required  makes  DNS  of  

engineering-­‐type   problems   prohibitively   expensive   in   the   foreseeable   future   [1].       Solutions   of  

RANS  equations,  on  the  other  hand,  are  now  widely  used  in  engineering  applications  to  predict  flow  

in  fairly  complex  configurations.    This  approach,  however,  suffers  from  one  principal  shortcoming;  

the  fact  that  the  turbulence  model  must  represent  a  very  wide  range  of  scales  reduces  its  reliability  

as  an  accurate  predictive  tool.    Among  the  three  approaches,  LES  is  an  attractive  simulation  method  

as   it   provides   a   compromise   between   accuracy   and   computational   cost.     LES   is   known   to   be   the  

optimal  means  of  capturing  the  detailed,  unsteady  physics  of  turbulent  flows.    [2]  

The  basic   idea   in  LES   is   to  resolve  the   large-­‐scale  turbulent  motions  and  to  model  the  small-­‐scale  

motions,   which   are   more   universal.     This   idea   can   be   explained   in   terms   of   the   energy   cascade  

concept.    The  turbulent  energy  is  transferred  from  large-­‐scale  motions  to  smaller  scales,  until  finally  

dissipated   into  heat  by  viscosity  at   the  molecular   level.    According  to  Kolmogorov’s  hypotheses,  a  

scale  separation  exists  within  the  energy  cascade  where  turbulent  energy  is  produced  in  the  largest  

scales,   transferred   to   decreasing   scales   by   the   energy   cascade   within   the   inertial   subrange,   and  

dissipated  through  viscosity  at  the  smallest  scales.    In  LES,  it  is  essential  to  resolve  about  80%  of  the  

turbulent  energy  of  the  large  scales  while  representing  about  20%  transferred  to  small  scales  using  

a  subgrid  scale  (SGS)  model.    RANS  has  an  inherent  shortcoming  since  it  averages  over  all  turbulent  

scales   and   thus,   provides   a   time-­‐averaged   field;   however,   LES   provides   time-­‐dependent   fields,  

which   offers   improved   accuracy   in   predicting  unsteady   turbulent  motions.     The  benefit   of   LES   is  

evident  in  flows  where  vortex  shedding  and  unsteady  separation  are  significant    [3].        

 3  

In  the  present  study,  LES  prediction  of  turbulent  flow  past  a  bluff  body  is  performed.    This  study  is  

motivated  by   the   importance  of  bluff-­‐body   flows  within  various   combustion  applications.     In   this  

context,  bluff  bodies  provide  a  simple  geometry  to  study  recirculation  zones  that  help  stabilize  the  

flame.    The  purpose  of  this  study  is  to  validate  the  hydrodynamic  solution  for  the  non-­‐reacting  bluff  

body   flow   studied   experimentally   by   the   Clean   Combustion   Research   Group   at   the   University   of  

Sydney  [4].        Previous  studies  of  this  geometry  include  works  of  Drozda  [5]  and  Drozda  et  al.    [6],  

which   involve   LES   based   on   filtered   density   function   (FDF)   methodology   for   prediction   of   non-­‐

reacting  and  reacting  flows  in  this  configuration.  

In   this   study,   we   use   the   OpenFOAM   software   package   to   conduct   simulation   of   the   same  

configuration.     Simulation   of   flows   around  bluff   bodies   using  OpenFOAM  has   been   the   subject   of  

several   studies.     Lysenko   et   al.     [7]   studied   turbulent   flow   around   triangular   bluff   bodies,   and  

compares  results  from  OpenFOAM  with  ANSYS  Fluent.    Salvador  et  al.    [8]  included  a  non-­‐reacting  

case  in  their  study  of  a  premixed  reacting  flame  using  OpenFOAM.    An  important  issue  in  accurate  

simulation  of   turbulent   flows   is  proper   generation  of   turbulent   inlet   boundary   conditions,  where  

various   statistics   are   not   only   time   varying,   but   also   physically   representative   of   turbulent   flow.    

Several  methods  have  been  developed  to  address  this  boundary  condition,  including  the  use  of  pre-­‐

compiled   data   from   a   precursor   study,   generating   synthetic   turbulence,   or   turbulence   mapping  

methods  [9].      Volavy  et  al.    [10]  provided  a  comparison  of  a  uniform  inlet  velocity  profile,  with  that  

of  a  mapped  turbulent  inlet,  and  demonstrated  the  need  for  proper  turbulent  inlet  for  flows  over  a  

reversed  step.     In   this   study,  we  developed  an   improved   inflow  boundary  condition  based  on   the  

OpenFOAM  mapping-­‐type  boundary  condition,  as  discussed  in  Section  3.4.  

2. Formulation  

The  basic  equations  governing   incompressible   isothermal   turbulent   flows  are   the  conservation  of  

mass   and   momentum   by   describing   variation   of   transport   variables   in   space  𝑥!(𝑖 = 1,2,3)     and  

 4  

time  𝑡.     The   transport   variables   used   are   fluid   density  𝜌(𝑥, 𝑡),   pressure  𝑝(𝑥, 𝑡),   and   the   velocity  

vector  𝑢!(𝑥! , 𝑡)    (𝑖 = 1,2,3).      

  𝜕𝜌𝜕𝑡+𝜕𝜌𝑢!𝜕𝑥!

= 0   (1)  

  𝜕𝜌𝑢!𝜕𝑡

+𝜕𝜌𝑢!𝑢!𝜕𝑥!

= −𝜕𝑝𝜕𝑥!

+ 𝜌𝜈𝜕!𝑢!𝜕𝑥!𝜕𝑥!

  (2)  

where  𝜈    denotes  the  kinematic  viscosity.  

LES   uses   a   spatial   filtering   method   which   is   essentially   the   convolution   integral   over   the   entire  

volume  [3]  ,  

 𝑢 𝑥! , 𝑡 = 𝐺 𝑟! , 𝑥! 𝑢(𝑥! − 𝑟! , 𝑡)𝑑𝑟!

!!

!!   (3)  

where  G  is  a  filter  that  satisfies  the  normalization  condition  

  𝐺 𝑟! , 𝑥! 𝑑𝑟! = 1   (4)  

and  U  is  any  function  of  space  and  time  and      denotes  the  filtered  field  variables.  

As  a  result  of  LES  filtering  we  obtain  a  residual  field  defined  as  

  𝑢′ 𝑥! , 𝑡 ≡  𝑢 𝑥! , 𝑡 −   𝑢 𝑥! , 𝑡   (5)  

We  thus  have  

  𝑢 𝑥! , 𝑡 = 𝑢 𝑥! , 𝑡 +  𝑢′ 𝑥! , 𝑡   (6)  

The  filtered  form  of  the  governing  equation  is  obtained  by  applying  the  filtering  operation  to  Eqs.    

(1)  and  (2).  

  𝜕𝜌𝜕𝑡+𝜕𝜌 𝑢!𝜕𝑥!

= 0  (7)  

  𝜕𝜌 𝑢!𝜕𝑡

+𝜕𝜌 𝑢!𝑢!𝜕𝑥!

= −𝜕 𝑝𝜕𝑥!

+ 𝜌𝜈𝜕! 𝑢!𝜕𝑥!𝜕𝑥!

  (8)  

 

 5  

We  can  decompose  the  second  term  on  the  left  side  of  Eq.    (8)  as  

  𝜕𝜌 𝑢!𝑢!𝜕𝑥!

=𝜕𝜌 𝑢! +  𝑢!! 𝑢! +  𝑢!!

𝜕𝑥!=𝜕𝜌 𝑢! 𝑢!

𝜕𝑥!+𝜕𝜌𝜏!"!

𝜕𝑥!   (9)  

where   𝜏!"! =   𝑢!𝑢! − 𝑢! 𝑢!  and   is   defined   as   the   residual   or   SGS   stress   tensor.     The   residual  

stress  tensor  is  unclosed  and  must  be  modeled,  since  we  have  no  representation  of   𝑢!𝑢!  available  

by  the  governing  equations.    The  residual  kinetic  energy  is  defined  as  

  𝑘! ≡12𝜏!!!   (10)  

The   residual   kinetic   energy   is   related   to   the   isotropic   part   of   the   residual   stress   tensor.     The  

anisotropic  component  of  the  residual  stress  tensor  is  responsible  for  momentum  transport.  

  𝜏!"! ≡ 𝜏!!! −23𝑘!𝛿!"   (11)  

where  𝛿!"  is  the  Kronecker  delta.    The  isotropic  component  can  be  combined  into  a  modified  filtered  

pressure  

  𝑝! ≡ 𝑝 +23𝜌𝑘!   (12)  

The  resulting  filtered  momentum  equation  is  obtained  as  

  𝜕 𝑢!𝜕𝑡

+𝜕 𝑢! 𝑢!𝜕𝑥!

= 𝜈𝜕! 𝑢!𝜕𝑥!𝜕𝑥!

−𝜕𝜏!"!

𝜕𝑥!−1𝜌𝜕 𝑝!𝜕𝑥!

  (13)  

The  LES  closure  problem  is  due  to  the  residual  stress  tensor.    The  SGS  modeling  of  this  stress  has  

been   addressed   in   many   investigations   and   several   closures   have   been   developed;   examples  

include  the  Smagorinsky  and  dynamic  models  [3].  

 6  

3. Simulation  

3.1.  OpenFOAM  

In  this  study,  we  use  the  OpenFOAM  software  package  to  perform  the  computational  fluid  dynamics  

(CFD)   simulation   of   the   flow   past   the   bluff   body.     OpenFOAM   has   been   offered   as   open-­‐source  

software   since   2004   [11].   The   software   contains   a   full   suite   of   numerical   methods,   solvers,  

boundary   conditions,  meshing,   and  plug-­‐ins   for   third  party   post-­‐processing  using   the   application  

ParaView  (also  open-­‐source).    In  addition  to  the  numerical  capabilities,  OpenFOAM  provides  built-­‐

in  parallelism   for  high-­‐performance   computing  using  Message  Passing   Interface   (MPI),  There   are  

several   studies   showing   good   scalability   of   OpenFOAM   on   various   computing   platforms   [12];  

however,  the  scalability  of  the  simulation  varies  based  on  the  details  of  the  numerics.  

3.2.  Numerical  Specification  

The  data  used  for  this  analysis  is  provided  by  the  experimental  studies  of  a  flat  faced,  cylindrical  

bluff  body  axially  centered  in  a  stream  of  air  (co-­‐flow),  performed  at  the  University  of  Sydney.    The  

bluff  body  contains  a  fuel  jet  at  the  center.    Details  of  the  geometry  are  shown  in  Figure  1,  and  the  

configuration  parameters  are  provided  in    

Table  1.    Additional  data  for  this  geometry  is  available  from  the  Clean  Combustion  Research  Group  

website  [4].  

The   boundary   conditions   applied   in   the   present   simulations   are   generally   consistent   with   the  

experimental  setup  at  the  University  of  Sydney;  zero-­‐gradient  velocity  and  pressure  are  applied  at  

the  exit  plane  of   the  domain,  and  the  sides  (co-­‐flow)  specified  as  symmetry  planes  to  provide  the  

representation  of   a  wind   tunnel.    The  bluff   face   is  modeled  by   specifying  zero  velocity,   and  zero-­‐

gradient  pressure.      

In  this  study,  we  select  the  non-­‐reacting  bluff  body  jet  configuration  B4C1.    Three  sets  of  data  are  

available   for   this  configuration,  denoted  as  B4C1-­‐S(1-­‐3).    Each  set  contains  data  at  various  points  

 7  

for  mean  streamwise  velocity,  mean  axial  velocity,  and  RMS  of  these  values.    Data  sets  B4C1-­‐S2  and  

B4C1-­‐S3  also  contain  value   for  Reynolds  shear  stress.    Data  points  are   located  radially  across   the  

bluff,   as   well   as   in   the   streamwise   direction.     The   B4C1-­‐S1   set   contains   data   up   to   1.2   bluff  

diameters,  while  B4C1-­‐S(2-­‐3)  provides  values  up  to  3.3  times  the  diameter  of  the  bluff.    The  data  set  

B4C1-­‐S2  is  used  for  this  study,  due  to  the  availability  of  data,  and  the  range  of  sampling  locations.  

 Figure  1:  Bluff  Body  Schematic  

 

 

Table  1:  Non  Reacting  Bluff  Body  Jet  Data  –  B4C1  

Burner  Description   Bluff-­‐Body  Bluff  Diameter   50  mm  Jet  Diameter   3.8  mm  Bulk  Mean  Jet  Velocity   61  m/s  Co-­‐flow  Velocity   20  m/s    

 8  

The   OpenFOAM   “pimpleFoam”   solver   is   used   for   this   study,   because   of   its   useful   control  

capabilities.    As  a  transient,  incompressible  solver  that  combines  the  PISO-­‐  and  SIMPLE-­‐  algorithms,  

pimpleFoam  provides  an  improvement  over  pisoFoam  by  allowing  automatic  control  over  the  time  

step  length  based  on  a  user-­‐provided  Courant  number.    This  results  in  improved  numerical  stability  

and   facilitates   the   initial   setup  of   the   simulations.     The  numerical   schemes   from   the   “motorBike”  

tutorial  are  used  as  the  starting  point  for  the  simulation.    The  discretization  scheme  applied  to  the  

temporal   derivative   is   a   second   order   implicit   backward   scheme.     The   spatial   derivatives   use  

second  order  central-­‐differencing  linear  and  Total  Variation  Diminishing  (TVD)  schemes.      

OpenFOAM  contains  several  LES  models   that  are  applicable   for   incompressible   flow.    The  models  

being   investigated   in  this  work  are  the  standard  OpenFOAM  implementations  of   the  Smagorinsky  

model   [13],   and   a   localized   Dynamic   One   Equation   Eddy   viscosity   model   “dynOneEqEddy”   [14].      

The  “LESProperties”  file  defines  the  LES  model  used  by  the  solver,  the  LES  filter  width  to  use,  and  

related  coefficients  and  parameters.    The  Smagorinsky  coefficients  used  are  𝐶! = 1.048  , and  𝐶! =

0.094.    The  Dynamic  One  Equation  Eddy  parameters  are  set  using  a  “simple”  filter,  and    𝐶! = 1.048  .    

Both  LES  model  configurations  use  the  “cubeRootVol”  LES  filter  width.  

3.3.  Grid  

The   computational   domain   considered   for   this   study   is   chosen   such   that   the   geometry   does   not  

influence  the  behavior  of  the  jet,  and  the  experimental  data  points  are  included  within  the  domain.    

The   grid   is   comprised   of   a   grid   151   cells   in   each   cross-­‐stream   direction,   and   201   cells   in   the  

streamwise   direction,   resulting   in   approximately   4.58   million   cells.     For   the   case   setup   in  

OpenFOAM,  X   is  used  as   the  streamwise  direction,  and  Y  and  Z  are  both  assigned  as  cross-­‐stream  

directions.    Figure  2  represents  the  geometric  extents  used,  which  are  0  to  216  mm  in  𝑋  direction,  

and   -­‐40.5mm   to   40.5   mm   in   both   𝑌and  𝑍  directions.     This   corresponds   to   a   cross-­‐stream   cell  

dimension  of   approximately  0.0536  mm  which   is   seven   cells   across   the   span  of   the   jet,   shown   in  

Figure  3.        A  structured  grid  of  constant  mesh  size  is  used  to  avoid  numerical  error  introduced  by  

 9  

non-­‐orthogonality,   as   well   as   LES   filtering   commutation   error.     A   three-­‐dimensional,   structured  

mesh  is  used  to  simulate  the  inherently  three-­‐dimensional  nature  of  turbulent  fluid  flow  variables.        

 

Figure  2:  Computational  Domain  Representation  using  ParaView  (dimensions  in  millimeters)  

   Figure  3:  Resolution  of  Circular  Jet  (37  cells)    

The  mesh  for  this  study  is  generated  using  the  OpenFOAM  meshing  utility  blockMesh.    The  overall  

geometry   is   specified   as   a   single   block   rectangular   prism   with   eight   vertices,   extending   in   the  

 10  

positive  X   direction   from   the  Y-­‐Z   plane,   and   centered   about   the   origin.     Each   side   of   the   block   is  

assigned  to  a  named  collection  of  cells  on  the  mesh  boundary  called  a  “patch.”      

A   mesh   utility   topoSet   is   used   to   create   groups   of   cells   called   “cellSets,”   based   on   a   specified  

geometric   condition   such   as   cylinders   of   diameter   Dbluff   and   Djet   corresponding   to   bluff   and   jet  

nozzle,  respectively.    These  groups  of  cells  are  used  to  determine  which  cells  lie  within  the  specified  

geometry,  and  are  coincident  with  the  boundary  faces  of  the  “inlet”  patch.    From  this  information,  

groups  of  faces  called  “faceSets”  are  used  to  create  patches  to  represent  the  bluff  geometry  on  the  

inlet  plane.    The  createPatch  utility  is  used  to  make  the  “jet”  and  “bluff”  patches,  by  reassigning  faces  

from  the  “inlet”  patch  using  “faceSets.”    This  results  in  a  well-­‐defined  circular  geometry  for  the  jet  

and  the  bluff.    The  resulting  “jet”  patch  area  is  approximately  4%  larger  than  the  actual  surface  area  

of   the   jet,  and  the  “bluff”  patch   is  approximately  1%  less   than  the  actual  surface  area  of   the  bluff.    

This  case  was  modeled  without  using  a  physical   representation  of   the  bluff  burner  geometry,  but  

rather  placing  this  geometry  specification  as  a  “patch”  on  the  mesh  boundary.        

3.4.  Turbulent  Inlet  Boundary  Condition  

As  described  by  Gabor  and  Baba-­‐Ahmadi  [15],  a  turbulent  inlet  boundary  condition  should  remain  

generic   and   allow   turbulent   properties   to   be   specified   easily,   so   the   boundary   condition   is  

applicable  for  a  variety  of  flow  conditions.    There  are  several  methods  to  describe  a  turbulent  inlet  

flow  boundary  condition  properly.    One  of  these  is  generation  of  synthetic  turbulence,  by  means  of  a  

forcing   frequency   derived   from   characteristic   flow   parameters.     This  method   can   provide   a   very  

precise  description  of  the  boundary  condition  for  the  specific  flow  configuration;  however,  it  is  case  

dependent  and  not  trivial   to  derive.    Another  method  involves  a  precursor  study  where  turbulent  

fields   are   created   and   stored   for   use   as   a   pre-­‐defined,   time-­‐dependent   turbulence   library.     This  

method   may   be   computationally   expensive   to   generate,   but   provides   a   reusable   boundary  

description   for   the   specific   flow   conditions   generated.     A   promising   method   is   to   simulate  

turbulence  by  creating  an  isolated  sub-­‐domain  with  cyclic  boundaries.    This  essentially  creates  an  

infinitely   long   domain,   which   allows   generation   of   fully   developed   turbulent   flow,   typically   wall  

 11  

bounded,  channel  or  pipe  flow.    This  requires  additional  computational  expense,  as  these  cells  are  

not  a  part  of  the  computational  domain  of  interest,  and  provide  no  other  use  other  than  generating  

inlet   turbulence.     The   resulting   turbulence   created   by   the   cyclic   boundary   method   reasonably  

represents   turbulent   flow,   which   is   not   simply   the   collection   of   random   signals,   but   rather   a  

collection  of  coherent  turbulent  structures  governed  by  transport  equations  (Eqs.  (1)  and  (2)).    A  

modification   to   this   method   is   the   mapped   turbulent   sub-­‐domain,   which   involves   providing   the  

sampling  plane  within  the  computational  domain  of  interest,  thereby  reducing  the  additional  cells  

required   for  generating   turbulence.    This  method   is   implemented   in  OpenFOAM,  as   the  “mapped”  

boundary  condition,  and  provides  control  by  using  an  optional  method  to  scale  the  sampled  field  to  

a  prescribed  mean  value.    The  mapping  method  has  shown  to  be  successful  for  channel  flow  [9],  and  

provides  a  simple,  yet  accurate  description  of  turbulent  flow.    The  boundary  condition  used  in  this  

study  is  based  on  the  mapped  boundary  condition  in  OpenFOAM.  

3.4.1.  Scaling  Method  

The  premise  of  the  mapped  boundary  condition  is  that  the  faces  of  a  boundary  patch  are  projected  

onto  an  arbitrary  plane  within   the  computational  domain.    These  values  are   then  collected   into  a  

list,  on  which  scaling  operations  are  then  performed.    Finally,  these  scaled  values  are  reassigned  to  

their  corresponding  boundary  faces.    The  OpenFOAM  mapped  boundary  condition  scales  the  mean  

velocity  of  the  sample  by  shifting  all  the  values,  or  by  multiplying  the  sampled  field  by  a  scale  factor,  

depending  on  the  relation  between  the  local  sampled  average  and  the  prescribed  value.    The  scaling  

method   uses   the   prescribed   value   divided   by   a   calculated   sample  mean   value   to   develop   a   scale  

factor.    When  a  multiplier  scales   the  velocity,   that  multiplier  also  scales   the  standard  deviation  of  

the  field.    This  is  undesirable  when  the  sampled  field  mean  value  is  significantly  different  from  the  

prescribed  mean   value.     For   the   bluff-­‐body   flow   the   sample   velocity   shall   always   be   less   than   or  

equal   to   the   inlet   velocity   for   a   jet   expanding   into   a   slower   moving   co-­‐flow.     This   limitation   is  

addressed  by  the  alternative  scaling  method  described  in  this  work,  where  the  standard  deviation  is  

controlled   by   scaling   the   sampled   field   to   a   prescribed   value.     This   sampling   results   in   better  

 12  

agreement  of  filtered  velocity  with  the  data.    To  match  the  RMS  values,  the  first  step  is  to  calculate  

the   standard   deviation   of   the   sampled   field.     Then   we   calculate   a   scaling   factor   by   dividing   the  

desired  standard  deviation  by  the  sampled  standard  deviation.  

  𝛾 =𝜎!"#$%"&'#(𝜎!"#$%&'

  (14)  

where  𝜎  denotes  the  standard  deviation.    Next,  each  value  of  the  sampled  field  is  multiplied  by  the  

scaling  factor  to  create  a  scaled  field.    

  𝑢 !"#$%& =  𝛾   𝑢  !"#$%&'     (15)  

Now  we  calculate   the   “mean  shift”  difference  between   the  mean  value  of   the   scaled   field  and   the  

desired  mean  value  

  𝛽 =   𝑢 !"#$%"!"#$ −   𝑢 !"#$%&   (16)  

where    denotes  the  local  mean  value  calculated  from  the  sampled  points.    Finally,  add  the  “mean  

shift”   to   each  value  of   the   scaled   field   to  obtain   the   final   field   that  has  been   scaled   in  both  mean  

value  and  standard  deviation.  

    𝑢 !"#$% =   𝑢 !"#$%& + 𝛽   (17)  

This  scaling  method  is  applied  to  each  component  of  the  velocity  for  the  boundary  condition  used  in  

this  research.    The  bulk  mean  velocity  and  fluctuations  from  the  experimental  data  initial  conditions  

are   used   to   define   the   boundary   condition   for   both   the   jet   and   co-­‐flow.     This   method   aims   to  

generate   realistic   turbulence   by   sampling   existing   fluctuations   from   the   internal   domain,   rather  

than   generating   a   synthetic   perturbed   inlet.     In   comparison   to   the   standard   boundary   condition,  

this   method   provides   additional   control   over   the   second   order   statistics.     The   current  

implementation   scales   the   distribution   of   each   vector   component   to   a   prescribed   standard  

deviation  and  mean  value.    The  intent  of  this  mapping/scaling  boundary  condition  is  to  provide  the  

bulk  mean  statistical  description  of  the  jet.      

 13  

4. Results  

The   computational  work   done   for   this   study   is   performed   using   the   Texas   Advanced   Computing  

Center  (TACC)  Stampede  Supercomputer  located  at  the  University  of  Texas  at  Austin.    Simulations  

are  also  performed  using  Northeastern  University’s  Discovery  Cluster  at   the  Massachusetts  Green  

High  Performance  Computing  Center   (MGHPCC).    OpenFOAM  2.3.0   is   used   at   both   facilities,  with  

modified  solvers  and  boundary  conditions.      

The  simulation  is  compared  against  the  non-­‐reacting  experimental  data  set  B4C1-­‐S2.    The  following  

figures   are  plotted  using   time-­‐averaged   results,  which   is   averaged  about   the   angular  direction   to  

provide   results   in   a   consistent   format   with   that   of   the   experiment.     The   results   compare   two  

different  LES  models  from  OpenFOAM  with  all  other  case  parameters  held  constant.    Figure  4  and  

Figure   5   demonstrate   the   turbulent   nature   of   the   bluff   body   flow,   showing   instantaneous   iso-­‐

surfaces   of   streamwise   filtered   velocity,   and   magnitude   of   vorticity,   respectively.     The   time-­‐

averaged  streamwise  velocity  shown  in  Figure  6  displays  the  recirculation  zone  of   the  bluff  body,  

with  an  inner  and  outer  vortex  structure  visible  by  Line  Integral  Convolution.      

A  limitation  of  the  mapping  boundary  condition  is  the  inability  to  define  the  profile  of  the  sampled  

field;   as   a   result,   a   single   value   must   be   specified.     The   mean   values   calculated   from   the   initial  

conditions   do   not   provide   good   agreement  with   the   data  when   used  with   the  mapped   boundary  

condition.     During   initial   studies,   the   mean   velocity   was   under-­‐predicted   and   the   turbulent  

fluctuations  were  over-­‐predicted  using  both  LES  models.    To  address  this  issue,  the  mean  value  and  

fluctuations  are  modified   to  maintain  closer  agreement  with   the  experimental  data.    The  mapped  

turbulent   inlet   boundary   condition   is   specified   to  maintain   a   scaled  mean   velocity   of   70.76  m/s,  

which  corresponds  to  the  centerline  velocity  specified  in  the  initial  conditions.    The  scaled  sampled  

standard  deviation  is  set  to  maintain  approximately  1%  of  the  scaled  mean  velocity.    The  velocity  

contours   predicted   by   the   Smagorinsky   model   shown   in   Figure   7   demonstrate   the   longer,   less  

turbulent   jet   than   that   shown   in   Figure   8   predicted   by   the   Dynamic   One   Equation   model.     The  

 14  

velocity   profiles   are   accurate   closer   to   the   bluff,   as   shown   in   Figure   9.     They   however   start   to  

diverge   from   the   experimental   values   at   Figure   10.     Downstream,   the   overall   agreement   is  

reasonable   but   near   the   centerline   the   Smagorinsky   model   over-­‐predicts,   and   the   Dynamic   One  

Equation  model   under-­‐predicts   the   streamwise   velocity.     In   both   cases,   the   jet   velocity   is   under-­‐

predicted  after  the  recirculation  zone,  or  at  approximately  one  bluff  diameter  (x  =  0.50),  as  shown  

in   Figure   11.     The   streamwise   velocity   fluctuations   are   shown   alongside   the   velocity   profiles   in  

Figure   9   through   Figure   11.     Both  models   follow   the   experimental   profile   reasonably   well,   with  

peaks   in   the  profile  consistent  with   the   inner,  central,  and  outer  mixing   layers.    The  Smagorinsky  

model  tends  to  under-­‐predict  the  fluctuations  close  to  the  face  of  the  bluff  as  shown  in  Figure  9,  and  

over-­‐predict  the  fluctuations  further  from  the  bluff  in  Figure  10.    The  Dynamic  One  Equation  model  

shows   closer   agreement  with   the   data   across   all   sample   points.     The   velocity   profile   in   contours  

greater   than  one  bluff  diameter   tend  to  under-­‐predict   the   free  stream  co-­‐flow  velocity  of  20  m/s,  

which   is   due   to   the  over-­‐prediction  of   the   spread   rate   of   the   jet.     The   radial   velocity  profiles   are  

shown  in  Figure  12  through  Figure  14.    Both  methods  provide  reasonable  prediction  of  the  radial  

velocity  at  all  locations.  

 Figure  4:  Instantaneous  Streamwise  Filtered  Velocity  Iso-­‐surfaces  (clipped  by  X-­‐Y  plane)  

 

 15  

 Figure  5:  Magnitude  of  Instantaneous  Vorticity  Iso-­‐surfaces  (clipped  by  X-­‐Y  plane)  

 

 

Figure  6:  Line  Integral  Convolution  of  Streamwise  Filtered  Velocity  in  Bluff  Region,  X-­‐Y  Plane  

 16  

 

Figure  7:  LES  filtered  Streamwise  Velocity  Contours  Predicted  by  the  Smagorinsky  Model.      

a) Smagorinsky,  X-­‐Y  plane,  Instantaneous  contours  of  ⟨𝑢⟩  

b) Smagorinsky,  X-­‐Z  plane,  Instantaneous  contours  of  ⟨𝑢⟩  

c) Smagorinsky,  X-­‐Y  plane,  Time-­‐Averaged  contours  of  ⟨𝑢⟩  

 17  

 

Figure  8:  LES  filtered  Streamwise  Velocity  Contours  Predicted  by  the  Dynamic  One  Equation  Model.      

 

 

 

a) Dynamic  One  Equation,  X-­‐Y  plane,  Instantaneous  contours  of  ⟨𝑢⟩  

b) Dynamic  One  Equation,  X-­‐Z  plane,  Instantaneous  contours  of  ⟨𝑢⟩  

c) Dynamic  One  Equation,  X-­‐Y  plane,  Time-­‐Averaged  contours  of  ⟨𝑢⟩  

 18  

 

Figure  9:  Radial  Profiles  of  the  Mean  and  Resolved  RMS  Streamwise  Velocity  (x=0.003,  0.01,  0.02  [m]).  

 

 19  

 

Figure  10:  Radial  Profiles  of  the  Mean  and  Resolved  RMS  Streamwise  Velocity  (x=0.03,  0.04,  0.05  [m]).  

 

 

 20  

 

Figure  11:  Radial  Profiles  of  the  Mean  and  Resolved  RMS  Streamwise  Velocity  (x=0.06  [m]).  

 

 

 

 

 

Figure  12:  Radial  Profiles  of  the  Mean  Radial  Velocity  (x=0.003,  0.01  [m]).  

 21  

 

Figure  13:  Radial  Profiles  of  the  Mean  Radial  Velocity  (x  =  0.02,  0.03,  0.04[m]).  

 22  

 

Figure  14:  Radial  Profiles  of  the  Mean  Radial  Velocity  (x  =  0.05,  0.06[m]).  

 

5. Summary  and  Concluding  Remarks  

Large   eddy   simulation   (LES)   is   conducted   of   a   turbulent   bluff   body   flow   using   OpenFOAM.     The  

results   are   comparable   with   the   experimental   data   provided   by   the   Clean   Combustion   Research  

Group  at  the  University  of  Sydney.    The  focus  of  this  study  is  on  implementing  an  improved  method  

in  OpenFOAM  to  generate  the  turbulent  inflow  condition.    This  is  handled  by  a  mapping  boundary  

condition,  which  samples  a  location  from  within  the  computational  domain  and  scales  the  field  to  a  

prescribed  mean  and  standard  deviation.      

LES   predictions   are   obtained   via   two   subgrid   scale  models:   Smagorinsky   and   Localized  Dynamic  

One  Equation  Eddy  Viscosity.    Both  models  provide  reasonable  overall  prediction  of  the  turbulent  

wake   behind   the   bluff   body.     In   general,   the   Smagorinsky   model   tends   to   under-­‐predict   the  

turbulent   fluctuations   and   the   Localized   Dynamic   One   Equation   Eddy   Viscosity  model   tends   to  

slightly   over-­‐predict   turbulent   fluctuations.     The   under-­‐prediction   of   fluctuations   by   the  

Smagorinsky   model   is   likely   due   to   the   dissipative   nature   of   this   LES   model.     As   the   velocity  

contours   are   generally   well   produced   in   the   region   closest   to   the   inlet,   but   deviate   from   the  

experimental  data  further  into  the  domain,  it  is  likely  that  numerical  schemes  and  LES  models  have  

 23  

a  significant  role  in  the  accuracy  of  these  simulations.    Altering  the  inlet  parameters  provides  some  

control   over   the   velocity   profiles;   however,   those   profiles   past   the   recirculation   zone   are   always  

under-­‐predicted   in   the   results.     The  predictions  obtained   from   the  Dynamic  One  Equation  model  

are  more   consistent  with   the   experimental   data,   for   both   first   and   second   order  moments.     This  

work   provides   a   preliminary   implementation   of   a   modified   mapped   boundary   condition   in  

OpenFOAM  by   including   additional   controls   to  match   a   prescribed   inflow   condition.     OpenFOAM  

along   with   the   new   inlet   boundary   condition   has   shown   to   provide   LES   prediction   of   turbulent  

flows  past  a  bluff  body  with  reasonable  accuracy.    The  near  field  of  the  flow  is  favorably  predicted.    

The  far  field  however  shows  less  accuracy  mainly  near  the  centerline  and  due  to  over-­‐prediction  of  

the  spread  rate  of  the  jet.    The  modified  mapping  boundary  condition  has  the  potential  for  further  

developments   including  additional   control  over   second  order   statistics,   and   scaling   to  maintain  a  

specified  distribution  as  well  as  parameters  to  control  the  behavior  of  the  jet  downstream.  

 

 

   

 24  

References  

[1]   D.  C.  Wilcox,  Turbulence  modeling  for  CFD  vol.  1.  La  Cañada,  CA:  DCW  Industries,  Inc.,  1993.  

[2]   W.  Rodi,  "Comparison  of  LES  and  RANS  calculations  of  the  flow  around  bluff  bodies,"  Journal  of  Wind  Engineering  and  Industrial  Aerodynamics,  vol.  69–71,  pp.  55-­‐75,  1997.  

[3]   S.  B.  Pope,  Turbulent  Flows.  New  York,  NY:  Cambridge  University  Press,  2000.  

[4]   The  University  of  Sydney.  (2014).  Bluff-­‐Body  Flows  and  Flames.  Available:  http://sydney.edu.au/engineering/aeromech/thermofluids/bluff.htm  

[5]   T.  G.  Drozda,  "Implementation  of  LES/SFMDF  for  prediction  of  non-­‐premixed  turbulent  flames,"  Ph.D.  Dissertation,  University  of  Pittsburgh,  2006.  

[6]   T.  Drozda,  M.  Sheikhi,  C.  Madnia,  and  P.  Givi,  "Developments  in  formulation  and  application  of  the  filtered  density  function,"  Flow,  Turbulence  and  Combustion,  vol.  78,  pp.  35-­‐67,  2007.  

[7]   D.  A.  Lysenko,  I.  S.  Ertesvåg,  and  K.  E.  Rian,  "Modeling  of  turbulent  separated  flows  using  OpenFOAM,"  Computers  &  Fluids,  vol.  80,  pp.  408-­‐422,  2013.  

[8]   N.  M.  C.  Salvador,  M.  T.  Mendonça,  and  W.  M.  da  Costa  Dourado,  "Large  Eddy  Simulation  of  bluff  body  stabilized  flame  with  turbulent  premixed  flame,"  Guidelines  of  Workshop  on  Space  Engineering  and  Technology,  June  2012.  

[9]   G.  Tabor,  M.  Baba-­‐Ahmadi,  E.  de  Villiers,  and  H.  Weller,  "Construction  of  inlet  conditions  for  LES  of  turbulent  channel  flow,"  in  Proceedings  of  the  ECCOMAS  congress,  Jyväskylä,  Finland,  2004.  

[10]   J.  Volavy,  M.  Forman,  and  M.  Jicha,  "Influence  of  the  turbulence  representation  at  the  inlet  on  the  downstream  flow  pattern  in  LES  of  backward-­‐facing  step,"  EPJ  Web  of  Conferences,  vol.  25,  2012.  

[11]   OpenFOAM  Foundation.  (2014).  OpenFOAM®    Release  History.  Available:  http://www.openfoam.org/download/history.php  

[12]   OpenFOAM  Foundation.  (2014).  Parallel  Computing.  Available:  http://openfoam.org/features/parallel-­‐computing.php  

[13]   OpenFOAM  Foundation.  (2014).  OpenFOAM®    Programmer's  C++  documentation  -­‐  Smagorinsky  Class  Reference.  Available:  http://foam.sourceforge.net/docs/cpp/a02317.html  

[14]   OpenFOAM  Foundation.  (2014).  OpenFOAM®    Programmer's  C++  documentation  -­‐  dynOneEqEddy  Class  Reference.  Available:  http://foam.sourceforge.net/docs/cpp/a00580.html  

[15]   G.  R.  Tabor  and  M.  H.  Baba-­‐Ahmadi,  "Inlet  conditions  for  large  eddy  simulation:  A  review,"  Computers  &  Fluids,  vol.  39,  pp.  553-­‐567,  2010.