• No results found

HIGH RESOLUTION MODELLING OF EVAPOTRANSPIRATION AND SENSIBLE HEAT FLUX IN THE NETHERLANDS Explorative research for a global thermal convection model for predicting bird behaviour

N/A
N/A
Protected

Academic year: 2021

Share "HIGH RESOLUTION MODELLING OF EVAPOTRANSPIRATION AND SENSIBLE HEAT FLUX IN THE NETHERLANDS Explorative research for a global thermal convection model for predicting bird behaviour"

Copied!
31
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

HIGH  RESOLUTION  MODELLING  OF  EVAPOTRANSPIRATION  AND  

SENSIBLE  HEAT  FLUX  IN  THE  NETHERLANDS  

Explorative   research   for   a   global   thermal   convection   model   for   predicting   bird  

behaviour  

 

ABSTRACT  -­‐  Many  bird  species  have  evolved  traits  to  increase  flight  efficiency  to  help  cover  the   long  migration  routes  taken  in  search  of  resources  required  for  breeding  or  survival.  For  birds   migrating   over   land,   soaring   flight   is   the   principle   adaptation   seen   in   a   number   of   species.   Soaring  flight  is  supported  by  rising  air,  which  is  largely  caused  by  thermal  lift.  For  this  reason,  a   model   that   accurately   describes   thermal   lift   through   calculated   proxies   (SHF,   w*)   at   a   high   spatial  and  temporal  resolution  and  across  large  spatial  domains  contributes  significantly  to  the   study   of   migratory   birds.   As   thermal   lift   is   part   of   the   energy   balance   of   the   surface,   it   is   dependent  on  atmospheric  and  surface  conditions.  Latent  heat  plays  a  major  role  in  the  energy   balance   of   the   Earth’s   surface   and   consequently   in   the   occurrence   of   atmospheric   conditions   suitable  for  soaring  flight.  As  the  latent  heat  flux  is  largely  dependent  on  soil  moisture  content,   this  research  sets  out  to  incorporate  existing  modelled  –and  measured  hydrological  data  in  the   Netherlands  into  a  thermal  lift  model  to  produce  output  of  high  spatial  and  temporal  resolution.   Furthermore,  performance  of  the  model  under  different  hydrological  input  is  evaluated.  This  is   done  in  order  to  determine  the  spatial  and  temporal  scale  at  which  hydrological  data  is  required   to   accurately   model   thermal   lift.   The   results   from   this   research   fit   in   a   larger   effort   to   successfully  model  migrating  bird  behaviour.  

                               

BSc.  Thesis  by:  Bart  Sweerts  

Supervisors:  Prof.  dr.  ir.  Willem  Bouten  &  dr.  Judy  Shamoun-­‐Baranes   Reviewer:  dhr.  Wouter  Vansteelant  MSc.  

Date:  July  7th  2016    

(2)

Table  of  contents    

1.  Introduction  ...  2  

2.  Theoretical  Framework  ...  4  

Energy  balance  and  thermal  convection  ...  4  

Joint  model  ...  5  

3.  Methods  ...  5  

Data  collection  ...  5  

Data  processing  ...  6  

Evapotranspiration  model  description  ...  7  

Parameterization  and  computation  of  soil  moisture  ...  9  

Sensitivity  analysis  ...  11  

Analysis  area  and  period  ...  12  

4.  Results  ...  12  

Model  output  with  data  from  NHI  ...  12  

Model  output  with  data  from  PCR-­‐GLOBWB  ...  14  

Variable  sensitivity  ...  16  

5.  Discussion  ...  17  

Model  output  with  data  from  NHI  ...  17  

Model  output  with  data  from  PCR-­‐GLOBWB  ...  18  

Variable  sensitivity  ...  18  

6.  Conclusion  ...  19  

5.  References  ...  20  

Appendix  ...  23  

Appendix  1  –  Variable  conversion  ...  23  

Appendix  2  –  MATLAB  evapotranspiration  script  ...  24  

Appendix  3  –  MATLAB  net  solar  radiation  script  ...  28    

1.  Introduction  

Most   birds   display   migratory   behaviour   on   varying   temporal   and   spatial   scales   to   exploit   resources  for  breeding  and  survival  (Newton,  2008).  To  deal  with  these  relatively  long  migration   routes,   birds   have   evolved   a   range   of   traits   to   increase   flight   efficiency   (Hedenström,   2002,   Norberg,  2012).    For  birds  migrating  over  land,  soaring  flight  is  the  principle  adaptation  seen  in   a  number  of  species  (Hedenström,  1993).  As  some  of  these  birds  are  not  able  to  sustain  flapping   flight  for  an  extended  period  of  time,  they  are  dependent  on  thermal  lift  that  supports  soaring   flight.  Consequently,  much  of  migrating  birds  behaviour  can  be  linked  to  atmospheric  conditions   (Vansteelant  et  al.,  2014).  For  this  reason,  studies  on  avian  migratory  behaviour  are  intrinsically   linked  to  research  on  atmospheric  dynamics  (Kunz  et  al.,  2008).  

  Extended   work   has   been   done   on   the   influence   of   atmospheric   dynamics   on   animal   migration.  This  includes  a  considerable  amount  of  models  that  simulate  migration  of  a  variety  of   avian  animals  based  on  meteorological  data  (Shamoun-­‐Baranes  et  al.,  2010).  Shamoun-­‐Baranes   et   al.   state   that   these   models   are   either   data-­‐driven   or   concept-­‐driven,   but   always   include   atmospheric  condition  data  as  input.  Furthermore,  they  note  that  data  quantity  and  quality,  the  

(3)

model  framework  and  communication  and  collaboration  with  other  researchers  or  institutes  are   important  factors  to  consider  when  integrating  atmospheric  conditions  into  migration  models.      

Tracking   data   for   migrating   birds   is   collected   on   high   spatial   and   temporal   resolution   and  across  large  spatial  domains  (Bouten  et  al.  2013).  Furthermore,  atmospheric  conditions  and   the   resulting   bird   behaviour   show   large   gradients   over   time   and   space.   For   this   reason,   developing   a   high-­‐resolution   thermal   convection   model   that   covers   large   areas   contributes   significantly  to  bird  behaviour  research  by  linking  bird-­‐tracking  data  to  atmospheric  conditions.  

To   model   thermal   convection   with   the   use   of   existing   data,   the   energy   balance   of   the   Earth’s  surface  has  to  be  considered.  When  solar  radiation  meets  the  Earth’s  surface,  the  energy   is  roughly  divided  into  four  components:  

 

                                1 − 𝑟 𝑅 ↓    = 𝑅𝑛𝑙 ↑   +  𝑆𝐻𝐹 +  𝜆𝐸 + 𝐺      (1)          

With  incoming  solar  radiation  (R↓),  albedo  (r),  net  longwave  radiation  (Rnl↑),  sensible  heat  flux   (SHF),  latent  heat  (λE)  and  soil  heat  exchange  by  conduction  (G)  (Bonan,  G.,  2002).  SHF  has  in   combination   with   boundary   layer   height   (BLH),   a   proxy   for   thermal   depth,   been   used   to   calculate   the   convective   scale   velocity   (w*),   which   acts   as   a   proxy   for   thermal   lift   velocity   (Bohrer  et  al,  2011,  Shamoun-­‐Baranes  et  al,  2016).  

In  moist  conditions  like  the  Netherlands,  latent  heat  is  the  most  important  component  of   the   energy   balance   of   the   surface,   as   solar   radiation   is   primarily   used   to   heat   and   evaporate   moisture   (Bonan,   2002).   Since   latent   heat   flux   is   largely   dependent   on   soil   water   content,   hydrological  models  provide  important  input  for  the  modelling  of  thermal  lift.  The  soil  moisture   content  parameterization  currently  used  in  meteorological  models  often  displays  high  degrees   of  uncertainty.  For  this  reason,  there  is  a  need  for  a  method  that  uses  existing  hydrological  data   from  varying  sources  to  accurately  model  thermal  lift  parameters  used  to  predict  soaring  bird   behaviour.    

  The   aim   of   the   research   is   twofold.   First   of   all,   it   aims   to   effectively   calculate   evapotranspiration  rates  to  determine  SHF  on  a  high  spatial  and  temporal  resolution  and  across   a  large  domain.  For  this,  daily  hydrological  data  will  be  used  from  the  complex  and  high  spatial   resolution  (1km)  Nationaal  Hydrologisch  Instrumentarium  (NHI).  Second,  it  aims  to  determine   the  possibilities  for  a  global  sensible  heat  flux  model  using  the  coarser  5’  (geographical  minutes)   PCRaster  Global  Water  Balance  model  (PCR-­‐GLOBWB).  In  both  cases,  emphasis  is  placed  on  the   impact   of   various   input   variables,   parameters   and   model   mechanics   to   determine   what   is   required  for  an  accurate  high-­‐resolution  sensible  heat  flux  model.  

(4)

2.  Theoretical  Framework  

Energy  balance  and  thermal  convection  

There   exists   a   constant   exchange   of   energy   between   the   Earth’s   surface   and   the   atmosphere   (Figure  1).  As  net  radiation  reaches  the  Earth’s  surface,  it  is  divided  in  the  latent-­‐heat  flux  (lE),   soil  heat  flow  (G)  and  the  sensible  heat  flux  (H  or  SHF).  

                   

Fig  1.  Surface  energy  budget.  Obtained  from  USGS  (2016)    

The  latent  heat  flux  is  the  energy  that  is  used  to  evaporate  water  and  can  be  further  divided  in   the   soil   evaporation,   transpiration   and   interception   evaporation   flux.   The   energy   flow   directly   influencing  thermal  lift  is  SHF,  which  can  be  quantified  by  calculating  and  subtracting  the  other   energy  fluxes  from  the  absorbed  solar  irradiation.  

  Sensible  heat  is  the  energy  that  is  used  to  raise  the  temperature  of  the  air  (Stull,  1988).   As  warm  air  is  lighter  than  cold  air,  the  airs  buoyancy  rises  resulting  in  thermal  lift.  However,   there   are   more   factors   that   influence   soaring   flight   behaviour.   An   important   factor   is   the   boundary  layer  height  (BLH).  The  BLH  is  the  border  of  the  lowest  layer  of  the  atmosphere  that  is   heavily   influenced   by   surface   characteristics.   Energy   fluxes   concerning   the   Earth   are   in   large   confined  to  this  layer.  Above  this  layer  is  the  free  atmosphere,  where  air  moves  with  little  to  no   influence  from  the  surface.  Shamoun-­‐Baranes  et  al.  (2016)  find  that  BLH  provides  a  good  proxy   for   where   soaring   birds   fly   and   BLH   is   a   useful   predictor   of   ground   speed   in   soaring   birds   engaged  in  migration  (Vansteelant  et  al.  2015).  However,  other  studies  propose  to  use  a  proxy   for  thermal  lift  velocity  provided  by  a  scaling  coefficient,  known  as  convective  scale  velocity  (w*)   (Stull,   1988;   Bohrer   et   al,   2011).   An   adapted   calculation   by   Shamoun-­‐Baranes   et   al.   (2016)   includes  several  atmospheric  parameters  (eq.  2).  

(5)

                                                                                             𝑤∗   = !"#$ !  . !"# !!! ! !                                                                                      (2)    

With  gravitational  constant  (g),  boundary  layer  height  (BLH),  temperature  (T),  sensible  heat  flux   (SHF),  density  of  air  (ρ)  and  specific  heat  of  air  (Cp).  In  this  research,  the  main  output  is  SHF.   However,  in  combination  with  atmospheric  variables  different  proxies  like  w*  can  be  calculated   with  relative  ease.  

 

Joint  model  

To   determine   values   for   SHF,   the   other   energy   fluxes   in   equation   1   have   to   be   quantified.   In   order  to  do  this,  a  model  will  be  written  that  contains  all  parts  of  the  energy  balance  equation.   The   computed   energy   fluxes   are   latent   heat   of   evapotranspiration,   latent   heat   of   interception   evaporation  and  soil  heat  flow.  Additionally,  the  parameterization  of  the  modules  is  computed   separately  on  a  100m  spatial  resolution.  Figure  2  provides  a  general  overview  of  the  total  model.     The  model  will  consist  of  a  main  script  that  contains  the  initialization  and  calls  the  other   script  to  calculate  the  energy  fluxes.  The  initialization  consists  of  data  loading  and  processing.   From  there,  the  different  energy  fluxes  are  calculated  in  separate  modules.  These  are  returned  to   the  main  script  to  produce  SHF,  which  can  then  be  used  to  calculate  other  thermal  lift  proxies.  

 

 

 

             Fig  2.  General  overview  of  the  total  model.  The  module  discussed  in  this  research  in  bold  

3.  Methods   Data  collection  

For  the  computation  of  evapotranspiration,  data  from  a  variety  of  sources  was  obtained  (Table   1).   First   off,   the   ERA-­‐interim   reanalysis   provides   the   atmospheric   conditions.   Secondly,   two   hydrological  models,  the  Nationaal  Hydrologisch  Instrumentarium  (NHI)  by  Deltares  (De  Lange   et   al.,   2014)   and   the   PCRaster   Global   Water   Balance   model   (PCR-­‐GLOBWB)   by   the   physical  

(6)

geography   department   of   the   University   of   Utrecht   (Van   Beek   et   al.,   2008),   provide   the   soil   moisture  content.  The  hydrological  models  are  discussed  in  more  detail  in  the  methods.  Thirdly,   land-­‐use   maps   were   acquired   from   the   CORINE   land   cover   programme.   Finally,   radiation   data   was  acquired  from  The  Satellite  Application  Facility  on  Climate  Monitoring  (CM-­‐SAF).  

Table  1.  

List  of  used  data  with  source  and  resolution  

Source   Data  used  

  Spatial  resolution   Temporal  resolution  

ERA-­‐interim  (reanalysis)   by  ECMWF  

-­‐  Temperature  at  2m     -­‐  Dewpoint  temperature   -­‐  Windspeed  at  10m     -­‐  Total  cloud  cover  

0.0125°   (degrees)   6-­‐hourly   National  Hydrologisch   Instrumentarium  (NHI)   by  Deltares  

-­‐  Soil  moisture  content   -­‐  Soil  moisture  deficit   -­‐  Rootzone  depth     -­‐  Soil  properties    

250m   Daily  

PCR-­‐GLOBWB  (model)  by   University  of  Utrecht  

-­‐  Soil  moisture  content   5’  (geographical  

minutes)  

Daily   CORINE  land  cover  

programme   -­‐  Land-­‐use   100m   -­‐  

The  Satellite  Application   Facility  on  Climate   Monitoring  (CM  SAF)  

-­‐  Surface  incoming  shortwave  radiation   0.05°  (degrees)   1-­‐hourly  

 

Data  processing  

Some  of  the  data  required  conversion  or  processing  to  be  used  in  the  model  calculations.    

  Formulas  related  to  evapotranspiration  rates  use  net  solar  radiation  (Rn).  To  determine   net  radiation  at  the  surface  (Rn),  equation  3-­‐5  are  used.  

  𝑅!=   (1 − 𝑎)𝑅!−   𝑅!"      (3)   𝑅!" = 𝜎 𝑇! 0.34 − 0.14   𝑒! 1.35   𝑅! 𝑅!"− 0.35      (4)   𝑅!"= 0.75  𝑅!      (5)    

With  net  radiation  at  the  surface  (Rn),  albedo  (a),  surface  incoming  shortwave  radiation  (Rs),  net   longwave   radiation   (Rnl),   hourly   Stefan-­‐Boltzman   constant   (𝜎  =   2.043   10-­‐10   MJ   m-­‐2   hour-­‐1),   average  hourly  temperature  (T),  actual  saturation  vapour  pressure  (ea)  (equation  9),  clear  sky   solar   radiation   (equation   5)   and   extra-­‐terrestrial   radiation   (Ra).   Hourly   values   for   extraterrestrial   radiation   were   determined   according   to   the   method   provided   by   Allen   et   al.   (2006)  (Annex  1,  eq.  1  -­‐  4).    

(7)

  Formulas  for  the  computation  of  potential  evapotranspiration  (ET0)  require  windspeed   at   2m.   Therefore,   ERA-­‐interim   windspeed   values   at   10m   are   converted   using   the   FAO   conversion  curve  (Allen  et  al.,  1998)  (Annex  1,  figure  1).    

  A   third   issue   arose   in   the   computation   of   evaporation   rates.   The   hydrological   models   used  in  this  research  do  not  separately  compute  the  soil  moisture  content  in  the  top  100mm  of   the  soil.  Therefore,  this  soil  moisture  content  is  estimated  by  determining  if  the  moisture  deficit   exceeds  100mm.    Any  soil  moisture  present  in  the  top  100mm  is  then  used  to  compute  top-­‐layer   soil  moisture  content.  

  To   be   able   to   use   all   variables   in   calculations,   interpolation   in   space   and   time   was   required.   For   this,   the   simplest   technique   of   ‘nearest   neighbour’   interpolation   was   used.   This   technique  was  chosen  because  it  makes  the  least  assumptions  regarding  the  newly  formed  query   points  and  their  assigned  value  (Hartkamp  et  al.,  1999).      

   

Evapotranspiration  model  description  

The  latent  heat  flux  module  calculates  latent  heat  flux  (W  m-­‐2)  by  multiplying  evapotranspiration   with  the  latent  specific  heat  of  water.  To  determine  evapotranspiration  rates  the  module  uses  a   crop   coefficient   –   reference   evapotranspiration   (Kc   -­‐   ET0)   procedure.   A   reference   evapotranspiration   (ET0)   is   computed   for   a   reference   crop   like   grass   or   alfalfa   and   is   then   multiplied  by  an  empirical  crop-­‐specific  coefficient  (Kc)  ranging  from  [0  to  1.4]  to  produce  crop   potential  evapotranspiration  (ETc).  To  calculate  evapotranspiration  rates,  the  American  Society   of  Civil  Engineers  (ASCE)  standardized  reference  evapotranspiration  equation  is  used,  which  is   an  adapted  version  of  the  FAO-­‐56  Penman-­‐Monteith  equation  (Allen  et  al.,  2005;  Walter  et  al.,   2001):     𝐸𝑇!= 0.408𝛥 𝑅!− 𝐺 +  𝛾𝑇 + 273 𝑈𝐶𝑛 ! 𝑒!−   𝑒! 𝛥 +  𝛾 1 + 𝐶𝑑  𝑈!       6        

With   slope   vapour   pressure   (𝛥)  (equation   7),   net   solar   radiation   on   the   surface   (Rn),   soil   heat   flux  density  (G),  psychometric  constant  (𝛾),  hourly  mean  temperature  at  2m  (T),  hourly  mean   wind  speed  at  2m  (U2),  saturation  vapour  pressure  (es)  (equation  8)  and  actual  vapour  pressure   (ea)  (equation  9)  (Allen  et  al.,  1998).    

𝛥 =  4098  (0.6108  𝑒   !".!"  ! !  !  !"!.!) (𝑇 + 237.3)!      (7)       𝑒! = 0.6108  𝑒   !".!"  ! !  !  !"#.!      (8)

(8)

     

𝑒! = 0.6108  𝑒  !!".!"  !!"#  !  !"#.!!"#      (9)

With  temperature  at  2m  (T)  and  dewpoint  temperature  at  2m  (Tdew).  

  Furthermore,  parameters  Cn  and   Cd  can  be  adjusted  to  account  for  different  vegetation   heights,  nighttime  calculations  and  to  allow  for  1-­‐hour  time-­‐step  calculations  (Table  2)  (Walter   et  al.,  2001).    

 

Table  2.    

Values  for  Cn  and  Cd  in  equation  1.  Adapted  from  Walter  et  al.  (2001).    

Calculation  Time  Step   Short  Reference  ET0s   Tall  Reference  ET0t  

  Cn   Cd   Cn   Cd  

Daily   900   0.34   1600   0.38  

Hourly  during  daytime   37   0.24   66   0.25  

Hourly  during  nighttime   37   0.96   66   1.7  

 

However,   Walter   et   al.   (2001)   state   that   short   vegetation   has   an   approximate   height   of   0.12m   and  tall  vegetation  an  approximate  height  of  0.50m.  As  this  method  was  specifically  developed   for   agricultural   needs,   it   does   not   take   into   account   natural   vegetation   that   can   grow   much   higher   than   the   tall   vegetation   class.   Nonetheless,   for   natural   vegetation   tall   reference   parameters  are  used  to  compute  potential  evapotranspiration.  Secondly,  nighttime  parameters   are  used  whenever  net  radiation  equals  or  is  lower  than  0.  

  To  determine  actual  evapotranspiration  (ETact),  ET0  is  multiplied  by  a  crop  coefficient  Kc.   However,  determining  values  for  Kc  requires  several  steps.  First  off,  values  for  Kc  vary  over  time   under   the   influence   of   the   seasons   and   growing   conditions.   In   the   FAO   dual   crop   coefficient   method  Kc  is  divided  in  a  transpiration  component  (Kcb)  and  a  soil  evaporation  component  (Ke).     Values   for   Kcb  follow   a   seasonal   curve   with   peak   values   occurring   during   the   growing   season   (figure   2).   Secondly,   values   of   Kcb   are   reduced   by   multiplication   with   a   water   stress   coefficient  Ks  (eq.  10).    

 

𝐾!=   𝑇𝐴𝑊 −   𝐷!

1 − 𝑝 𝑇𝐴𝑊      (10)    

Where  TAW  is  total  available  water  in  the  rootzone  at  field  capacity,  Dr  is  water  shortage  relative   to  field  capacity  and  p  is  the  fraction  of  TAW  a  crop  can  extract  without  suffering  water  stress.   Values  for  p  were  obtained  from  the  FAO  (Allen  et  al.,  1998).  

  Evaporation  from  the  soil  is  predicted  by  estimating  the  amount  of  energy  that  is  able  to   penetrate  the  canopy  and  reach  the  ground.  This  is  done  by  subtracting  the  current  vegetation  

(9)

coefficient  (Kcb)  from  the  maximum  value  the  vegetation  coefficient  can  have  at  optimal  growing   conditions  (Kc  max)  (eq.  11).  

 

𝐾! =   𝐾!  !"#−   𝐾!"      (11)    

When  the  soil  is  wet,  evaporation  is  only  limited  by  the  empirically  derived  maximum  value  for   total   crop   coefficient   Kc   max.   When   the   soil   moisture   contents   drop   beneath   ideal   levels,   a   reducing  coefficient  Kr  is  introduced  (eq.  12).  For  this  coefficient,  soil  moisture  content  in  the  top   100mm  of  the  soil  is  considered,  from  which  evaporation  occurs.  

     

𝐾! =  𝑇𝐸𝑊 −   𝐷!,!!!

𝑇𝐸𝑊 − 𝑅𝐸𝑊      (12)    

Where  total  evaporable  water  (TEW)  is  the  amount  of  water  in  the  top  100mm  of  soil  at  field   capacity,  readily  evaporated  water  (REW)  is  the  amount  of  water  that  can  be  extracted  through   evaporation   before   water   stress   and   De,j  -­‐1   is   the   cumulative   moisture   depletion   from   the   soil   surface   layer   at   the   previous   day.   Combining   the   crop-­‐coefficients   and   potential   evapotranspiration   rates   eventually   produces   actual   evapotranspiration   ETc   act   in   millimetres   (eq.  13).  

 

𝐸𝑇!  !"# = 𝐾!𝐾!"+   𝐾!𝐾! 𝐸𝑇!=   𝐾!  !"#𝐸𝑇!      (13)        

To   convert   to   an   energy   flux,   the   evapotranspiration   [mm]   is   multiplied   by   the   latent   evaporation  heat  of  water  (l  =  2257000  J  kg-­‐1)  and  divided  by  the  amount  of  seconds  in  1  hour   (3600)   to   produce   the   latent   heat   flux   (λE)   [W   m-­‐2].   Additionally,   the   latent   heat   flux   can   be   subtracted  from  the  net  solar  radiation  to  estimate  sensible  heat  flux  (SHF).  

 

Parameterization  and  computation  of  soil  moisture  

The  NHI  is  used  to  calculate  the  impact  of  precipitation  and  other  water  flows  on  water  levels  in   the   Netherlands.   In   this   fashion   it   is   used   as   a   decision   making   tool   for   water   management   strategies.   Therefore,   for   the   scope   of   this   research   the   output   from   the   NHI   is   considered   adequate.  However,  since  the  NHI  only  spans  the  Netherlands,  a  different  hydrological  input  is   required  for  the  model  to  be  used  on  larger  scales.  For  this,  the  global  PCR-­‐GLOBWB  model  is   used.   To   evaluate   whether   hydrological   input   from   the   PCR-­‐GLOBWB   may   be   used   for   the   purpose   of   high   resolution   modelling   of   evapotranspiration,   the   parameterization,   resolution   and  accuracy  of  the  soil  moisture  computation  has  to  be  determined.    

(10)

  This  section  starts  with  a  description  of  the  parameters  used  in  the  evapotranspiration   model.  After  this,  the  computation  of  the  soil  moisture  content  in  the  NHI  and  PCR-­‐GLOBWB  is   briefly  discussed.  

 

Parameters  

Not   only   the   soil   moisture   content,   but   also   the   parameterization   of   the   NHI   model   is   not   available  on  a  global  scale.  Therefore,  it  is  important  to  determine  the  differences  between  the   parameters  used  in  the  NHI  and  the  parameters  used  in  the  global  PCR-­‐GLOBWB  model.  Table  3   contains   the   model-­‐specific   parameters   used   in   the   evapotranspiration   model.   Furthermore,   it   contains  the  parameters  that  were  empirically  derived  on  a  100  by  100m  scale  for  this  study.  A   more  detailed  description  of  their  computation  can  be  found  in  Kooij  (2016).  

 

Table  3.    

Model-­‐specific  and  100  by  100m  parameters.   Model-­‐

specific   Parameter  

NHI  

Resolution   NHI  computation   PCR-­‐GLOBWB  resolution   PCR-­‐GLOBWB  computation   Rootzone  

depth   250m   Empirically  derived  from  250x250(m)  land-­‐use  map  of   the  Netherlands.    

5’  (geographical  

minutes)   Obtained  from  Global  Crop  Water  Model   (Siebert  and  Döll,  2010)   Total  

available   water  (TAW)  

250m   Derived  from  rootzone  depth   and  250x250(m)  soil  type  map   of  the  Netherlands.  

5’  (geographical  

minutes)   Derived  from  rootzone  depth  and  aggregated  (5’)   soil  type  map.  

Readily   available   water  (RAW)  

250m   Fraction  of  TAW,  depends  on  

soil  type   5’  (geographical  minutes)   Fraction  of  TAW,  depends  on  soil  type   Total  

evaporable   water  (TEW)  

250m   Derived  from  soil  type  and   evaporation  layer  depth   (100mm)  

5’  (geographical  

minutes)   Derived  from  soil  type  and  evaporation  layer  depth   (100mm)  

Readily   evaporable   water  (REW)  

250m   Fraction  of  TEW,  depends  on   soil  type  

5’  (geographical   minutes)  

Fraction  of  TEW,  depends   on  soil  type  

Parameter  (100  by  100m)   Computation   Description  

Crop  depletion  factor  (p)   Assigned  to  CORINE  land-­‐use   according  to  Allen  et  al.   (1998)  

Fraction  of  water  a  crop  can  extract  from  the   rootzone  without  suffering  from  water-­‐stress   Maximum  vegetation  

coefficient     (Kc  max)  

Derived  from  climate   conditions  and  CORINE  land-­‐ use  map  

The  coefficient  for  a  vegetation  type  under   optimal  growing  conditions  

Current  vegetation   coefficient  (Kcb)  

Derived  from  leaf  area  index  

(LAI)  and  Kc  max   The  current  vegetation  coefficient  

 

Soil  moisture  –  NHI  

The   NHI   contains   five   hydrological   models   that   combine   to   solve   the   water   balance   in   the   Netherlands.  Concerning  soil  moisture  content,  the  Soil  Vegetation-­‐Atmosphere  water  Transfers   (SVAT)   model   computes   the   vertical   water   transfer   between   the   groundwater   and   the  

(11)

atmosphere.   On   a   daily   basis   and   a   250m   resolution,   evapotranspiration   and   unsaturated   transfer  of  water  (infiltration,  percolation  and  capillary  rise)  are  computed  to  form  the  essential   processes  that  determine  water  flow  and  content  in  the  unsaturated  soil  layer  (Van  Walsum  and   Supit,  2012;  Van  der  Bolt  et  al.,  2015).  Moisture  content  levels  in  the  unsaturated  zone  are  then   combined   with   rootzone   depth   and   TAW   to   compute   soil   moisture   content   and   soil   moisture   deficit  in  the  rootzone.  

 

Soil  moisture  –  PCR-­‐GLOBWB  

PCR-­‐GLOBWB  computes  moisture  content  levels  for  two  stacked  soil  layers  and  an  underlying   groundwater   layer,   where   the   upper   soil   layer   is   defined   as   the   rootzone   layer.   Exchange   between   the   layers   consists   unsaturated   transfer.   Additionally,   water   is   extracted   from   the   upper   soil   layer   through   evapotranspiration.   In   general,   the   mechanics   of   PCR-­‐GLOBWB   are   similar   to   those   of   the   NHI.   However,   the   resolution   is   considerably   lower   and   the   parameterization  is  more  simplified.  

 

Sensitivity  analysis  

Due   to   the   complex   nature   of   the   model   and   the   large   amount   of   parameters   and   variables   involved,   the   model   has   many   degrees   of   freedom.   Consequently,   one   could   argue   that   such   a   model  can  be  made  to  produce  virtually  any  desired  behaviour  (Hornberger  and  Spear,  1981).   For   this   reason,   a   sensitivity   analysis   (SA)   is   conducted   to   determine   the   variation   in   model   output  that  can  be  attributed  to  the  different  input  variables  (Saltelli,  2002;  Saltelli  et  al.,  2000).   Through   the   SA,   it   is   possible   to   determine   which   variables   have   the   highest   influence   on   the   output   and   thus   require   the   most   attention   concerning   quality   of   input   data   and   parameterization  in  the  model.  Due  to  the  limited  scope  of  this  research,  only  the  input  variables   were  analysed.  

  The  first  step  of  a  SA  is  to  analyse  the  input  variables  on  scale  and  shape.  This  includes   the   range   of   the   input   variation   and   the   form   of   its   probability   density   function   (Saltelli   et   al.,   2008).  Secondly,  the  input  variables  are  visualized  by  plotting  them  against  the  model  output.   Through   inspection,   these   plots   can   yield   general   response   trends   and   to   some   extent   their   strength  (Saltelli  et  al.,  2008).  The  third  step  is  to  statistically  analyse  and  quantify  the  influence   input  variables  exert  on  the  output.  For  this,  regression  curves  are  fitted  to  the  data.  Resulting   from   the   regression,   the   explained   variance   (R2)   and   residual   mean   square   error   (RMSE)   are   analysed.  High  values  for  R2  and  low  values  for  RMSE  point  to  strong  response  of  the  output  to   the  variable  in  question  (Judd  et  al.,  2011).  Unlike  done  in  the  standard  procedures  of  a  SA,  this   research  does  not  use  variable  values  randomly  drawn  from  the  variables’  respective  probability  

(12)

density   functions   but   conducts   a   similar   analysis   on   actual   model   input   data.   For   the   analysis,   soil  moisture  data  from  the  NHI  is  used.  

 

Analysis  area  and  period  

The   research   aims   to   provide   insight   in   the   question   whether   a   high-­‐resolution   model   can   be   extrapolated  on  European  or  Global  scale.  For  this  reason,  any  analysis  of  the  model  output  in   the  Netherlands  should  cover  a  large  variety  of  lower  and  upper  boundary  conditions.  However,   due   to   constraints   in   computing   power   and   time,   a   selection   is   made   concerning   the   analysed   input  and  output  data.  Consequently,  the  analysis  focuses  on  an  area  with  relatively  wet  polders   (Flevoland)  and  dry  sandy  soils  (Veluwe)  (Fig.  3A)).  Furthermore,  it  spans  the  last  week  of  June   2012,   a   period   that   contained   warm   sunlit   days   as   well   as   cooler   cloudy   days   (Fig.     3B)).   Additionally,   in   this   period   the   growing   season   has   sufficiently   started   to   show   the   effects   of   different  vegetation  types.  

 

Fig  3.  A)  Soil  moisture  deficit  (NHI),  B)  Analysis  period  hourly  net  radiation  and  temperature  

4.  Results  

The   model   results   are   discussed   in   three   parts.   In   the   first   part,   the   model   output   using   soil   moisture   data   from   the   NHI   is   discussed   in   terms   of   probable   outcome,   temporal   and   spatial   gradients  and  systematic  errors.  In  the  second  part,  a  similar  procedure  is  followed  with  output   resulting   from   PCR-­‐GLOBWB   soil   moisture   input.   In   the   third   part,   the   sensitivity   analysis   provides  the  influence  of  the  different  input  variables  on  the  model  output.    

 

Model  output  with  data  from  NHI  

To  determine  whether  the  model  generates  probable  output  in  general,  the  energy  balance  was   computed  for  the  last  week  of  June  2012  (Fig.  4).      

(13)

  Fig  4.  Average  hourly  energy  balance  of  the  land  surface  of  the  Netherlands.  

 

Figure  4  shows  net  radiation  peaks  between  150-­‐450,  evapotranspiration  75-­‐200,  transpiration   50-­‐150   and   evaporation   25-­‐100   Wm-­‐2.   Additionally,   subtracting   evapotranspiration   from   net   radiation  produces  the  sensible  heat  flux  peaking  between  80-­‐220  Wm-­‐2.    

  However,  sensible  heat  flux  does  not  just  vary  through  time  but  also  through  space.  To   gain   insights   in   this,   Figure   5   shows   the   average   daytime   sensible   heat   flux   for   June   21st.   Furthermore,   it   contains   the   energy   balance   in   a   grid   cell   with   high   and   low   soil   moisture   content  for  June  21-­‐23rd.    

Fig  5.  A)  Average  daytime  sensible  heat  flux  June  21st,  high  sensible  heat  flux  (red)  and  low  sensible  heat  

flux  (blue).  Energy  balance  June  21-­‐23  in  B)  dry  and  C)  wet  cell.  With  soil  moisture  data  from  NHI.  

 

There  are  clear  differences  between  areas  with  high  sensible  heat  flux  like  the  Veluwe,  Flevoland   and  urban  areas  and  areas  with  low  sensible  heat  flux.  In  terms  of  spatial  resolution,  differences   in  surface  characteristics  are  easily  recognizable  in  the  sensible  heat  flux.  Furthermore,  Figure   3B-­‐C   illustrate   that   large   differences   exist   in   energy   balance   dynamics   between   relatively   wet   and  dry  grid  cells.  Wet  grid  cells  show  highly  dynamic  behaviour,  with  significant  fluctuation  in  

(14)

the   percentage   of   solar   radiation   that   is   expended   as   sensible   heat   flux.   Dry   grid   cells   on   the   other   hand   can   be   considered   relatively   inactive,   only   producing   a   small   amount   of   transpiration.    

 

Model  output  with  data  from  PCR-­‐GLOBWB  

Figure  6  shows  the  energy  balance  in  the  Netherlands  in  the  last  week  of  June  2012.  The  slightly   higher   values   of   net   solar   radiation   are   caused   by   differences   in   clipping   extent.   Furthermore,   evapotranspiration  values  seem  to  account  for  a  larger  portion  of  the  energy  balance  resulting  in   lower   values   for   SHF.   Nonetheless,   the   magnitude   of   the   energy   fluxes   is   similar   to   those   previously  found  in  the  computation  with  soil  moisture  data  from  the  NHI.    

 

Fig   6.  Average  hourly  energy  balance  of  the  land  surface  of  the  Netherlands  with  soil  moisture  content  

data  from  PCR-­‐GLOBWB.  

 

 

Fig   7.   Average   daytime   sensible   heat   flux   June   21st,   computed   using   soil   moisture   data   and  

(15)

 

However,  the  spatial  differences  in  computed  SHF  yielded  less  promising  results.  Figure  7  shows   that  computed  SHF  differs  greatly  from  SHF  computed  using  data  from  the  NHI.  First  off,  the  low   spatial  resolution  of  PCR-­‐GLOBWB  is  clearly  visible  in  the  results,  with  lower  resolution  spatial   gradients   in   SHF   as   a   result.   For   some   areas   the   results   can   still   be   generally   interpreted   concerning  lower  and  higher  values  for  SHF,  but  for  determining  high-­‐resolution  border  values   these  results  are  not  suitable.  Secondly,  Figure  7  shows  that  there  are  large  areas  with  negative   average   values   for   SHF   occurring   throughout   The   Netherlands.   Analysing   soil   moisture   deficit   data   as   computed   by   PCR-­‐GLOBWB   (Figure   8)   shows   that   those   areas   have   a   negative   soil   moisture  deficit.                    

Fig  8.  Soil  moisture  deficit  (PCR-­‐GLOBWB).  Negative  values  

(16)

Variable  sensitivity  

Figure   9   shows   input   variables   plotted   against   model   output   with   linear   and   quadratic   curves   fitted  through  the  data.  Table  4  contains  values  for  explained  variance  (R2)  and  residual  mean   square  error  (RMSE).    

 

 

Fig   9.   Scatterplots   of   input   variables   A)   net   solar  

radiation  B)  temperature  C)  dewpoint  temperature   D)  soil  moisture  deficit  E)  windspeed  against  model   output.  With  linear  (yellow)  and  quadratic  (purple)   regression  lines.                

(17)

Table  4.  

Correlation  between  input  variables  and  computed  evapotranspiration  over  the  Netherlands  (N  =  18000).   R2  and  RMSE  denote  the  coefficient  of  determination  and  residual  mean  square  error  respectively.    

Variable   Polynomial     R2   RMSE  

Net  radiation  (Rn)   Linear   Quadratic   0.7349   0.7558   0.0426   0.0409   Temperature  (T)   Linear   Quadratic   0.5814  0.5982   0.0534  0.0523   Dewpoint  

temperature  (D)   Linear  Quadratic   0.1033  0.1652   0.0961  0.0927  

Soil  moisture  deficit  

(Sd)   Linear  Quadratic   0.7274  0.7891   0.0031  0.0027  

Windspeed  (U2)   Linear  

Quadratic   0.2057  0.2057   0.0694  0.0694  

 

From  figure  7  it  is  obvious  that  radiation  (A)  and  temperature  (B)  have  a  strong  positive  effect   on  the  actual  evapotranspiration.  Furthermore,  the  soil  moisture  deficit  (D)  has  a  clear  negative   effect   on   the   actual   evapotranspiration.   Windspeed   (E)   has   a   mild   positive   effect   on   evapotranspiration.  The  effect  of  dewpoint  temperature  (C)  is  the  least  pronounced,  especially   when  computed  through  a  quadratic  regression.    

  The   regression   results   of   net   radiation   and   soil   moisture   are   similar   in   terms   of   explained  variance,  R2.  However,  the  smaller  variation  in  output  response  to  the  soil  moisture   deficit   causes   a   significantly   lower   RMSE.   Therefore,   soil   moisture   deficit   has   the   strongest   relation  to  actual  evapotranspiration.  Temperature  also  has  fairly  high  R2  and  low  RMSE  values   and  observation  of  the  scatterplots  indicates  a  strong  correlation  with  net  radiation.  Windspeed   yields   a   fairly   low   R2   and   average   RMSE.   R2   and   RMSE   are   left   unchanged   when   computing   a   quadratic  relation,  meaning  the  relation  between  windspeed  and  evapotranspiration  is  strongly   linear.   Dewpoint   temperature   has   a   significantly   lower   R2   and   higher   RMSE   compared   to   the   other   variables.   The   negative   relation   is   expected   as   higher   dewpoint   temperatures   coincide   with  higher  relative  humidity,  which  negatively  impacts  evapotranspiration  (Wallace  &  Hobbs,   2006).    

5.  Discussion  

Model  output  with  data  from  NHI  

The  energy-­‐balance  values  seen  in  Figure  4  are  similar  to  values  produced  by  studies  measuring   and  modelling  the  components  of  the  surface  energy  balance  (Meijninger  et  al.,  2002;  Wilson  et   al.,  2002;  Jia  et  al.,  2003;  Santos  et  al.,  2009).  

  Looking  at  the  spatial  differences  and  wet  versus  dry  grid  cell  in  Figure  5,  the  results  are   largely  realistic,  as  evaporation  from  the  top  layer  quickly  drops  during  prolonged  dry  periods.  

(18)

However,  the  topsoil  layer  is  also  the  first  to  receive  water  from  the  atmosphere,  a  characteristic   that  is  not  included  in  the  hydrological  models.  Another  systematic  error  of  the  model  becomes   evident   when   analysing   the   energy   balance   in   Figure   5B.   In   the   absence   of   a   significant   evapotranspiration   flux,   all   energy   is   immediately   expended   as   sensible   heat   flux.   In   reality,   a   portion  of  the  radiation  is  used  to  heat  the  soil  as  soil  heat  exchange  conduction  (G).  However,   this  variable  was  set  to  0  in  the  model.  The  effect  of  this  energy  flux  would  be  most  pronounced   during  the  start  and  end  of  the  day,  when  it  respectively  extracts  and  adds  energy  to  the  system.      

Model  output  with  data  from  PCR-­‐GLOBWB  

Large  areas  showed  negative  average  values  for  SHF.  In  terms  of  model  mechanics,  this  is  caused   by   the   fact   that   according   to   moisture   data   from   PCR-­‐GLOBWB,   these   areas   have   a   negative   water  deficit  meaning  they  contain  more  water  than  the  total  amount  present  at  field  capacity.   The  consequence  of  this  unlikely  phenomenon  is  that  evaporation  rises  dramatically,  drowning   out  most  spatial  differences  in  SHF  related  to  vegetation  and  soil  type.  In  reality  this  may  occur   after   periods   of   extreme   rainfall   in   poorly   drained   areas.   This   issue   can   either   be   caused   by   faulty  computation  of  soil  moisture  content  or  by  wrong  parameterization.  Either  way,  it  makes   the  soil  moisture  data  from  PCR-­‐GLOBWB  unreliable  in  the  computation  of  SHF.  

 

Variable  sensitivity  

The  results  found  in  the  sensibility  analysis  were  largely  expected.  However,  the  extent  of  the   effect  of  soil  moisture  was  less  expected.  In  the  Netherlands  fairly  wet  conditions  are  generally   dominant.  Furthermore,  the  analysis  period  was  not  particularly  dry  or  moist.  Therefore,  it  was   expected   that   the   effect   of   soil   moisture   content   on   actual   evapotranspiration   would   be   less   pronounced  than  observed.  However,  this  effect  may  be  caused  by  a  variety  of  factors.    

  First  off,  there  exists  a  strong  correlation  between  soil  type,  vegetation  and  soil  moisture   deficit.   For   this   reason,   soil   moisture   deficit   may   be   strongly   correlated   to   other   parameters   related  to  evapotranspiration  rates  rather  than  being  a  cause  on  its  own.  To  fully  understand  the   differences   in   output   and   their   cause,   the   parameters   used   in   the   model   have   to   be   analysed.   However,  the  computation  and  therefore  the  analysis  of  these  parameters  are  beyond  the  scope   of  this  research.    

  Secondly,  it  is  important  to  note  that  under  wet  conditions  evaporation  plays  a  large  role   in  the  total  evapotranspiration  (Fig.  5C).  The  current  computation  for  the  soil  moisture  content   related  to  evaporation  (top  100mm)  is  basic  and  lacks  vital  dynamics.  It  is  computed  on  a  daily   basis   from   data   obtained   from   the   hydrological   models   and   remains   constant   over   a   24-­‐hour   computation   period.   However,   in   reality   evaporation   extracts   and   precipitation   adds   water   to  

(19)

the   top   layer   throughout   the   day   thereby   influencing   evaporation   rates.   Also,   rootzone   extraction,   water   percolation   and   dew   formation   further   influence   the   top   layer   moisture   content   on   a   small   timescale.   To   successfully   capture   the   influence   of   moisture   on   evapotranspiration  on  a  1-­‐hour  resolution,  a  top  layer  water  balance  that  is  computed  per  hour   should  be  added  to  the  model.  

6.  Conclusion  

With  the  available  data  and  formulas  it  is  possible  to  model  sensible  heat  flux  on  a  100m  spatial   and   1   hour   temporal   resolution   in   the   Netherlands.   The   results   are   probable   concerning   the   order   of   magnitude   and   gradients   over   different   surface   characteristics.   However,   the   output   was   heavily   influenced   by   soil   moisture,   a   characteristic   of   the   model   confirmed   in   the   sensitivity  analysis.  Since  soil  moisture  data  is  only  computed  on  a  daily  basis,  this  characteristic   flattens  the  energy  balance  dynamics  throughout  the  day.  

  Furthermore,   running   the   model   with   input   data   from   PCR-­‐GLOBWB   proved   to   be   problematic.  The  large  influence  of  soil  moisture  caused  the  low  resolution  of  PCR-­‐GLOBWB  to   resonate   strongly   in   the   sensible   heat   flux   output.   Furthermore,   issues   related   to   the   parameterization  of  PCR-­‐GLOBWB  affected  the  output.  However,  it  is  not  entirely  clear  whether   this  results  from  errors  in  or  insufficient  study  of  the  parameterization  in  question.  

  A  possible  solution  to  these  problems  consists  of  two  efforts.  First  off,  adding  a  1-­‐hour   dynamic   topmost   soil   moisture   layer   to   the   model   would   help   to   capture   fluctuations   in   evaporation  throughout  the  day.  A  similar  layer  can  be  constructed  for  the  rootzone,  with  basic   formulas   for   root   uptake,   percolation   and   capillary   rise   to   catch   the   hourly   fluctuations   in   moisture   content.   The   second   effort   consists   of   creating   high-­‐resolution   parameterization   for   areas  outside  of  the  Netherlands.  With  the  availability  of  land-­‐use  and  soil  characteristics  on  a   high  resolution,  parameterization  can  be  computed  in  a  similar  fashion  to  the  NHI  to  force  the   low-­‐resolution  soil  moisture  input.  

  With   the   input   of   realistic   soil   moisture   content   data,   the   evapotranspiration   module   produces   an   acceptable   base   for   more   complex   computation   of   SHF.   With   the   addition   of   soil   heat   exchange   by   conduction   and   interception   latent   heat   of   evaporation,   the   occurrence   and   size  of  the  SHF  can  be  modelled  in  more  detail.  However,  more  research  is  required  to  apply  this   method   to   larger   spatial   domains,   as   the   current   parameterization   is   incapable   of   coping   with   the   soil   moisture   data   provided   by   PCR-­‐GLOBWB.   Consequently,   the   base   evapotranspiration   most   likely   drowns   out   any   further   complexity   in   SHF   dynamics   produced   by   the   soil   heat   exchange  and  interception  evaporation  module.  

(20)

5.  References  

Allen,   R.G.,   Pereira,   L.S.,   Raes,   D.,   Smith,   M.   (1998).   Crop   evapotranspiration   –   Guidelines   for   computing   crop   water   requirements   FAO   Irrigation   and   drainage   papers,   56.   FAO   –   Food   and   Agriculture  Organization  of  the  United  Nations,  Rome.  

 

Allen,   R.G.,   Pereira,   L.S.,   Smith,   M.,   Raes,   D.,   Wright,   J.L.   (2005).   FAO-­‐56   Dual   crop   coefficient   method   for   estimating   evaporation   from   soil   and   application   extensions.   Journal   of   Irrigation   and  Drainage  Engineering,  1(2),  pp.  2-­‐13.  

 

Van   Beek,   L.P.H.   and   M.F.P.   Bierkens   (2008),   The   Global   Hydrological   Model   PCR-­‐GLOBWB:   Conceptualization,  Parameterization  and  Verification,  Report  Department  of  Physical  Geography,   Utrecht   University,   Utrecht,   The   Netherlands,   http://vanbeek.geo.uu.nl/suppinfo/vanbeekbierkens2009.pdf    

Van   der   Bolt,   F.,   Van   Walsum,   P.,   Veldhuizen,   A.   (2015).   Toelichting   op   berekenen   van   verdamping  in  MetaSWAP  in  LHM.    

 

Bohrer,   G.   et   al.   (2011).   Estimating   updraft   velocity   components   over   large   spatial   scales:   contrasting  migration  strategies  of  golden  eagles  and  turkey  vultures.  Ecology  Letters,  15(2),  pp.   96-­‐103.  

 

Bonan,  G.  (2002).  Ecological  Climatology.  New  York:  Cambridge  University  Press    

Bouten,   W.,   Baaij,   E.W.,   Shamoun-­‐Baranes,   J.,   Camphuysen,   C.J.   (2013).  A   flexible   GPS   tracking  

system  for  studying  bird  behaviour  at  multiple  scales.  J.  Ornithology,  154,  pp.  571-­‐580.  

 

De   Lange,   W.J.,   Prinsen,   G.F.,   Hoogewoud,   J.C.,   Veldhuizen,   A.A.,   Verkaik,   J.,   Oude   Essink,   G.H.P.,   van  Walsum,  P.E.V.,  Delsman,  J.R.,  Hunink,  J.C.,  Massop,  H.Th.L.,  Kroon,  T.  (2014).  An  operational   multi-­‐scale,  multi-­‐model  system  for  consensus-­‐based,  integrated  water  management  and  policy   analysis:  The  Netherlands  Hydrological  Instrument.  Environmental  Modelling  &  Software,  59,  pp.   98-­‐108.  

 

Hartkamp,  A.D.,  De  Beurs,  K.,  Stein,  A.,  White,  J.W.  (1999).  Interpolation  techniques  for  climate   variables.  Mexico,  DF  (Mexico):  CIMMYT.  Series:  CIMMYT  NRG-­‐GIS  Series.  

 

Hedenström,   A.   (1993).   Migration   by   Soaring   or   Flapping   Flight   in   Birds:   The   Relative   importance   of   Energy   Cost   and   Speed.   Philosophical   Transactions   of   the   Royal   Society   B,   342(1302).  

 

Hedenström,  A.  (2002).  Aerodynamics,  evolution  and  ecology  of  avian  flight.  Trends  in  Ecology  &   Evolution,  17(9),  pp.  415-­‐422.  

 

Hornberger,   G.M.,   Spear,   R.C.   (1981).   An   approach   to   preliminary   analysis   of   environmental   systems.  Journal  of  Environmental  Management,  12,  pp.  7-­‐18.  

 

Jacovides,   C.P.,   Kontoyiannis,   H.   (1995).   Statistical   procedures   for   the   evaluation   of   evapotranspiration  computing  models.  Agricultural  Water  Management,  27,  pp.  365-­‐371.  

 

Jia,  L.,  Su,  Z.,  van  den  Hurk,  B.,  Menenti,  M.,  Moene,  A.,  De  Bruin,  H.A.R.,  Yrisarry,  J.J.B.,  Ibanez,  M.,   Cuesta,   A.   (2003).   Estimation   of   sensible   heat   flux   using   the   Surface   Energy   Balance   System  

(21)

(SEBS)   and   ATSR   measurements.   Physics  and  Chemistry  of  the  Earth,  Parts  A/B/C,   28(1-­‐3),   pp.   75-­‐88.  

 

Judd,  C.M.,  McClelland,  G.H.,  Ryan,  G.S.  (2011).  Data  analysis:  A  model  comparison  approach.  New   York:  Routledge  

 

Kunz,   T.   et   al.   (2008).   Aeroecology:   probing   and   modelling   the   aerosphere.   Integrative   &   Comparative  Biology,  48(1),  pp.  1-­‐11.  

 

Meijninger,  W.M.L.,  Hartogensis,  O.K.,  Kohsiek,  W.,  Hoedjes,  J.C.B.,  Zuurbier,  R.M.,  De  Bruin,  H.A.R.   (2002).  Determination  of  area-­‐averaged  sensible  heat  fluxes  with  large  aperture  scintillometer   over  heterogeneous  surface  –  Flevoland  field  experiment.  Boundary-­‐Layer  Meteorology,  105(1),   pp.  37-­‐62.  

 

Newton,  I.  (2008).  The  migration  ecology  of  birds.  Oxford:  Academic  press.    

Norberg,   U.M.   (1990).   Vertebrate   flight:   mechanics,   physiology,   morphology,   ecology   and   evolution.  Berlin,  Springer-­‐Verlag.  

 

Rutter,   A.J.,   Kershaw,   K.A.,   Robins,   P.C.,   Morton,   A.J.   (1971).   A   predictive   model   of   rainfall   interception  in  forests,  1.  Derivations  of  the  model  from  observations  in  a  plantation  of  Corsican   pine.  Agricultural  Meteorology,  9,  pp.  367-­‐384.  

 

Saltelli,  A.  (2002).  Sensitivity  Analysis  for  Importance  Assessment.  Risk  Analysis,  22(3),  pp.  579-­‐ 590.  

 

Saltelli,  A.,  Chan,  K.,  Scott,  M.  (2000).  Sensitivty  analysis,  New  York:  John  Wiley  &  Sons.    

Saltelli,   A.,   Ratto,   M.,   Campolongo,   F.,   Cariboni,   J.,   Gatelli,   D.,   Saisana,   M.,   Taratola,   S.   (2008).   Global  Sensitivity  Analysis:  The  Primer,  Chichester,  West  Sussex:  John  Wiley  &  Sons.  

 

Santos,  C.A.C.,  Silva,  B.B.,  Rao,  T.V.R.,  Neale,  C.M.U.  (2009).  Energy  balance  measurements  over  a   banana   orchard   in   the   Semiarid   region   in   the   Northeast   of   Brazil.   Pesquisa   Agropecuaria   Brasileira,  44(11).  

 

Shamoun-­‐Baranes,  J.,  Bouten,  W.,  van  Loon,  E.,  Meijer,  C.,  Camphuysen,  C.  (2016).  Flap  or  soar?   How   a   flight   generalist   responds   to   its   aerial   environment.   Philosophical   Transactions   of   the   Royal  Society  B.  

 

Shamoun-­‐Baranes,  J.,  Bouten,  W.,  van  Loon,  E.  (2010).  Integrating  meteorology  into  research  on   migration.  Integrative  &  Comparative  Biology,  pp.  1-­‐14.  

 

Siebert,   S.,   Döll,   P.   (2012).   Quantifying   blue   and   green   virtual   water   contents   in   global   crop   production  as  well  as  potential  production  losses  without  irrigation,  Journal  of  Hydrology,  384,   pp.  198-­‐217.  

 

Stull,  R.B.  (1988).  An  Introduction  to  Boundary  Layer  Meteorology.  Dordrecht:  Kluwer  Academic   Publishers.  

 

Vansteelant,   W.,   Bouten,   W.,   Klaassen,   R.,   Koks,   B.,   Schlaich,   A.,   van   Diermen,   J.,   van   Loon,   E.,   Shamoun-­‐Baranes,   J.   (2015).   Regional   and   seasonal   flight   speeds   of   soaring   migrants   and   the   role  of  weather  conditions  at  hourly  and  daily  scales.  Journal  of  Avian  Biology,  46(1),  pp.  25-­‐39.    

Referenties

GERELATEERDE DOCUMENTEN

Instead of dwelling on historical and ethical issues, this article asserts that using the narrative analysis of the Greimassian approach, narrative syntax in

Making these observations requires a new genera- tion of satellite sensors able to sample with these combined characteristics: (1) spatial resolution on the order of 30 to 100-m

Finally, the projected forcing conditions thus obtained can be used to drive appropriate validated coastal impact models (e.g. Delft3D, Mike21, CMS, GENESIS, SBEACH, XBeach) to

Owing to their small size and use of biocompatible DNA as a component, these nanocarriers feature advantages including: (1) high loading capacity attributed to large

Computerkunde heeft een vormende waarde die deels in hetzelfde vlak ligt als die van de wiskunde, maar de vormende waarde van computerkunde komt naar mijn overtuiging beter tot

The primary consumer protection rights legislation (CPA), in particular section 5(1)(a) as well as section 47 of ECTA, are both very clear on the applicable legislation

We analyzed sleep quantity and quality in patients with hyperacute ischemic stroke on stroke units, and related sleep to measures of functional recovery..

The positive effect of complementarity on value creation is higher in the case of opening variety of practice compared to the case of opening variety of context. Finally,