The!Tropical!Subseasonal!Variability! Simulatedinthe!NASA...

63
1 The Tropical Subseasonal Variability Simulated in the NASA GISS General Circulation Model Daehyun Kim 1 , Adam H. Sobel 1,2,3 , Anthony D. Del Genio 4 , Yonghua Chen 2,4 , Suzana J. Camargo 1 , MaoSung Yao 4,5 , Maxwell Kelley 6 , and Larissa Nazarenko 6 1 LamontDoherty Earth Observatory of Columbia University, Palisades, NY 2 Department of Applied Physics and Applied Mathematics, Columbia University, New York, NY 3 Department of Earth and Environmental Sciences, Columbia University, New York, NY 4 NASA Goddard Institute for Space Studies, New York, NY 5 Sigma Space Partners, Institute for Space Studies, New York, NY 6 Center for Climate System Research, Columbia University, New York, NY Journal of Climate Submitted August 5, 2011 *Corresponding author: Daehyun Kim, LamontDoherty Earth Observatory, Palisades, New York; email: [email protected]

Transcript of The!Tropical!Subseasonal!Variability! Simulatedinthe!NASA...

  • 1

       

       

    The  Tropical  Subseasonal  Variability  Simulated  in  the  NASA  GISS  General  Circulation  Model  

         

    Daehyun  Kim1,  Adam  H.  Sobel1,2,3,  Anthony  D.  Del  Genio4,  Yonghua  Chen2,4,  Suzana  J.  Camargo1,  Mao-‐‑Sung  Yao4,5,  Maxwell  Kelley6,  and  

    Larissa  Nazarenko6      

    1Lamont-‐‑Doherty  Earth  Observatory  of  Columbia  University,  Palisades,  NY    2Department  of  Applied  Physics  and  Applied  Mathematics,  Columbia  University,    New  York,  

    NY    3Department  of  Earth  and  Environmental  Sciences,  Columbia  University,  New  York,  NY    

    4NASA  Goddard  Institute  for  Space  Studies,  New  York,  NY    5Sigma  Space  Partners,  Institute  for  Space  Studies,  New  York,  NY  

    6Center  for  Climate  System  Research,  Columbia  University,  New  York,  NY        

    Journal  of  Climate    

    Submitted  August  5,  2011        

       *Corresponding  author:  Daehyun  Kim,  Lamont-‐‑Doherty  Earth  Observatory,  Palisades,  New  York;  e-‐‑mail:  [email protected]      

  • 2

    Abstract    

    The  tropical  subseasonal  variability  simulated  by  the  Goddard  Institute  for  Space  

    Studies   (GISS)   general   circulation   model   (GCM)   –   Model   E2   is   examined.   Several  

    versions   of   the   Model   E2   were   developed   with   changes   to   the   convective  

    parameterization   in   order   to   improve   the   simulation   of   the  Madden-‐‑Julian   oscillation  

    (MJO).  Consistent  with  previous  studies,  the  simulation  of  the  MJO  is  found  to  depend  

    strongly   on   cumulus   parameterization.   When   the   convective   scheme   is   modified   to  

    have   a   greater   fractional   entrainment   rate,   Model   E2   is   able   to   simulate   MJO-‐‑like  

    disturbances   with   proper   spatial   and   temporal   scales.     The   improvement   in   MJO  

    simulation   comes   at   the   cost   of   increased   biases   in   the   mean   state,   consistent   in  

    structure   and   amplitude   with   those   found   in   other   GCMs   when   tuned   to   have   a  

    stronger  MJO.  Increasing  the  rate  of  rain  re-‐‑evaporation  has  additional  positive  impacts  

    on  the  simulated  MJO.  MJO  life-‐‑cycle  composites  of  various  model  fields  in  a  selected  

    version   of   Model   E2   demonstrate   realistic   features   of   the   simulated   MJO.   By   re-‐‑

    initializing  a  relatively  poor-‐‑MJO  version  with  restart  files  from  a  relatively  better-‐‑MJO  

    version,   a   series   of   30-‐‑day   integrations   is   constructed   to   examine   the   impacts   of   the  

    parameterization   changes   on   the   organization   of   tropical   convection.   The   poor-‐‑MJO  

    version  with   smaller   entrainment   rate   has   a   tendency   to   reduce   the   contrast   between  

    dry   and  wet   regimes,   so   that   tropical   convection  becomes   less   organized.  Besides   the  

    MJO,   the   number   of   tropical   cyclone-‐‑like   vortices   simulated   by   the   model   is   also  

  • 3

    affected  by  the  changes  in  convection  scheme.  With  a  larger  entrainment  rate,  the  model  

    simulates  a  smaller  number  of  such  storms  globally.        

  • 4

    1.  Introduction    

    In  the  tropics,  there  are  a  number  of  distinct  modes  of  variability  with  time  scales  

    shorter   than   a   season.   On   synoptic   time   scales   (say,   2-‐‑10   days)   these   include  

    convectively  coupled  waves  (e.g.,  Wheeler  and  Kiladis  1999)  and  tropical  cyclones.    On  

    a  somewhat  longer  time  scale,  the  Madden-‐‑Julian  oscillation  (MJO,  Madden  and  Julian  

    1971;   Madden   and   Julian   1972)   is   the   dominant   mode   of   tropical   intraseasonal  

    variability   (ISV),   characterized   by   its   planetary   spatial   scale,   30-‐‑60   day   period,   and  

    eastward  propagation.  As  it  modulates  deep  convection  over  the  tropics,   the  MJO  has  

    large   impacts   on   a   wide   variety   of   climate   phenomena   across   different   spatial   and  

    temporal   scales.   Some   examples   include   the   onsets   and   breaks   of   the   Indian   and  

    Australian   summer   monsoons   (e.g.   Yasunari   1979;   Wheeler   and   McBride   2005),   the  

    formation  of  tropical  cyclones  (e.g.  Liebmann  et  al.  1994;  Maloney  and  Hartmann  2000b;  

    Maloney  and  Hartmann  2000a;  Bessafi  and  Wheeler  2006;  Camargo  et  al.  2009)  and  the  

    onset   of   some  El  Niño   events   (e.g.   Takayabu   et   al.   1999;   Bergman   et   al.   2001;  Kessler  

    2001).    

    Previous  studies  have  demonstrated  that  simulation  of  the  MJO  can  be  improved  

    by  changing  specific  aspects  of  the  cumulus  parameterization  of  the  GCM.  The  changes  

    which  have  been  made  to  this  end  differ  in  detail,  but  have  in  common  that  they  tend  to  

    inhibit   deep   cumulus   convection   (Tokioka   et   al.   1988;   Wang   and   Schlesinger   1999;  

    Maloney   and   Hartmann   2001;   Lee   et   al.   2003;   Zhang   and  Mu   2005;   Lin   et   al.   2008).  

  • 5

    Simulation  of  the  MJO,  however,  has  been  a  difficult  test  for  most  climate  models,  from  

    the  Atmospheric  Model  Intercomparison  Project  (AMIP,  Slingo  et  al.  1996)  to  the  more  

    recent  Coupled  Model   Intercomparison  Project-‐‑3   (CMIP3,  Lin   et   al.   2006)  used   in   the  

    Fourth   Assessment   Report   (AR4)   of   the   United   Nations   Intergovernmental   Panel   on  

    Climate  Change  (IPCC).  Lin  et  al.  (2006)  showed  that  only  2  among  14  models  in  AR4  

    had  MJO  variance  comparable  to  observations,  with  even  those  lacking  realism  in  many  

    other  MJO  features.  Kim  et  al.  (2011)  attributed  this  common  symptom  to  a  systematic  

    relationship  between  MJO  strength  and  mean  bias.    They  showed  that  a  specific  set  of  

    mean   state   biases   worsen   as   a   result   of   the   same   parameterization   changes   which  

    strengthen   the  MJO  mode   in   a   number   of  different  models.     This   suggests   that   those  

    parameterization  changes  may  have  been  rejected  because  of  higher  priorities  placed  on  

    mean  state  simulation  compared  to  MJO  simulation.  

    Lin  et  al.  (2006)  showed  that  the  National  Aeronautics  and  Space  Administration  

    (NASA)   Goddard   Institute   for   Space   Studies   (GISS)   GCM   (GISS-‐‑ER   in   AR4)   had   no  

    MJO.    Here,  we  use  the  AR5  version  of   the  GISS  GCM  –  Model  E2,  which   is  different  

    from   AR4   version   in   many   aspects   including   its   parameterization   of   cumulus  

    convection   and   clouds   (see   Section   2a).  As  will   be   shown   later   in   Section   3,   the  AR5  

    version   is   not   able   to   simulate   the  MJO   either.   The   first   question  we   address   in   this  

    study  is  whether  parameterization  changes  similar  to  those  used  to  improve  simulation  

    of  the  MJO  in  other  models  are  also  effective  in  the  GISS  model.      

  • 6

    The   second   question   we   address   is   how   changes   in   the   cumulus   convection  

    scheme   which   improve   the   simulation   of   the   MJO   do   so.     To   this   end,   we   present  

    several   diagnostics   aimed   at   providing   insight   into   physical   processes.    We   also   use  

    short-‐‑term  integrations  in  which  instantaneous  model  states  from  a  simulation  with  one  

    model  version  are  used  as  initial  conditions  for  simulations  with  another  model  version.    

    This   approach   is   similar   to   that   of   Boyle   et   al.   (2008),   who   performed   short-‐‑term  

    integrations   of  GCMs  during   a   strong  MJO  period.     They   initialized   the  model   every  

    day  using  reanalysis  data,  and  aimed  to  identify  the  physical  processes  responsible  for  

    the  deviation  of  model  fields  from  the  observed  state  by  studying  the  growth  of  those  

    deviations  over  short  time  periods.  We  adapt  the  methodology  of  Boyle  et  al.  (2008)  to  

    compare   different   versions   of   the   same   GCM,   rather   than   to   compare   one   GCM   to  

    observations.     By   initializing   a   version   of   the  GCM  which   simulates   the  MJO   poorly  

    with  initial  conditions  from  one  that  simulates  it  better,  our  purpose  is  to  identify  key  

    differences  between  those  two  versions  of  the  same  GCM.    

    Tuning   a   convection   scheme   to   improve   the   intraseasonal   variability   -‐‑   MJO  

    affects   longer  as  well  as   shorter   time  scales  of   climate  as  well.    We  address  here  how  

    parameterization  changes  which  change   the  MJO  simulation  simultaneously   influence  

    both  the  mean  state  and  the  number  of  tropical  cyclone-‐‑like  vortices  which  form  in  the  

    simulation.  The  systematic  variation  of  mean  state  with  MJO  simulation  in  other  GCMs  

    was  reported   in  Kim  et  al.   (2011a).  Several  studies  have   investigated  sensitivity  of   the  

  • 7

    TC   activity   in   low-‐‑resolution  GCMs   to   physical   parameterizations   (Vitart   et   al.   2001;  

    Yoshimura  et  al.  2006;  LaRow  et  al.  2008),  but  to  our  knowledge  a  direct  comparison  of  

    MJO  and  TC  simulation  changes  has  not  been  presented.    

    In   this   study,  we   investigate  one  particular  model   –  Model  E2   -‐‑   to   address   the  

    above  issues.  For  this  purpose,  we  use  several  versions  of  Model  E2,  differing  in  details  

    of  convective  parameterization.  Model  E2  and  the  observational  data  used  are  described  

    briefly  in  Section  2,  followed  by  the  description  of  sensitivity  experiments  performed  in  

    Section  3.  One  version  of  the  model   is  selected  and  analyzed  in  detail   in  Section  4.  To  

    understand   the   difference   between   versions   with   relatively   better   and   worse  

    representations  of  the  MJO,  a  series  of  30-‐‑day  integrations  is  performed  by  initializing  

    the  worse-‐‑MJO  version  using  restart  files  obtained  from  the  better-‐‑MJO  version  (Section  

    5).   The   changes   in   tropical   cyclone   characteristics   caused   by   the   modification   in  

    convection  scheme  to  improve  the  MJO  simulation  are  analyzed  in  Section  6.  Summary  

    and  conclusions  are  given  in  Section  7.  

       2.  Model  and  Data    a.  Model  E2  GCM  

    The  atmospheric  component  of  the  GISS  Model  E  GCM  (Schmidt  et  al.  2006)  was  

    used  in  the  IPCC  AR4.    The  operational  version  used  for  AR4  simulations  was  run  at  a  

    resolution  of  4°x5°x20L.  

  • 8

    For  the  Fifth  Assessment  Report  (AR5),  simulations  are  being  conducted  with  the  

    newer  Model   E2   (Schmidt   et   al.,   in   preparation)   at   a   resolution   of   2°x2.5°x40L.   This  

    serves   as   the   control   version   of   the   model   in   this   paper,   to   which   other   modified  

    versions   will   be   compared.   Model   E2   contains   numerous   changes   to   the   cumulus  

    parameterization  relative   to  Model  E.    Among   the  most   important  are:   (1)  Convective  

    entrainment   and  updraft   speed   are  diagnosed  using   the  parameterization   of  Gregory  

    (2001),   as   described   in   Del   Genio   et   al.   (2007).   (2)   Convective   condensate   in   small  

    particles  whose  fall  speeds  are  significantly   less   than  the  updraft  speed   is   transported  

    upward  rather  than  immediately  detraining  as  in  Del  Genio  et  al.  (2005);  the  portion  of  

    frozen  condensate   in   the   form  of  graupel   extends  up   to  a  minimum  temperature   that  

    depends  on  updraft  speed.    (3)  Downdrafts  originate  from  multiple  levels  above  cloud  

    base   and   detrain   at   all   lower   levels   including   below   cloud   base;   downdrafts  

    entrain/detrain  momentum   as  well   as   heat   and  moisture;   downdraft  mass   fluxes   are  

    used   to  calculate  a  gustiness  correction   to   surface   fluxes.     (4)  The  convective  pressure  

    gradient   is   assumed   to   reduce   convective   momentum   transport   as   in   Gregory   et   al.  

    (1997).     (5)   The   adjustment   time   for   convection   to   adjust   the   cloud   base   to   neutral  

    buoyancy   is   set   to   1   hour,   twice   the   physics   timestep.   Readers   are   referred   to   the  

    appendix  for  a  brief  description  of  the  cumulus  parameterization  in  Model  E2.  

    Model  E2  also  includes  changes  to  the  stratiform  cloud  parameterization,  among  

    them:  (1)  The  threshold  relative  humidity  for  cloud  formation  is  a  function  of  the  large-‐‑

  • 9

    scale  vertical  velocity  above  the  boundary  layer,  with  a  scale-‐‑aware  correction  for  layer  

    thickness;   within   the   boundary   layer   the   threshold   relative   humidity   is   based   on   an  

    assumed   Gaussian   distribution   of   saturation   deficit   as   in   Siebesma   et   al.   (2003);  

    stratiform   clouds   do   not   form   in   subsaturated   air   below   cloud   top   in   the   convective  

    portion  of  the  gridbox  or  below  the  cloud  base  of  a  boundary  layer  convective  cloud.    (2)  

    The   phase   in   which   cloud   forms   is   maintained   until   the   cloud   dissipates   unless  

    supercooled   liquid   is  glaciated  by   the  Bergeron-‐‑Findeisen  process;   convective  snow   is  

    no   longer   permitted   to   glaciate   a   supercooled   stratiform   cloud.     (3)   The   critical  

    supersaturation  for  homogeneous  nucleation  of   ice   is  based  on  Kärcher  and  Lohmann  

    (2002).    (4)  The  optical  thickness  of  precipitation  is  accounted  for  by  the  radiation.    (5)  In  

    unfavorable   conditions,   stratiform   cloud   erosion   by   evaporation   up   to   the   threshold  

    relative  humidity  is  allowed.    (6)  Various  changes  to  the  values  of  parameters  that  affect  

    autoconversion,   maximum   cloud   particle   size,   and   the   temperature   dependence   of  

    liquid  vs.  ice  formation  have  been  made.  

     

    b.  Sensitivity  experiments  -‐‑-‐‑  modifications  to  AR5  version  

    i)  Removal  of  a  mass  flux  limiter  in  the  convection  scheme  

    As  will   be   shown   in   Figure   3,   the  AR5   version   of  Model   E2   (AR5a)   lacks   any  

    representation  of  the  MJO.  This  is  likely  related  to  the  fact  that  deep  convection  occurs  

    in   the   model   too   frequently,   and   is   not   sufficiently   inhibited   when   tropospheric  

  • 10

    moisture   is   low.   To   make   this   point,   Figure   1   shows   composited   daily   precipitation  

    based  on  daily  precipitable  water.  The  inset  in  Figure  1a,  where  we  zoom  in  and  show  

    only   the   bins   with   precipitable   water   between   20   and   45   mm,   highlights   the   earlier  

    onset   of   deep   convection   in   AR5a   compared   to   the  model   with   greater   entrainment,  

    AR5a_Ent1.   Figure   1b,   which   shows   the   probability   density   function   of   daily  

    precipitable  water,   shows   that   these   events   are   not   rare   cases.    Only   grid   points   that  

    have   precipitable   water   between   35   and   45   mm   are   used   in   the   composite   of   the  

    convective   heating   based   on   precipitation   in   Figure   2a,   to   see   the   mean   depth   of  

    cumulus  clouds  simulated  in  this  range  of  precipitable  water.  The  figure  exhibits  a  deep  

    structure  when   log10(precipitation)   is   greater   than   0.5   (precipitation   >   3.2  mm  day-‐‑1).  

    This  means   that   deep   convection   occurs   even  when   precipitable  water   is   between   35  

    and   45  mm,   on   the   order   of   50%   of   the   saturation   value.   In   observations,   cloud   top  

    height  is  strongly  modulated  by  tropospheric  moisture  (Holloway  and  Neelin  2009)  so  

    that  –  unlike  in  the  results  shown  in  figure  2  -‐‑  deep  convection  with  cloud  tops  near  the  

    tropopause   is   inhibited   unless   the   column   is   close   to   saturation.   In   the   AR5_Ent1  

    version,  which  has  an  increased  entrainment  rate  and  simulates  a  better  MJO  (Figure  3)  

    than   that   in  AR5,   the   frequency   of   these  deep   convection   events   reduces   (Figure   2b).  

    Also,  the  composited  precipitation  is  weaker  than  that  in  AR5a  when  precipitable  water  

    is  lower  than  about  50  mm,  while  it  is  much  stronger  when  precipitable  water  is  higher  

    than  50  mm  (Figure  1b).  

  • 11

    This  model   deficiency   is   partly   caused   by   a   limiter   in   the   convection   scheme.  

    When  the  cloud  mass  (which  grows  vertically  when  environmental  air  is  incorporated  

    into   the   plume   from   the   environment   through   entrainment)   exceeds   the   cloud   base  

    layer  mass  at  any  level  above  cloud  base,  the  limiter  sets  the  entrainment  rate  to  zero.  

    Because   the   masses   of   model   layers   (the   pressure   difference   between   model   levels)  

    increase  with  height  up  to  the  mid-‐‑troposphere,  the  limiter  makes  the  entrainment  rate  

    become  zero  frequently.  With  zero  entrainment,   the  cloud  parcel  can  easily  rise  to  the  

    tropopause   level   without   losing   its   buoyancy.   The   frequent   activation   of   deep  

    convection   in   this   model   would   inhibit   large-‐‑scale   organization   of   convection   in   the  

    tropics,   such   as   the   MJO.   In   all   sensitivity   experiments   described   below,   the   limiter  

    described  above   is  modified   so   that   it   is   activated  only  when  cloud  mass  exceeds   the  

    mass  of  the  layer  through  which  the  cloud  is  passing,  rather  than  the  cloud  base  layer.  

    The  “zero  entrainment  rate  event”  never  happens  with  this  modification.  

    ii)  Additional  modifications  to  the  convection  scheme  

    A   number   of   sensitivity   experiments   have   been   conducted   in   an   attempt   to  

    optimize  the  model’s  simulation  of  the  MJO,  convectively  coupled  waves,  and  the  mean  

    climate,   and   the   number   of   TCs.   The   entrainment   rate   coefficient   (which   controls   the  

    overall   magnitude   of   the   entrainment   rate),   and   rain   re-‐‑evaporation   constants   (the  

    fraction   of   convective   condensate   available   for   rain   re-‐‑evaporation)   have   both   been  

    varied  in  these  sensitivity  experiments.  Specifically,  in  AR5a_Ent1,  a  one-‐‑plume  model  

  • 12

    with  entrainment   rate   coefficient   (!!  in  Appendix,   equation  A6)  0.6   is  used   instead  of  

    the  original   two-‐‑plume  structure  with  entrainment  rate  coefficients   (0.3,  0.6).  With  the  

    modifications  in  the  limiter,  this  change  effectively  increases  the  entrainment  rate  in  the  

    model   thereby   making   the   plumes   in   the   convection   scheme   more   sensitive   to  

    environmental   humidity.   Based  on  AR5_Ent1,  more   rain   re-‐‑evaporation   is   allowed   in  

    AR5a_Ent1_Re,  by  limiting  the  maximum  amount  of  condensate  used  in  the  downdraft  

    to  half  of  the  total  condensate.  In  AR5a,  a  maximum  of  all  condensate  can  be  used  in  the  

    downdraft  as  a  source  of  negative  buoyancy.  Allowing  more  rain  re-‐‑evaporation  makes  

    it   difficult   for   the   convection   scheme   to   produce   precipitation   in   a   dry   column.  

    Entrainment   rates   are   adjusted   to   have   values   between   AR5a   and   AR5a_Ent1   in  

    AR5a_Ent2_Re,  by  reviving  the  two-‐‑plume  model  but  with  slightly  larger  entrainment  

    rate   (0.4,   0.6).   AR5a_Ent2_Re   also   allows   water   vapor   and   condensate   to   affect  

    downdraft  buoyancy,  to  offset  the  suppression  of  downdrafts  caused  by  increased  rain  

    evaporation,   and   has   been   adjusted   to   global   radiative   balance   by   tuning   threshold  

    relative  humidity  value  for  cloud  formation.  

    Table   1   summarizes   the   versions   of   Model   E2   used   in   this   study.   A   series   of  

    AMIP-‐‑type  simulations  is  performed  with  these  models,  prescribing  Hadley  Centre  sea  

    surface  temperature  data  (Rayner  et  al.  2003)  as  a  boundary  condition.  The  simulation  

    period   is   1989-‐‑2008,   and   the   vertical   resolution   of   the   model   is   40   levels   for   all  

  • 13

    simulations  while  the  horizontal  resolution  is  2.5  degree  longitude  x  2.0  degree  latitude,  

    except  for  C_AR5a,  where  1.0  degree  longitude  x  1.0  degree  latitude  resolution  is  used.  

     

    c.  Observational  data  

    We   validate   the   simulations   of   rainfall   against   the   Global   Precipitation  

    Climatology  Project  (GPCP)  dataset  (Huffman  et  al.  2001).  We  use  outgoing  longwave  

    radiation   (OLR)   from   the   Advanced   Very   High   Resolution   Radiometer   (AVHRR)  

    (Liebmann   and   Smith   1996).   The   upper   (200-‐‑hPa)   and   lower   (850-‐‑hPa)   tropospheric  

    zonal  winds  are  taken  from  the  National  Centers  for  Environmental  Prediction  (NCEP)–

    National  Center  for  Atmospheric  Research  (NCAR)  Reanalysis  data  (Kalnay  et  al.  1996).  

    For   the   surface   latent   heat   flux   we   also   use   the   objectively   analyzed   air–sea   fluxes  

    (OAFlux)   from   Yu   and   Weller   (2007).   The   specific   humidity   and   925-‐‑hPa   moisture  

    convergence  based  on  the  40-‐‑yr  European  Centre  for  Medium-‐‑Range  Weather  Forecasts  

    (ECMWF)  ReAnalysis   (ERA-‐‑40)   (Uppala   et   al.   2005)   are   also   included   in   our   analysis  

    since   Tian   et   al.   (2006)   indicated   possible   shortcomings   in   the   MJO-‐‑relevant   specific  

    humidity   fields   from   the   NCEP–NCAR   reanalysis.   The   Special   Sensor  

    Microwave/Imager  (SSM/I)  -‐‑  Tropical  Rainfall  Measuring  Mission  (TRMM)  Microwave  

    Imager   (TMI)   combined   precipitable   water   data   is   also   used.   The   observed   tropical  

    cyclone  data  are  from  the  IBTrACS  dataset  (Knapp  et  al.  2010).  

     3.  Simulations  of  the  MJO  using  Model  E2  

  • 14

     a.  Simulations  of  the  MJO  in  AR4  and  AR5  versions  of  Model  E2  

    Following   Wheeler   and   Kiladis   (1999),   wavenumber-‐‑frequency   diagrams   are  

    constructed  to  determine  the  capability  of  the  models  to  simulate  convectively  coupled  

    equatorial  waves  and  the  MJO.  Figure  3  shows  the  symmetric  wavenumber-‐‑frequency  

    power  spectra  (normalized  by  estimated  background  power,  Wheeler  and  Kiladis  1999)  

    of   equatorial   precipitation   from   observations   and   several   versions   of   Model   E2.   In  

    observations   (Figure   3a),   the   power   in   the  Kelvin   and   equatorial  Rossby  wave   bands  

    (with  equivalent  depth  of  25m)  as  well  as  in  the  MJO  band  (wavenumber  1-‐‑3,  period  30-‐‑

    60   days)   are   distinguished   from   the   background   spectrum.   The   most   significant  

    improvement   that  AR5a  has   compared   to  AR4a   is   its   simulation  of   the  Kelvin  mode.  

    The  Kelvin  mode  in  AR5a  is  similar  to  that  in  observations  in  its  strength  (how  much  it  

    is  distinguished  from  the  background  spectrum)  and  phase  speed  (25m  line).  Compared  

    to   AR5a,   AR4a   has   a   much   weaker   and   faster   Kelvin   mode,   which   is   also   mostly  

    confined   to  high   frequency   (i.e.  period   less   than  7days).  Despite   these   improvements,  

    AR5a  still  lacks  the  MJO  mode.  The  MJO  in  AR4a,  coupled  with  an  oceanic  GCM,  was  

    investigated  in  Lin  et  al.  (2006),  who  showed  the  lack  of  an  MJO  in  that  model.    

    Figure   3   also   contains   symmetric   components   of   the   wavenumber-‐‑frequency  

    spectra  of  the  different  versions  of  Model  E2.  C_AR5a  (Figure  3d)  represents  a  version  

    of  Model  E2   that  uses  higher  horizontal   resolution   (1ox1o)   than   that   in  AR5a  by  using  

    the   cubed   sphere   dynamical   core,   which   replaces   the   longitude-‐‑latitude   grid  

  • 15

    configuration   in  AR5a.   In  AR5c   (Figure   3e),  we   coupled  AR5a  with   an  oceanic  GCM.  

    Results   show   that   increasing   horizontal   resolution   in   the   atmospheric   model   and  

    enabling  air-‐‑sea  coupling  make  almost  no  change   in   its  MJO  simulation.  We  conclude  

    that   the   horizontal   resolution   and   air-‐‑sea   coupling   are   not   crucial   factors   for   the  

    representation  of  the  MJO  in  Model  E2.  

     

    b.  Simulations  of  the  MJO  with  changes  in  convection  scheme  

    Figure   3   shows   that   unlike   the   control   model   AR5a,   models   AR5a_Ent1,  

    AR5a_Ent1_Re,  and  AR5a_Ent2_Re  are  able   to   simulate   the  MJO   to   some  extent.  This  

    suggests   that,  as  shown  in  previous  studies,   the  moist  convection  scheme   is  crucial   in  

    the  simulation  of  the  MJO  using  GCMs,  with  entrainment  rate  (Lee  et  al.  2003;  Bechtold  

    et  al.  2008;  Lin  et  al.  2008)  and  rain  re-‐‑evaporation  (Maloney  and  Hartmann  2001)  being  

    particularly   important.   Increasing   entrainment   rate   and   allowing   more   rain   re-‐‑

    evaporation  strengthens  the  MJO  in  the  model.    

    Enhancing   rain   re-‐‑evaporation   also   reduces   the   speed   of   the   simulated   Kelvin  

    wave.   The   convectively   coupled   Kelvin   waves   simulated   in   AR5a_Ent1_Re   and  

    AR5a_Ent2_Re   fit   between  dispersion   curves  with   equivalent   depths   of   12   and   25  m,  

    while   in   AR5a_Ent1   and   observations   the   equivalent   depth   is   about   25~50   m.   The  

    equivalent   depth   decreases   with   increasing   water   vapor   amount   in   the   atmosphere.  

    When   we   increase   the   amount   of   rain   re-‐‑evaporation   in   the   convection   scheme,  

  • 16

    precipitable   water   increases   due   to   increased   condensate   evaporation.   For   example,  

    November-‐‑April   mean   precipitable   water   averaged   over   the   warm   pool   region   (40-‐‑

    180oE,   20oS-‐‑20oN),   increases   from   about   40   mm   in   AR5a_Ent1   to   about   45   mm   in  

    AR5a_Ent1_Re.  Note  that   the  amount  of  precipitable  water  becomes  more  close  to  the  

    observed   value   (about   46  mm,   from   the   Special   Sensor  Microwave/Imager   -‐‑   Tropical  

    Rainfall  Measuring  Mission  Microwave  Imager  combined  precipitable  water  data)  with  

    increased  rain  re-‐‑evaporation.  

    To   select   one   model   version   from   those   used   in   the   sensitivity   experiments  

    described   above   for   more   detailed   analysis,   we   looked   at   other   aspects   of   the  

    simulations   in  addition   to   the  wavenumber-‐‑frequency  power   spectra.   In  Figure  4,   the  

    relationship   between   the   strength   and   the   dominant   propagation   direction   of   ISV   is  

    summarized  in  a  scatter  diagram.  A  similar  plot  was  shown  in  Kim  et  al.  (2011a)  using  

    simulations  with  other  models  (grey  circles  in  Figure  4,  see  Kim  et  al.  2011a  for  further  

    description).  The  ISV  strength  metric  is  obtained  by  averaging  the  standard  deviation  of  

    20-‐‑100   day   filtered   precipitation   over   the   tropics   (30oS-‐‑30oN,   0-‐‑360oE).   For   the  

    propagation   direction   metric,   the   eastward/westward   ratio   of   ISV   is   calculated   from  

    space-‐‑time  power   spectra,   by   dividing   the   sum  of   the   spectral   power   over   eastward-‐‑

    propagating  zonal  wavenumbers  1-‐‑3  and  frequency  range  30-‐‑70  days  by  its  westward-‐‑

    propagating   counterpart.   Figure   4   shows   that   there   is   an   approximate   linear  

    relationship   between   the   two   metrics,   which   is   consistent   with   Kim   et   al.   (2011a)’s  

  • 17

    results.  Overall,  versions  with  ISV  strength  equal  to  or  less  than  the  observed  value  of  

    2.5  mm2/day2  (open  circles)  underestimate  the  eastward/westward  ratio  metric.  Strong-‐‑

    ISV   versions   (ISV   strength   larger   than   2.5   mm2/day2,   closed   circles)   generally   show  

    eastward/westward  ratios  comparable   to   that  of  observations  and  larger   than  those  of  

    the  weak-‐‑ISV  versions.  

    Kim  et  al.  (2011a)  also  showed  that  the  strength  of  the  ISV  varies  systematically  

    with   certain   aspects   of   the   mean   state   in   AGCM   simulations.   To   investigate   the  

    relationship   between   ISV   characteristics   and   the   mean   state   in   Model   E2,   the  

    relationship  between  ISV  and  mean  state  simulation  is  summarized  in  a  scatter  diagram  

    (Figure   5),  with   boreal   summer   on   the   left   and  boreal  winter   on   the   right   (again,   the  

    models  used  in  Kim  et  al.   (2011b)  are   included  as  grey  circles).  The  transverse  axes   in  

    Figure   5   show   the   seasonally-‐‑stratified   ISV   strength   metric,   while   the   ordinate   axes  

    show   a   May-‐‑October   and   November-‐‑April   mean   precipitation   index,   respectively,  

    defined  as  the  mean  precipitation  difference  between  the  west  Pacific  and  Indian  Ocean  

    regions   (see  caption).   It   is   shown   in  Figure  5   that  a  model  with  stronger   ISV   tends   to  

    simulate  a  larger  difference  in  mean  precipitation  between  the  selected  regions  in  both  

    seasons,   with   even   more   precipitation   occurring   over   the   wetter   WP   warm   pool.  

    Strengthening   the  magnitude   of   ISV   and   improving   the   propagation   direction   of   ISV  

    result  in  a  degradation  of  the  mean  state  by  this  metric  (an  increase  of  the  difference  of  

    the  mean  precipitation  index  from  its  observed  value).  

  • 18

    The  above   results   suggest   that   the   improvements   in   the   simulation  of   the  MJO  

    obtained  by  increasing  entrainment  rates,  and  by  allowing  more  rain  re-‐‑evaporation  in  

    the  convection  scheme  are  accompanied  by  a  larger-‐‑than-‐‑observed  ISV  magnitude  and  

    a  systematic  mean  state  bias  consistent  with  that  found  in  other  models.  Keeping  this  in  

    mind,   we   select   one   version   -‐‑   AR5a_Ent1   -‐‑   and   analyze   the   MJO   simulated   in   that  

    version   in   the   following  section.  Version  AR5a_Ent1  has   the  eastward/westward   ratio  

    closest   to   the   observed   value   among   all   our   simulations   (Figure   4)   while   having   a  

    smaller   mean   state   bias   than   other   strong-‐‑ISV   versions   (Figure   5).   Also,   AR5a_Ent1  

    shows  a  Kelvin  wave  speed  (Figure  3),  which  is  closer  to  the  observed  value  than  those  

    in  AR5a_Ent1_Re  and  AR5a_Ent2_Re.  

       4.  Analysis  of  the  MJO  simulated  in  a  selected  version  of  Model  E2    

    In   this   section,   we   investigate   the   characteristics   of   the  MJO   simulated   in   the  

    experiment  AR5a_Ent1.  The  combined  empirical  orthogonal  function  (CEOF)/MJO  life-‐‑

    cycle   composite   approach  of  Wheeler   and  Hendon   (2004)   is   adopted   for   this  purpose  

    (Kim   et   al.   2009;  Waliser   et   al.   2009).  Using   this  method,   the  MJO   is   extracted   as   the  

    dominant  mode  of  intraseasonal  variability  using  OLR  and  zonal  wind  at  850  and  200  

    hPa.  Figure  6  shows  the  first  two  CEOF  modes  of  20–100-‐‑day  15oS–15oN  averaged  850-‐‑

    hPa  and  200-‐‑hPa  zonal  wind  and  OLR  from  observations  and  the  AR5a_Ent1  simulation.  

    The  MJO  mode  extracted  from  AR5a_Ent1  (Figure  6b)  captures  the  gross  features  of  the  

  • 19

    leading   mode   in   observations   (Figure   6a),   such   as   the   location   of   the   maximum   in  

    convection   (minimum   OLR),   baroclinic   wind   structure,   and   planetary   spatial   scale,  

    although   the   fractional   variance   explained   by   the   MJO   mode   (sum   of   variances  

    explained  by  1st  and  2nd  modes)  is  about  29%,  smaller  than  that  observed  (about  44%).  

    To   investigate   the   frequency  of   the  MJO,  we  constructed  unfiltered  PCs  by  projecting  

    the   CEOFs   onto   unfiltered   anomalies,   and   then   calculating   power   spectra   of   the  

    resulting  fields.  The  power  spectrum  of  the  unfiltered  PC  shows  that  the  MJO  extracted  

    by   CEOF   has   spectral   peaks   near   the   observed   MJO   time   scale   that   are   physically  

    meaningful  and  distinct  from  a  red  noise  process.    

    The  MJO   life-‐‑cycle   composites   of   various   variables   are   constructed   using   PCs  

    from  the  CEOF,  to  examine  the  physical  structure  of  the  simulated  MJO.  The  MJO  life-‐‑

    cycle  composite  of  a  variable  for  a  specific  phase  is  constructed  by  averaging  20-‐‑100  day  

    filtered   anomalies   on   the  days  when  PCs   indicate   that   the  MJO   is   in   that  phase  with  

    strong  enough  amplitude   (Wheeler  and  Hendon  2004).  The  phase  of   the  MJO  directly  

    links   to   the   location   of   the   anomalous   convection.   For   example,   Figure   7   shows  

    composites   of   tropospheric   zonal   wind   and   temperature   (10oS-‐‑10oN   averaged)   in   a  

    longitude-‐‑height  diagram  when  the  MJO-‐‑related  convection  is  located  near  the  western  

    Pacific.   Note   that   it   is   phase   7   in   the   model,   while   it   is   phase   6   in   observations.  

    AR5a_Ent1  exhibits  a  deep  baroclinic  structure  in  the  zonal  wind  field  similar  to  that  in  

    observations.  The  structure  of   composited   temperature   is  also  similar   to  observations,  

  • 20

    showing  upper  level  warming  near  the  anomalous  convective  region  and  cooling  below.  

    The  upper-‐‑tropospheric   cool  anomaly  over   the   Indian  Ocean   is   stronger   in   the  model  

    than  observations.  

    Figure   8   shows   MJO   life-‐‑cycle   composites   of   tropospheric   specific   humidity  

    anomalies.  Here,  we  pick  up  three  phases  during  which  MJO-‐‑related  convection  moves  

    from  the  Maritime  continent  to  the  western  Pacific.  We  see  from  Figure  8  how  moisture  

    anomalies  that  accompany  MJO  convection  develop  globally.  In  observations,  there  is  a  

    signature  of  moisture  preconditioning  ahead  (east)  of  current  deep  convection,  which  is  

    well  captured  by  AR5a_Ent1.  The  moisture  anomaly  simulated  in  AR5a_Ent1,  however,  

    shows  a  distinct  minimum  near  850hPa,  which  is  not  observed.    

    We   construct   the  MJO   life-‐‑cycle   composites   of   OLR,   surface   evaporation,   and  

    925hPa  moisture   convergence   for   all   phases   (1-‐‑8)   and  plot   them   in   a   longitude-‐‑phase  

    diagram   in   Figure   9.   The   contours   in   Figure   9   show   the   OLR   anomaly,   which   is  

    propagating   east   during   the   whole   life   cycle   from   the   western   Indian   Ocean   to   the  

    eastern  Pacific.  In  observations,  evaporation  anomalies  slightly  lag  the  MJO  convection,  

    but  by  less  than  90  degrees  of  phase  so  that  the  two  fields  are  correlated  with  each  other.  

    This   suggests   that   positive   latent   heat   flux   anomalies   supply   energy   to   convecting  

    columns.  This   characteristic   is   partly   captured   in  AR5a_Ent1   over   the  western   Indian  

    Ocean  and  the  western  Pacific  with  the  phase  lag  larger  than  observations.  On  the  other  

    hand,  the  convection  anomaly  is  correlated  with  negative  evaporation  anomaly  over  the  

  • 21

    eastern  Indian  Ocean.  This  suggests  that,  in  this  region,  evaporation  damps  convective  

    anomalies.   The   moisture   convergence   within   the   boundary   layer   observed   and  

    simulated   using   AR5a_Ent1   both   slightly   lead   the   convection   anomaly   during   its  

    passage.   The   boundary   layer   moisture   convergence   therefore   contributes   to   the  

    moisture  preconditioning  shown  in  Figure  8,  especially  below  850hPa.    

    Above   results   show   that   the  MJO   simulated   in   AR5a_Ent1   has   structures   that  

    compare   qualitatively   well   to   observations   in   various   respects,   such   as   zonal   wind,  

    temperature,  specific  humidity,  and  925hPa  moisture  convergence.    

       5.  Re-‐‑initialization  experiments  

     To  better  understand  the  improvement  in  MJO  simulation  in  AR5a_Ent1,  a  series  

    of   30-‐‑day   integrations   is   performed  with  AR5a  using  multiple   initial   conditions   from  

    the  better-‐‑MJO  version  (AR5a_Ent1).  Figure  10  shows  a  schematic  of  the  re-‐‑initialization  

    experiment.  A  similar  experimental  design  is  used  in  Boyle  et  al.  (2008),  who  initialized  

    climate  models  using   reanalysis  data   to  diagnose   the   time   evolution  of  model   biases.  

    Here  we  want   to   diagnose   the   difference   between   two   versions   of   one  model,   so  we  

    initialize  one  version  with  data   from  another  version.  We  aim  to  understand  how  the  

    large-‐‑scale  organization  of  tropical  convection  (i.e.  MJO)  is  destroyed  when  we  abruptly  

    change  the  physics.  For  the  experiments,  a  strong-‐‑MJO  event  is  depicted  from  a  20-‐‑year  

    integration   of   AR5a_Ent1   using   PCs   from   the   CEOF.   The   life   cycle   of   an   MJO-‐‑like  

  • 22

    disturbance  simulated  in  AR5a_Ent1  is  shown  in  a  2D  phase  space  in  Figure  11a.  In  this  

    2D  phase  space,  distance   from  the  origin  determines  magnitude  of   the  MJO.  A  strong  

    MJO  event  occurs  during  March-‐‑April   2000.  Hovmöller  diagrams  of   total,   anomalous  

    (deviations   from   the   seasonal   cycle),   and   20-‐‑100   day   filtered   equatorial   (15oS-‐‑15oN)  

    precipitation   show   an   eastward   propagation   of   precipitation   anomalies   with   phase  

    speed   ~5  m   s-‐‑1   during   this   period   (Figure   11).  Daily   restart   files   from   the   four-‐‑month  

    (extended   to   February-‐‑May   2000)   period   of   the   strong-‐‑MJO   event   in   the   AR5a_Ent1  

    version  are  used  to  initialize  30-‐‑day  simulations  of  the  AR5a  version  of  the  Model  E2.    

    During   the   course   of   the   30-‐‑day   integration,   the   AR5a   version   systematically  

    deviates  from  the  AR5a_Ent1  version.  Figure  12a  shows  the  composite  deviations  of  the  

    temperature  from  the  first  day  of  simulation.  It  indicates  that  the  tropical  atmosphere  is  

    stabilized   (warmer   upper/colder   lower   troposphere)   gradually   until   day   30.   The  

    warming  aloft   is   larger   than   the   cooling  below,   so   that   the  mass-‐‑weighted  average  of  

    tropospheric   temperature   increases   (Figure   12b).   The   precipitable   water   (Figure   12c)  

    also   increases,   suggesting   that   the   relative   humidity   is   retained.   Therefore,   the  

    systematic  changes  caused  by  decreasing  entrainment  rate   (from  AR5a_Ent1   to  AR5a)  

    are   characterized   as   enhanced   stability,   and   increasing   temperature   and   precipitable  

    water   in   the   tropics.     These   are   consistent   with   Kim   et   al.   (2011a)   who   showed,   for  

    example,  that  models  with  stronger  MJOs  also  had  a  cold  bias  in  the  upper  troposphere  

    relative  to  those  with  weaker  MJOs.  

  • 23

    With   the   greater   stability,   temperature   and   precipitable   water,   tropical  

    convection  becomes  less  organized.  In  Figure  13,  the  Hovmöller  diagram  of  equatorial  

    precipitation   is   plotted,   representing   the   sequences   during   the   30-‐‑day   integration.  

    Figure   13a   displays   rainfall   simulated   using   AR5_Ent1.   Figures   13b   and   13c   are  

    corresponding  Hovmöller  plots   constructed  using   the  data   after   10   and  20  days   from  

    initialization,  respectively.  We  use  the  AR5a  simulation  to  construct  Figure  13d,  which  

    we  can  regard  as  an  infinite-‐‑day-‐‑after-‐‑initialization  version.  Therefore,  from  left  to  right  

    we  see  how  the  changes  in  convection  scheme  alter  the  distribution  of  convection  in  the  

    tropics.    

    In   AR5a_Ent1,   tropical   convection   organizes   into   planetary   scale   super-‐‑cloud  

    clusters   and   the   organized   convection   moves   eastward   (Figure   13a).   Here   the  

    organization   is   roughly  defined   in   the  Hovmöller  diagram  as   the   formation  of  closely  

    located   strong   rainfall   blobs.   After   10   days   from   initialization,   there   are   still   some  

    organized,  eastward  propagating  features,  but  they  are  much  less  prominent  than  those  

    simulated  using  AR5a_Ent1.  The  precipitating  (>  2mm  day-‐‑1)  area  spreads  to  the  whole  

    tropics   after   20   days,  making   the   pattern   diffuse.   The   signature   of   organization   as   in  

    AR5a_Ent1  is  hardly  seen  after  this  point.  This  illustrates  how  reducing  the  entrainment  

    rate  disorganizes  tropical  convection  and  thereby  degrades  the  MJO.    

    Figure  14  demonstrates  how  reduced  entrainment  destroys  the  organization,  by  

    showing   the   change   in   the   local   moisture-‐‑convection   relationship   during   the   initial  

  • 24

    period   of   the   serial   integrations.   The   shading   in   Figure   14   represents   composited  

    pressure  velocity  and  diabatic  heating  rate  based  on  precipitable  water  over  the  warm  

    pool  region  (ocean  only).  Note  that  we  use  data  from  the  AR5a_Ent1  simulation  for  the  

    period   of   the   strong-‐‑MJO   event   to   construct   Figure   14.   The   composites   show   rising  

    motion   in  moist  columns  and  sinking  motion   in  dry  columns,   representing  a  notional  

    overturning  circulation  between  them.  This  overturning  circulation  is  maintained  by  the  

    differential   diabatic   heating   rate   shown   in   Figure   14b;   diabatic   heating   preferentially  

    occurs   in   the   moist   columns.   It   is   reasonable   to   think   that   the   strength   of   the  

    overturning  circulation  reflects  the  degree  of  organization;  when  the  tropical  convection  

    is  more  organized,  we  expect  a  stronger  overturning  circulation  between  moist  and  dry  

    columns.  

    Next  we  collect  only  the  first  days  from  the  30-‐‑day  integrations  with  AR5a  and  

    subtract  corresponding  AR5a_Ent1  simulations  from  them.  The  difference  calculated  in  

    this  way  reflects  the  impact  of  decreasing  entrainment  rate.  The  deviations  of  pressure  

    velocity   and   diabatic   heating   rate   are   composited   based   on   AR5a_Ent1   precipitable  

    water  (Figure  14,  contours).  Figure  14b  shows  that  anomalous  deep  convection  quickly  

    (within  1  day)  develops  after  changing  the  parameterization,  especially  in  the  regime  of  

    precipitable  water  35-‐‑45  mm.  This  means  that  the  activation  of  deep  convection  in  this  

    region  is  suppressed  in  AR5a_Ent1  due  to  the  larger  entrainment  rate.  At  the  same  time,  

    diabatic  heating  in  the  moist  columns  reduces,  making  the  differential  heating  between  

  • 25

    wet  and  dry  columns  smaller.  The  reduced  differential  heating  results  in  the  weakening  

    of  the  overturning  circulation,  as  shown  in  Figure  14a;  both  rising  and  sinking  motions  

    weaken.  The  weakening  of   the  overturning   circulation   implies   less-‐‑organized   tropical  

    convection   and   a   poor   MJO,   as   seen   in   Figure   13.   This   result   suggests   that   the  

    strengthening  of   the   local  moisture-‐‑convection  relationship   (i.e.  depth  and  strength  of  

    convection  more  strongly  depends  on  column  integrated  moisture  amount)  is  one  of  the  

    reasons  for  the  improvements  in  the  simulation  of  the  MJO  in  Model  E2.  

       6.  Impacts  on  tropical  cyclone  activity      

    To  detect  and  track  TC-‐‑like  structures  in  the  Model  E2  simulations,  we  used  the  

    algorithm  described  in  Camargo  and  Zebiak  (2002).  Using  the  6-‐‑hourly  model  output,  

    we   localize   structures   with   a   local   minimum   sea   level   pressure,   local   maximum  

    vorticity  (850hPa)  and  wind  speed,  and  a  warm  core.  We  use  thresholds  for  these  local  

    maxima  and  minima  based  on  the  model  statistics.  For  example,  over  the  western  North  

    Pacific,   the   specific   values   of   the   thresholds   are   3.8x10-‐‑5   s-‐‑1   for   vorticity,   9.2  m   s-‐‑1   for  

    wind  speed,  and  2.3oC  for  the  local  temperature  anomaly,  respectively.  The  thresholds  

    depend  on  basin,  but  the  difference  between  values  for  different  basins  are  small.  Once  

    all  the  conditions  are  satisfied  the  cyclones  are  tracked  using  the  low-‐‑level  vorticity.  To  

    our  knowledge,  this  is  the  first  time  that  the  TC  activity  of  any  version  of  the  GISS  GCM  

    has   been   examined.   Figure   15   shows   the   first   positions   and   track  density   (number   of  

  • 26

    times  a  TC  passes  by  a  grid  point)  for  the  TCs  in  AR5a  (top),  AR5a_Ent1  (middle),  and  

    observations   (bottom)   for   the   period   1989-‐‑2008.  Overall,   the   regions   of   TC   formation  

    correspond  to  those  in  observations  to  a  degree  roughly  typical  of  other  low-‐‑resolution  

    climate  models   (see   e.g.  Camargo   et   al.   2005).    Closer   examination   reveals   significant  

    deficiencies.    The   total   absence  of  TCs   in   the  North  Atlantic   in   the  model   stands  out.    

    Other   cases   of   low-‐‑resolution   models   with   similarly   low   biases   in   the   Atlantic   have  

    been   noted   (Camargo   et   al.   2005).   There   are   also   very   few  model   TCs   in   the   eastern  

    North   Pacific,   again   a   bias   commonly   present   in  many   low-‐‑resolution  models.   In   the  

    AR5a  model,   there   is   also   apparent   formation   of   TCs   on   the   coast   of   California   and  

    South   Africa,   a   phenomenon   which   has   no   counterpart   in   the   observations.   The  

    formation  of  storms  over  Australia  is  an  interesting  characteristic  of  this  model,  as  land  

    formation  is  not  observed  in  other  regions  of  the  world  (with  the  exception  of  very  near  

    the  coastlines,  which  could  probably  be  improved  by  tuning  the  track  algorithm).  In  the  

    southern  Hemisphere,   the  model   is  more  active   in   the  South  Pacific   than   in   the  South  

    Indian   Ocean,   while   in   observations   the   reverse   occurs.   Finally,   the   southern  

    hemisphere   storm   tracks   tend   to   go   further   toward   the   poles   than   in   observations,  

    which  could  be  due  to  differences  in  the  thresholds  on  tracking  storms  in  observations  

    and  in  the  model.    

    The  number  of  TCs  (NTC)  in  the  model  is  lower  than  that  observed,  as  shown  in  

    Figure  16.  The  left  panels  show  the  mean  NTC  per  month  globally  and  per  hemisphere  

  • 27

    while  the  right  panels  show  NTC  per  year  (globally  and  northern  hemisphere)  and  per  

    season  (southern  hemisphere;  season:  July  to  June).  Due  to  the  lack  of  TCs  in  the  North  

    Atlantic  and  eastern  North  Pacific,  the  low  bias  of  the  model  NTC  is  much  stronger  in  

    the   northern   hemisphere   than   in   the   southern   hemisphere.   Furthermore,   while   the  

    numbers  of  storms  produced  in  the  off  seasons  in  both  hemispheres  are  similar  to  those  

    in  the  observations,  the  model  tends  to  produce  too  few  storms  in  the  peak  season  in  all  

    regions.    

    There  is  a  significant  reduction  in  the  number  of  TCs  in  the  model  when  using  a  

    larger  entrainment  rate  in  the  convection  scheme  (AR5a_Ent1).  We  analyzed  large-‐‑scale  

    variables  (e.g.  vorticity,  vertical  shear)  in  the  AR5a  and  AR5a_Ent1  versions,  and  there  

    are  no  significant  differences  between  the  2  simulations  in  the  large-‐‑scale  environment  

    (not   shown).     This   strikes   us   as   an   interesting   result;   changes   to   the   convective  

    parameterization  which  strengthen  one  form  of  convective  organization  seem  to  inhibit  

    another.     We   do   not   understand   this   result   at   present,   nor   do   we   know   if   it   will  

    generalize  to  other  models.    If  it  does,  understanding  the  physics  behind  it  would  seem  

    to  be  a  worthy  goal.

       7.  Summary  and  Conclusions  

     In  this  study,  we  examined  the  tropical  subseasonal  variability  simulated  in  the  

    several   versions  of   the   atmospheric   component   of   the  NASA  GISS  GCM   -‐‑  Model_E2,  

  • 28

    which  is  being  used  for  the  IPCC  AR5.  The  AR5  version  of  Model_E2  was  superior  to  its  

    ancestor   –   Model_E,   in   terms   of   the   magnitude   of   intraseasonal   variability,   and   the  

    amplitude   and   phase   speed   of   convectively   coupled   Kelvin   waves.   Despite   these  

    improvements,   the   AR5   version   still   lacked   the   MJO   mode,   which   dominates  

    intraseasonal   variability   over   the   tropics   and   interacts   with   various   other   climate  

    components  in  nature.    

    Consistent  with  previous  studies  (Maloney  and  Hartmann  2001;  Lee  et  al.  2003;  

    Lin   et   al.   2008),   simulation   of   the   MJO   was   found   to   depend   on   cumulus  

    parameterization.  By  tuning  the  convective  scheme  to  have  a  greater  entrainment  rate  

    and  more  rain  re-‐‑evaporation,  we  were  able  to  simulate  the  MJO  using  Model_E2  with  a  

    fidelity  typical  of  other  models  similarly  tuned  to  strengthen  their  MJO  simulations,  at  a  

    cost  of  mean  state  biases  also  typical  of  such  models   (Kim  et  al.  2011a).    Based  on  the  

    overall  magnitude  of  tropical  intraseasonal  variability,  mean  state  and  equivalent  depth  

    of   the   tropical   atmosphere,   we   selected   one   version   (AR5a_Ent1)   from   the   set   of  

    sensitivity  experiments  to  investigate  the  MJO  in  that  simulation.  

    We  used  a  CEOF  approach  (Wheeler  and  Hendon  2004)  to  extract  the  MJO  mode,  

    which  was  defined  here  as  the  leading  mode  of  coherent  variability  between  anomalies  

    of   upper   and   lower   tropospheric   zonal  wind,   and   convection,   from  observations   and  

    AR5a_Ent1   simulations.   The   leading  pair   of  CEOFs   in  AR5a_Ent1   captured   the   gross  

    features  of   the   leading  mode   in  observations,  such  as   the   location  of   the  maximum  in  

  • 29

    convection  (minimum  OLR),  baroclinic  wind  structure,  and  planetary  spatial  scale.  The  

    power  spectrum  of  the  unfiltered  PC  showed  that  the  MJO  isolated  by  the  CEOFs  had  

    spectral  peaks  near  the  observed  MJO  time  scale  that  were  physically  meaningful  and  

    distinct  from  a  red  noise  process.    

    The  MJO  life-‐‑cycle  composite  constructed  using  PCs  of  the  leading  pair  of  CEOFs  

    showed  that  the  MJO  simulated  in  AR5a_Ent1  had  structures  that  compare  qualitatively  

    well   to   observations   in   various   respects,   such   as   zonal   wind,   temperature,   specific  

    humidity,   and   925hPa   moisture   convergence.   In   both   observations   and   AR5a_Ent1,  

    positive   moisture   anomalies   in   the   lower   troposphere   develop   ahead   (east)   of   the  

    convective  anomaly  related  to  the  MJO.  Anomalous  925hPa  moisture  convergence  leads  

    the   convective   anomaly,   suggesting   its   role   in  moistening   the   lower   troposphere.  The  

    tropospheric  moisture  anomalies  between  900  and  800hPa  are  not  sufficiently  coherent  

    with  those  at  other  levels  compared  to  observations.  

    By  re-‐‑initializing  a  relatively  poor-‐‑MJO  version  (AR5a)  with  restart   files   from  a  

    relatively  better-‐‑MJO  version  (AR5a_Ent1),  a  series  of  30-‐‑day  integrations  is  performed  

    to  examine  the  impacts  of  the  parameterization  changes  on  the  organization  of  tropical  

    convection.   During   the   course   of   the   30   days,   the   simulated   tropical   atmosphere  

    deviated  from  the  AR5a_Ent1  with  a  greater  stability,  a  higher  temperature,  and  higher  

    precipitable  water.  At  the  same  time,  the  contrast  between  dry  and  wet  regimes  became  

    reduced,   so   that   the   tropical   convection   became   diffuse   and   less   organized.   This  

  • 30

    suggests  that  the  lack  of  an  MJO  in  AR5a  was  partly  due  to  the  inadequate  relationship  

    between  tropospheric  moisture  and  precipitation  (diabatic  heating).  

    TC-‐‑like  structures  in  selected  Model  E2  simulations  (AR5a  and  AR5a_Ent1)  were  

    detected  and  tracked  using  the  algorithm  described  in  Camargo  and  Zebiak  (2002).  The  

    regions   of   TC   formation   correspond   to   those   in   observations,   to   a   degree   roughly  

    typical  of  other   low-‐‑resolution  climate  models,   although   there  were  almost  no  TCs   in  

    the  North  Atlantic.  The  number  of  TCs  in  the  model  was  lower  than  that  observed.    The  

    frequency   of   TCs   (number)   in   the   model   was   affected   by   the   changes   in   convection  

    scheme.  With  the  larger  entrainment  rate,  the  model  simulates  an  even  smaller  number  

    of   TCs   globally.     The   fact   that   a   change   in   the   parameterization   of   entrainment  

    strengthens  one  form  of  convective  organization  (MJO)  while  weakening  another  (TCs)  

    suggests  that  the  parameterizations’  flaws  are  likely  structural,  rather  than  solvable  by  

    parameter  tuning.    On  a  practical  level,  it  illustrates  the  conundrum  faced  by  modelers  

    in   selecting   the   “best”   physics   for   operational   versions   of   GCMs   and   the   need   for  

    further  insights  into  these  processes.  

     Acknowledgements.    This work was supported by NASA grant NNX09AK34G,  by  the  

    NASA  Modeling  and  Analysis  Program  model  development  RTOP  at  GISS,  and  by  the  

    NASA  Precipitation  Science  Program.

     

       

  • 31

    Appendix.  Moist  convection  scheme  in  Model  E2  

     

    1)  Convective  cloud  model  

    The  moist  convection  scheme  in  Model  E2  uses  the  entraining-‐‑detraining  plume  

    model   for   updrafts   and   downdrafts.   For   each   updraft   or   downdraft   plume,   the  

    following  equations  are  used  to  diagnose  the  properties  of  the  plume.  

     

    Mass  flux  (!):  

     

    !!!,!

    !!= ε− δ M!,!,                                                                                              (A1)  

     

    where   superscripts  !  and  !  indicates   the   updraft   and  downdraft,   respectively.  Here,  !  

    and  !  represent  fractional  entrainment  and  detrainment  rate,  respectively.  

     

    Potential  temperature(!),  specific  humidity  (!),  and  horizontal  momentum  (!, !):  

     

    !!!,!∅!,!

    !!= εM!,!∅− δM!,!∅!,! + S∅!,!,                                                        (A2)  

     

  • 32

    where   the  overbar   indicates  a  gridbox-‐‑mean  quantity.  The  source/sink   term  !  for  each  

    variable  ∅  is   summarized   in   Table   A1.   When   condensation   occurs,   the   resulting  

    condensate  is  used  in  cumulus  microphysics  described  in  the  next  section.  

     

    Table  A1.  Source/sink  terms  in  Eq.  (A2)  

     

    ∅   θ   q   u, v  

    Source  

    /sink  

    Latent  heat  of  

    condensation  

    /evaporation,  

    freezing  

    /melting  

    Evaporation  

    /condensation  

    Convective-‐‑scale  

    pressure  gradient  

    force  

     

    Note  that  the  grid-‐‑scale  budget  equation  for  any  generic  prognostic  variable  ∅  is  

    given  by,  

     

    !∅!!= − !

    !!!!

    M!∅! +M!∅! − M! +M! ∅ + S∅! + S∅!.                        (A3)  

     

    For  the  updraft  plume,  the  vertical  velocity  of  the  cloud  parcel  (!)  is  calculated  

    following  Gregory  (2001):  

  • 33

     

    !!! !! !

    !!= agB! − bδ w! ! − ε w! !,                                                                    (A4)  

     

    where  ! = 1 6 , ! = 2/3,    !  is  gravitational  acceleration  and  !  is  parcel  buoyancy.  At  

    each  level,  the  buoyancy  of  the  cloud  parcel  is  diagnosed  using  

     

    B! = !!!!

    T!! − T!! − µμ!,                                                                                    (A5)  

     

    where  !!  is  virtual  temperature  and  !  is  the  adiabatic  cloud  water  mixing  ratio.  

    The  fractional  entrainment  and  detrainment  rate  used  in  (A1),  (A2),  and  (A4)  are  

    determined   differently   in   the   updraft   and   downdraft.   In   updraft   calculations,   if   the  

    buoyancy  of  the  parcel  is  positive,  the  fractional  entrainment  rate  is  determined  by,  

     

    ε = !!!"!!

    !! !,                                                                                                                  (A6)  

     

    where  !!  is  a  constant  introduced  by  Gregory  (2001)  as  a  fraction  of  the  kinetic  energy  

    gained  by  buoyancy  that  is  transferred  to  the  air  mass  entrained  from  the  environment.  

    The  Model   E2   convection   scheme   launches   two   plumes  with   different   values   of  !!  to  

    allow  for  instantaneous  variability  of  convection  depth  within  a  gridbox  (e.g.,  deep  and  

  • 34

    shallow,  deep  and  congestus,  etc.).  Above  the  level  of  neutral  buoyancy,  the  model  no  

    longer  entrains,  but  it  detrains  convective  air  mass  to  the  environment  and  assumes  the  

    fractional  detrainment  rate   to  also  given  by   (A6),  but  multiplied  by  −1.  The   fractional  

    entrainment  rate  for  the  downdraft  is  set  to  2×10!!  (m-‐‑1);  there  is  no  downdraft  vertical  

    velocity  equation.  The  downdraft  does  not  detrain  until  it  becomes  positively  buoyant,  

    with  75%  of  its  mass  detraining  at  each  subsequent.  

    The   updraft   calculation   is   triggered   from   a   conditionally   unstable   layer  where  

    the   virtual  moist   static   energy   of   the   layer   exceeds   the   saturation   virtual  moist   static  

    energy   of   the   level   above   (Yao   and  Del   Genio   1989).   The  mass   flux   at   cloud   base   is  

    determined  as   the  mass  flux  required  to  remove  the   instability  at  cloud  base  during  a  

    convective   adjustment   time   currently   specified   to   be   1   hour.   This   total   mass   flux   is  

    partitioned   into   two   parts   for   the   less-‐‑entraining   (i.e.   smaller  !!  value)   and   more-‐‑

    entraining  plumes,  depending  on  the  grid-‐‑scale  sigma  velocity  at   the   level  above.  The  

    stronger  the  upward  motion  at  the  level  above  cloud  base,  which  represents  lower-‐‑level  

    convergence,  the  larger  the  mass  fraction  of  the  less-‐‑entraining  plume.  For  the  potential  

    temperature,   specific   humidity,   and   horizontal   momentum   of   the   updraft,   gridbox-‐‑

    mean  values  at  the  cloud-‐‑base  level  are  used.  The  updraft  vertical  velocity  is  initialized  

    using   the   turbulent   kinetic   energy   (TKE)   from   the   planetary   boundary   layer   scheme  

    (2 !!TKE  for  less-‐‑entraining  plume  and   !

    !TKE  for  more-‐‑entraining  plume).  

  • 35

    A  downdraft  can  be  triggered,  during  the  course  of  calculations  of  the  rise  of  the  

    updraft  plume,  from  any  level  where  an  equal  mixture  of  cloudy  and  environmental  air  

    has  negative  buoyancy.  A  downdraft  calculation  is  initialized  with  a  fixed  fraction  (1/6)  

    of   the  mass   of   the   updraft   and   an   equal   amount   of  mass   from   the   environment.   For  

    potential   temperature,   specific   humidity,   and   horizontal   momentum,   cloudy   and  

    environmental  quantities  are  averaged  for  the  initial  downdraft  properties.  Evaporation  

    of   convective   condensate  produced   and   carried  by   the  updraft  modifies   the  potential  

    temperature   and   specific   humidity   of   the   downdraft   plume.   In   the   AR5   version,   all  

    condensate  can  evaporate  in  the  downdraft  until  it  saturates.  The  downdraft  is  allowed  

    to   descend   below   cloud   base   if   it   is   negatively   buoyant.   (In   AR5a   the   buoyancy   is  

    simply   based   on   the   temperature   difference   between   the   downdraft   plume   and  

    environment;   in   experiment   AR5a_Ent2_Re   the   buoyancy   was   changed   to   include  

    effects  of  water  vapor  and  convective  condensate).  

     

    2)  Cloud  microphysics  

    At  each  level,  the  fractions  of  condensate  to  be  i)  advected  upward,  ii)  detrained  

    out   of   the   updraft,   and   iii)   precipitated   are   diagnosed   based   on   the   assumed   cloud  

    particle   size   distribution,   terminal   velocity   of   particles,   and   vertical   velocity   of   the  

    updraft   from   Eq.   (A4),   following   Del   Genio   et   al.   (2005).   Note   that   the   updrafts   are  

    currently  the  only  transport  mechanism  for  condensate.  Specifically,  mass  distribution  

  • 36

    (! ! )   of   the   condensate   is   assumed   to   be   given   by   the   Marshall-‐‑Palmer   size  

    distribution,  

     

    ! ! = !"!!!/6 !!!!!",                                                                                (A6)  

     

    where  !  is   the   particle   diameter,  !! = 8×10!!!! ,   ! = !!!!!/! !/! ,   !! = !! ,!!,!!  

    represents  the  densities  of  liquid  (!!),  graupel  (!!)  and  ice  (!!).  

    Based  on  parcel  temperature,  different  formulas  are  used  for  liquid,  graupel,  and  

    ice/snow   to   calculate   the   amount   of   condensate   that   precipitates,   detrains,   and   is  

    advected  upward  

     

    f! = 1−!!!!!"!!!!!"

    ,                                                                                                                (A7)  

     

    where  !!  is  the  fraction  of  frozen  condensate  existing  as  ice/snow  (rather  than  graupel),  

    !!  is   parcel   temperature,  !!  is   temperature   at   the   freezing   level,  !!" = !! − 4!!"#  is   the  

    temperature  where  all  condensate  becomes  ice/snow,  and  !!"#  is  the  vertical  velocity  of  

    the   parcel   at   the   freezing   level.  Above   0o   all   convective   condensate   is   assumed   to   be  

    liquid.  

  • 37

    Particle   size-‐‑fall   speed   relationships   are   fits   to   the   terminal   velocity   ( !!)  

    measurements  from  field  experiments,  adjusted  for  pressure  variations  with  respect  to  

    surface  pressure  !!,  given  by  

     

    v!" D = −0.267+ 5.15×10!D− 1.0225×10!D! + 7.55×10!D!!!!

    !.!,    (A8)  

     

    v!" D = 19.3D!.!"!!!

    !.!.                                                                              (A9)  

     

    v!" D = 11.72D!.!"!!!

    !.!.                                                                          (A10)  

     

    for  liquid,  graupel,  and  ice/snow,  respectively.  

    Equations   above   are   solved   for   the   critical   values   of   diameter  !!±  at   which  

    !!± = !! ± ∆!,  where  ∆! =!!∆!∆!, ∆!  is  layer  thickness,  and  ∆!  is  the  model  physics  time  

    step.  A  particle  whose  Doppler  vertical   velocity  !! − !!  would  not   carry   it   out  of   the  

    layer  in  one  physics  time  is  detrained,  while  particles  with  positive  (negative)  Doppler  

    vertical  velocity  are  advected  upward  (precipitated).  

    The  amount  of  condensate   in  each  category   is  calculated  by   integrating   the  Eq.  

    (A6)   over   the   appropriate   particle   size   range.   For   example,   the  mass   of   precipitating  

    condensate  (the  part  of  the  mass  distribution  with  ! > !!!),  which  is  

  • 38

     

    µμ! =!!!!!

    !D!e!!!dD!!!! = πρ!N!/6 e

    !!!!!λ!! D!!! λ! + 3D!!

    ! λ! + 6D!!λ+ 6 .  (A12)  

     

    Any  precipitating  condensate  that  is  not  evaporated  in  the  downdraft  process  

    can  be  re-‐‑evaporated  in  the  environment  if  possible.  

       

  • 39

    Reference    Bechtold,  P.,  M.  Kohler,  T.  Jung,  F.  Doblas-‐‑Reyes,  M.  Leutbecher,  M.  J.  Rodwell,  F.  Vitart,  

    and  G.  Balsamo,  2008:  Advances  in  Simulating  Atmospheric  Variability  with  the  ECMWF  Model:   From  Synoptic   to  Decadal  Time-‐‑Scales.  Quarterly   Journal   of   the  Royal  Meteorological  Society,  134,  1337-‐‑1351.  

    Bergman,   J.   W.,   H.   H.   Hendon,   and   K.   M.   Weickmann,   2001:   Intraseasonal   Air-‐‑Sea  Interactions  at  the  Onset  of  El  Nino.  Journal  of  Climate,  14,  1702-‐‑1719.  

    Bessafi,   M.   and   M.   C.   Wheeler,   2006:   Modulation   of   South   Indian   Ocean   Tropical  Cyclones  by  the  Madden-‐‑Julian  Oscillation  and  Convectively  Coupled  Equatorial  Waves.  Monthly  Weather  Review,  134,  638-‐‑656.  

    Boyle,   J.,   S.   Klein,   G.   Zhang,   S.   Xie,   and   X.   Wei,   2008:   Climate   Model   Forecast  Experiments  for  TOGA  COARE.  Monthly  Weather  Review,  136,  808-‐‑832.  

    Camargo,  S.  J.  and  S.  E.  Zebiak,  2002:  Improving  the  Detection  and  Tracking  of  Tropical  Cyclones  in  Atmospheric  General  Circulation  Models.  Weather  and  forecasting,  17,  1152-‐‑1162.  

    Camargo,   S.   J.,   A.   G.   Barnston,   and   S.   E.   Zebiak,   2005:   A   Statistical   Assessment   of  Tropical  Cyclone  Activity  in  Atmospheric  General  Circulation  Models.  Tellus  A,  57,  589-‐‑604.  

    Camargo,  S.  J.,  M.  C.  Wheeler,  and  A.  H.  Sobel,  2009:  Diagnosis  of  the  MJO  Modulation  of   Tropical   Cyclogenesis   Using   an   Empirical   Index.   Journal   of   the   Atmospheric  Sciences,  66,  3061-‐‑3074.  

    Del  Genio,  A.  D.,  M.  S.  Yao,  and  J.  Jonas,  2007:  Will  Moist  Convection  Be  Stronger  in  a  Warmer  Climate?  Geophysical  Research  Letters,  34,  L16703.  

    Del  Genio,  A.  D.,  W.  Kovari,  M.  S.  Yao,  and  J.  Jonas,  2005:  Cumulus  Microphysics  and  Climate  Sensitivity.  Journal  of  climate,  18,  2376-‐‑2387.  

    Gregory,  D.,  R.  Kershaw,  and  P.  Inness,  1997:  Parametrization  of  Momentum  Transport  by   Convection.   II:   Tests   in   Single   Column   and   General   Circulation   Models.  Quarterly  Journal  of  the  Royal  Meteorological  Society,  123,  1153-‐‑1183.  

    Gregory,   D.,   2001:   Estimation   of   Entrainment   Rate   in   Simple   Models   of   Convective  Clouds.  Quarterly  Journal  of  the  Royal  Meteorological  Society,  127,  53-‐‑72.  

    Huffman,   G.   J.,   R.   F.   Adler,   M.   M.   Morrissey,   D.   T.   Bolvin,   S.   Curtis,   R.   Joyce,   B.  McGavock,   and   J.   Susskind,   2001:   Global   Precipitation   at   One-‐‑Degree   Daily  Resolution  from  Multisatellite  Observations.  Journal  of  Hydrometeorology,  2,  36-‐‑50.  

    Kalnay,  E.,  M.  Kanamitsu,  R.  Kistler,  W.  Collins,  D.  Deaven,  L.  Gandin,  M.   Iredell,   S.  Saha,   G.  White,   J.  Woollen,   Y.   Zhu,  M.   Chelliah,  W.   Ebisuzaki,  W.  Higgins,   J.  Janowiak,  K.  C.  Mo,  C.  Ropelewski,  J.  Wang,  A.  Leetmaa,  R.  Reynolds,  R.  Jenne,  and  D.  Joseph,  1996:  The  NCEP/NCAR  40-‐‑Year  Reanalysis  Project.  Bulletin  of  the  American  Meteorological  Society,  77,  437-‐‑471.  

  • 40

    Karcher,   B.   and   U.   Lohmann,   2002:   A   Parameterization   of   Cirrus   Cloud   Formation:  Homogeneous   Freezing   of   Supercooled  Aerosols.   Journal   of   geophysical   research,  107,  4010.  

    Kessler,   W.   S.,   2001:   EOF   Representations   of   the   Madden-‐‑Julian   Oscillation   and   Its  Connection  with  ENSO.  Journal  of  Climate,  14,  3055-‐‑3061.  

    Kim,   D.,   K.   Sperber,   W.   Stern,   D.   Waliser,   I.   S.   Kang,   E.   Maloney,   W.   Wang,   K.  Weickmann,   J.   Benedict,   and   M.   Khairoutdinov,   2009:   Application   of   MJO  Simulation  Diagnostics  to  Climate  Models.  Journal  of  climate,  22,  6413-‐‑6436.  

    Kim,   D.,   A.   H.   Sobel,   E.   D.   Maloney,   D.   M.   W.   Frierson,   and   I.-‐‑S.   Kang,   2011a:   A  Systematic  Relationship  between  Intraseasonal  Variability  and  Mean  State  Bias  in  AGCM  Simulations,  Journal  of  Climate,  In  press  

    Kim,  D.,  A.  H.  Sobel,  and  I.-‐‑S.  Kang,  2011b:  A  Mechanism  Denial  Study  on  the  Madden-‐‑Julian  Oscillation,  J.  Adv.  Model.  Earth  Syst,  Revised  

    LaRow,  T.,  Y.  Lim,  D.  Shin,  E.  Chassignet,  and  S.  Cocke,  2008:  Atlantic  Basin  Seasonal  Hurricane  Simulations.  Journal  of  climate,  21,  3191-‐‑3206.  

    Lee,   M.   I.,   I.   S.   Kang,   and   B.   E.   Mapes,   2003:   Impacts   of   Cumulus   Convection  Parameterization   on  Aqua-‐‑Planet  AGCM  Simulations   of   Tropical   Intraseasonal  Variability.  Journal  of  the  Meteorological  Society  of  Japan,  81,  963-‐‑992.  

    Liebmann,  B.  and  C.  A.  Smith,  1996:  Description  of  a  Complete  (Interpolated)  Outgoing  Longwave   Radiation   Dataset.   Bulletin   of   the   American  Meteorological   Society,   77,  1275-‐‑1277.  

    Liebmann,  B.,  H.  H.  Hendon,  and  J.  D.  Glick,  1994:  The  Relationship  between  Tropical  Cyclones   of   the   Western   Pacific   and   Indian   Oceans   and   the   Madden-‐‑Julian  Oscillation.  J.  Meteor.  Soc.  Japan,  72,  401-‐‑411.  

    Lin,   J.   L.,  M.   I.   Lee,  D.  Kim,   I.   S.  Kang,   and  D.  M.  W.   Frierson,   2008:   The   Impacts   of  Convective   Parameterization   and   Moisture   Triggering   on   AGCM-‐‑Simulated  Convectively  Coupled  Equatorial  Waves.  Journal  of  Climate,  21,  883-‐‑909.  

    Lin,   J.  L.,  G.  N.  Kiladis,  B.  E.  Mapes,  K.  M.  Weickmann,  K.  R.  Sperber,  W.  Lin,  M.  C.  Wheeler,  S.  D.  Schubert,  A.  Del  Genio,  L.   J.  Donner,  S.  Emori,   J.  F.  Gueremy,  F.  Hourdin,  P.  J.  Rasch,  E.  Roeckner,  and  J.  F.  Scinocca,  2006:  Tropical  Intraseasonal  Variability  in  14  IPCC  AR4  Climate  Models.  Part  I:  Convective  Signals.  Journal  of  Climate,  19,  2665-‐‑2690.  

    Madden,  R.  A.  and  P.  R.  Julian,  1971:  Detection  of  a  40-‐‑50  Day  Oscillation  in  the  Zonal  Wind  in  the  Tropical  Pacific.  J.  Atmos.  Sci.,  28,  702-‐‑708.  

    ——,  1972:  Description  of  Global-‐‑Scale  Circulation  Cells  in  the  Tropics  with  a  40-‐‑50  Day  Period.  Journal  of  the  Atmospheric  Sciences,  29,  1109-‐‑1123.  

    Maloney,   E.   D.   and   D.   L.   Hartmann,   2000a:   Modulation   of   Eastern   North   Pacific  Hurricanes  by  the  Madden-‐‑  Julian  Oscillation.  Journal  of  Climate,  13,  1451-‐‑1460.  

    ——,  2000b:  Modulation  of  Hurricane  Activity   in   the  Gulf  of  Mexico  by   the  Madden-‐‑Julian  Oscillation.  Science,  287,  2002-‐‑2004.  

  • 41

    ——,  2001:  The  Sensitivity  of  Intraseasonal  Variability  in  the  NCAR  CCM3  to  Changes  in  Convective  Parameterization.  Journal  of  Climate,  14,  2015-‐‑2034.  

    Maloney,   E.   D.   and   A.   H.   Sobel,   2004:   Surface   Fluxes   and   Ocean   Coupling   in   the  Tropical  Intraseasonal  Oscillation.  Journal  of  Climate,  17,  4368-‐‑4386.  

    Rayner,  N.A.,  D.E.  Parker,  E.B.  Horton,  C.K.  Folland,  L.V.  Alexander,  D.P.  Rowell,  E.C.  Kent,   and  A.  Kaplan,   2003:  Global  analyses  of   sea   surface   temperature,   sea   ice,  and   night  marine   air   temperature   since   the   late   nineteenth   century.   J.  Geophys.  Res.  (Atmospheres),  108(14),  2-‐‑1  

    Siebesma,  A.  P.,  C.  S.  Bretherton,  A.  Brown,  A.  Chlond,  J.  Cuxart,  P.  G.  Duynkerke,  H.  Jiang,  M.   Khairoutdinov,   D.   Lewellen,   and   C.   H.  Moeng,   2003:   A   Large   Eddy  Simulation  Intercomparison  Study  of  Shallow  Cumulus  Convection.  Journal  of  the  Atmospheric  Sciences,  60,  1201-‐‑1219.  

    Slingo,   J.  M.,  K.  R.  Sperber,   J.  S.  Boyle,   J.  P.  Ceron,  M.  Dix,  B.  Dugas,  W.  Ebisuzaki,   J.  Fyfe,  D.  Gregory,  J.  F.  Gueremy,  J.  Hack,  A.  Harzallah,  P.  Inness,  A.  Kitoh,  W.  K.  M.   Lau,   B.  McAvaney,   R.  Madden,  A.  Matthews,   T.  N.   Palmer,   C.  K.   Park,  D.  Randall,   and   N.   Renno,   1996:   Intraseasonal   Oscillations   in   15   Atmospheric  General   Circulation   Models:   Results   from   an   AMIP   Diagnostic   Subproject.  Climate  Dynamics,  12,  325-‐‑357.  

    Sobel,  A.  H.,   E.  D.  Maloney,  G.   Bellon,   and  D.  M.   Frierson,   2010:   Surface   Fluxes   and  Tropical  Intraseasonal  Variability:  A  Reassessment.  J.  Adv.  Model.  Earth  Syst,  2.  

    Takayabu,   Y.   N.,   T.   Iguchi,   M.   Kachi,   A.   Shibata,   and   H.   Kanzawa,   1999:   Abrupt  Termination  of  the  1997-‐‑98  El  Nino  in  Response  to  a  Madden-‐‑Julian  Oscillation.  Nature,  402,  279-‐‑282.  

    Tian,  B.,  D.  E.  Waliser,  E.  J.  Fetzer,  B.  H.  Lambrigtsen,  Y.  L.  Yung,