• No results found

Designing Thermal Management Systems For Lithium-Ion Battery Modules Using COMSOL

N/A
N/A
Protected

Academic year: 2021

Share "Designing Thermal Management Systems For Lithium-Ion Battery Modules Using COMSOL"

Copied!
91
0
0

Loading.... (view fulltext now)

Full text

(1)

Designing Thermal

Management Systems For

Lithium-Ion Battery

Modules Using COMSOL

Emma Bergman

(2)

Abstract  

In  this  thesis,  a  section  of  a  lithium  ion  battery  module,  including  five  cells  and  an  indirect   liquid  cooling  system,  was  modelled  in  COMSOL  Multiphysics  5.3a.  The  purpose  of  this  study   was  to  investigate  the  thermal  properties  of  such  a  model,  including  heat  generation  per  cell   and  temperature  distribution.  Additionally,  the  irreversible  and  reversible  heat  generation,   the  cell  voltage  and  the  internal  resistance  were  investigated.  The  study  also  includes  the   relation  between  heat  generation  and  C-­‐rates,  and  an  evaluation  of  COMSOL  Multiphysics   5.3a  as  a  software.  

It  was  found  that  having  liquid  cooling  is  beneficial  for  the  thermal  management,  as  the   coolant  flow  helps  to  transfer  away  the  heat  generated  within  the  battery.  The  results  also   show  that  it  is  important  to  not  go  below  a  set  cell  voltage  at  which  the  cell  is  considered   fully  discharged.  If  a  control  mechanism  to  stop  the  battery  is  not  implemented,  the   generated  heat,  and  consequently  the  temperature,  increase  drastically.  COMSOL  

Multiphysics  5.3a  was  considered  a  suitable  software  for  the  modelling.  For  future  research   it  is  of  interest  to  expand  the  model  to  a  full  scale  module  to  fully  investigate  the  

temperature  distribution  where  more  cells  are  being  cooled  by  the  same  coolant  loop.    

(3)

Acknowledgements  

First  I  would  like  to  thank  my  supervisor  at  Northvolt  Ehsan  Haghighi  for  all  the  help  and   guidance  during  the  project.  I  want  to  thank  my  supervisor  at  KTH  Göran  Lindbergh  for   valuable  help  during  the  project.  I  would  also  like  to  thank  Henrik  Ekström  at  KTH  and  

COMSOL  for  all  the  help  and  explanations  regarding  the  construction  of  the  battery  model  in   COMSOL.  I  would  also  like  to  thank  Per  Backlund  and  Daniel  Ericsson  at  COMSOL  for  

additional  help  with  the  model.  Finally  I  would  like  to  thank  everyone  at  Northvolt  who   helped  me  out  in  various  ways  during  the  project.    

(4)

Table  of  Contents  

Abstract   1  

Acknowledgements   2

Table  of  Contents   3

Introduction  and  Project  Description   4

Background   5

Li(Ni1/3Mn1/3Co1/3)O2  (NMC)  Li-­‐Ion  Batteries   5

Generated  Heat   6

Cooling  systems   11

Methodology   15

Model   15

Study   16

Results  and  Discussion   20

Inlet  Flow  Rate   20

Driving  Cycle   23

Current   41

Conclusions  and  future  work   42

Nomenclature  and  Abbreviations   44

Abbreviations   44

Nomenclature   44

References   46

Appendix   48

Appendix  0:  Calculations  complement   48

Appendix  1:  Scale  adjusted  plots  and  3D  temperature  model  for  Driving  2  cycle  1   51

Appendix  2:  Plots  for  the  C-­‐Rate  Measurements   54

Appendix  3:  Parameter  values  used  in  COMSOL  model   57

Appendix  4:  Lithium  ion  cell  variables  used  in  COMSOL  model   59 Appendix  5:  Coolant  heat  capacity  interpolation  used  in  COMSOL  model   60 Appendix  6:  Coolant  density  interpolation  used  in  COMSOL  model   61 Appendix  7:  Coolant  dynamic  viscosity  interpolation  used  in  COMSOL  model   62 Appendix  8:  Coolant  thermal  conductivity  interpolation  used  in  COMSOL  model   63

Appendix  9:  Modelling  Instructions   64

(5)

Introduction  and  Project  Description  

Lithium  ion  batteries  are  lightweight  energy-­‐dense  batteries  ideal  for  both  portable  and   stationary  uses.  The  interest  in  lithium  ion  batteries  is  high  both  from  the  appliance  and  the   vehicle  industries.  However,  one  issue  with  lithium  ion  batteries  is  that  they  are  badly   affected  by  high  temperatures.  A  prolonged  exposure  to  high  temperatures  can  decrease   the  life  time  of  the  battery.  If  the  battery  is  exposed  to  high  enough  temperatures,  a  thermal   runaway  might  occur,  which  can  cause  the  battery  to  explode.  Consequently,  the  area  of   thermal  control  is  of  high  interest  for  battery  system  producers.    

The  goal  of  this  project  is  to  determine  the  effectiveness  of  liquid  cooling  system  for  a  set  of   five  21700  lithium  ion  batteries,  which  are  a  part  of  a  bigger  module.  This  will  be  

investigated  through  simulations  in  COMSOL  Multiphysics  5.3a.  A  model  is  constructed,   which  will  calculate  the  heat  generation  from  the  electrochemistry  given  certain  input   parameters.  To  analyse  the  system,  some  parameters  such  as  fluid  velocity,  current,  and   driving  cycle  are  varied  to  investigate  the  effect  of  temperature  distribution  and  heat   generation.    

(6)

Background  

Li(Ni1/3Mn1/3Co1/3)O2  (NMC)  Li-­‐Ion  Batteries  

Lithium  ion  batteries  are  relatively  new  on  the  market,  and  the  demand  for  them  have   grown  rapidly.  They  have  high  energy  density,  which  make  them  ideal  for  portable  devices   requiring  small  and  lightweight  batteries.  The  interest  in  Li-­‐ion  batteries  is  also  large  from   the  automobile  industry.  The  properties  of  the  energy  dense  Li-­‐ion  batteries  are  desired  for   electric  and  hybrid  vehicles.  

Most  lithium  ion  batteries  typically  utilize  a  graphite  material  for  the  negative  electrode.  The   positive  electrode  consists  of  the  lithium  in  a  metal  oxide  form,  usually  mixed  with  other   metals.  One  such  positive  electrode  material  is  NMC,  or  Li(Ni1/3Mn1/3Co1/3)O2.  It  is  showing   promising  properties  in  regards  to  being  used  for  electric  vehicle,  as  it  has  a  high  energy   density  and  a  large  rechargeable  capability.  The  combinations  of  metals  in  the  positive   electrode  adds  stability  to  the  cell,  in  addition  to  making  the  electrode  high  performing  and   cost-­‐effective  [1].  The  recommended  electrolyte  for  Li-­‐ion  cells  is  1M  LiPF6  3:7  EC-­‐EMC  [2].    

Cylindrical  lithium-­‐ion  batteries  consist  of  a  can  containing  a  jellyroll  of  the  cathode,  anode,   and  separator,  in  addition  to  terminals,  current  collectors,  insulation  plates,  and  safety   features.  This  is  illustrated  in  Fig.1  [3].  The  jellyroll  is  constructed  by  alternating  cathode  and   anode  sheets,  with  separator  sheets  between  them.  It  has  been  given  its  nickname  by  how  it   is  rolled  up  to  fit  the  cylindrical  cell,  similar  to  the  jelly  roll  pastry.  The  current  generated  by   the  batteries  is  collected  through  the  current  collectors,  the  positive  and  negative  ends  [4].    

 

  Fig  1:  Configuration  of  a  cylindrical  lithium  ion  battery  cell  [4].  

(7)

Generated  Heat  

Being  able  to  measure  the  generated  heat  is  an  important  step  of  battery  system  modelling.  

If  the  generated  heat  is  too  high,  it  may  cause  damage  to  the  system,  and  possibly  even   cause  the  system  to  explode  in  the  event  of  a  rapid  thermal  runaway.  A  thermal  runaway  is   caused  by  high  temperatures  in  the  battery  allowing  for  undesired  exothermic  reactions  to   occur.  This  in  turn  causes  even  higher  temperatures,  which  allow  for  even  more  undesired   reactions.  At  high  enough  temperatures,  this  phenomenon  is  irreparable  and  might  even   cause  an  explosion  due  to  heat  and  phase  shift  to  combustable  gas.  It  is  initiated  by  the   melting  of  the  protective  solid  electrolyte  interface  layer  at  90  °C.  Once  the  layer  is  gone,  the   electrolyte  and  the  negative  electrode  are  able  to  react  with  each  other,  causing  the  first   exothermic  reaction  at  a  temperature  of  100  °C  [4].    However,  already  before  that  level,  the   battery  calendar  life  is  estimated  to  decrease  significantly  for  temperatures  at  40  °C  and   higher  [5].    

Consequently,  battery  developers  are  looking  into  methods  of  measuring  the  general  heat,   in  order  to  produce  sufficient  cooling  systems.    

 

Measuring  the  heat  directly  

The  first  method  uses  an  equation  developed  by  Bernardi  et.  al.  [6],  which  separates  the   generated  heat  into  reversible  and  irreversible  heating.  In  its  simplified  form  the  equation  is:  

𝑄 = 𝐼 𝑈%&− 𝑉 − 𝐼𝑇*+*, =   𝐼.𝑅 − 𝑇 ∙ ∆𝑆 ∙43         (Eq.1)  

Where  𝑄  is  the  heat  generation,  I  is  the  current,  UOC  is  the  open-­‐circuit  potential,  V  is  the  cell   potential,  T  is  the  temperature,  R  is  the  overpotential  resistance,  dU/dT  is  the  entropic  heat   coefficient,  ∆S  is  the  entropy  change,  I  is  the  current,  and  F  is  the  Faraday  constant.  The  first   half  of  the  equation  represents  the  irreversible  heating,  or  Joule  heating.  The  second  part  of   the  equation  represents  the  reversible  heating,  which  is  due  to  entropy  changes.  To  use  this   equation,  it  is  necessary  to  assume  no  heat  generation  from  mixing  or  phase  changes,  no   spatial  variations  in  temperature  or  SOC,  only  one  electrochemical  reaction  occurring  at   each  electrode,  and  that  the  Joule  heating  in  the  current  collectors  is  negligible  [7].  

The  method  is  quite  accurate,  but  difficult  and  time  consuming  to  measure  in  practice.  Onda   et.  al.  [8]  gives  four  methods  of  how  to  perform  experimental  measurements  of  the  

overpotential  resistance,  R.  These  are  to  measure  the  resistance  by  V-­‐I  characteristics,  by  

(8)

difference  between  OCV  and  cell  voltage,  by  intermittent  discharge,  and  by  an  ac  meter.  

Measuring  the  difference  between  OCV  and  cell  voltage  is  the  most  common  method,   however,  it  can  take  a  long  time  to  conduct  the  experiment,  since  the  OCV  needs  to  stabilize   after  the  change  in  SOC.  Onda  et.  al.  also  report  inconsistencies  for  the  last  two  measuring   methods.  To  measure  the  entropy  Onda  et.  al.  list  two  suggested  methods.  The  first  method   measures  the  entropy  change  by  temperature  gradient  of  OCV,  and  the  second  method   measures  entropy  change  by  heat  production.  This  measurement  is  another  time  consuming   operation,  as  the  OCP  must  stabilize  again.  Karimi  and  Li  [9]  performs  a  computational  study   using  Eq.1  as  an  alternative.  

 

Cooling  medium  heat  removal  

Another  method  looks  instead  to  the  heat  removed  by  the  cooling  system.  Calculating  the   heat  transported  away  by  the  system  requires  a  simpler  measurement,  but  neglect  heat   remaining  in  the  battery,  and  it  is  needed  to  be  calculated  separately  [10].  An  equation  for   the  cooling  medium  heat  removal  is:  

𝑄 = 𝑚𝐶7(𝑇9− 𝑇:)           (Eq.2)  

Where  𝑄  is  rate  of  heat  generation,  𝑚  is  the  mass  flow  rate  of  coolant,  Cp  is  the  specific  heat   capacity  of  coolant,  To  is  the  outlet  temperature,  and  Ti  is  the  Inlet  temperature.  The  

advantage  of  this  method  is  the  simple  measurement.  All  that  is  needed  is  to  measure  the   inlet  and  outlet  temperatures  of  the  cooling  medium.  However,  the  system  disregards  losses   and  heat  remaining  in  the  battery.  

 

Computational  Modelling  

The  methods  mentioned  are  the  main  experimental  methods  to  calculate  the  battery  heat   generation.  However,  since  the  experimental  methods  are  complex  and  can  be  inefficient   battery  developers  often  look  into  computational  models.    

COMSOL  Multiphysics  is  an  engineering  simulations  software.  It  allows  the  user  to  define  a   geometry  and  set  material  properties  and  physics  to  describe  a  process.  COMSOL  then   solves  the  system  through  built  in  or  defined  equations  [11].  In  addition  to  the  base  package,   the  add-­‐on  modules  Heat  Transfer  in  Solids  and  Fluids  and  Batteries  &  Fuel  Cells  are  useful   when  modelling  battery  heat  management.    

(9)

 

The  Pseudo-­‐Two-­‐Dimensional  model  (P2D)  

COMSOL  [12]  looks  at  the  electrochemical  reactions  directly,  based  on  the  model  developed   by  Newman’s  team  [13,  14].  The  model  is  the  most  widely  used  for  the  purpose  and  

approaches  the  cell  from  a  homogeneous  and  isothermal  point  of  view  [15].  It  considers  the   one-­‐dimensional  transport  from  the  negative  electrode  to  the  positive  electrode,  through   the  separator,  according  to  Ohm’s  law.  The  model  is  based  on  concentrated  solution  theory   and  porous  electrode  theory.  The  concentrated  solution  theory  works  in  the  way  that  it   treats  the  electrolyte  as  a  binary  salt  with  polymer  solvent.  This  theory  describes  the  mass-­‐  

and  charge  transport  in  the  electrolyte  phase.  The  porous  cathode  theory  works  in  that  the   composite  negative  electrode  can  be  modelled  looking  at  both  resistance  to  the  solid  state   transport  considering  both  kinetic  and  diffusional  effects.  It  adds  an  extra  dimension  to  the   model  to  describe  the  lithium  transport  according  to  Fick’s  law.  Furthermore,  the  use  of   Butler-­‐Volmer  kinetics  to  describe  the  reversible  process  allows  for  an  effective  method  of   describing  both  the  discharge  and  the  charge  processes.  Following  these  theories,  the  model   can  look  at  the  major  features  of  the  system,  without  making  it  too  complex  [12,  13].  

Although  it  is  not  stated  by  Newman’s  team  in  their  reports,  this  model  is  generally  referred   to  as  the  P2D,  or  Pseudo-­‐Two-­‐Dimensional,  model  in  literature  [4,  16].  The  pseudo-­‐

dimension  part  of  the  name  refers  to  how  the  equation  for  lithium  conservation  is  solved  in   the  particle  r-­‐dimension.    

 

Diffusion  in  Porous  Media    

Fick’s  second  law  is  one  of  the  governing  equations  for  the  P2D  model  [4].  It  describes   diffusion  with  a  linear  equation  that  assumes  a  constant  diffusion  coefficient,  D.  In  Li-­‐ion   batteries  this  equation  describes  the  transport  of  solid  Li  in  the  solid  electrode  phase.  Fick’s   second  law  is  written  [1]:  

<=

<> =  𝛻 ∙ (𝐷𝐿𝑖𝛻𝑐𝑒) (Eq.3)

Where  c  is  the  lithium  ion  concentration,  t  is  the  time,  DLi  is  the  lithium  diffusion  coefficient,   and  ce  is  the  lithium  ion  concentration  in  the  electrolyte  phase.  To  accurately  describe  the  

(10)

lithium  ion  diffusion,  Fick’s  law  also  requires  a  set  of  boundary  and  initial  conditions  for  the   time,  and  location  relative  to  the  radius.  These  are  as  follows:  

𝑐 = 𝑐E  𝑎𝑡  𝑡 = 0   𝐷I:<=

<J= 0  𝑎𝑡  𝑟 = 0    

𝐷I:𝜕𝑐

𝜕𝑟=𝑖M

𝐹  𝑎𝑡  𝑟 = 𝑟E    

COMSOL’s  battery  module  also  utilizes  the  Bruggeman  model,  tF  =  ep-­‐1/2,  as  a  correction   factor  for  the  porous  media  mass  transfer  [12].    

 

Lithium  material  balance  

The  lithium  material  balance  in  the  polymer  and  salt  phases  is  another  of  the  governing   equations  [13].  It  uses  the  transport  equation  for  concentrated  solutions.  

It  is  given  by  [4]:  

𝜀P<>< 𝑐P− ∇ ∙ 𝐷I:∇𝑐P:R4ST+ 𝑎V𝑗M 1 − 𝑡YE = 0       (Eq.4)   Where  e  is  the  electrode  porosity,  ce  is  the  lithium  ion  concentration  in  the  electrolyte   phase,  DLi  is  the  lithium  diffusion  coefficient,  ie  is  the  electrolyte  phase  current  density,  t+  is   the  transfer  number  of  lithium  ions,  F  is  the  Faraday  constant,  as  is  the  solid  phase  specific   interfacial  area,  and  jn  is  the  pore-­‐wall  flux  across  interface.  The  boundary  conditions  are   dependent  on  the  location  in  the  phase,  and  are  as  follows:  

𝜕𝑐P

𝜕𝑥 = 0  𝑎𝑡  𝑥 = 0  𝑎𝑛𝑑  𝑥 = 𝐿    

Butler-­‐Volmer  kinetics  

Butler-­‐Volmer  kinetics  are  included  in  the  P2D  model.  They  describe  the  charge  transfer   kinetics  process  at  the  interface  between  the  solid  electrode  and  the  electrolyte.  Using   Butler-­‐Volmer’s  equation  requires  setting  up  a  set  of  boundary  conditions.  The  expression   assumes  no  potential  gradients  at  the  interface  between  the  current  collector  and  the   electrolyte,  for  the  electrolyte,  or  at  the  interface  between  the  separator  and  the  electrode,   for  the  solid.  The  electrolyte  does  not  have  a  concentration  gradient  in  the  interface  either.  

The  concentration  gradient  at  the  surface  of  the  solid  particle  is  proportional  to  the  lithium   pore  wall  flux,  and  there  is  a  symmetry  for  the  Li-­‐ion  concentration  in  the  middle  of  the  

(11)

particles.  The  boundary  conditions  also  assume  that  the  applied  current  discharge  is   constant  [7].  

The  Butler-­‐Volmer  equation  is  written  [4]:  

𝑖 = 𝑖E exp a𝑎𝑅𝑇𝐹h𝑠 − exp a𝑅𝑇𝑐𝐹h𝑠         (Eq.5)  

where  i  is  the  current  density,  i0  is  the  exchange  current  density,  aa  anode  transfer  

coefficient,  F  is  the  Faraday  constant,  hs  is  the  surface  overpotential,  R  is  the  overpotential   resistance,  and  T  is  the  temperature.

The  exchange  current  density,  i0,  is  defined:  

𝑖E =𝐹𝑘𝑎a𝑐𝑘𝑐a𝑎 𝑐𝑚𝑎𝑥𝑠 − 𝑐𝑠 a𝑐𝑐𝑒a𝑎         (Eq.6)   where  F  is  the  Faraday  constant,  ka  is  the  anodic  reaction  rate  constant,  kc  is  the  cathodic   reaction  rate  constant,  csmax  is  the  maximal  concentration  of  lithium  ions  in  solid  phase,  cs  is   the  concentration  of  lithium  ions  in  solid  phase,  and  ce  is  the  concentration  of  lithium  ion  in   the  electrolyte  phase.  

The  overpotential,  hs,  is  defined:  

hV = ØV− ØP− 𝑈%&           (Eq.7)  

where  Øs  is  the  solid  phase  potential,  Øe  is  the  electrolyte  phase  potential,  and  UOC  is  the   open-­‐circuit  potential.    

   

Concentrated  solution  theory  

Concentrated  solution  theory  describes  another  of  the  governing  equations  for  the  P2D   model.  The  equation  predicts  the  potential  variation  in  the  separator  from  the  material   balance  of  the  lithium  salt.  With  the  assumption  that  solvent  concentration  is  independent   of  electrolyte  concentration,  the  equation  can  be  derived  [13].  

The  equation  read  [4]:  

𝑖P+ 𝑘PccÑØP.de,f4 Rgg 1 +<hM=<hMc±

R 1 − 𝑡YE Ñ𝑙𝑛𝑐P = 0     (Eq.8)  

where  ie  is  the  current  density  in  the  electrolyte  phase,  keff  is  the  effective  ionic  conductivity,   Øe  is  the  electrolyte  phase  potential,  Rg  is  the  universal  gas  constant,  T  is  the  temperature,  F   is  the  Faraday  constant,  f±  is  the  molecular  salt  activity  coefficient,  ce  is  the  concentration  of   lithium  ion  in  the  electrolyte  phase,  and  t+  is  the  transfer  number  of  lithium  ions.  

It  is  controlled  by  the  following  location  dependent  boundary  condition  [4]:  

(12)

𝜕ØP  

𝜕𝑥 = 0  𝑎𝑡  𝑥 = 0  𝑎𝑛𝑑  𝑥 = 𝐿    

Porous  electrode  theory  

The  porous  electrode  theory  describes  the  Li-­‐ion  battery’s  internal  changes  and  status,   depending  on  its  electrodes  and  electrolytes,  as  well  as  the  battery’s  structure  [17].  The   theory  consists  of  Ohm’s  law,  which  administers  the  movement  of  electrons  [4,  18]:  

𝑖V = −sPccÑØV             (Eq.9)  

where  is  is  the  solid  phase  current  density,  seff  is  the  solid  phase  effective  electronic   conductivity,  and  Øs  is  the  solid  phase  potential.  

It  is  regulated  by  a  set  of  location  dependent  boundary  conditions  [4]:  

−sPcc𝜕ØV  

𝜕𝑥 = 𝐼

𝐴  𝑎𝑡  𝑥 = 0  𝑎𝑛𝑑  𝑥 = 𝐿  

𝜕ØV  

𝜕𝑥 = 0  𝑎𝑡  𝑥 = 𝐿M  𝑎𝑛𝑑  𝑥 = 𝐿M+ 𝐿VP7    

Cooling  systems  

To  dissipate  the  heat  generation  out,  cooling  systems  are  widely  used.  Passive  systems  that   let  ambient  air  reach  the  battery  are  the  simplest,  while  liquid  and  combined  phase  change   cooling  systems  can  be  quite  complex.  The  desired  temperature  range  is  usually  between   25°C  and  35°C,  as  too  high  or  low  temperatures  can  reduce  the  effectiveness  of  the  battery   [15].  For  air  and  liquid  cooling,  the  thermal  management  system  could  also  heat  up  the   battery  cells  in  events  of  low  temperatures.  

 

Air  Cooling  

Cooling  with  air  is  a  traditional  and  widely  used  thermal  regulation  approach.  Most  systems   are  passively  air  cooled  if  they  have  at  least  one  interface  in  contact  with  surrounding   ambient  air.  Besides  passive  air  cooling,  systems  can  also  be  cooled  actively  with  the  use  of   fans  and  methods  to  cool  the  air  to  lower  than  ambient.  This  is  due  to  passive  air  cooling   having  a  lower  convective  heat  transfer  coefficient  than  active  air  cooling.  Since  the  passive   air  cooling  convective  heat  transfer  is  so  low,  it  is  only  effective  for  really  small  systems  that   do  not  produce  large  amounts  of  heat  [4].  However,  a  huge  advantage  of  air  is  its  light   weight,  and  ease  to  circulate  [19].  Active  air  cooling  was  adapted  early  by  the  first  electric  

(13)

hybrid  vehicles,  2000  Honda  Insight  and  2001  Toyota  Prius.  They  take  advantage  of  the  car’s   available  air  conditioning  system  to  take  the  cooled  air  from  the  cabin  to  cool  the  battery   before  it  is  exhausted.  A  blower  that  can  operate  at  various  speeds  is  utilised  to  draw  air   from  the  cabin  to  the  battery.  The  two  cars  use  slightly  different  systems  to  ensure  even   heat  distribution  [7].  However,  for  a  fully  electric  vehicle  there  is  doubt  that  an  air  cooling   system  would  be  sufficient  [15].  

 

Liquid  Cooling  

Cooling  the  battery  with  liquid  is  an  area  of  interest.  Liquids  have  higher  thermal  

conductivities  than  air  and  can  cool  more  effectively  and  uniformly.  It  is  also  possible  to   decrease  the  size  of  the  battery  pack  as  the  cells  can  be  placed  closer  to  each  other.  

However,  liquids  can  be  less  flexible  than  air  and  it  is  of  high  importance  that  they  do  not   leak  out  onto  the  battery  [19].  There  are  three  major  liquid  cooling  methods.  The  first  is   direct  submersion  into  the  liquid.  The  second  is  indirect  cooling  through  placing  the  battery   modules  on  a  cooling  plate  that  cools  the  batteries  from  the  bottom  up.  The  third  method  is   also  indirect,  where  cooling  tubes  or  jackets  are  placed  around  the  batteries  [15].  All  three  of   the  methods  have  their  own  advantages  and  disadvantages.    

 

Direct  Submersion  

Direct  submersion  of  the  battery  in  the  cooling  liquid  is  the  most  straightforward  liquid   cooling  method.  The  battery  unit  surfaces  are  directly  in  contact  with  the  coolant,  which   minimizes  the  thermal  resistance  between  battery  and  coolant.  However,  a  disadvantage  of   using  direct  submersion  is  that  it  requires  the  coolants  to  be  dielectric.  Consequently,  highly   viscous  fluids  are  common.  These  viscous  coolants  cause  a  higher  power  consumption  for   circulating  the  fluid  [4].  The  reason  dielectric  fluids  are  needed  is  to  avoid  short  circuit.  Since   these  fluids  often  are  oil  based,  it  is  also  important  to  consider  other  properties,  such  as   toxicity  and  flammability,  to  not  switch  one  problem  for  another.  There  are  coolants   available  that  can  reduce  the  maximum  temperature  if  thermal  runaway  occurs.  Although   for  these  fluids,  economical  factors  might  also  play  in,  as  the  fluid  cost  can  greatly  increase   the  cost  of  the  entire  system  [20].  When  considering  direct  submersion  cooling  for  batteries   to  be  used  in  electrical  vehicles,  the  risk  of  short  circuit  is  still  significant  since  the  cells  are  in   direct  contact  with  the  fluid.  Having  direct  submersion  also  makes  it  more  difficult  to  replace  

(14)

faulty  cells.  With  this  in  mind,  direct  submersion  cooling,  although  effective,  is  not  ideal  for   batteries  to  be  used  in  electrical  vehicles.    

 

Indirect  cooling  with  cooling  plates  

With  indirect  cooling,  an  extra  factor  arises  in  form  of  a  layer  between  the  coolant  and  the   battery  cell.  It  is  important  to  consider  convection  heat  transfer  and  thermal  contact  

resistance  here,  as  those  properties  affect  the  effectiveness  of  the  cooling  system  more  than   the  extra  thermal  resistance  itself  [20].  Cooling  plates  are  placed  under  the  battery  cells.  The   plates  are  thin,  with  a  cooling  media  transported  between  the  plates  [15].  The  cooling   channels  in  between  the  cooling  plates  can  be  of  various  styles,  ranging  from  straight   channels  to  complex  structures.  The  batteries  are  cooled  from  the  bottom  up,  causing  a   temperature  difference  between  top  and  bottom.  Adding  an  additional  heat  plate  at  the  top   can  prevent  this.  A  further  step  to  obtain  more  uniform  cooling  is  to  add  fins,  which  are   additional  plates  going  between  the  cells.  However,  these  measures  add  weight  to  the   cooling  system  [20].    

 

Indirect  cooling  with  cooling  tubes  or  jackets  

Cooling  tubes  and  cooling  jackets  are  two  indirect  methods  of  cooling  the  battery.  Cooling   tubes  usually  consists  of  a  series  of  wavy  tubes  placed  alongside  the  battery  cells,  so  that   they  have  a  contact  area  on  one  side.  Cooling  jackets  usually  form  casings  around  the   battery,  and  then  allows  the  coolant  to  flow  through  the  compartment  surrounding  the   casings.  Cooling  jacket  can  thus  cover  the  entire  battery,  while  cooling  tubes  have  a  smaller   contact  area.  However,  a  great  advantage  of  cooling  tubes  over  cooling  jackets  is  that  the   safety  increase,  as  the  possibility  of  liquid  leaking  out  is  smaller.  With  cooling  tubes,  it  is   possible  to  have  all  fluid  connections  to  the  tube  outside  of  the  battery  model.  Cooling  tubes   also  have  fewer  welding  spots.  Another  advantage  of  cooling  tubes  is  that  they  have  a   smaller  volume  and  the  total  weight  is  less,  which  is  advantageous  for  mobile  uses  [20].  

However,  it  should  be  noted  that  the  risk  of  leakage  is  still  smaller  for  a  heating  jacket  than   many  other  solutions  [19].  Electric  vehicle  producer  Tesla  has  patented  both  cooling  tubes   and  cooling  jackets  for  use  in  their  vehicles  [21,  22].  The  Tesla  solutions  are  shown  in  fig.  2   and  fig.  3.  To  improve  thermal  conductivity  while  electrically  insolating  the  transfer,  a   thermal  interface  material,  TIM,  is  added  between  the  cell  and  cooling  container  [23].    

(15)

  Fig  2:  Tesla  cooling  case  patent  [22]     Fig  3:  Tesla  cooling  tubes  patent  [21]  

 

Phase  Change  Material  (PCM)  Cooling    

Cooling  through  phase  change  materials  utilizes  the  PCM  as  a  heat  sink  during  battery   discharge.  The  PCM  collects  the  heat  and  goes  through  a  phase  change,  from  a  less  energy   dense  phase  to  a  phase  with  more  free  energy,  such  as  solid  to  liquid  or  liquid  to  gas.  PCM   cooling  is  a  passive,  rather  than  active,  cooling  method.  The  PCM  cycles  through  the  phases   between  battery  discharge  and  standby.  A  disadvantage  of  PCM  cooling  systems  are  that   they  do  not  function  well  in  extreme  weather.  If  it  is  too  hot,  the  PCM  might  melt  completely   and  stop  working  as  a  heat  sink.  In  too  cold  weather  the  PCM  will  be  difficult  to  melt,  adding   a  large  thermal  inertia  which  requires  significant  energy  to  warm  up  [20].  It  is  extremely   important  to  consider  the  melting  point  of  the  PCM.  The  ideal  melting  point  should  be  within   the  temperature  range  the  battery  operates.  The  ideal  PCM  has  the  perfect  balance  of   thermal  conductivity.  Low  thermal  conductivities  cause  uneven  melting,  which  can  lower  the   effective  PCM  cooling.  Too  high  thermal  conductivities  cause  the  entire  PCM  to  melt,  which   renders  it  unable  to  function  properly.  Other  properties  such  as  toxicity,  stability,  and   flammability  are  important  to  consider  for  safety  aspects  [15].    PCM  systems  are  not  used  in   commercial  electric  vehicle  battery  systems  today,  as  research  within  the  area  still  have   some  grounds  to  cover.    

 

 

(16)

Methodology  

Model  

The  study  was  conducted  through  creating  a  model  of  a  five-­‐cell  battery  and  cooling  tube   thermal  system  in  COMSOL  Multiphysics  5.3a.  The  model  includes  electrochemistry,  heat   transfer,  and  liquid  flow  physics.  The  electrochemistry  is  based  on  the  P2D  model,  through   COMSOL’s  Lithium-­‐Ion  Battery  interface.  The  model  uses  a  coupling  function  to  pair  the  1D   electrochemical  model  with  a  3D  thermal  model.  Interpolations  are  set  up  to  describe  the   temperature  affected  properties  of  the  coolant.  For  the  studied  effects,  equations  utilizing   COMSOL’s  built  in  equations  where  formulated  where  they  could  not  be  collected  directly.  

The  equation  to  describe  the  internal  resistance  of  the  model  is  written:  

𝐼𝑛𝑡𝑒𝑟𝑛𝑎𝑙  𝑟𝑒𝑠𝑖𝑠𝑡𝑎𝑛𝑐𝑒 = 𝐸𝑂𝐶𝑉79V− 𝐸𝑂𝐶𝑉MPm − 𝐸=Phh

𝐼n77  

This  equation  is  derived  over  time  to  plot  the  time  dependent  internal  resistance  for  set   parameters  and  drive  cycle.  EOCVpos  and  EOCVneg  are  variables  calculated  from  the  

equilibrium  potential  for  the  current  state  of  charge  for  that  electrode.  Ecell  is  calculated  by   integrating  the  cell  voltage  over  the  positive  current  collector.  

The  reversible  heat  production  is  calculated  by  integrating  the  reversible  heat  source  in   W/m3  over  the  two  electrodes,  and  then  multiplying  it  with  the  area  of  the  jellyroll.    

The  irreversible  heat  is  all  heat  that  is  not  reversible,  so  it  was  calculated  by  subtracting  the   reversible  heat  from  the  total  heat  production.    

The  model  assumes  that  heat  transfer  only  occurs  between  contact  surfaces  in  the  model.  

No  heat  is  being  transferred  to  the  air  or  between  batteries  separated  by  air.  Heat  is  defined   as  being  produced  uniformly  in  the  battery  cylinders.  However,  heat  is  being  transferred   with  different  heat  transfer  coefficient  depending  on  direction  in  the  battery  cell.  The  heat   transfer  vertically,  or  along  the  jellyroll  layers  is  higher  than  the  radial  heat  transfer  that   crosses  the  jellyroll  layers.    

Parameter  data  used  in  the  model  were  obtained  from  fact  sheets,  through  in-­‐person   communication,  through  calculations,  and  through  built-­‐in  COMSOL  data  when  it  was   deemed  comparable.    

The  3D  model  for  thermal  modelling  was  created  in  the  CAD  software  SolidWorks  and   imported  to  COMSOL.  This  model  is  shown  in  Fig.  4.  

(17)

A  full  description  to  re-­‐create  the  model  and  a  list  of  constant  parameters  can  be  found  in   Appendix  9.    

 

Fig.  4:  Geometrical  model  of  5-­‐cell  battery  module  section  

Study  

To  conduct  the  study  of  the  cooling  tube  cooling  system’s  effectiveness,  two  areas  of   interest  were  identified.  These  are  the  inlet  flow  rate  and  the  driving  cycle.  In  addition  to   those,  it  was  also  investigated  what  effect  the  current  has  on  the  heat  generation.  

 

Inlet  Flow  Rate  

The  suggested  normal  flow  rate  for  the  system  is  1  liter  per  minute  of  coolant  entering  the   inlet.  Flow  rates  of  0.5  liter  per  minute  higher  and  lower  than  the  suggested  flow  rate  were   tested  to  determine  the  effect  of  changing  the  velocity.  A  simulation  was  also  run  with  an   extremely  low  flowrate  to  illustrate  the  situation  at  no  flow.  All  of  these  were  run  for  the   specific  driving  cycle  Driving  1  Cycle  1,  which  is  described  in  the  Driving  Cycle  section.    

For  each  flowrate,  the  maximum  temperature  reached  and  minimum  temperature  in  a   battery  cell  at  the  same  time  was  obtained.    

 

(18)

Driving  Cycle  

For  the  set  flowrate  of  1  liter  per  minute,  the  effect  of  five  different  driving  cycles  was   simulated.  The  simulations  included  temperature,  cell  voltage,  internal  resistance,  total  heat   production,  and  how  the  total  heat  is  split  into  reversible  and  irreversible  heat.  This  is  to  see   how  different  uses  affect  the  battery  system.  The  five  driving  cycles  are  pictured  in  Fig.  5-­‐9   with  the  y-­‐axis  as  C-­‐rate,  and  the  exact  data  is  given  in  Table  1.  

 

  Fig.  5:  Driving  1  Cycle  1                      Fig.  6:  Driving  1  Cycle  2  

 

  Fig.  7:  Driving  2  Cycle  1                      Fig.  8:  Driving  2  Cycle  2  

 

(19)

  Fig.  9:  Driving  3  

 

Table  1:  Driving  cycle  data  

Current  

An  area  of  additional  interest  is  the  effect  of  current  on  heat  production.  To  analyze  this   effect,  simulations  were  run  for  various  C-­‐rates.  A  1C  C-­‐rate  is  set  to  3.2  A.  With  this  as  a   base,  eight  different  C-­‐rates  were  run  for  both  a  charge  and  discharge  process,  until  they   were  either  fully  charged  of  fully  discharged.  This  was  defined  as  reached  the  cell  voltage   limits  of  3.0  V  as  fully  discharged  and  4.2  V  for  fully  charged.  

The  C-­‐rates  included  in  the  study  are  listed  in  Table  2.  

Driving  1  Cycle  1   Driving  1  Cycle  2   Driving  2  Cycle  1   Driving  2  Cycle  2   Driving  3   Duration  

(s)   C-­‐rate   Duration  

(s)   C-­‐rate   Duration  

(s)   C-­‐rate   Duration  

(s)   C-­‐rate   Duration  

(s)   C-­‐rate   270   -­‐1.381   1300   0.581   20   -­‐0.306   2550   -­‐0.972   480   -­‐0.9   180   -­‐0.025   180   -­‐0.025   25   -­‐0.4   2550   0.638   10   -­‐1.028   350   1.078   2600   -­‐1.472   20   -­‐0.419         900   0.791  

20   -­‐0.134   20   -­‐0.134   10   -­‐0.122         120   0.119  

              20   -­‐0.306         120   0.238  

              25   -­‐0.4         120   0.475  

              20   -­‐0.419         120   0.791  

              10   -­‐0.122         120   0.119  

              20   -­‐0.306         1200   -­‐1.263  

              25   -­‐0.4         10   -­‐1.028  

              20   -­‐0.419            

                180   -­‐0.028              

Total:  820  s   Total:  4100  s   Total:  395  s   Total:  5100  s   Total:  3200  s  

(20)

 

0.1C     Charge  

Discharge  

0.3  C   Charge  

Discharge  

0.5C     Charge  

Discharge  

0.7C    

  Charge  

Discharge  

1C     Charge  

Discharge  

1.2C    

  Charge  

Discharge  

1.4C    

  Charge  

Discharge  

1.5C    

  Charge  

Discharge  

 Table  2:  C-­‐rates  included  in  study    

The  discharging  models  where  given  initial  lithium  ion  concentrations  of  26814  mol/m3  in   the  negative  electrode,  and  22995  mol/m3  in  the  positive  electrode,  values  assumed  to   represent  the  defined  fully  charged  battery.  The  charging  models  were  given  initial  lithium   ion  concentrations  of  7921.3  mol/m3  in  the  negative  electrode,  and  41318  mol/m3  in  the   positive  electrode,  values  assumed  to  represent  the  defined  fully  discharged  battery.  The   defined  battery  statuses  are  not  the  same  as  the  completely  drained  or  charged  battery,  but   set  as  limits  for  desired  usage.    

 

 

(21)

Results  and  Discussion  

 

Inlet  Flow  Rate  

Temperature  data  and  curves  were  obtained  for  the  four  flowrates.  As  the  driving  cycle   remained  constant  Driving  1  cycle  1,  temperature  is  the  only  changing  factor  in  this  

experiment.  Fig.  10-­‐13  show  how  the  temperature  inside  the  battery  varies  over  time  for  the   four  different  velocities.  Table  2  summarizes  the  data  through  the  maximum  temperature   reached  at  any  point  of  time,  the  minimum  temperature  in  a  battery  cell  at  that  time,  and   the  coolant  outflow  temperature.  All  figures  are  modelled  on  the  same  time  and  

temperature  scale  for  simplified  comparison.    

 

  Fig.  10:  Temperature  over  time  at  flowrate  0  liter  per  minute  

(22)

  Fig.  11:  Temperature  over  time  at  flowrate  0.5  liter  per  minute  

 

  Fig.  12:  Temperature  over  time  at  flowrate  1  liter  per  minute  

(23)

  Fig.  13:  Temperature  over  time  at  flowrate  1.5  liter  per  minute  

 

Test   Max  temp.   Min  temp.     Coolant  Outflow  Temp.    

Flow  Rate   0  L/min   25.4  °C   24.6  °C   24.6  °C  

   

0.5  L/min   22.8  °C   20.8  °C   20.03  °C   1  L/min   22.7  °C   20.8  °C   20.02  °C   1.5  L/min   22.6  °C   20.7  °C   20.01  °C   Table  2:  Temperature  data  for  the  different  flow  rates.    

 

The  data  shows  that  having  no  coolant  flow  gives  a  noticeable  effect,  where  the  

temperature  is  significantly  higher  than  for  even  the  lowest  coolant  flowrate  at  0.5  liter  per   minute.  Between  the  different  flow  rates,  the  difference  is  smaller.  Between  0.5  liter  per   minute  (fig.  11)  and  1.5  liter  per  minute  (fig.  13),  the  coolant  flow  rate  has  been  tripled.  

However,  the  difference  in  maximum  temperature  in  the  cells  is  only  0.2  °C.  The  coolant   outflow  temperature  has  decreased  by  0.02  °C  between  the  two  simulations.    

The  temperature  rises  when  the  battery  is  in  use,  as  heat  is  being  generated.  When  the   battery  is  not  in  no  additional  heat  is  being  generated,  and  two  different  scenarios  are  visible   from  the  figures.  In  Figure  10  with  the  0  liter  per  minute  flowrate,  the  temperature  stabilizes  

(24)

at  this  time  stamp.  For  the  other  three  flow  rates,  the  temperature  decreases,  as  the  coolant   can  transfer  away  more  heat  than  what  is  being  produced.    

 

Driving  Cycle  

For  the  five  different  driving  cycles,  the  data  obtained  includes  cell  voltage,  heat  generation,   heat  generation  separated  into  reversible  and  irreversible  heat,  temperature,  and  internal   resistance.  They  were  all  given  a  coolant  flow  rate  of  1  liter  per  minute.  Thus,  the  

temperature  obtained  for  the  Driving  1  cycle  1  cycle  is  the  same  as  for  the  1  liter  per  minute   flow  rate  given  above.  

 

Cell  Voltage  

The  cell  voltage  is  shown  in  Fig.  14-­‐18.  The  plots  also  contain  the  drive  cycle,  to  illustrate   how  they  depend  on  each  other.  A  bottom  limit  of  3.0  V  and  an  upper  limit  of  4.2  V  has  been   defined  in  product  sheets  as  the  cell  voltages  that  equals  to  0  %  SOC  and  100  %  SOC.  In  the   cycle  Driving  1  cycle  2,  the  drive  cycle  has  been  allowed  to  continue  past  this  limit  to   illustrate  what  would  happen.  All  cycles  have  been  put  on  the  same  y-­‐axes  for  simplified   comparison.  The  Driving  1  cycle  2  cycle  will  be  difficult  to  read  due  to  the  wide  range  of  the   axis,  and  a  scale-­‐adjusted  version  is  available  in  Appendix  1.  

 

  Fig.  14:  Cell  voltage  over  time  for  Driving  1  cycle  1  

(25)

 

  Fig.  15:  Cell  voltage  over  time  for  Driving  1  cycle  2  

  Fig.  16:  Cell  voltage  over  time  for  Driving  2  cycle  1  

(26)

  Fig.  17:  Cell  voltage  over  time  for  Driving  2  cycle  2  

  Fig.  18:  Cell  voltage  over  time  for  Driving  3  

 

In  Fig.  14-­‐18,  it  can  be  clearly  seen  how  the  cell  voltage  is  related  to  the  C-­‐rate.  When  the  C-­‐

rate  is  negative,  implying  that  the  battery  is  being  discharged,  the  cell  voltage  drops.  When   the  C-­‐rate  is  zero,  implying  that  the  battery  is  not  in  use,  the  cell  voltage  stabilizes.  When  the   C-­‐rate  is  positive,  implying  that  the  battery  is  being  charged,  the  cell  voltage  increases.  For   all  the  tested  driving  cycles,  except  for  Driving  1  cycle  2  (Fig.15),  the  cell  voltage  stays  within   the  given  limits.  Within  the  limits,  the  cell  voltage  does  not  increase  or  decrease  as  rapidly  as  

(27)

it  can  be  seen  dropping  once  the  cell  voltage  goes  below  3.0  V  in  Driving  1  cycle  2,  without   changing  the  C-­‐rate.    

 

Heat  Production  per  battery  cell,  including  reversible  and  irreversible  heat  

The  heat  production,  and  the  heat  separated  reversible  and  irreversible  heat  for  an   individual  battery  cell  in  the  model  is  shown  in  Fig.  19-­‐23,  respectively  Fig.  24-­‐28.  In   addition,  cut  sections  of  the  cycle  Driving  3  has  been  added  to  show  how  the  heat  is  

dissipated  from  inside  the  battery.  The  total  heat  adds  the  reversible  and  irreversible  heat.  It   is  assumed  that  the  cells  produce  equal  heat.  The  driving  cycle  is  included  to  illustrate  the   relation.  Again,  the  cycle  Driving  1  cycle  2  is  out  of  the  ordinary  range,  and  will  thus  produce   more  heat  than  the  other  cycles.  All  cycles  have  been  put  on  the  same  y-­‐axes  for  simplified   comparison.  The  Driving  1  cycle  2  cycle  will  be  difficult  to  read  due  to  the  wide  range  of  the   axis,  and  a  scale-­‐adjusted  version  is  available  in  Appendix  1.  

 

  Fig.  19:  Total  heat  production  over  time  for  a  single  cell  over  Driving  1  cycle  1  

(28)

  Fig.  20:  Total  heat  production  over  time  for  a  single  cell  over  Driving  1  cycle  2  

  Fig.  21:  Total  heat  production  over  time  for  a  single  cell  over  Driving  2  cycle  1  

(29)

  Fig.  22:  Total  heat  production  over  time  for  a  single  cell  over  Driving  2  cycle  2  

  Fig.  23:  Total  heat  production  over  time  for  a  single  cell  over  Driving  3    

(30)

 

Fig.  24:Reversible  and  irreversible  heat  production  over  time  for  a  single  cell  over  Driving  1  cycle1  

 

Fig.  25:Reversible  and  irreversible  heat  production  over  time  for  a  single  cell  over  Driving  1  cycle2  

(31)

 

Fig.  26:Reversible  and  irreversible  heat  production  over  time  for  a  single  cell  over  Driving  2  cycle1  

 

Fig.  27:Reversible  and  irreversible  heat  production  over  time  for  a  single  cell  over  Driving  2  cycle2  

(32)

  Fig.  28:  Reversible  and  irreversible  heat  production  over  time  for  a  single  cell  over  Driving  3    

In  the  figures  it  can  be  seen  that  high  C-­‐rates,  either  positive  or  negative,  cause  higher  heat   generation,  while  lower  C-­‐rates  causes  less  heat  production.  Looking  at  the  reversible  and   irreversible  heat,  we  can  see  that  more  irreversible  heat  is  produced  when  the  battery  is   discharging.  At  discharge,  the  reversible  heat  generated  is  negative,  which  means  that  it  is   actually  cooling  the  system.  However,  as  the  irreversible  heat  also  increases  during  the  same   time  period,  the    makes  reversible  heat  loss  is  evened  out  when  looking  at  the  total  heat   generation.  For  the  cycle  Driving  1  cycle  2  (Fig.20,  Fig.  25),  the  heat  increases  rapidly  at  the   timestamp  of  when  the  potential  drops  below  3.0  V.  The  battery  struggles  with  the  situation,   and  more  heat  is  produced.  

 

Temperature  

The  modelled  temperatures  for  the  driving  cycles  are  given  in  Table  3  and  Fig.  29-­‐33.  Driving   1  cycle  1,  Driving  2  cycle  1,  Driving  2  cycle  2,  and  Driving  3  are  also  given  as  heat  colored  3D   figures  in  Fig.  34-­‐37.  The  3D  figures  are  given  at  the  point  of  time  where  the  highest  

temperature  is  measured  and  placed  on  the  same  temperature  scale.  Driving  1  cycle  2  is   excluded  to  better  illustrate  the  temperature  distribution  in  the  other  3D  models.  The   obtained  temperatures  are  strongly  related  to  the  heat  productions  for  the  driving  cycles.  

The  more  demanding  drive  cycle,  the  more  heat  is  being  produced.  The  additional  heat   results  in  higher  temperatures.  This  is  extra  notable  for  Driving  1  Cycle  2,  where  heat  

(33)

production  is  allowed  to  run  higher  for  the  sake  of  the  study.  All  cycles  have  been  put  on  the   same  y-­‐axes  for  simplified  comparison.  The  Driving  1  cycle  2  cycle  will  be  difficult  to  read  due   to  the  wide  range  of  the  axis,  and  a  scale-­‐adjusted  versions  are  available  in  Appendix  1.  

 

Test   Max  temp.   Min  temp.     Coolant  Outflow  Temp.    

Driving  cycles   Driving  1  cycle  1   22.7  °C   20.8  °C   20.02  °C  

     

Driving  1  cycle  2   27.7  °C   22.1  °C   20.05  °C   Driving  2  cycle  1   20.1  °C   20.04  °C   20.001  °C   Driving  2  cycle  2     23.6  °C   21.0  °C   20.02  °C   Driving  3   23.8  °C   21.1  °C   20.02  °C   Table  3:  Temperature  data  for  the  different  driving  cycles  

 

  Fig.  29:  Temperature  over  time  for  Driving  1  cycle  1  

 

(34)

  Fig.  30:  Temperature  over  time  for  Driving  1  cycle  2  

  Fig.  31:  Temperature  over  time  for  Driving  2  cycle  1  

(35)

  Fig.  32:  Temperature  over  time  for  Driving  2  cycle  2  

  Fig.  33:  Temperature  over  time  for  Driving  3  

 

(36)

  Fig.  34:  Temperature  distribution  for  Driving  1  cycle  1  

  Fig.  35:  Temperature  distribution  for  Driving  2  cycle  1  

(37)

  Fig.  36:  Temperature  distribution  for  Driving  2  cycle  2  

 

  Fig.  37:  Temperature  distribution  for  Driving  3  

 

References

Related documents

Thus, when considering the control, the system have a single input (in actual usage this is the current I sys but could also be the voltage V sys ) and multiple outputs in the form

• How much weight and amounts of lithium-ion batteries for vehicles are put on the market in the Nordic countries and how many are available for recycling or second life

The thermal model combined with the lifetime models are used to investigate both static and dynamic charging with regards to the temperature of the ASSE and the expected lifetime

5.2.1 Difference deformation density for LiV6O13 Since the LiV6O13 structure is described in terms of two independent atoms for each original V6O13 atom, it is possible to compare

From the literature study in Chapter 2, it can be argued that it is not realis- tic to increase the energy density of iron based Li-ion insertion materials by more than 5-10% based

(a) Original graph (b) Ordered during the scheduling process for the time point that comes just after ܵ ଴ǡ଴ has been scheduled.. The number of nodes taken into account at a time

Three crucial thermal issues related to temperature variations, temperature gradients, and temperature cycling are addressed in this thesis.. Existing test scheduling techniques

Key words: battery management system, electric vehicle, Kalman Filter, Li-ion battery cell model, state