curnow SI revised proofs - pnas.org · ! 3!! Figure, S3:!Relationship! between! equilibrium and!...

13
1 SUPPLEMENTARY INFORMATION FOR CURNOW ET AL Figure S1: Equilibrium unfolding curves for alanine mutants in bR helix G. Raw data, with Yaxis showing chromophore absorbance at 560 nm.

Transcript of curnow SI revised proofs - pnas.org · ! 3!! Figure, S3:!Relationship! between! equilibrium and!...

  1  

SUPPLEMENTARY  INFORMATION  FOR  CURNOW  ET  AL  

 

 

 

Figure  S1:  Equilibrium  unfolding  curves  for  alanine  mutants  in  bR  helix  G.  Raw  data,  with  Y-­axis  showing  chromophore  absorbance  at  560  nm.  

 

 

 

  2  

 

Figure  S2:  Kinetic  chevron  plots  for  systematic  mutations  through  helix  G  (a-­‐k)  

and   mutations   introduced   for   double-­‐mutant   cycles   (l-­‐r).   L201A   (a)   is   a  

stabilizing  mutation   and   all   the   others   are   destabilizing.   V210A   (h),   V217A   (i),  

L223A  (k),  D96A  (l),  S226A  (n),  S193A/E204A  (p),  D212A  (q)  and  Y185F/D212A  

(r)  showed  substantial  changes  in  kfH2O  and  mkf.    

  3  

 

Figure   S3:   Relationship   between   equilibrium   and   kinetic   measurements   of  

ΔΔGuH2O  for  Ala  mutations  in  helix  G.    

 

 

  4  

 

 

Figure  S4:  Overall  m-­‐values  derived  from  kinetic  and  equilibrium  data  do  not  agree  well,  probably  because  background  effects  limit  the  accuracy  of  curve  fitting  for  equilibrium  data.  

  5  

 

 

Figure  S5:  Relationship  of  ΔΔGuH2Okin  to  m-­‐values.  ΔΔGuH2Okin  is  not  strongly  correlated  with  mkf  (a)  but  is  approximately  linear  with  mku  (b)  and  muH2Okin  (c).  There  is  a  clear  correlation  between  ΔΔGTS-­‐U  and  β  (d),  a  measure  of  the  global  position  of  the  transition  state  calculated  from  m-­values  (equation  S10).  

  6  

SUPPLEMENTARY  TABLE  1  

  Kinetics   Equilibrium   Φ-­value  

Mutant   Helix  kfH2O  (s-­1)  

 mkf  (kcal.mol-­1)   kuH2O  (s-­1)   mku  (kcal.mol-­1)  

ΔGuH2Okin  (kcal.mol-­1)  

muH2Okin  (kcal.mol-­1)  

ΔΔGuH2Okin  (kcal.mol-­1)  

[D]50%  (cSDS)  muH2Oeqm  

(kcal.mol-­1)  ΔΔGuH2Oeqm  (kcal.mol-­1)  

Φ fkin   Φ feqm   Φ f0.401  

WT   -­‐   0.30  ±  0.06a   2.05  ±  0.2a   6.2e-­‐16  ±  1.6e-­‐16a   23.0  ±  0.3a   19.9  ±  7b   25.1  ±  0.6b   -­‐   0.748  ±0.002c   25.2  ±  1.2c   -­‐   -­‐   -­‐   -­‐  

L201A   G   0.81  ±  0.20   2.86  ±  0.3   5.2e-­‐16  ±  2.0e-­‐16   22.2  ±  0.2   20.6  ±  9   25.1  ±  0.6   +0.7  ±  0.4b   0.773  ±  0.002   34.3  ±  1.2   +0.7  ±  0.03b   Stabilizedd  

N202A   G   0.12  ±  0.04   1.51  ±  0.4   8.4e-­‐16  ±  6.0e-­‐16   22.9  ±  0.4   19.2  ±  15   24.4  ±  1.0   -­‐0.7  ±  0.5   0.732  ±  0.006   32.8  ±  1.8   -­‐0.5  ±  0.02   0.8  ±  0.6b   1.1  ±  0.4   0.9  ±  0.1  

I203A   G   0.15  ±  0.04   1.65  ±  0.4   9.0e-­‐16  ±  3.0e-­‐16   23.1  ±  0.2   19.3  ±  8   24.7  ±  0.7   -­‐0.6  ±  0.2   0.711  ±  0.013   24.2  ±  1.9   -­‐0.9  ±  0.07   0.7  ±  0.3   0.5  ±  0.2   1.2  ±  0.1  

E204A   G   0.29  ±  0.18   2.57  ±  0.7   6.9e-­‐14  ±  5.0e-­‐14   21.7  ±  0.6   17.2  ±  16   24.2  ±  1.6   -­‐2.7  ±  2.6   0.650  ±  0.008   17.8  ±  2.4   -­‐2.0  ±  0.03   0   0   0.3  ±  0.1  

T205A   G   0.31  ±  0.09   2.33  ±  0.3   2.3e-­‐15  ±  9.2e-­‐16   21.9  ±  0.2   19.2  ±  10   24.3  ±  0.6   -­‐0.7  ±  0.3   0.726  ±  0.007   22.6  ±  0.2   -­‐0.5  ±  0.01   0   0   0.2  ±  0.1  

L207A   G   -­‐   -­‐   -­‐   -­‐   -­‐   -­‐   -­‐   Not  two-­‐statee   -­‐   -­‐   -­‐  

F208A   G   0.16  ±  0.05   1.59  ±  0.4   8.0e-­‐14  ±  7.1e-­‐14   20.2  ±  0.6   16.7  ±  16   21.7  ±  1.2   -­‐3.2  ±  3.2   0.640  ±  0.007   35.1  ±  0.3   -­‐3.3  ±  0.03   0.1  ±  0.1   0.1  ±  0.1   0.2  ±  0.1  

M209A   G   0.27  ±  0.06   2.31  ±  0.2   5.3e-­‐15  ±  5.4e-­‐15   21.8  ±  0.6   18.6  ±  19   24.1  ±  1.2   -­‐1.3  ±  1.3   0.700  ±  0.002   28.6  ±  1.0   -­‐1.3  ±  0.05   0.1  ±  0.1   0   0.2  ±  0.1  

V210A   G   0.07  ±  0.02   0.50  ±  0.3   1.7e-­‐15  ±  1.7e-­‐15   22.2  ±  0.7   18.5  ±  19   22.7  ±  1.3   -­‐1.4  ±  1.5   0.724  ±  0.002   23.8  ±  1.4   -­‐0.6  ±  0.04   0.6  ±  0.7   1.4  ±  0.5   0.4  ±  0.1  

S214A   G   0.26  ±  0.06   2.17  ±  0.2   7.6e-­‐16  ±  2.0e-­‐16   22.3  ±  0.2   19.8  ±  7   24.5  ±  0.5   -­‐0.2  ±  0.1   0.739  ±  0.009   21.4  ±  0.4   -­‐0.2  ±  0.004   Not  significantly  destabilized  

V217A   G   0.05  ±  0.01   0.41  ±  0.2   9.6e-­‐16  ±  3.9e-­‐16   22.8  ±  0.2   18.6  ±  8   23.3  ±  0.6   -­‐1.3  ±  0.6   0.697  ±  0.02   19.3  ±  2.5   -­‐1.1  ±  0.2   0.8  ±  0.4   1.0  ±  0.3   0.9  ±  0.1  

F219A   G   -­‐   -­‐   -­‐   -­‐   -­‐   -­‐   -­‐   Not  two-­‐state   -­‐   -­‐   -­‐  

L221A   G   0.26  ±  0.06   1.81  ±  0.2   1.1e-­‐16  ±  3.0e-­‐17   23.5  ±  0.2   20.9  ±  8   25.3  ±  0.5   +1.0  ±  0.4   0.760  ±  0.002   26.2  ±  1.0   +0.3  ±  0.01   Not  significantly  destabilized  

I222A   G   0.14  ±  0.03   1.56  ±  0.3   2.9e-­‐15  ±  1.0e-­‐15   21.8  ±  0.2   18.6  ±  8   23.4  ±  0.6   -­‐1.3  ±  0.5   0.707  ±  0.009   16.8  ±  3.7   -­‐0.9  ±  0.2   0.3  ±  0.2   0.5  ±  0.2   0.4  ±  0.1  

L223A   G   0.09  ±  0.02   0.80  ±  0.2   2.6e-­‐15  ±  1.0e-­‐15   22.2  ±  0.2   18.4  ±  8   23.0  ±  0.6   -­‐1.5  ±  0.7   0.656  ±  0.005   20.9  ±  1.3   -­‐2.1  ±  0.1   0.5  ±  0.3   0.3  ±  0.1   0.5  ±  0.1  

L224A   G   -­‐   -­‐   -­‐   -­‐   -­‐   -­‐   -­‐   Not  two-­‐state   -­‐   -­‐   -­‐  

R225A   G   0.16  ±  0.08   1.03  ±  0.6   6.3e-­‐16  ±  8.8e-­‐16   22.7  ±  0.8   19.6  ±  39   23.7  ±  1.7   -­‐0.3  ±  0.7   0.740  ±  0.001   25.1  ±  0.3   -­‐0.2  ±  0.002   Not  significantly  destabilized  

T170A   F   0.18  ±  0.10   1.21  ±  0.7   2.1e-­‐15  ±  2.3e-­‐15   24.0  ±  0.6   18.9  ±  23   25.2  ±  1   -­‐1  ±  1.2   0.683  ±  0.001   28.8  ±  8.7   -­‐2.5  ±  0.3   0.3  ±  0.4   0.1  ±  0.1   N/Df  

S226A   G   0.10  ±  0.03   0.41  ±  0.4   1.8e-­‐14  ±  1.0e-­‐14   21.1  ±  0.4   17.3  ±  11   21.6  ±  0.5   -­‐2.6  ±  1.6   0.684  ±  0.03   -­‐20.3  ±  2.7   -­‐2.1  ±  0.4   0.3  ±  0.2   0.3  ±  0.1   0  

T170A/  S226A   F/G   0.15  ±  0.10   1.42  ±  0.4   6.8e-­‐15  ±  3.0e-­‐15   21.4  ±  0.3   18.1  ±  15   22.8  ±  0.5   -­‐1.8  ±  1.4   0.660  ±  0.001   -­‐12.9  ±  2.1   -­‐2.4  ±  0.1   0.2  ±  0.2   0.2  ±  0.1   0.2  ±  0.1  

S193A/  E204A   F/G   0.10  ±  0.03   0.95  ±  0.4   1.6e-­‐14  ±  2.0e-­‐14   21.2  ±  1.0   17.4  ±  23   22.2  ±  1   -­‐2.5  ±  3.3   0.663  ±  0.005   -­‐13.8  ±  0.9   -­‐2.4  ±  0.1   0.3  ±  0.3   0.3  ±  0.1   0.2  ±  0.1  

D96A   C   0.13  ±  0.04   0.89  ±  0.4   2.3e-­‐14  ±  1.9e-­‐14   21.4  ±  0.6   17.3  ±  15   22.3  ±  0.7   -­‐2.6  ±  2.3   0.681  ±  0.01   -­‐22.4  ±  0.7   -­‐2.3  ±  0.3   0.2  ±  0.2   0.2  ±  0.1   0.3  ±  0.1  

D212A   G   0.06  ±  0.02   0.55  ±  0.4   6.0e-­‐16  ±  3.0e-­‐16   23.7  ±  0.3   19.0  ±  11   24.2  ±  0.5   -­‐0.9  ±  0.5   0.744  ±  0.004   -­‐21.0  ±  1.9   -­‐0.5  ±  0.1   1.1  ±  0.7   1.9  ±  0.6   N/Df  

Y185F/  D212A  

F/G   0.06  ±  0.02   0.22  ±  0.4   6.8e-­‐16  ±  1.0e-­‐15   23.9  ±  0.9   18.9  ±  1.5   24.1  ±  1   -­‐1.0  ±  1.4   0.705  ±  0.007   -­‐17.83  ±  0.1   -­‐1.5  ±  0.2   1.0  ±  1.5   0.6  ±  0.2   N/Df  

 

aErrors  are  standard  errors  from  curve  fitting.  bErrors  are  propagated  by  quadrature;  errors  in  ΔΔGuH2Okin  are  calculated  from  fractional  errors  in  ΔGuH2Okin  .  cErrors  are  deviation  from  the  mean  average  of  at  least  two  experiments.  dΦ-­‐value  analysis  requires  that  the  mutation  destabilizes  bR  by  at  least  -­‐0.6  kcal.mol-­‐1.  See  text.  eEquilibrium  data  for  L207A,  F219A  and  L224A  did  not  fit  to  the  standard  two-­‐state  equation  (see  text).  fΔΔGU0.401  not  sufficiently  large  for  phi-­‐value  analysis.  Mutants  in  rows  2-­‐18  inclusive  are  from  a  systematic  alanine  scan  of  TM  helix  G;  mutants  in  rows  19-­‐25  inclusive  are  positive  results  from  double-­‐mutant  cycle  analysis.    

  7  

SUPPLEMENTARY  TABLE  2  

Mutant   kfH2O  (s-­1)  mkf  

(kcal.mol-­1)   kuH2O  (s-­1)  

mku  (kcal.mol-­1)  

ΔGuH2Okin  (kcal.mol-­1)  

muH2Okin  (kcal.mol-­1)  

ΔΔGuH2Okin  (kcal.mol-­1)  

ΔGintkin  

(kcal.mol-­1)  [D]50%  (χSDS)  

muH2Oeqm  (kcal.mol-­1)  

ΔΔGuH2Oeqm  (kcal.mol-­1)  

ΔGinteqm  (kcal.mol-­1)

Φ fkin   Φ fintkin   Φ finteqm  

T170A   0.18  ±  0.1   1.21  ±  0.7   2.10e-­‐15  ±  2.3e-­‐15   23.96  ±  0.6   18.9  ±  23.2   25.17  ±  1   -­‐1.0    ±  1.2   -­‐   0.683  ±  

0.001   28.8  ±  8.7   -­‐2.5  ±  0.3   -­‐   0.3  ±  0.4   -­‐   -­‐  

S226A   0.10  ±  0.03   0.41  ±  0.4   1.79e-­‐14  ±  

9.9e-­‐15   21.14  ±  0.4   17.3  ±  10.9   21.55  ±  0.5   -­‐2.6  ±  1.6   -­‐   0.684  ±  0.03     -­‐20.29  ±  2.7   -­‐2.1  ±  0.4   -­‐   0.3  ±  0.2   -­‐   -­‐  

T170A/  S226A   0.15  ±  0.1   1.42  ±  0.4   6.80e-­‐15  ±  

3.0e-­‐15   21.39  ±  0.3   18.1  ±  14.5   22.82  ±  0.5   -­‐1.8  ±  1.4   -­‐1.8  ±  1.4   0.660  ±  0.001   -­‐12.88  ±  2.1   -­‐2.4  ±  0.1   -­‐2.2  ±  0.4   0.2  ±  0.2   0.3  ±  0.2   0.2  ±  0.1  

S193A   0.27  ±  0.1   2.05  ±  0.6   1.98e-­‐16  ±  2.0e-­‐16   23.28  ±  0.8   20.6  ±  22.1   25.33  ±  1   +0.7  ±  0.7   -­‐   0.761  ±  

0.02   -­‐26.69  ±  0.5   -­‐0.1  ±  0.4   -­‐   Not  significantly  destabilized  

E204A   0.29  ±  0.2   2.57  ±  0.7   6.91e-­‐14  ±  5.0e-­‐14   21.64  ±  0.6   17.1  ±  16.3   24.21  ±  1   -­‐2.8  ±  2.6   -­‐   0.668  ±  

0.01     -­‐14.16  ±  4.0   -­‐2.3  ±  0.7   -­‐   0   -­‐   -­‐  

S193A/  E204A  

0.10  ±  0.03   0.95  ±  0.4   1.58e-­‐14  ±  

2.0e-­‐14   21.20  ±  1.0   17.4  ±  22.6   22.15  ±  1   -­‐2.5  ±  3.3   0.5  ±  0.7   0.663  ±  0.005   -­‐13.83  ±  0.9   -­‐2.4  ±  0.1    0   0.3  ±  0.3   -­‐   -­‐  

T46A   0.03  ±  0.02   0.22  ±  0.7   7.9e-­‐15  ±  

3.9  e-­‐14   21.05  ±  3.3   17.2  ±  86     21.30  ±  3.4     -­‐2.7  ±  13   -­‐   0.648  ±  0.006   -­‐13.92  ±  0.1   -­‐2.7  ±  0.2   -­‐   0.5  ±  2.3   -­‐   -­‐  

D96A   0.13  ±  0.04   0.89  ±  0.4   2.3e-­‐14  ±  

1.9e-­‐14   21.38  ±  0.6   17.3  ±  15.3   22.27  ±  0.7   -­‐2.6  ±  2.3   -­‐   0.681  ±  0.01   -­‐22.43  ±  0.7   -­‐2.3  ±  0.3   -­‐   0.2  ±  0.2   -­‐   -­‐  

T46A/  D96A   Unusual  denaturant  response:  large  increase  in  ku   0.664  ±  0.001   -­‐25.66  ±  2.5   -­‐2.9  ±  0.1   -­‐2.1  ±  0.3   -­‐   -­‐   -­‐  

T90A   Unusual  denaturant  response:  large  increase  in  ku   0.700  ±  0.005   -­‐38.66  ±  0.7   -­‐2.3  ±  0.2   -­‐   -­‐   -­‐   -­‐  

D115A   Folding  and  unfolding  data  do  not  correspond     0.812  ±  0.07     -­‐18.32  ±  5.2   1.3  ±  0.1   -­‐   -­‐   -­‐   -­‐  

T90A/  D115A   High  stability  prevents  collection  of  chevron  data   0.846  ±  0.01   -­‐9.37  ±  1.0   1.8  ±  0.2   -­‐2.8  ±  0.2   -­‐   -­‐   -­‐  

Y185F   Shows  apparent  curvature  in  unfolding  chevron  arm   0.749  ±  0.01   -­‐28.26  ±  1.5   -­‐0.4  ±  0.3   -­‐   -­‐   -­‐   -­‐  

D212A   0.06  ±  0.02   0.55  ±  0.4   6.02e-­‐16  ±  

3e-­‐16   23.69  ±  0.3   19.0  ±  11.4   24.24  ±  0.5   -­‐0.9  ±  0.5   -­‐   0.744  ±  0.004   -­‐21.02  ±  1.9   -­‐0.5  ±  0.1   -­‐   1.1  ±  0.7   -­‐   -­‐  

Y185F/  D212A  

0.06  ±  0.02   0.22  ±  0.4   6.8e-­‐16  ±  

1.0e-­‐15   23.90  ±  0.9   18.9  ±  1.5   24.11  ±  1   -­‐1.0  ±  1.4   -­‐   0.705  ±  0.007   -­‐17.83  ±  0.1   -­‐1.5  ±  0.2   0.6  ±  0.5   1.0  ±  1.5   -­‐   -­‐  

  8  

SUPPLEMENTARY  METHODS  

Molecular  biology  

Site-­‐directed  mutations  were  introduced  to  recombinant  WT  bacteriorhodopsin  in  a  plasmid  

originally  constructed  by  Krebs  and  coworkers  [1]  that  can  be  propagated  in  strains  of  both  

Escherichia   coli   and   Halobium   salinarum.   Targeted   single   Ala   mutations   were   introduced  

using  the  Quikchange  method  (Stratagene)  and  the  identity  of  each  mutant  was  confirmed  by  

sequencing.   Each   mutant   was   then   transformed   into   the   bR-­‐deficient  Halobium   strain   L33  

according  to  the  method  of  Cline  and  Doolittle  [2].      

Protein  expression  and  purification  

Mutants   were   prepared   in   the   purple   membrane   form   by   standard   methods   [3]   after  

constitutive  expression   from   the   recombinant  plasmid.    Briefly,   transformed  L33  cells  were  

grown  to  stationary  phase  in  1L  of  high  salt  media  (250  g/L  NaCl,  20  g/L  MgSO4.2H2O,  3  g/L  

Sodium  citrate,  2  g/L  KCl,  10  g/L  Peptone  LP0037  (Oxoid))  in  the  presence  of  4  µg/ml  of  the  

selective   antibiotic   mevinolin   before   being   lysed   under   low   salt   conditions.   The   purple  

membrane  was  then  isolated  by  successive  rounds  of  ultracentrifugation.  Typical  yields  were  

2-­‐3  mg  mutant  protein  per  litre  Halobium  culture.  The  identity  and  purity  of  each  mutant  was  

confirmed  by  mass  spectrometry  [4].  

Equilibrium  thermodynamics  

Equilibrium  unfolding   curves  were   constructed   as   previously   described   [5-­‐7]   using   a   CARY  

UV/Vis  spectrophotometer  (Varian)  at  25  °C.  bR  was  diluted  to  6  µM  in  mixed  lipid/detergent  

micelles   of   15  mM  1,2-­‐Dimyristoyl-­‐sn-­‐Glycero-­‐3-­‐phosphocholine   (DMPC)   and   16  mM  3-­‐[(3-­‐

Cholamidopropyl)dimethylammonio]-­‐1-­‐propanesulfonate   (CHAPS)   in   10   mM   sodium  

phosphate   buffer,   pH   6.0.   These   conditions   give   one   bR   monomer   per   micelle   [8].   Small  

aliquots  of  20%  sodium  dodecylsulfate  (SDS)  were  introduced  to  this  sample  and  SDS-­‐induced  

changes   to   the   folded   state   of   the   protein   were   assessed   by   monitoring   the   chromophore  

absorbance   band   at   560   nm.   The   characteristic   retinal   chromophore   band   is   sensitive   to  

conformational   change   and   disappears   outright   when   bR   is   unfolded.   Unfolding   curves  

generated  from  this  method  exhibit  a  non-­‐linear  background  signal  that  can  complicate  curve-­‐

fitting  procedures.  To  obtain  free  energy  values,  we  adopted  a  similar  approach  to  that  used  

recently  by  Bowie  and  coworkers  [9]and  calculated  changes  in  free  energies  by  determining  

the  midpoint  and  slope  of  the  major  unfolding  transition  from  fitting  to  equation  S1  [10-­‐12]:  

  9  

Y  =  (AF  +  mF*[D])  +  (AU  +  mU*[D])exp(muH2O*([D]-­‐[D]50%)/RT)/  1+  exp(muH2O*([D]-­‐[D]50%)/RT)

                  [Equation  S1]  

Where  AF  and  AU  are  the  absorbance  of  the  folded  and  unfolded  state  respectively  and  mF  and  

mU  are   the   associated   gradients;  muH2O   is   the   slope  of   the  major  unfolding   transition;   [D]   is  

denaturant  concentration;  and  [D]50%  is  the  denaturant  concentration  at  which  the  protein  is  

50%  unfolded.    

Differences   in   m-­‐value   and   [D]50%   between   native   and   mutant   proteins   can   be   used   to  

determine   the   overall   energetic   peturbation   caused   by   the   mutation   at   equilibrium  

(ΔΔGuH2Oeqm)  according  to  equation  S2:  

ΔΔGuH2Oeqm  =  0.5  (muH2Owt  +  muH2Omut)*Δ[D]50%       [Equation  S2]  

This  approach  reduces  errors  associated  with  extrapolation  and  a  particular  advantage  is  the  

high  accuracy  with  which  [D]50%  can  be  determined  (See  Table  2).    

Kinetics  

Kinetics  of  folding  and  unfolding  were  determined  using  a  stopped-­‐flow  instrument  (Applied  

Photophysics)   at   25   °C   by   following   SDS-­‐induced   changes   in   absorbance   at   560   nm   as  

previously  described  [7].  The  major   folding  and  unfolding   transitions  can  be  attributed   to  a  

single  exponential  phase,  giving  classical   two-­‐state  behaviour  under  appropriate  conditions.  

For   unfolding,   the   protein   was   reconstituted   in   DMPC/CHAPS   and   diluted   1:5   into  

SDS/DMPC/CHAPS.  For   refolding,   sequential  mixing  was  used  because   the  unfolded  state   is  

not  stable;  the  covalently  bound  retinal  is  lost  with  t1/2  of  ~13  min  [7]  and  the  resulting  SDS-­‐

unfolded  apoprotein  shows  complex  folding  behaviour  with  multiple  phases  [13,  14].  bR  was  

thus   first   unfolded   for   two   minutes,   in   order   to   reach   an   SDS-­‐unfolded   state   that   retains  

covalently-­‐bound   retinal,   then   refolded   into   concentrated   DMPC/CHAPS   to   reduce   the  

effective  denaturant  concentration  and  refold  the  protein.  

A  classical  kinetic  chevron  plot  was  constructed  by  combining  kinetic  data  from  folding  and  

unfolding  experiments  at  different  concentrations  of  SDS.  The  observed  rate  constant  at  any  

denaturant  concentration,  kobs,  is  the  sum  of  the  folding  and  unfolding  rates:    

kobs  =  kf  +  ku                 [Equation  S3]  

Denaturant  concentration  is  expected  to  be  linear  with  the  logarithm  of  the  rate  of  unfolding,  

giving  equation  S4:  

  10  

ln  ku  =  ln  kuH2O  +  mu*χSDS             [Equation  S4]  

Where  kuH2O  is  the  rate  of  unfolding  in  the  absence  of  denaturant  and  mu  is  the  slope.  The  

equivalent  expression  for  the  rate  of  folding  is:  

ln  kf  =  kfH2O  –  mf*χSDS               [Equation  S5]  

Where   kfH2O   is   the   rate   of   folding   in   the   absence   of   denaturant   and  mf   is   the   associated  

gradient.   Combining   folding   and   unfolding   rate   data   on   a   plot   of   kobs   vs.   χSDS   gives   a  

downward-­‐pointing  arrow,  or  chevron,  that  can  be  fit  to  the  sum  of  equations  S4  and  S5.  This  

gives   the   fundamental   rate  constants  of   folding  and  unfolding   that  are  used   to  calculate   the  

overall  free  energy  of  unfolding  from  kinetics  (ΔGuH2Okin):  

ΔGuH2Okin =  -­‐RT*ln(kfH2O/kuH2O)           [Equation  S6]  

The  overall  m-­‐value,  muH2O,  is  obtained  from:  

muH2O  =  mkf  +  mku               [Equation  S7]  

Where:  

 mkf  =  -­‐RT*mf                 [Equation  S8]    

 mku  =  RT*mu                 [Equation  S9]  

Note  that  mkf  and  mku  are  equivalent  to  mTS-­U  and  mTS-­N,  respectively.  The  m-­‐values  can  be  used  

to  calculate  β,  a  measure  of  the  overall  position  of  the  transition  state  on  the  reaction  

coordinate,  using:  

β  =  mkf  /  muH2O               [Equation  S10]  

Φ-­value  analysis  

Φ-­‐values  were  determined  from  the  ratio  of  ΔΔGTS-­‐U  and  ΔΔGuH2O  according  to  Equation  S11:  

Φfkin  =  ΔΔGTS-­‐U  /  ΔΔGuH2O             [Equation  S11]  

Where  ΔΔGTS-­‐U  is  calculated  from  the  intrinsic  folding  rates  for  wild-­‐type  and  mutant  proteins:  

ΔΔGTS-­‐U  =  RT*ln(kfH20mut/  kfH2Owt)           [Equation  S12]  

And  ΔΔGuH2Okin  was  determined  from  kinetic  data:  

  11  

ΔΔGuH2Okin  =  ΔGuH2Okinwt  -­‐  ΔGuH2Okinmut         [Equation  S13]  

Where   mutant   and   WT   data   are   denoted   with   superscript   mut   and   wt,   respectively.  Alternatively,  complementary  Φ-­‐value  calculations  were  performed  substituting  ΔΔGuH2O  from  equilibrium  experiments  (Φfeqm)  or  using  data  at  a  single  denaturant  concentration  of  0.401  χSDS   (Φf0.401)   where   ΔΔGTS-­‐U0.401   is   calculated   from   folding   rate   constants   at   0.401   χSDS   and  ΔΔGu0.401  is  determined  from  linear  extrapolation.  

Double  mutant  cycles  The   double-­‐mutant   cycle   technique   uses   pairwise   mutagenesis   to   provide   information   on  specific   interaction   events   in   a   complex   background   [15].   Some   of   the   mutations   that   are  found  to  destabilize  bR  involve  the  replacement  of  bulky  hydrophobic  amino  acid  sidechains  with  alanine.  This  can  violate  some  of  the  assumptions  underlying  Φ-­‐value  analysis  because  it  introduces  a  cavity  within  the  protein  so  that  complex  solvent  effects,  reorganisation  energies  and   the   potential   for   introducing   new   interactions   in   the   unfolded   state   need   to   be  considered.  

A  double  mutant  cycle  overcomes  these  uncertainties  by  providing  information  on  a  specific  interaction   that   avoids   complications   from   the   protein   or   solvent   background.   In   a   given  protein,  E,  two  residues  that  are  presumed  to  interact  from  structural  data,  X  and  Y,  are  each  mutated  separately  and  together.  This  gives  a  family  of  protein  variants:  E-­‐XY  (wild-­‐type),  E-­‐X  (Y  mutated),  E-­‐Y  (X  mutated)  and  E  (both  X  and  Y  mutated)  as  shown  in  Figure  S5.  

 

 

Figure  S5:  Representation  of  a  double  mutant  cycle.  

 

Calculating  the  difference  in  the  total  free  energy  change  of  unfolding  at  equilibrium  for  each  mutant  versus  wild-­‐type  (ΔΔGU-­‐F)  gives  a  double-­‐mutant  cycle  (Figure  S5)  that  determines  the  interaction  energy  between  X  and  Y,  ΔΔGU-­‐Fint,  according  to:  

ΔΔGU-­‐Fint  =  (ΔΔGU-­‐FE-­‐XY→E-­‐X  +  ΔΔGU-­‐FE-­‐XY→E-­‐Y)  -­‐  ΔΔGU-­‐FE-­‐XY→E         [Equation  S14]    

Similarly,  one  can  calculate  the  interaction  energy  in  the  transition  state  from  kinetic  double  mutant  cycles  by  first  calculating  the  energetic  peturbation  for  each  mutation:  

ΔΔGTS-­‐U  =  RT  ln(kfH2OWT/kfH2OMUT)             [Equation  S15]  

  12  

And  use  this  to  calculate  the  coupling  energy  between  X  and  Y  according  to:  

ΔΔGTS-­‐Uint  =  (ΔΔGTS-­‐UE-­‐XY→E-­‐X  +  ΔΔGTS-­‐UE-­‐XY→E-­‐Y)  -­‐  ΔΔGTS-­‐UE-­‐XY→E   [Equation  S16]      

This  allows  us  to  calculate  a  Φ-­‐value  based  on  specific  interaction  energy,  Φint,  according  to:  

Φint  =  ΔΔGTS-­‐Uint  /  ΔΔGU-­‐Fint             [Equation  S17]    

If  Φint  =  0,  the  native  interaction  energy  is  lost  in  the  transition  state  and  the  two  residues  are  no   longer   in  contact.   If  Φint  =  1   then   the  residues  are   in  native-­‐like  contact   in   the   transition  state.  The  main  advantage  of  Φint  is  that  by  isolating  the  interaction  energy  between  X  and  Y  errors   from   solvent   and   reorganization   effects   are  much   smaller   and   unfolded   state   effects  cancel   out;   the  method   can   thus   unambiguously   resolve  whether  mutations   destabilise   the  protein   by   changes   in   solvent   interactions   as   well   as   through   the   loss   of   structural  interactions.   Importantly,   this   method   is   well-­‐suited   to   proteins   with   partially-­‐structured  unfolded   states   as   for   bR.   Double  mutant   cycles   have   now  been  widely   used   to   dissect   the  transition   state   structure   of   several   small   aqueous   proteins   [16,   17]   and   to   analyse   the  contribution  of  electrostatic  interactions  to  folding  and  stability  [18,  19].  Additionally,  double  mutant  Φ-­‐values  were  used  as  a  benchmark  for  computer  simulations  [20]  of  Barnase  folding.  With  regard  to  integral  membrane  proteins,  double  mutant  cycles  have  recently  been  used  to  examine  the  contributions  of  aromatic  side-­‐chains  [21]  and  hydrogen  bonding  groups  [22]    to  membrane  protein  folding  and  stability  but  have  not  yet  been  applied  in  mechanistic  studies  or  in  the  context  of  transition  state  analysis.  

SUPPLEMENTARY  REFERENCES  

1.   Krebs,  M.P.,  et  al.,  Expression  of  the  bacterioopsin  gene  in  Halobacterium  halobium  using  a  multicopy  plasmid.  Proc.  Natl.  Acad.  Sci.  USA,  1991.  88(3):  p.  859-­‐863.  

2.   Cline,   S.W.   and   F.W.   Doolittle,   Efficient   transfection   of   the   archaebacterium  Halobacterium  halobium.  J.  Bacteriol.,  1987.  169(3):  p.  1341-­‐1344.  

3.   Oesterhelt,   D.   and   W.   Stoeckenius,   Isolation   of   the   cell   membrane   of   Halobacterium  halobium  and  its  fractionation  into  red  and  purple  membrane.  Methods  Enzymol.,  1974.  31:  p.  667-­‐678.  

4.   Whitelegge,   J.P.,   C.B.   Gundersen,   and   K.Y.   Faull,   Electrospray-­ionization   mass  spectroscopy  of   intact   intrinsic  membrane  proteins.   Protein  Science,  1998.  7:   p.   1423-­‐1430.  

5.   Faham,  S.,  et  al.,  Side-­chain  contributions  to  membrane  protein  stability  and  structure.  J.  Mol.  Biol.,  2004.  335(1):  p.  297-­‐305.  

6.   Chen,   G.Q.   and   E.   Gouaux,  Probing   the   folding   and   unfolding   of  wild-­type   and  mutant  forms   of   bacteriorhodopsin   in   micellar   solutions:   evaluation   of   reversible   folding  conditions.  Biochemistry,  1999.  38:  p.  15380-­‐15387.  

7.   Curnow,   P.   and   P.J.   Booth,  Combined   kinetic   and   thermodynamic   analysis   of  α-­helical  membrane   protein   unfolding.   Proc.   Natl.   Acad.   Sci.   USA,   2007.   104(48):   p.   18970-­‐18975.  

8.   Brouillete,  C.G.,  et  al.,  Structure  and  thermal  stability  of  monomeric  bacteriorhodopsin  in  mixed   phospholipid/detergent   mixtures.   Proteins:   Structure,   Function   and   Genetics,  1989.  5:  p.  38-­‐46.  

9.   Joh,  N.H.,  et  al.,  Modest  stabilization  by  most  hydrogen-­bonded  side-­chain  interactions  in  membrane  proteins.  Nature,  2008.  453(7199):  p.  1266-­‐1270.  

10.   Serrano,  L.,  A.  Matouschek,  and  A.R.  Fersht,  The   folding  of  an  enzyme.   III.  Structure  of  the   transition   state   for   unfolding   of   barnase   analysed   by   a   protein   engineering  procedure.  J.  Mol.  Biol.,  1992.  224(3):  p.  805-­‐818.  

  13  

11.   Clarke,  J.  and  A.R.  Fersht,  Engineered  disulfide  bonds  as  probes  of  the  folding  pathway  of  barnase:   increasing   the   stability   of   proteins   against   the   rate   of   denaturation.  Biochemistry,  1993.  32:  p.  4322-­‐4329.  

12.   Fersht,  A.R.,  Structure  and  Mechanism  in  Protein  Science.  1998,  New  York,  NY,  USA:  W.  H.  Freeman  and  Co.  

13.   Booth,   P.J.,   et   al.,   Intermediates   in   the   folding   of   the   membrane   protein  bacteriorhodopsin.  Nat.  Struct.  Biol.,  1995.  2(2):  p.  139-­‐143.  

14.   Booth,  P.J.,  A.  Farooq,  and  S.L.  Flitsch,  Retinal  binding  during  folding  and  assembly  of  the  membrane  protein  bacteriorhodopsin.  Biochemistry,  1996.  35:  p.  5902-­‐5909.  

15.   Horovitz,  A.,  Double-­mutant  cycles:  a  powerful   tool   for  analysing  protein  structure  and  function.  Fold.  Des.,  1996.  1(6):  p.  R121-­‐126.  

16.   Grantcharova,  V.P.,  D.S.  Riddle,   and  D.  Baker,  Long-­range  order   in   the   src   SH3   folding  transition  state.  Proc.  Natl.  Acad.  Sci.  USA,  2000.  97(13):  p.  7084-­‐7089.  

17.   Horovitz,   A.,   L.   Serrano,   and   A.R.   Fersht,  COSMIC   analysis   of   the  major   alpha-­helix   of  barnase  during  folding.  J.  Mol.  Biol.,  1991.  219(1):  p.  5-­‐9.  

18.   Tissot,  A.C.,  S.  Vuilleumier,  and  A.R.  Fersht,  Importance  of  two  buried  salt  bridges  in  the  stability  and  folding  pathway  of  barnase.  Biochemistry,  1996.  35(21):  p.  6786-­‐6794.  

19.   Serrano,   L.,   et   al.,   Estimating   the   contribution   of   engineered   surface   electrostatic  interactions   to   protein   stability   by   using   double   mutant   cycles.   Biochemistry,   1990.  29(40):  p.  9343-­‐9352.  

20.   Salvatella,  X.,  et  al.,  Determination  of  the  folding  transition  states  of  barnase  by  using  ΦI-­value-­restrained  simulations  validated  by  double  mutant  ΦJ-­values.  Proc.  Natl.  Acad.  Sci.  USA,  2005.  102(35):  p.  12389-­‐12394.  

21.   Hong,  H.,  et  al.,  Role  of  aromatic  side  chains  in  the  folding  and  thermodynamic  stability  of  integral  membrane  proteins.  J.  Am.  Chem.  Soc.,  2007.  129:  p.  8320-­‐8327.  

22.   HyunJoong   Joh,   N.,   et   al.,   Modest   stabilization   by   most   hydrogen-­bonded   side-­chain  interactions  in  membrane  proteins.  Nature,  2008.  453(7199):  p.  1266-­‐1270.