cloning and expression of three lada-type alkane

22
Instructions for use Title Cloning and expression of three ladA-type alkane monooxygenase genes from an extremely thermophilic alkane- degrading bacterium Geobacillus thermoleovorans B23 Author(s) Boonmak, Chanita; Takahashi, Yasunori; Morikawa, Masaaki Citation Extremophiles, 18(3), 515-523 https://doi.org/10.1007/s00792-014-0636-y Issue Date 2014-05 Doc URL http://hdl.handle.net/2115/59072 Rights The final publication is available at Springer via http://dx.doi.org/10.1007/s00792-014-0636-y Type article (author version) File Information Boonmak Extremophiles 2014.pdf Hokkaido University Collection of Scholarly and Academic Papers : HUSCAP

Upload: others

Post on 24-Dec-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Cloning and expression of three ladA-type alkane

Instructions for use

Title Cloning and expression of three ladA-type alkane monooxygenase genes from an extremely thermophilic alkane-degrading bacterium Geobacillus thermoleovorans B23

Author(s) Boonmak, Chanita; Takahashi, Yasunori; Morikawa, Masaaki

Citation Extremophiles, 18(3), 515-523https://doi.org/10.1007/s00792-014-0636-y

Issue Date 2014-05

Doc URL http://hdl.handle.net/2115/59072

Rights The final publication is available at Springer via http://dx.doi.org/10.1007/s00792-014-0636-y

Type article (author version)

File Information Boonmak Extremophiles 2014.pdf

Hokkaido University Collection of Scholarly and Academic Papers : HUSCAP

Page 2: Cloning and expression of three ladA-type alkane

 

 

1  

1  

Cloning  and  expression  of  three  ladA-­‐type  alkane  monooxygenase  genes  1  

from  an  extremely  thermophilic  alkane-­‐degrading  bacterium  Geobacillus  2  

thermoleovorans  B23  3  

 4  

Chanita  Boonmak,  Yasunori  Takahashi  and  Masaaki  Morikawa  5  

 6  

Division  of  Biosphere  Science,  Graduate  School  of  Environmental  Science,  Hokkaido  7  

University,  Sapporo  060-­‐0810,  Japan  8  

 9  

Correspondence:  Masaaki  Morikawa  10  

Tel/Fax:  +81-­‐11-­‐706-­‐2253;  Email:  [email protected]  11  

 12  

 13  

Abbreviations:  C12,  dodecane;  C16,  hexadecane  14  

 15  

16  

Page 3: Cloning and expression of three ladA-type alkane

 

 

2  

2  

Abstract  1  

An  extremely  thermophilic  bacterium,  Geobacillus  thermoleovorans  B23,  is  capable  of  2  

degrading  a  broad  range  of  alkanes  (with  carbon  chain  lengths  ranging  between  C11  and  3  

C32)  at  70  °C.  Whole-­‐genome  sequence  analysis  revealed  that  unlike  most  alkane-­‐degrading  4  

bacteria,  strain  B23  does  not  possess  an  alkB-­‐type  alkane  monooxygenase  gene.  Instead,  it  5  

possesses  a  cluster  of  three  ladA-­‐type  genes,  ladAαB23,  ladAβB23,  and  ladBB23,  on  its  6  

chromosome,  whose  protein  products  share  significant  amino  acid  sequence  identities,  49.8,  7  

34.4,  and  22.7%,  respectively  with  that  of  ladA  alkane  monooxygenase  gene  found  on  a  8  

plasmid  of  Geobacillus  thermodetrificans  NG  80-­‐2.  Each  of  the  three  genes,  ladAαB23,  9  

ladAβB23,  and  ladBB23,  was  heterologously  expressed  individually  in  an  alkB1  deletion  mutant  10  

strain,  Pseudomonas  fluorescens  KOB2Δ1.  It  was  found  that  all  three  genes  were  functional  11  

in  P.  fluorescens  KOB2Δ1,  and  partially  restored  alkane  degradation  activity.  In  this  study,  12  

we  suggest  that  G.  thermoleovorans  B23  utilizes  multiple  LadA-­‐type  alkane  13  

monooxygenases  for  the  degradation  of  a  broad  range  of  alkanes.  14  

 15  

Keywords:  LadA  alkane  monooxygenase,  Geobacillus  thermoleovorans  B23,  alkane  16  

degradation,  genome  17  

 18  

   19  

Page 4: Cloning and expression of three ladA-type alkane

 

 

3  

3  

Introduction  1  

Petroleum  hydrocarbons  are  toxic  to  animals  and  plants;  however,  more  than  a  small  2  

number  of  microorganisms  are  capable  of  degrading  and  using  them  as  carbon  and  energy  3  

sources  (Rojo  2009;  Das  and  Chandran  2011).  The  first  step  of  alkane  degradation  by  aerobic  4  

microorganisms  is  the  hydroxylation  reaction  catalyzed  by  alkane  monooxygenases.  One  of  5  

the  best  studied  enzymes  is  an  AlkB-­‐type  alkane  monooxygenase,  which  is  widely  6  

distributed  in  bacteria  such  as  Gram-­‐negative  Acinetobacter,  Alcanivorax,  Burkholderia,  7  

Pseudomonas,  and  Stenotrophomonas  and  Gram-­‐positive  Mycobacterium,  Nocardia,  8  

Prauserella,  Rhodococcus,  and  Geobacillus  (van  Beilen  and  Funhoff  2007;  Takei  et  al.  2008;  9  

Rojo  2009).  Tourova  et  al.  (2008)  reported  that  eight  different  alkB  gene  homologs,  alkB-­‐10  

geo1  to  alkB-­‐geo8,  are  distributed  among  the  members  of  the  genus  Geobacillus.  These  11  

homologs  occur  in  each  strain  in  various  combinations  of  four  to  six  alkB  gene  homologs.  12  

Analysis  of  the  codon  usage  in  the  various  alkB  homologs  suggested  that  the  pool  of  these  13  

genes  is  common  to  most  Gram-­‐positive  and  certain  Gram-­‐negative  bacteria  (Tourova  et  al.  14  

2008).  15  

An  extremely  thermophilic  bacterium,  Geobacillus  thermoleovorans  strain  B23,  was  16  

previously  isolated  from  the  production  waters  of  a  subterranean  petroleum  reservoir  at  a  17  

depth  of  2150  m  and  a  temperature  of  105  °C,  in  Niigata,  Japan.  Strain  B23  degrades  a  broad  18  

range  of  alkanes  (C11–32)  at  70  °C  by  a  terminal  oxidation  pathway,  followed  by  a  β-­‐19  

oxidation  pathway  (Kato  et  al.  2001a).  To  understand  the  metabolism  of  alkane  degradation  20  

in  this  strain,  alcohol  dehydrogenase  and  aldehyde  dehydrogenase  have  been  previously  21  

characterized  and  demonstrated  their  high  thermostability  and  preference  for  long-­‐chain  22  

substrates  (Kato  et  al.  2001b;  Kato  et  al.  2010).  Moreover,  an  investigation  of  alkane-­‐23  

inducible  proteins  in  this  strain  revealed  that  the  synthesis  of  two  unique  membrane  24  

proteins  (P16  and  P21),  together  with  catalase  and  superoxide  dismutase,  was  significantly  25  

increased  during  alkane  degradation.  An  acyl-­‐CoA  oxidase  activity  of  the  type  found  in  26  

eukaryotes  was  also  suggested  in  strain  B23  (Kato  et  al.  2009).  However,  the  identity  of  the  27  

alkane  monooxygenase,  which  catalyzes  the  initial  step  of  the  alkane  degradation  pathway,  28  

has  remained  unknown.  To  obtain  further  knowledge  of  the  unique  alkane  degradation  29  

pathway  in  G.  thermoleovorans  B23,  whole  genome  sequence  analysis  was  conducted,  and  30  

the  candidate  alkane  monooxygenase  genes  were  characterized,  as  detailed  below.  31  

 32  

Page 5: Cloning and expression of three ladA-type alkane

 

 

4  

4  

Materials  and  methods  1  

 2  

Strains,  plasmids,  and  media  3  

Geobacillus  thermoleovorans  B23,  was  grown  in  Luria  broth  (LB)  medium  for  maintainance  4  

at  60  °C.  Escherichia  coli  DH5α  was  grown  in  LB  medium  at  37˚C  and  used  as  the  host  strain  5  

for  gene  cloning  using  the  cloning  vector  pUC19  (Yanisch-­‐Perron  et  al.  1985).  The  6  

transformants  harboring  pUC19  and  its  derivatives  were  grown  in  LB  medium  supplemented  7  

with  100  µg/ml  ampicillin.  E.  coli  MegaX  DH10BTM  T1R  ElectrocompTM  cells  [Δ(mrr-­‐hsdRMS-­‐8  

mcrBC);  Invitrogen,  Carlsbad,  CA)  were  used  for  cloning  the  gene  expression  plasmids,  9  

pCom8:ladAαB23,  pCom8:ladAβB23,  and  pCom8:ladBB23.  Pseudomonas  fluorescens  KOB2Δ1  is  10  

an  alkB1  deletion  mutant  of  P.  fluorescens  CHA0,  with  dramatically  reduced  C12–C16  alkane  11  

degradation  activity  (Smits  et  al.  2001).  This  strain  can  thus  be  used  for  alkane  degradation  12  

and  growth  complementation  experiments  using  alkane  monooxygenase  genes  from  other  13  

bacteria  (Smits  et  al.  2001).  pCom8  is  an  E.  coli–Pseudomonas  shuttle  expression  vector  14  

based  on  pUCP25,  with  a  P.  putida  GPo1  alkB  promoter  and  a  positive  regulator,  alkS  (Smits  15  

et  al.  2001).  The  transformants  harboring  pCom8  and  its  derivatives  were  grown  on  LB  16  

medium  supplemented  with  the  appropriate  antibiotics  (10  µg/ml  gentamicin  for  E.  coli  and  17  

100  µg/ml  gentamicin  for  P.  fluorescens;  Smits  et  al.  2002).  The  recombinant  E.  coli  strains  18  

and  P.  fluorescens  KOB2Δ1  were  grown  in  LB  medium  at  37  °C  and  30  °C,  respectively,  19  

unless  otherwise  noted.    20  

 21  

Draft  genome  sequencing  and  assembly  22  

Chromosomal  DNA  from  G.  thermoleovorans  B23  was  prepared  using  Marmur  method  23  

(Marmur  1961).  Paired-­‐end  libraries  were  generated  from  the  chromosomal  DNA,  and  24  

sequencing  was  carried  out  using  Roche  454  GS  FLX  (Basel,  Switzerland)  at  Hokkaido  System  25  

Science  Co.,  Ltd.  (Sapporo,  Japan).  The  raw  sequence  data  was  assembled  using  GS  De  Novo  26  

Assembler  v2.8.  The  sequence  reads  were  also  compared  with  Geobacillus  kaustophilus  27  

HTA426  reference  genome  using  the  multiple  genome  alignment  program  Mauve  (Darling  et  28  

al.  2010).  Genes  were  annotated  by  xBASE  bacterial  genome  annotation  service  29  

(http://www.xbase.ac.uk/annotation/).  The  draft  genome  sequence  of  G.  thermoleovorans  30  

B23  is  available  in  DDBJ/EMBL/GenBank  under  the  accession  numbers  BATY01000001–31  

Page 6: Cloning and expression of three ladA-type alkane

 

 

5  

5  

BATY01000209  (Boonmak  et  al.  2013).  The  nucleotide  sequence  between  ORF2  and  ORF11  1  

(11,986  bp),  namely  the  “ladAB  gene  island,”  has  also  been  deposited  in  2  

DDBJ/EMBL/GenBank  under  the  accession  number  AB727923.  The  nucleotide  sequence  for  3  

alkanesulfonate  monooxygenase  gene  cluster  ssuCBAED  has  been  also  deposited  under  4  

AB898656.  5  

 6  

Phylogenetic  tree  analysis  7  

Phylogenetic  trees  were  constructed  by  Kimura’s  two-­‐parameter  correction  (Kimura  1980)  8  

using  the  neighbor-­‐joining  method  (Saitou  and  Nei  1987)  of  MEGA  version  5  (Tamura  et  al.  9  

2011).  Confidences  for  the  phylogenetic  tree  were  estimated  by  bootstrap  analysis  (100  10  

replicates;  Felsenstein  1985).  Homology  search  against  the  protein  database  was  performed  11  

using  protein  BLAST  (http://blast.ncbi.nlm.nih.gov/Blast.cgi;  Altschul  et  al.  1997)  12  

 13  

Cloning  of  ladA  alkane  monooxygenase  gene  homologs  and  construction  of  expression  14  

plasmids  15  

ladAαB23,  ladAβB23,  and  ladBB23  genes  amplified  by  PCR  using  KOD-­‐Plus-­‐Neo  DNA  polymerase  16  

(Toyobo,  Kyoto,  Japan)  in  a  C1000  TouchTM  Thermal  Cycler  (Bio-­‐Rad,  Hercules,  CA),  17  

according  to  the  standard  protocol  from  the  supplier.  The  primers  used  for  PCR  18  

amplification  of  these  genes  were  designed  based  on  the  genome  sequence  of  the  strain  19  

B23,  and  the  restriction  sites  for  NdeI  (CATATG)  and  SmaI  (CCCGGG)  were  introduced  in  the  20  

primers  to  facilitate  cloning.  The  primer  sequences  are  as  follows:  αF  21  

(AATTCCATATGGATCAACCCCTATTATTC)  and  αR  22  

(TAGGGCCCGGGCTAGCCATCTGTTCGTGAGC)  for  ladAαB23,  βF  23  

(AATGGCATATGAGTGTGCGGCAAATG)  and  βR  (GATTCCCGGGTCACACGCGGATCGACTT)  for  24  

ladAβB23,  and  BF  (GGAAACATATGGTTGAATTTATTACGATGG)  and  BR  25  

(CGACCCGGGTTAGATCCATACTTCCGTTTGT)  for  ladBB23.  The  sequences  of  the  genes  were  26  

confirmed  by  first  cloning  the  PCR  products  obtained  for  the  target  genes  ladAαB23,  ladAβB23,  27  

and  ladBB23,  into  the  SmaI  site  of  pUC19,  followed  by  transformation  of  E.  coli  DH5α  using  28  

the  heat  shock  method  (Inoue  et  al.  1990).  The  nucleotide  sequence  of  the  clones  obtained  29  

was  then  determined  using  the  BigDye  Terminator  Cycle  Sequencing  Kit  v3.1  on  an  ABI  3100  30  

DNA  sequencer  (Perkin-­‐Elmer  Applied  Biosystems,  Wellesley,  MA),  and  the  sequences  were  31  

Page 7: Cloning and expression of three ladA-type alkane

 

 

6  

6  

confirmed  to  be  correct.  ladAαB23,  ladAβB23,  and  ladBB23  were  then  separately  subcloned  as  1  

NdeI-­‐SmaI  fragments  into  the  corresponding  sites  of  pCom8  downstream  of  the  alkB  2  

promoter  and  used  for  the  transformation  of  E.  coli  MegaX  DH10BTM  T1R  ElectrocompTM  cells  3  

according  to  the  standard  protocol.  The  resultant  clones  of  pCom8:ladAαB23,  4  

pCom8:ladAβB23,  or  pCom8:ladBB23  were  then  used  for  the  transformation  of  P.  fluorescens  5  

KOB2Δ1  (Højberg  et  al.  1999).  6  

 7  

Heterologous  gene  expression  of  ladAαB23,  ladAβB23,  and  ladBB23  in  P.  fluorescens  KOB2Δ1  8  

In  vivo  growth  complementation  test  of  P.  fluorescens  KOB2Δ1  recombinants  containing  9  

only  pCom8  (as  negative  control),  pCom8:ladAαB23,  pCom8:ladAβB23,  and  pCom8:ladBB23  was  10  

performed  in  5  ml  of  E2  minimal  medium  (Lageveen  et  al.  1988)  supplemented  with  1%  11  

(v/v)  dodecane  (C12)  or  hexadecane  (C16),  0.01%  Tween  80,  and  2%  (v/v)  pre-­‐culture  in  E2  12  

medium.  The  cultures  were  incubated  under  shaking  conditions  (120  rpm)  at  either  30  or  13  

35  °C  for  8  days.  The  experiments  were  performed  in  triplicate.  Turbidimetry  was  not  useful  14  

for  measuring  cell  growth  due  to  high  background  values  because  of  the  formation  of  oil–15  

water  emulsion  in  the  cultures.  The  growth  of  P.  fluorescens  KOB2Δ1  recombinants  was  16  

therefore  determined  by  total  protein  measurement  using  the  Pierce®BCA  Protein  Assay  Kit  17  

(Thermo  Scientific,  Waltham,  MA).    18  

Alkane  degradation  test  was  performed  in  triplicate  in  5  ml  E2  minimal  medium  19  

supplemented  with  1%  (v/v)  standard  gas  oil  (Tokyo  Kasei,  Tokyo,  Japan)  and  0.01%  Tween  20  

80  in  50  ml  screw  cap  centrifuge  tubes  (Catalog  no.  18422CTF50;  IWAKI,  Tokyo,  Japan).  The  21  

cultures  were  incubated  under  shaking  conditions  (120  rpm)  at  35  °C  for  8  days.  The  entire  22  

culture  volume  (5  ml)  was  extracted  with  5  ml  of  acetone:  hexane  (1:1).  After  centrifugation  23  

at  30,000  xg  for  20  min,  the  amount  of  alkanes  remaining  in  the  solvent  layer  was  directly  24  

analyzed  by  gas  chromatography.  A  portion  (1  µl,  split  ratio  50:1)  of  the  sample  was  25  

analyzed  by  GC  system  HP6890  (Agilent,  Palo  Alto,  CA)  with  a  340  µm  x  30  m  non-­‐polar  26  

capillary  HP-­‐1  column  and  FID  detector  (GC/FID)  or  JEOL  JMS-­‐DX303  mass  spectrometer  27  

(JEOL,  Tokyo,  Japan).  Helium  was  used  as  carrier  gas  at  a  flow  rate  of  25  ml/min.    28  

 29  

Results  and  discussion  30  

 31  

Page 8: Cloning and expression of three ladA-type alkane

 

 

7  

7  

Whole  genome  sequence  analysis  of  G.  thermoleovorans  B23  1  

We  had  tried  all  the  best  to  amplify  alkB-­‐type  genes  by  PCR  utilizing  template  genomic  DNA  2  

from  G.  thermoleovorans  B23  with  several  combinations  of  degenerate  and  single  sequence  3  

primers  for  alkB  in  Geobacillus  bacteria  (Tourova  et  al.  2008).  However,  no  alkB  gene  4  

homolog  was  obtained.    In  order  to  identify  the  genes  responsible  for  the  initial  oxidation  of  5  

alkanes  in  G.  thermoleovorans  B23,  whole  genome  sequence  analysis  was  conducted.  The  6  

genome  sequencing  generated  210,521  raw  reads  covering  a  total  of  87,444,156  bp,  7  

achieving  approximately  >25x  genome  coverage.  Q40  plus  bases  were  99.94%  reliable.  The  8  

total  G+C  mol%  was  52.3%,  which  is  the  same  in  G.  thermoleovorans  CCB_US3_UF5  9  

(NC_016593),  and  is  equivalent  to  the  value  of  52.09%  in  G.  kaustophilus  HTA426  10  

(NC_006510).  The  raw  sequence  data  was  assembled  into  209  contigs  (total  size  3,353,053  11  

bp).  An  analysis  of  16S  rRNA  and  recN  gene  sequences  of  the  strain  B23  confirmed  that  this  12  

strain  belonged  to  G.  thermoleovorans,  with  the  two  genes  showing  99.7%  and  100%  13  

identity,  respectively,  with  G.  thermoleovorans  DSM  5366T.    14  

 15  

ladA  alkane  monooxygenase  homologs  in  G.  thermoleovorans  B23  and  related  Geobacillus  16  

strains  17  

G.  kaustophilus  HTA426  does  not  degrade  alkanes  and  does  not  contain  genes  such  as  alkB  18  

that  encode  alkane  monooxygenases.  To  identify  candidate  genes  encoding  alkane  19  

monooxygenases  in  the  strain  B23,  we  first  subtracted  the  commonly  shared  orthologous  20  

gene  regions  from  the  genome  sequence  of  strain  B23  by  utilizing  G.  kaustophilus  HTA426  21  

genome  data.  Surprisingly,  the  resulting  DNA  regions,  which  were  unique  to  the  strain  B23,  22  

did  not  have  any  alkB-­‐like  alkane  monooxygenases,  unlike  the  strain  G.  thermoleovorans  23  

DSM  5366T  and  almost  all  the  other  species  in  the  genus  Geobacillus.  We  then  carefully  24  

analyzed  the  DNA  regions  unique  to  strain  B23  by  Blast  25  

(http://blast.ncbi.nlm.nih.gov/Blast.cgi)  and  found  that  it  carries  three  possible  ladA-­‐type  26  

alkane  monooxygenase  genes,  namely  ladAαB23,  ladAβB23,  and  ladBB23,  in  the  11,834  bp  27  

region  named  the  “ladAB  gene  island.”  The  three  genes  encode  proteins  of  463  (LadAαB23  28  

[BAM76377]),  437  (LadAβB23    [BAM76372]),  and  372  (LadBB23  [BAM76371])  amino  acids,  29  

respectively.  The  amino  acid  sequences  of  the  three  proteins  showed  significant  similarity  30  

(49.8%,  34.4%,  and  22.7%  identical,  respectively;  Fig.  1),  with  the  FMN-­‐dependent  alkane  31  

monooxygenase,  LadA  [YP_001127577,  440  amino  acids],  whose  gene  was  located  in  the  32  

Page 9: Cloning and expression of three ladA-type alkane

 

 

8  

8  

plasmid  pLW1071  of  Geobacillus  thermodenitrificans  NG80-­‐2  (Feng  et  al.  2007;  Li  et  al.  1  

2008).    2  

It  is  generally  accepted  that  the  proteins  whose  mutual  sequence  identity  over  30%  3  

belong  to  the  same  enzyme  group.  Above  a  cut-­‐off  roughly  corresponding  to  30%  sequence  4  

identity,  90%  of  the  pairs  are  homologous;  below  25%  less  than  10%  are  homologous  (Rost  5  

1999).  Two  proteins  share  50%  or  higher  sequence  identity,  their  backbones  differ  by  less  6  

than  1  Å  RMS  deviation.  We  thus  propose  the  name  LadAαB23  and  LadAβB23  for  apparent  7  

LadA  homologues.  While  LadBB23  shares  less  identity  (22.7%)  with  LadA  and  even  higher  8  

similarity  with  putative  alkanesulfonate  monooxygenases  (SsuD)  of  G.  thermoleovorans  9  

CCB_US3_UF5  [YP_004983996]  and  well-­‐studied  SsuD  of  Escherichia  coli  K-­‐12  [P80645],  10  

100%  and  39.9%  respectively.  SsuD  is  a  member  of  bacterial  luciferase  family  same  as  LadA-­‐11  

alkane  monooxygenase  based  on  structural  properties.  It  differs  from  LadA  in  catalytic  12  

activity  that  involves  in  C-­‐S  bond  cleavage,  not  the  hydroxylation  of  a  long-­‐chain  alkane  (Ellis  13  

2010).  Highly  conserved  His12  in  SsuD,  which  is  suggested  to  stabilize  negatively  charged  14  

sulfo  group  of  substrate  alkanesulfonate  binding,  is  replaced  by  neutral  Ser11  in  all  LadB  15  

alkane  monooxygenases.  Interestingly,  at  another  locus  on  chromosome  of  B23,  there  is  16  

ssuCBAED  gene  cluster  [AB898656]  similar  to  ssuEADCB  gene  cluster  of  E.  coli  which  is  17  

required  for  utilization  of  aliphatic  sulfonates  (van  der  Ploeg  et  al.  1999).  The  SsuD  encoded  18  

in  this  gene  cluster  shares  63.8%  identity  with  E.  coli  SsuD  [P80645]  and  completely  19  

conserves  all  the  active  site  residues  (Eichhorn  et  al.  2002).  A  phylogenetic  tree  based  on  20  

amino  acid  sequences  clearly  showed  the  position  of  LadBB23  in  a  unique  evolutionary  21  

lineage  distinct  from  SsuD  clade  (Fig.  2).  Above  knowledge  allowed  us  to  propose  LadBB23  22  

representing  a  novel  alkane  monooxygenase  domain  group  that  is  independent  of  LadA  and  23  

SsuD  clade.    24  

Genome  alignment  analysis  of  the  relevant  Geobacillus  strains  revealed  that  the  25  

gene  organization  in  the  “ladAB  gene  island”  of  strain  B23  was  identical  to  that  of  G.  26  

thermoleovorans  CCB_US3_UF5  and  quite  similar  to  Geobacillus  sp.  G11MC16,  Geobacillus  27  

sp.  Y4.1MC1,  and  Geobacillus  thermoglucosidarius  C56-­‐YS93,  whereas  the  “ladAB  gene  28  

island”  region  was  completely  missing  in  Geobacillus  kaustophilus  HTA426  and  G.  29  

thermodenitrificans  NG80-­‐2  (Fig.  1a).  LadA-­‐type  alkane  monooxygenase  is  a  member  of  the  30  

luciferase  family  of  enzymes,  which  is  classified  as  a  two-­‐component  enzyme  system.  It  31  

comprises  of  (i)  a  monooxygenase  enzyme  that  catalyzes  the  oxidation  of  various  substrates  32  

Page 10: Cloning and expression of three ladA-type alkane

 

 

9  

9  

such  as  luciferin,  aromatic  and  polycyclic  compounds,  long-­‐chain  alkanes,  and  sulfonated  1  

compounds  and  (ii)  an  NADPH-­‐dependent  FMN  reductase.  The  FMN  reductase  provides  2  

reduced  riboflavin  molecule  as  a  prosthetic  group  of  the  monooxygenase  enzyme  to  3  

activate  an  oxygen  molecule  for  the  oxidation  reaction  (Li  et  al.  2008;  Ellis  2010;  Morikawa  4  

2010).  We  found  an  FMN  reductase  (ORF10)  gene  located  upstream  of  the  “ladAB  gene  5  

island,”  as  shown  in  Fig.  1.  6  

 7  

Distribution  of  alkB  and  ladA  gene  homologs  among  strains  belonging  to  the  genus  8  

Geobacillus  9  

Figure  3  shows  a  phylogenetic  tree  of  the  alkane-­‐degrading  Geobacillus  strains  based  on  10  

their  16S  rRNA  gene  sequences.  It  is  interesting  to  note  that  there  is  no  clear  relationship  11  

between  the  species  harboring  alkB  and  ladA.  For  example,  G.  thermoleovorans  strains  DSM  12  

5366T  and  vw3-­‐1n  harbor  six  alkB  homologs  (Tourova  et  al.  2008),  whereas  the  G.  13  

thermoleovorans  strains  B23  and  CCB_US3_UF5  harbor  three  ladA  and  no  alkB  homologs,  14  

suggesting  that  alkane  monooxygenase  genes  are  not  species-­‐specific  but  strain-­‐specific.  No  15  

strain  in  the  genus  Geobacillus  has  ever  been  reported  to  harbor  both  alkB  and  ladA  alkane  16  

monooxygenase  genes.  There  are  variations  in  the  copy  number  of  alkB  homologs  among  17  

various  Geobacillus  strains  and  other  bacteria,  but  the  copy  number  of  chromosome-­‐18  

encoded  ladA  homologs  is  three,  as  reported  till  date  (Fig.  1).  It  is  proposed  that  the  19  

multiple  alkB  homologs  found  in  Geobacillus  are  derived  from  Rhodococcus  bacteria,  based  20  

on  their  significantly  high  sequence  similarities  and  similar  G+C  mol%  (Tourova  et  al.  2008).    21  

G.  thermodenitrificans  NG80-­‐2  maintains  a  single  copy  of  ladA  on  the  plasmid  22  

pLW1071;  the  presence  of  transposase  genes  on  the  plasmid  pLW1071  suggests  that  ladA  is  23  

an  alien  gene  (Fig.  1b;  Feng  et  al.  2007).  It  should  be  worth  to  note  that  the  G+C  mol%  of  the  24  

ladA  is  unusually  low  34.8%  when  compared  with  that  of  the  NG80-­‐2  whole  genome  49.0%.  25  

In  contrast,  G+C  mol%  of  ladAαB23,  ladAβB23,  ladBB23,  and  the  whole  genome  of  B23  are  26  

52.9%,  53.0%,  53.3%,  and  52.3%,  respectively.  It  is  probable  that  the  ‘ladAB  gene  island’  is  27  

originated  and  evolved  in  Geobacillus  bacteria.  Recently,  a  LadA  protein  homologue  28  

[YP_004495128]  has  been  reported  from  a  mesophilic  halotolerant  bacterium,  29  

Amycolicicoccus  subflavus  DQS-­‐9A1  (Nie  et  al.  2013).  This  enzyme  shares  significant  amino  30  

acid  sequence  similarity  with  the  LadA  in  G.  thermodenitrificans  NG80-­‐2  (51.5%  identity),  31  

suggesting  an  evolutionary  close  relationship.    32  

Page 11: Cloning and expression of three ladA-type alkane

 

 

10  

10  

 1  

Heterologous  expression  of  ladAαB23,  ladAβB23,  and  ladBB23  in  P.  fluorescens  KOB2Δ1  2  

Because  targeted  gene  disruption  was  not  successful  in  G.  thermoleovorans  B23,  we  tested  3  

the  alkane  monooxygenase  activities  by  heterologous  gene  expression  method.  The  growth  4  

of  P.  fluorescens  KOB2Δ1  recombinants  containing  either  the  vector  pCom8  alone  (negative  5  

control)  or  pCom8:ladAαB23,  pCom8:ladAβB23,  and  pCom8:ladBB23  was  examined  on  E2  6  

minimal  medium  supplemented  with  1%  (v/v)  of  dodecane  (C12)  or  hexadecane  (C16)  and  7  

0.01%  Tween  80  (Polyoxyethylene  sorbitan  monooleate)  at  30  and  35  °C.  At  30  °C,  which  is  8  

the  optimal  temperature  for  the  growth  of  KOB2Δ1,  the  growth  of  the  strains  harboring  9  

pCom8:ladAαB23,  pCom8:ladAβB23,  or  pCom8:ladBB23  was  slightly  better  on  a  medium  10  

containing  C12  or  C16  when  compared  to  the  strain  harboring  pCom8,  though  the  11  

difference  was  not  significant  (Fig.  4a).  The  growth  of  the  control  strain  (harboring  the  12  

vector  alone)  on  dodecane  and  hexadecane  was  probably  a  result  of  the  degradation  of  the  13  

C5–C16  alkanes  by  the  putative  P450-­‐type  oxygenase  (YP_007999854)  in  the  KOB2Δ1  14  

(CHA0)  strain  (van  Beilen  and  Funhoff  2007).  Since  ladAαB23,  ladAβB23,  and  ladBB23  were  15  

obtained  from  G.  thermoleovorans  B23,  a  thermophilic  bacterium  with  an  optimal  alkane  16  

degradation  temperature  of  70  ˚C  (Kato  et  al.  2001a),  the  low  assay  temperature  of  30  ˚C  17  

may  have  caused  a  structural  change  or  misfolding  of  the  recombinant  LadA  proteins,  18  

resulting  in  their  malfunction.  Therefore,  we  assessed  the  growth  of  the  cells  at  35  °C,  which  19  

is  the  highest  growth  temperature  for  KOB2Δ1.  We  found  that  the  recombinant  strains  20  

harboring  pCom8:ladAβB23  or  pCom8:ladBB23  showed  2.23-­‐  or  1.57-­‐fold  better  growth,  21  

respectively,  than  the  negative  control  (harboring  pCom8)  on  C16  (Fig.  4b).  It  was  also  not  22  

surprising  that  Tween  80  supported  the  growth  of  ladA-­‐homolog-­‐harboring  strains  because  23  

Tween  80  contains  oleyl  chains  in  its  fatty  acid  moiety,  which  could  also  be  a  substrate  of  24  

LadA-­‐type  enzymes.    25  

 26  

Alkane  degradation  activity  27  

Alkane  degradation  experiments  at  35  °C  revealed  that  the  recombinants  harboring  the  28  

ladAαB23,  ladAβB23,  and  ladBB23  degraded  more  C12–C23  alkanes  than  those  harboring  29  

vector  only  (Fig.  5),  with  the  following  alkane  degradation  activity:  pCom8:ladAαB23<  30  

pCom8:ladAβB23  <  pCom8:ladBB23.  No  significant  differences  in  the  substrate  specificity  31  

Page 12: Cloning and expression of three ladA-type alkane

 

 

11  

11  

(carbon  chain  length  of  the  alkane)  were  observed  between  LadAαB23,  LadAβB23,  and  LadBB23  1  

in  mesophilic  P.  fluorescens  KOB2Δ1  recombinants  at  35˚C.  Moreover,  there  was  no  lag  time  2  

for  the  alkane  degradation  activity  (data  not  shown),  suggesting  different  gene  expression  3  

regulation  and/or  different  substrate  alkane  uptake  systems  between  G.  thermoleovorans  4  

B23  and  P.  fluorescens  KOB2Δ1.  It  is  interesting  to  note  that  the  previously  identified  LadA  5  

of  G.  thermodenitrificans  NG80-­‐2  shares  the  highest  sequence  identity  49.8%  with  the  least  6  

active  enzyme,  LadAαB23  (Figs.  1  and  2).    7  

 8  

Conclusions  9  

Unlike  the  FMN-­‐dependent  ladA  alkane  monooxygenase  gene  which  exists  in  G.  10  

thermodenitrificans  NG80-­‐2  as  a  single  copy  on  the  plasmid  pLW1071,  whole  genome  11  

analysis  of  G.  thermoleovorans  B23  revealed  three  ladA  gene  homologs  ladAαB23,  ladAβB23,  12  

and  ladBB23,  which  formed  a  cluster  the  “ladAB  gene  island”  on  the  chromosome.  A  gene  13  

encoding  FMN  reductase,  which  reduces  FMN  at  the  active  site  of  each  LadA  alkane  14  

monooxygenase,  was  also  found  in  the  “ladAB  gene  island.”  Genome  sequence  alignment  15  

between  closely  related  Geobacillus  species  indicated  that  the  “ladAB  gene  island”  was  also  16  

found  on  the  chromosome  of  phylogenetically  distant  Geobacillus  strains,  suggesting  that  it  17  

is  present  universally.  Heterologous  gene  expression  experiments  in  P.  fluorescens  KOB2Δ1  18  

indicated  that  ladAαB23,  ladAβB23,  and  ladBB23  are  the  functional  alkane  monooxygenase  19  

genes  of  G.  thermoleovorans  B23  for  degrading  a  broad  range  of  alkanes.  20  

 21  

Acknowledgments  22  

Pseudomonas  fluorescens  KOB2Δ1  and  pCom8  were  kind  gifts  from  Dr.  Theo  H.  M.  Smits  23  

(Agroscope,  Wädenswil,  Switzerland).  This  work  was  supported  by  KAKENHI  (19380189)  to  24  

M.M.  25  

 26  

References  27  

Altschul  SF,  Madden  TL,  Schäffer  JZ,  Zhang  J,  Zhang  Z,  Miller  W,  Lipman  DJ  (1997)  Gapped  28  

BLAST  and  PSI-­‐BLAST:  a  new  generation  of  protein  database  search  programs.  Nucleic  29  

Acids  Res  25:3389-­‐3402.  30  

Page 13: Cloning and expression of three ladA-type alkane

 

 

12  

12  

Boonmak  C,  Takahashi  Y,  Morikawa  M  (2013)  Draft  Genome  Sequence  of  Geobacillus  1  

thermoleovorans  B23.  Genome  A  in  press.  2  

Darling  AE,  Mau  B,  Perna  NT  (2010)  Progressive  Mauve:  Multiple  Genome  Alignment  with  3  

Gene  Gain,  Loss,  and  Rearrangement.  PLoS  One  5:e11147.    4  

Das  N,  Chandran  P  (2011)  Microbial  degradation  of  hydrocarbon  contaminants:  An  overview.  5  

Biotechnol  Res  Int,  Article  ID  941810  6  

Eichhorn  E,  Davey  CA,  Sargent  DF,  Leisinger  T,  Richmond  J  (2002)  Crystal  structure  of  7  

Escherichia  coli  alkanesulfonate  monooxygenase  SsuD.  J  Mol  Biol  324:  457-­‐468.  8  

Ellis  HR  (2010)  The  FMN-­‐dependent  two-­‐component  monooxygenase  systems.  Arch  9  

Biochem  Biophys  497:  1–12.  10  

Felsenstein  J  (1985)  Confidnece  limits  on  phylogenesis:  An  approach  using  the  bootstrap.  11  

Evolution  39:783-­‐791.  12  

Feng  L,  Wang  W,  Cheng  J,  Ren  Y,  Zhao  G,  Gao  C,  Tang  Y,  Liu  X,  Han  W,  Peng  X,  Liu  R,  Wang  L  13  

(2007)  Genome  and  proteome  of  long-­‐chain  alkane  degrading  Geobacillus  14  

thermodenitrificans  NG80-­‐2  isolated  from  a  deep-­‐subsurface  oil  reservoir.  Proc  Natl  15  

Acad  Sci  USA  104:5602-­‐5607.    16  

Højberg  O,  Schnider  U,  Winteler  HV,  Sørensen  J,  Haas  D  (1999)  Oxygen-­‐sensing  reporter  17  

strain  of  Pseudomonas  fluorescens  for  monitoring  the  distribution  of  low-­‐oxygen  18  

habitats  in  soil.  Appl  Environ  Microbiol  65:4085-­‐4093.  19  

Inoue  H,  Nojima  H,  Okayama  H  (1990)  High  efficiency  transformation  of  Escherichia  coli  with  20  

plasmids.  Gene  96:23-­‐28.  21  

Kato  T,  Haruki  M,  Imanaka  T,  Morikawa  M,  Kanaya  S  (2001a)  Isolation  and  characterization  22  

of  long-­‐chain-­‐alkane  degrading  Bacillus  thermoleovorans  from  deep  subterranean  23  

petroleum  reservoirs.  J  Biosci  Bioeng  91:64-­‐70.  24  

Kato  T,  Miyanaga  A,  Haruki  M,  Imanaka  T,  Morikawa  M,  Kanaya  S  (2001b)  Gene  cloning  of  25  

an  alcohol  dehydrogenase  from  thermophilic  alkane  degrading  Bacillus  26  

thermoleovorans  B23.  J  Biosci  Bioeng  91:100-­‐102.  27  

Kato  T,  Miyanaga  A,  Kanaya  S,  Morikawa  M  (2009)  Alkane  inducible  proteins  in  Geobacillus  28  

thermoleovorans  B23.  BMC  Microbiol  9:60.  29  

Kato  T,  Miyanaga  A,  Kanaya  S,  Morikawa  M  (2010)  Gene  cloning  and  characterization  of  an  30  

aldehyde  dehydrogenase  from  long-­‐chain  alkane-­‐degrading  Geobacillus  31  

thermoleovorans  B23.  Extremophiles  14:33-­‐39.  32  

Page 14: Cloning and expression of three ladA-type alkane

 

 

13  

13  

Kimura  M  (1980)  A  simple  method  for  estimating  evolutionary  rate  of  base  substitutions  1  

through  comparative  studies  of  nucleotide  sequences.  J  Mol  Evol  16:111–120.  2  

Lageveen  RG,  Huisman  GW,  Preusting  H,  Ketelaar  P,  Eggink  G,  Witholt  B  (1988)  Formation  of  3  

polyesters  by  Pseudomonas  oleovorans:  Effect  of  substrates  on  formation  and  4  

composition  of  poly-­‐(R)-­‐  3-­‐hydroxyalkanoates  and  poly-­‐(R)-­‐  3-­‐hydroxyalkenoates.  Appl  5  

Environ  Microbiol  54:2924-­‐2932.  6  

Li  L,  Liu  X,  Yang  W,  Xu  F,  Wang  W,  Feng  L,  Bartlam  M,  Wang  L,  Rao  Z  (2008)  Crystal  structure  7  

of  long-­‐chain  alkane  monooxygenase  (LadA)  in  complex  with  coenzyme  FMN:  8  

unveiling  the  long-­‐chain  alkane  hydroxylase.  J  Mol  Biol  376:453-­‐465.  9  

Marmur  J  (1961)  A  procedure  for  the  isolation  of  deoxyribonucleic  acid  from  10  

microorganisms.  J  Mol  Biol  3:208-­‐218.  11  

Morikawa  M  (2010)  Dioxygen  activation  responsible  for  oxidation  of  aliphatic  and  aromatic  12  

hydrocarbon  compounds:  current  state  and  variants.  Appl  Microbiol  Biotechnol  13  

87:1595–1603.  14  

Nie  Y,  Fang  H,  Li  Y,  Chi  CQ,  Tang  YQ,  Wu  XL  (2013)  The  genome  of  the  moderate  halophile  15  

Amycolicicoccus  subflavus  DQS3-­‐9A1T  reveals  four  alkane  hydroxylation  systems  and  16  

provides  some  clues  on  the  genetic  basis  for  its  adaptation  to  a  petroleum  17  

environment.  PLoS  ONE  8,  e70986.  18  

Rojo  F  (2009)  Degradation  of  alkanes  by  bacteria.  Environ  Microbiol  11:2477-­‐2490.    19  

Rost  B  (1999)  Twilight  zone  of  protein  sequence  alignments.  Prot  Eng  12:85-­‐94.  20  

Saitou  N,  Nei  M.  (1987).  The  neighbor-­‐joining  method:  a  new  method  for  reconstructing  21  

phylogenetic  trees.  Mol  Biol  Evol  4:406–425.  22  

Smits  THM,  Balada  SB,  Witholt  B,  van  Beilen  JB  (2002)  Functional  analysis  of  alkane  23  

hydroxylase  from  Gram-­‐negative  and  Gram-­‐positive  bacteria.  J  Bacteriol  184:1733-­‐24  

1742.    25  

Smits  THM,  Seeger  MA,  Witholt  B,  van  Beilen  JB  (2001)  New  alkane-­‐responsive  expression  26  

vectors  for  Escherichia  coli  and  Pseudomonas.  Plasmid  46:16-­‐24.  27  

Takei  D,  Washio  K,  Morikawa  M  (2008)  Identification  of  alkane  hydroxylase  genes  in  28  

Rhodococcus  sp.  strain  TMP2  that  degrades  a  branched  alkane.  Biotechnol  Lett  29  

30:1447-­‐1452.  30  

Page 15: Cloning and expression of three ladA-type alkane

 

 

14  

14  

Tamura  K,  Peterson  D,  Peterson  N,  Stecher  G,  Nei  M,  Kumar  S  (2011)  MEGA5:  Molecular  1  

Evolutionary  Genetics  Analysis  using  Maximum  Likelihood,  Evolutionary  Distance,  and  2  

Maximum  Parsimony  Methods.  Molecular  Biology  and  Evolution  28:  2731-­‐2739.    3  

Tourova  TP,  Nazina  TZ,  Mikhailova  EM,  Rodionova  TA,  Ekimov  AN,  Mashukova  AV,  Poltaraus  4  

AB  (2008)  alkB  homologs  in  thermophilic  bacteria  of  the  genus  Geobacillus.  Mol  Biol  5  

42:217-­‐226.  6  

van  Beilen  JB,  Funhoff  E  (2007)  Alkane  hydroxylases  involved  in  microbial  alkane  7  

degradation.  Appl  Microbiol  Biotechnol  74:13–21.    8  

van  der  Ploeg  JR,  Iwanicka-­‐Nowickka  R,  Bykowski  T,  Hryniewicz  MM,  Leisinger  T  (1999)  The  9  

Escherichia  coli  ssuEADCB  gene  cluster  is  required  for  the  utilization  of  sulfur  from  10  

aliphatic  sulfonates  and  is  regulated  by  the  transcriptional  activator  CbI.  J  Biol  Chem  11  

274(41):29358-­‐65.  12  

Yanisch-­‐Perron  C,  Vieira  J,  Messing  J  (1985)  Improved  M13  phage  cloning  vectors  and  host  13  

strains:  nucleotide  sequences  of  the  M13mp18  and  pUC19  vectors.  Gene  33:103-­‐109.    14  

 15  

   16  

Page 16: Cloning and expression of three ladA-type alkane

 

 

15  

15  

Fig.  1  a.  Alignment  of  the  gene  organization  of  ladAαB23,  ladAβB23,  and  ladBB23  in  the  1  

chromosome  of  G.  thermoleovorans  B23  with  closely  related  species  from  GenBank;  and  b  2  

ladA  on  plasmid  pLW1071.    3  

ORF1,  2,  5,  6,  ABC  transporter  (permease);  ORF3,  Lysophospholipase-­‐like  protein;  ORF4,  4  

NLPA  lipoprotein;  ORF7,  Acyl-­‐CoA  dehydrogenase  type  2  domain;  ORF8,  12  hypothetical  5  

protein;  ORF9,  LysR  family  transcriptional  regulator;  ORF10,  NADPH-­‐dependent  FMN  6  

reductase;  ORF11,  fructokinases;  ORF13,  TnpR  resolvase;  ORFx/ORFy/ORFz,  hypothetical  7  

protein.  The  region  of  “ladAB  gene  island”  is  shown  in  shaded  box  in  (b).  Amino  acid  8  

sequence  similarity  scores  between  the  ladA  products  of  G.  thermodenitrificans  and  G.  9  

thermoleovorans  B23  are  shown  on  arrows.  Other  similarity  scores  on  arrows  are  between  10  

the  respective  gene  products  of  B23  and  other  species.  11  

 12  

Fig.  2  A  phylogenetic  tree  showing  positions  of  LadAαB23,  LadAβB23,  and  LadBB23  alkane  13  

monooxygenases  relative  to  closely  related  monooxygenases  and  bacterial  luciferase  14  

family  enzymes.    15  

The  phylogenetic  tree  was  constructed  from  evolutionary  distance  data  corrected  by  two-­‐16  

parameter  transformation,  using  the  neighbor-­‐joining  method.  Numbers  indicate  17  

percentages  of  bootstrap  sampling  derived  from  100  samples.  Adf,  F420-­‐dependent  18  

secondary  alcohol  dehydrogenase;  LuxA,  alkanal  monooxygenase;  kMerA,  F420-­‐dependent  19  

tetrahydromethanopterin  reductase;  tMerA,  F420-­‐dependent  tetrahydromethanopterin  20  

reductase  chain  A;  SsuD,  F420-­‐dependent  alkanesulfonate  monooxygenase.  Geobacillus  sp.  21  

WSUCF1  also  contains  a  ladAβ  homolog,  but  the  sequence  data  (WP_020755469)  is  22  

insufficient  to  be  included  in  the  tree.  23  

 24  

Fig.  3  A  phylogenetic  tree  of  Geobacillus  species  based  on  16S  rRNA  gene  sequences.  25  

The  phylogenetic  tree  was  constructed  from  evolutionary  distance  data  corrected  by  two-­‐26  

parameter  transformation  using  the  neighbour-­‐joining  method.  Numbers  indicate  27  

percentages  of  bootstrap  sampling  derived  from  100  samples.  Solid  and  open  arrows  28  

indicate  the  strains  harboring  ladA-­‐type  and  alkB-­‐type  genes,  respectively.  Geobacillus  sp.  29  

[ABV66256]  harboring  ladA  was  not  included  in  the  tree  because  of  the  unavailability  of  its  30  

16S  rRNA  gene  sequence.  31  

 32  

Page 17: Cloning and expression of three ladA-type alkane

 

 

16  

16  

Fig.  4  Complementary  growth  of  P.  fluorescens  KOB2Δ1  recombinant  strains  harboring  1  

pCom8  and  its  derivatives  containing  ladAαB23,  ladAβB23,  or  ladBB23.    2  

Cells  were  grown  on  E2  minimal  medium  supplemented  with  1%  (v/v)  of  n-­‐alkane  (C12  or  3  

C16)  and  0.01%  Tween  80  at  30  °C  (a)  and  35  °C  (b)  for  8  days.  Cell  growth  was  measured  by  4  

measuring  total  protein  concentrations.  *P  <  0.05  compared  with  control  pCom8  only  5  

recombinant.  6  

 7  

Fig.  5  Alkane  degradation  activity  of  P.  fluorescens  KOB2Δ1  recombinant  strains  harboring  8  

pCom8  and  its  derivatives  containing  ladAαB23,  ladAβB23,  or  ladBB23.    9  

Culture  conditions  were  the  same  as  those  in  Fig.  4.  The  amount  of  each  alkane  remaining  in  10  

the  culture  was  measured  by  GC-­‐FID.  Closed  bar,  pCom8  only;  gray  bar,  pCom8:ladAαB23;  11  

open  bar,  pCom8:ladAβB23;  shaded  bar,  pCom8:ladBB23.    *P  <  0.05  compared  with  control  12  

pCom8  only  recombinant.  13  

 14  

 15  

Page 18: Cloning and expression of three ladA-type alkane

For Peer Review

Fig. 1 a. Alignment of the gene organization of ladAαB23, ladAβB23, and ladBB23 in the chromosome of G. thermoleovorans B23 with closely related species from GenBank; and b ladA on plasmid pLW1071.

ORF1, 2, 5, 6, ABC transporter (permease); ORF3, Lysophospholipase-like protein; ORF4, NLPA lipoprotein;

ORF7, Acyl-CoA dehydrogenase type 2 domain; ORF8, 12 hypothetical protein; ORF9, LysR family transcriptional regulator; ORF10, NADPH-dependent FMN reductase; ORF11, fructokinases; ORF13, TnpR

resolvase; ORFx/ORFy/ORFz, hypothetical protein. The region of “ladAB gene island” is shown in shaded box in (b). Amino acid sequence similarity scores between the ladA products of G. thermodenitrificans and G. thermoleovorans B23 are shown on arrows. Other similarity scores on arrows are between the respective

gene products of B23 and other species.

530x376mm (147 x 147 DPI)

Page 21 of 25 Extremophiles

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 19: Cloning and expression of three ladA-type alkane

For Peer Review

Fig. 2 A phylogenetic tree showing positions of LadAαB23, LadAβB23, and LadBB23 alkane monooxygenases relative to closely related monooxygenases and bacterial luciferase family enzymes.

The phylogenetic tree was constructed from evolutionary distance data corrected by two-parameter

transformation, using the neighbor-joining method. Numbers indicate percentages of bootstrap sampling derived from 100 samples. Adf, F420-dependent secondary alcohol dehydrogenase; LuxA, alkanal

monooxygenase; kMerA, F420-dependent tetrahydromethanopterin reductase; tMerA, F420-dependent tetrahydromethanopterin reductase chain A; SsuD, F420-dependent alkanesulfonate monooxygenase. Geobacillus sp. WSUCF1 also contains a ladAβ homolog, but the sequence data (WP_020755469) is

insufficient to be included in the tree.

363x264mm (150 x 150 DPI)

Page 22 of 25Extremophiles

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 20: Cloning and expression of three ladA-type alkane

For Peer Review

Fig. 3 A phylogenetic tree of Geobacillus species based on 16S rRNA gene sequences. The phylogenetic tree was constructed from evolutionary distance data corrected by two-parameter

transformation using the neighbour-joining method. Numbers indicate percentages of bootstrap sampling

derived from 100 samples. Solid and open arrows indicate the strains harboring ladA homologs and alkB homologs, respectively. Geobacillus sp. [ABV66256] harboring ladA was not included in the tree because of

the unavailability of its 16S rRNA gene sequence.

220x179mm (150 x 150 DPI)

Page 23 of 25 Extremophiles

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 21: Cloning and expression of three ladA-type alkane

For Peer Review

Fig. 4 Complementary growth of P. fluorescens KOB2∆1 recombinant strains harboring pCom8 and its derivatives containing ladAαB23, ladAβB23, or ladBB23.

Cells were grown on E2 minimal medium supplemented with 1% (v/v) of n-alkane (C12 or C16) and 0.01%

Tween 80 at 30 °C (a) and 35 °C (b) for 8 days. Cell growth was measured by measuring total protein concentrations. *P < 0.05 compared with control pCom8 only recombinant.

160x173mm (150 x 150 DPI)

Page 24 of 25Extremophiles

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

Page 22: Cloning and expression of three ladA-type alkane

For Peer Review

Fig. 5 Alkane degradation activity of P. fluorescens KOB2∆1 recombinant strains harboring pCom8 and its derivatives containing ladAαB23, ladAβB23, or ladBB23.

Culture conditions were the same as those in Fig. 4. The amount of each alkane remaining in the culture was measured by GC-FID. Closed bar, pCom8 only; gray bar, pCom8:ladAαB23; open bar, pCom8:ladAβB23;

shaded bar, pCom8:ladBB23. *P < 0.05 compared with control pCom8 only recombinant. 204x110mm (150 x 150 DPI)

Page 25 of 25 Extremophiles

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960