• No results found

Before  staining,  all  samples  were  diluted  (10-­‐1  and  10-­‐2)  in  sterile  4X  PBS,  and  100  µl  of  suspensions  were   filtered  thought  GTTP  Isopore  filters  of  0.22  µm  pore  size  and  16  mm  diameter  (Milipore).  Filters  were   stored  at  -­‐20ºC  until  use.  

 

11.2.  4’-­‐6-­‐diamidino-­‐2-­‐phenylindole  stain  (DAPI)  

In  order  to  determine  the  number  of  cells  present  in  each  culture,  a  piece  of  filter  was  stained  with  25-­‐

30  µl  of  4'-­‐6-­‐diamidino-­‐2-­‐phenylindole  solution  (1  µg  ml-­‐1)(Porter  &  Feig,  1980)  during  1.5  min  at  room   temperature.   Then,   washed   with   sterile   MiliQ   water   and   absolute   ethanol   and   dried   at   room   temperature   and   darkness.   Finally,   filter   was   mounted   with   a   drop   of   Citiflour   AF1   (Citifluor   ltd)   and   covered  in  a  microscope  slide.  Cells  were  quantified  using  a  fluorescence  microscope  (Axio  imager.A1,   Zeiss)  with  filter  set  49  (G  365,  FT  395,  BP  445/50,  Zeiss).  Counts  are  reported  as  means  calculated  from   15   randomly   chosen   microscope   fields.   Fifteen   microscope   fields   of   1200   µm2   were   the   optimum   number   of   fields   with   the   lowest   standard   deviation.   A   number   of   fields   >   15   did   not   produced   significant  modifications  in  the  means  and  the  standard  deviations.    

 

11.3.  Fluorescence  in  situ  hybridization  (FISH)  

To  evaluate  the  integrity  and  ribosome  containing  of  Salinibacter  cells,  a  piece  of  filter  was  hybridized   with   EHB-­‐412   monolabel   probe   (Antón   et   al.,   1999;   Antón   et   al.,   2000).   EHB-­‐412   probe   (5’-­‐

TACGCCCCATAGGGGTGT-­‐3’;  50  µg  ml-­‐1)  was  diluted  in  sterile  MiliQ  water  to  a  final  concentration  of  1  µg   ml-­‐1.  The  hybridization  was  performed  at  45%  formamide  and  the  hybridization  buffer  was  prepared  as   follows:  360  µl  5M  NaCl,  40  µl  1M  Tris-­‐HCl  pH  8.0,  904.  5  µl  formamide,  695.5  µl  Mili  Q  water,  and  2  µl  of   10%  SDS.  Filters  were  placed  on  a  clean  slide  and  each  was  hybridized  with  20  µl  of  hybridization  mix  (4   µl  probe,  16  µl  hybridization  buffer).  Slides  were  placed  into  a  hybridization  chamber  and  incubated  at   46ºC   during   2   h.   Then,   to   eliminate   the   unspecific   hybridizations,   filters   were   immersed   in   a   washing   buffer  (300  µl  5M  NaCl,  1  ml  1M  Tris-­‐HCl  pH  8.0,    500  µl  0.5  M  EDTA,  MiliQ  water  to  complete    a  final   volume  of  50  ml,    and    50  µl  of  10%  SDS)  and  incubated  at  48ºC  during  15  min.    After  washing,  filters   were  dried  at  room  temperature  and  darkness  and  then,  stained  with  DAPI.  Finally,  hibridized  cells  were   quantified  using  a  fluorescence  microscope  (Axio  imager.A1,  Zeiss)  with  filter  set  49  (G  365,  FT  395,  BP   445/50,  Zeiss)  for  DAPI,  and  the  HQ:  Cy3  filter  set  (AF  analysentechnik;  HQ  545/30,  Q  570  lp,  HQ610/75).  

Counts   were   reported   as   means   calculated   from   15   randomly   chosen   microscope   fields   and   the   percentage  of  hibridized  cells  was  calculated  based  on  the  total  of  DAPI  counts  of  each  sample.  

 

                                           

    IV.  RESULTS  AND  DISCUSSION  

           

 

Results  and  Discussion:  Chapter  1

 

-­‐69-­‐  

 

CHAPTER  1:  Intraspecific  diversity  and  biogeography  of  S.ruber  strains    

1.1.Background  

Growth  of  the  extremely  halophilic  bacterium  S.  ruber  (Antón  et  al.,  2000)  is  constrained  to  relatively   small  water  bodies  in  restricted  areas  on  Earth.  S.  ruber  has  been  isolated  from  different  areas  of  the   world,  and  in  sites  as  diverse  as  Mediterranean  coastal  solar  salterns  (Peña  et  al.,  2005)  or  the  remote   Andean  Peruvian  salterns  of  Maras  at  3,380  m  above  sea  level  (Maturrano  et  al.,  2006a).  The  extreme   conditions   and   geographical   isolation   of   its   environments   are   optimal   circumstances   for   observing   allopatric  speciation  (Coyne  &  Orr,  2004;  Whitaker,  2006).  Preliminary  analyses  based  on  fingerprinting   genomic   traits,   such   as   PFGE   or   RAPD,   although   indicating   a   certain   incipient   trend,   did   not   render   a   clear   cut   geographical   discrimination   among   isolates   (Peña   et   al.,   2005).   In   order   to   discern   biogeographical   patterns   in  S.   ruber,   ten   strains   from   five   different   locations   were   selected   to   study,   through  MLSA,  the  intraspecific  diversity  within  the  same  group.  For  this,  twelve  protein-­‐coding  genes,   which  had  been  observed  as  phylogenetically  informative,  were  selected  (Sória-­‐Carrasco  et  al.,  2007).  In   addition,   a   metabolomic   approach   by   Ion   Cyclotron   Resonance   Fourier   Transform   Mass   Spectrometry   (ICR-­‐FT/MS)   was   performed   to   evaluate   some   phenotypic   evidence   for   allopatric   segregation   of   members   of  S   ruber,   by   identification   of   phenotypic   patterns   of   the   chemical   extracts   of   this   strain   collection  (as  detailed  in  Materials  and  Methods  section).    

   

1.2.Multilocus  sequence  analysis  (MLSA)  

Multilocus  sequence  analysis  was  applied  to  study  the  intraspecific  diversity  of  S.  ruber  strains  isolated   from  different  geographic  localizations.  The  obtained  sequences  of  specific  coding-­‐protein  genes  from   each   strain   were   concatenated   and   analyzed   to   calculate   the   total   number   of   synonymous   or   nonsynonymous  substitutions.  Finally,  different  phylogenetic  reconstructions  were  applied  in  order  to   evaluate  stability  of  the  genealogies  by  including  and  excluding  the  16S  rRNA  gene  sequences.  

 

1.2.1.Amplification  of  protein-­‐coding  genes  

A  total  of  ten  S.  ruber  strains  (Table  3)  were  selected  in  order  to  study  their  intraspecific  diversity  by   MLSA.   The   selected   strains   were   representative   of   three   main   geographic   areas:   Mediterranean   (M8,   M31,  P13,  P18,  E3,  E7),  Atlantic  (C9,  C14)  and  Peruvian  (PR1,  PR3).    

Results  and  Discussion:  Chapter  1

 

-­‐70-­‐  

 

The   study   was   conducted   using   12   coding-­‐protein   genes   that   were   selected   according   to   previous   phylogenetic  S.  ruber  studies  (Soria-­‐Carrasco  et  al.,  2007).    

DNA   amplifications   were   performed   following   standard   procedures   as   indicated   in   Materials   and   Methods.  Amplified  products  of  proper  size  were  obtained  with  all  primers  tested  and  in  most  of  the   analyzed   strains.   Figure   12   shows   examples   of   these   amplifications   in   10   tested   strains,   where   the   expected  size  of  amplicons  for  groEL  (1293  bp)  and  tuf  (996  bp)  genes  can  be  seen.  Excluding  ES4  strain,   all  strains  showed  a  good  signal  and  mostly  optimal  concentrations  to  be  sequenced  (Fig.  12).  In  order  to   obtain   a   good   concentration   for   sequencing   purposes,   DNAs   yielding   lower   concentrations   were   re-­‐

amplified   (Fig.   12).   PCR   products   were   purified   using   standard   procedures,   quantified   and,   finally   sequenced  (see  Materials  and  Methods).    

                 

Figure  12:    Amplified  products  of  coding-­‐  protein  genes  in  S.  ruber  strains.  The  picture  shows  groEL  (A)  and  tuf  (B)   genes  separated  in  1.5%  agarose  gel  and  visualized  by  ethidium  bromide  staining.    Size  of  amplicons  and  names  of   strains  are  also  specified  (M:  Lambda  DNA/PstI  marker;  NC:  negative  control).  

 

1.2.2.Sequencing  of  protein-­‐coding  genes  

Sequencing  of  amplicons  was  performed  with  the  same  primers  used  for  gene  amplification,  and  under   the   conditions   explained   in   the   Materials   and   Methods   section.   Only   seven   out   of   the   twelve   tested   genes  yielded  good  sequences  and  could  be  used  for  phylogenetic  studies.  The  sequences  obtained  for   each   gene   were   checked   and   corrected   using   the   program   Sequencher   v4.7   (Gene   Codes   Corp.).  

Subsequently,   the   sequences   were   aligned   using   the   ClustalX   1.83   program.   The   alignment   was   then   manually   improved.   Finally,   hypervariable   or   unalignable   positions   were   removed   using   Gblocks   program  (http://molevol.ibmb.csic.es/Gblocks-­‐_server.html).  

The  total  number  of  substitutions  independently  of  whether  they  were  synonymous  or  nonsynonymous,   as  well  as  the  number  of  insertions-­‐deletions,  were  calculated  for  each  gene  (Table  8).    

Results  and  Discussion:  Chapter  1

 

present  any  type  of  substitution,  corresponding  to  the  most  conserved  gene  in  these  strains.  Insertions-­‐

deletions  were  not  present  in  any  of  the  analyzed  genes.  

Results  and  Discussion:  Chapter  1

 

-­‐72-­‐  

 

1.2.3.Phylogenetic  reconstruccions  

Phylogenetic  reconstructions  were  based  on  multiple  concatenated  genes  from  different  sequences  of   each  strain  data  sets  (Table  8).  Multiple  analyses  were  performed  in  order  to  find  topology  changes  due   to   the   effect   of   the   gene   composition   of   the   alignments,   and   to   evaluate   tree   topology   stabilities,   as   previously   recommended   (Ludwig   &   Klenk,   2001).     Bootstrap   values   were   performed   by   using   the   PHYML  program  with  a  total  of  100  replicates.    

As  mentioned  before,  phylogenetic  analyses  were  performed  by  including  and  excluding  indels,  as  well   as  by  using  different  data  sets:  one  including  the  16S  rRNA  gene  in  the  concatenate  (7,995  nucleotide   alignment,   Fig.   13)   or   disregarding   it   (6,513   nucleotide   alignment,   Fig.   14).   In   the   first   case,   it   was   remarkable  that  Ebro  strains  (Mediterranean)  were  affiliated  with  the  Peruvian  and  Atlantic  strains  (Fig.  

13).    

Similar  results  were  also  observed  when  removing  the  16S  rRNA  gene  from  the  analysis,  and  were  even   less   related   to   the   rest   of   the   Mediterranean   strains.   Besides,   Atlantic   strains   did   not   show   a   clear   phylogenetic  trend  (Fig.  14).  In  any  case,  most  of  the  trees  gave  congruent  topologies,  independently  of   the  use  of  PHYML  or  ARB  programs,  or  the  used  algorithms  (ML,  maximum  likelihood  and  NJ,  neighbour   joining).  Only  ML  showed  different  topologies  when  including  the  indels  in  the  analysis.  However,  none   of  the  tree  topologies  obtained  showed  a  clear  geographic  trend.  Despite  the  relatively  low  bootstrap   value   for   the   position   of   strain   E7,   the   remaining   topology   was   robust   with   bootstrap   values,   always   above  80%.  Contrarily  to  the  same  reconstruction,  where  the  SSU  rRNA  gene  was  included,  bootstrap   values  were  lower.  However,  despite  a  lower  robustness  of  the  tree  topology,  there  was  no  doubt  about   the  common  affiliations  between  the  Peruvian  and  Atlantic  strains,  and  between  C9  and  E3  strains  (Fig.  

13).  

In  general,  the  trees  agreed  with  regard  to  their  topologies,  since  only  M8  acquired  a  stable  position   when  including  the  16S  rRNA  gene  sequence  in  the  analysis.  Altogether,  both  reconstructions  did  not   show   a   clear   geographical   segregation   of   the   selected   strains,   in   contrast   to   observations   made   with   other  extremophiles  (Whitaker  et  al.,  2003).  Strains  from  Alicante  (P13  and  P18)  affiliated  together  with   those  of  Majorca  (M8  and  M31).  However,  the  Mediterranean  strains  E3  and  E7  affiliated  together  with   those   from   the   Atlantic   (C9   and   C14)   and   Peru   (PR1   and   PR3).   Neither   previous   studies   with   fingerprinting  techniques  (Peña  et  al.,  2005),  nor  those  with  an  MLSA  of  gene  stretches  of  nearly  8,000   homologous  positions  (Fig  13),  were  informative  enough  to  resolve  biogeographical  segregation.    

 

Results  and  Discussion:  Chapter  1

 

-­‐73-­‐  

 

This   could   be   an   indication   that   the   process   of   genetic   divergence   may   be   still   at   an   early   stage,   not   rendering  clearly  resolvable  trends.  However,  given  that  the  size  of  the  S.  ruber  genome  is  about  3,000   open  reading  frames  (Mongodin  et  al.,  2005),  and  despite  the  fact  that  the  genes  were  selected  from  a   set  of  putative  phylogenetic  markers  (Soria-­‐Carrasco  et  al.,  2007),  those  may  not  be  adequate  for  the   understanding  of  a  subtle  geographical  segregation.    

In  the  future,  intraspecific  whole-­‐genome  comparisons  with  S.  ruber  strains  might  indicate  which  genes   could  be  useful  for  understanding  allopatric  differentiation  based  on  genetic  drift.    

   

       

Figure   13:   Phylogenetic   reconstruction   based   on   a   PHYML   algorithm   corresponding   to   8   housekeeping   genes   including  SSU  rRNA  gene.  Strains  from  different  geographical  areas  are  marked  with  their  respective  colors.  The  bar   indicates  1%  of  sequence  divergence.    

           

 

Figure  14:  Phylogenetic  reconstruction  based  on  a  PHYML  corresponding  to  the  7  housekeeping  genes.  Strains  of   different  geographical  areas  are  marked  with  their  respective  colors.  The  bar  indicates  1%  sequence  divergence.    

 

Results  and  Discussion:  Chapter  1

 

-­‐74-­‐  

 

1.3.Metabolomic  comparisons  of  strains  

As  stated  above,  standard  genotyping  techniques  may  not  always  help  in  clearly  resolving  intraspecific   diversity,  so  there  is  a  need  to  apply  new  approaches  to  understand  allopatric  segregation  of  members   of   the   same   species   (Ramette   &   Tiedje,   2007).   For   this   reason,   a   nontargeted   metabolite   profiling   approach,   using   high-­‐field   ICR-­‐FT/MS,   was   evaluated   in   the   chemical   extracts   of   our  S.   ruber   strain   collection.  

 

1.3.1.Metabolome  composition  analyses    

The  first  experiment  was  performed  with  28  isolates  of  S.  ruber  coming  from  seven  geographical  areas  in   the   world   (Table   3),   which   were   divided   into   three   regions:   Mediterranean   (10   strains),   Atlantic   (13   strains)   and   Peruvian   (5   strains).   All   of   them   were   grown   simultaneously   with   the   same   medium   and   under  identical  environmental  conditions  to  avoid  culture-­‐dependent  differences.    Each  cell  suspension   was  processed  for  the  metabolite  extraction  resulting  in  three  cellular  fractions:  extracellular  (E),  cellular   soluble  (CS)  and  cellular  insoluble  (CI)  fractions,  which  were  grouped  and  analyzed  by  ICR-­‐FT/  MS  (see   Materials  and  Methods).    

The  complete  set  of  metabolomes  rendered  a  total  of  247,255  m/z,  from  which  11,880  were  attributed   to  known  elementary  compositions   containing  the  elements  C,  H,  O,  N  and  S.  Single-­‐peak  occurrence   was   reduced   from   11,880   (verified   by   isotopic   assignments   of   elementary   composition)   to   a   total   of   8,873  metabolites  at  a  m/z  lower  than  550  a.m.u.  (highest  probable  assignments),  which  were  used  for   statistical  analysis  (Table  9).    

 

1.3.2.Statistical  analyses  and  proposed  models  

The   statistical   model   used   for   data   processing   was   the   partial   least   squares   for   discriminant   analysis   (PLS-­‐DA),   which   revealed   statistically   significant   differences   between   the   different   geographical   areas   (P<0.05)  in  the  three  analyzed  fractions  (Table  9).  Models  were  calculated  independently  for  the  three   cellular  fractions  (extracellular,  cellular  soluble  and  cellular  insoluble),  and  the  cellular  insoluble  fraction   (CI)  was  chosen  as  the  most  descriptive  of  the  model  (Fig.  15).    

In  addition,  the  Pareto  scaling  of  the  intensity  values  with  a  logarithmic  transformation  of  the  data  was   chosen  to  consider  all  masses  equally,  including  those  with  medium-­‐  and  low-­‐intensity  values  (van  den   Berg  et  al.,  2006).  PLS-­‐DA  using  four  significant  components,  R2Y  (cum)  was  equal  to  0.98  and  Q2  (cum)   was  equal  to  0.45,  both  with  values  indicating  high  predictive  power.  

Results  and  Discussion:  Chapter  1

 

-­‐75-­‐  

 

Although  most  of  the  analyses  were  performed  using  the  whole  metabolome,  the  common  metabolome   in   all   strains   (also   called   core   metabolome)   was   formed   by   about   2,550   unique   masses   distinctly   expressed  in  the  different  analyzed  groups.  Moreover,  the  discriminative  metabolome  (i.e.  not  common   masses  to  all  extracts)  consisted  in  6,323  unique  masses  (Table  9).  The  number  of  discriminative  masses   considered  for  statistical  analysis  and  for  geographical  discrimination,  all  of  them  analyzed  in  positive  mode.  

 

Results  and  Discussion:  Chapter  1

 

-­‐76-­‐  

 

The   score   scatter   plot   and   the   loading   plots   revealed   statistically   significant   differences   (P<0.05)   between   the   different   samples.   The   score   scatter   plot   (Fig.   15a)   presented   a   view   of   how   well   the   different  geographical  origins  (classes)  were  separated  on  the  basis  of  their  metabolomic  composition  (X   variables).     Thus,   metabolome   comparisons,   focusing   on   geographically   discriminative   data,   yielded   clear-­‐cut  allopatric  differences.  In  the  same  way,  nontargeted  analysis  revealed  unique  features  for  each   group  of  isolates.  Fig.  15b  shows  the  loading  plot  in  where  the  different  masses  characteristic  to  each   geographical   area   were   differently   coloured   (red   for   Atlantic,   green   for   Mediterranean   and   blue   for   Peruvian   strains).   The   discriminative   masses   (variables   m/z)   for   each   origin   of   isolation   were   chosen   according  to  their  correlation  coefficient  value.  Those  having  the  highest  coefficients  were  considered  to   be  relevant  (that  is,  variables  (m/z)  with  a  correlation  value  higher  than  |0.002|).    

 

 

 

 

 

 

 

 

         

Figure  15:  PLS-­‐DA  models  of  all  cellular  insoluble  fractions  analyzed  with  electrospray-­‐  positive  mode  ICR-­‐FT/MS.  

(a)  Score   plot   model   shows   the   differentiation   based   on   the   geographical   origin   of   the   analyzed   samples.  (b)   Loading   plot   represents   the   masses   of   known   elementary   composition   (C,   H,   O,   N,   S   and   m/z   <550)   used   for   discriminating   analysis,   and   correlating   to   the   geographical   origins.   The   masses   with   a   high   correlation   with   geographical   origin   are   highlighted   with   their   corresponding   colour   depending   on   the   area,   while   the   non-­‐

discriminating  masses  are  represented  in  yellow.  

 

A  

B  

Results  and  Discussion:  Chapter  1

 

-­‐77-­‐  

 

Those  masses  associated  with  the  highest  correlation  coefficients  were  represented  in  the  van  Krevelen   projection   (H/C   versus   O/C   on   the   basis   of   their   elementary   composition   values;   Fig.   16)(Wu   et   al.,   2004).  In  this  regard,  the  most  relevant  markers  were  C,  H  and  O  molecules,  with  only  a  few  metabolites   containing   sulfur   or   nitrogen   (Fig.   16a).   The   comparisons   of   these   molecules   with   the   total   metabolic   spaces  showed  that  the  discriminative  metabolites  may  be  structurally  aliphatic  and  depleted  in  oxygen   (Fig.   16b).   Thus,   those   components   generally   associated   to   cell   membranes,   such   as   fatty   acids   and   terpenoids,  may  be  responsible  for  the  geographic  differentiation.  

 

 

 

 

 

 

 

 

 

 

 

Figure  16:  Representations  of  all  discriminating  m/z  values  from  metabolome  analysis.  (a)  Independent  of  their   geographical  origin  but  coloured  only  as  a  function  of  their  attributed  elementary  composition  (CHO,  CHON,  CHOS   or   CHONS)   and   visualized   in   a   van   Krevelen   diagram   (H/C   versus   O/C).   Most   of   the   discriminative   metabolites   contain  only  C,  H  and  O  and  these  are  compared  within  a  van  Krevelen  type  of  diagram  to  the  CHONS-­‐containing   metabolites  of  general  metabolome  databases  (www.metabolome.jp,  www.genome.jp/kegg/)  shown  in  gray  in  the   figure.   Note   that   the   triangular   region   corresponds   to   peptides   (CHON   and   CHONS);  (b)   All   discriminating   CHO   metabolites  in  a  van  Krevelen  diagram  coloured  as  a  function  of  their  origin  and  compared  to  the  metabolites  of   general  metabolome  databases  (www.metabolome.jp,  www.genome.jp/kegg/)  shown  in  gray  in  the  figure.  

   

A  

B  

Results  and  Discussion:  Chapter  1

 

-­‐78-­‐  

 

Besides,  between  the  two  principal  components,  a  relationship  for  geographical  distance  between  the   isolation  sites  could  be  found.  It  seemed  that  for  the  second  principal  component,  the  Atlantic  strains   showed   intermediate   differences   with   those   of   the   Mediterranean   and   Peruvian.   When   specifically   directing   the   recognition   of   discriminative   metabolites   among   geographically   distinct   metabolomes,   a   set   of   conspicuous   compounds   were   unambiguously   assigned   to   a   sulfonolipid   family   (Table   10).   The   compounds   forming   this   family   have   been   observed   to   be   major   components   of   the   cell   envelope   of   Cytophaga  (Godchaux  III  &  Leadbetter,  1984),  a  member  of  the  same  phylum  that  S.  ruber  (Antón  et  al.,   2002).  These  compounds,  which  could  account  for  10%  of  total  cellular  lipids,  have  been  proposed  as   signatures   for   S.   ruber   identification   (Corcelli   et   al.,   2004).   In   fact,   one   of   them   (C35H67NO8S,   m/z=660.4505),  whose  higher  intensity  was  observed  in  the  Peruvian  strains,  has  been  reported  to  be   characteristic   of  S.   ruber  (Table   10)  (Corcelli   et   al.,   2004).   The   ICR-­‐FT/MS   approach,   with   a   mass   precision  lower  than  600  p.p.b.,  revealed  that  S.  ruber  may  contain  at  least  nine  additional  sulfonolipids   analogous   to   C35H67NO8S   with   a   mass   range   644–688.   These   compounds   differ   from   the   originally   described   sulfonolipid   in   their   elementary   composition,   with   variations   in   their   side   chain   length,   insaturation  or  hydroxylation  degree,  with  variations  in  CH2,  H2  and  O,  respectively  (Table  10).  All  these   compounds  were  found  in  all  of  the  analyzed  samples  (in  negative  mode)  and  with  identical  intensity   ratios   between   isolates   from   the   same   location,   except   for   m/z   676.4454   (C35H68NO9S)   and   m/z   688.4455  (C36H68NO9S),  which  seemed  to  be  exclusive  of  the  Atlantic  strains  (Table  10).    

 

In   addition,   the   metabolomic   approach   allowed   the   targeted   search   for   special   metabolic   traits   considered  to  be  relevant  in  the  organisms’  phenotype.  Previous  biochemical  studies  on  S.  ruber-­‐type   strain   M31   revealed   the   presence   of   an   active,   hitherto   unreported,   rhodopsin   type   of   membrane   proton  translocation  system,  the  xanthorhodopsin,  responsible  for  the  putative  phototrophy  of  S.  ruber   (Balashov   et   al.,   2005).   In   the   same   way,   the   genome   sequence   of   the   same   organism   revealed   the   coding  region  of  one  halorhodopsin  (Peña  et  al.,  2005)  and  two  sensory  rhodopsin  homologous  genes   (Mongodin  et  al.,  2005).    

Searching  for  an  indication  of  the  presence  of  retinal,  the  chromophore  that  bound  to  rhodopsins,  an   experimental  positive  mass  285.22125  (theoretical  285.22129)  was  found  in  all  samples.  However,  the   m/z  value  was  only  discriminative  for  the  Mediterranean  strains.    

     

Results  and  Discussion:  Chapter  1

 

Table  10:  Proposal  elemental  composition  of  masses  assigned  to  sulfonolipids  with  their  structural  variations  from   C35H67NO8S,  originally  described  by  Corcelli  et  al.  (2004)  as  C35H66NO8S.  All  these  compounds  were  found  in  analyzed   samples  in  negative  mode,  where  n  indicates  the  number  of  analyzed  strains,  and  ND,  not  detected.  

 

   

1.3.3.Discriminative  analysis  of  Mediterranean  strains    

An  independent  experiment  was  performed  to  evaluate  the  metabolomic  composition  of  four  replicates   from  five  Mediterranean  strains:  P13  and  P18  from  Alicante,  M8  and  M31  from  Majorca,  and  IL3  from  

Results  and  Discussion:  Chapter  1

 

elaborated  to  the  discriminative  analysis  of  the  Mediterranean  strains.  The  soluble  cellular  fraction  was   chosen  as  representative  in  the  OPLS-­‐DA  model,  which  rendered  equivalent  but  clearer  results  than  PLS-­‐

DA,   showing   a   clear   separation   between   the   three   groups   of   strains   (Fig.   17).   Therefore,   these   differences  observed  between  different  strains  may  be  attributed  to  strain-­‐specific  metabolisms  rather   than  sample-­‐to  sample  variations.   pathways  (confirmed  by  KEGG  and  Japanese  metabolome  database).  

   

In   contrast   to   previous   results   (Peña   et   al.,   2005),   when   searching   for   discriminative   phenotypes   at   a   more  reduced  geographical  scale  using  the  ICR-­‐FT/  MS  approach,  a  phenotypic  segregation  in  individual   locations  was  observed  (Fig.  17a).  The  main  discriminative  metabolomics  profile  features  were  different   from  those  giving  resolution  at  a  larger  geographical  scale.    

 

A  

B  

Results  and  Discussion:  Chapter  1

 

-­‐81-­‐  

 

In  such  cases,  geographical  differences  were  associated  to  strain-­‐specific  compositions  of  N-­‐containing   molecules   (Fig.   17b).   The   confrontation   of   their   exact   masses   with   the   KEGG   and   the   Japanese   metabolome  database  (www.metabolome.jp)  indicated  that  the  discriminative  molecules  were  involved   primarily   in   the   core   metabolism   (that   is,   carbohydrate,   amino   acid,   and   fatty   acid   biosynthesis   and   metabolism).  

   

1.4.Conclusions  

These  findings  revealed  that  intraspecific  metabolic  diversity  of  S.  ruber  can  be  readily  detected  by  the   ICR-­‐FT/MS   approach   and   that   such   diversity   can   be   associated   to   different   geographical   patterns   at   different  metabolic  levels,  which  could  not  be  revealed  by  standard  genetic  methods  previously  used  to   assess  biogeography  of  prokaryotes  (Ramette  &  Tiedje,  2007;  Whitaker  et  al.,  2003).  

 

MLSA  approach,  based  on  different  gene  data  sets,  did  not  resolve  putative  genetic-­‐geographic  patterns,   as  the  genetic  divergence  may  be  too  subtle  for  the  given  selection  of  genes.  However,  one  must  take   into  account  that,  despite  the  fact  that  large  sets  of  concatenated  genes  tend  to  reflect  the  organismal   phylogeny  (Soria-­‐Carrasco  et  al.,  2007),  perhaps  only  full  genome  sequences  may  reflect  geographical   isolation  in  the  strain  collection  of  S.  ruber.  This  could  result  in  accordance  with  the  taxa  segregation   that   correlates   with   the   average   nucleotide   or   amino-­‐acid   identity   of   shared   genes   (Konstantinidis   &  

Tiedje,  2005).However,  the  still  sparse  database  of  full  genomes,  makes  the  metabolomic  approach  a   fast   and   less   expensive   alternative   for   revealing   prokaryotic   biogeography,   with   the   added   value   of   being  discriminative  at  different  levels  at  the  geographical  scale.  

It  seems  clear  that  different  studied  regions  led  to  the  isolation  of  strains  sharing  common  metabolic   traits,   such   as,   the   distinct   production   of   sulfonolipid   derivates.   However,   differences   were   generally   related   to   quantitative   composition   yields,   rather   than   qualitative   production   of   distinct   compounds.  

Nevertheless,   at   the   molecular   level,   these   facts   could   be   also   attributed   to   transcriptional   or   posttranscriptional   regulations   rather   than   composition   changes   in   genes   at   the   genomic   level.   In   addition,   the   metabolic   differences   correlated   with   the   geographical   areas,   influenced   perhaps   by   environmental  conditions  such  as  climate  and  distance,  since  Peruvian  and  Mediterranean  strains  were   found  as  the  most  different.    

   

Results  and  Discussion:  Chapter  1

 

-­‐82-­‐  

 

The  major  forces  for  these  differences  may  be  related  to  their  distinct  response  to  the  environmental  

The  major  forces  for  these  differences  may  be  related  to  their  distinct  response  to  the  environmental