biochimica et biophysica acta - ulisboaweb.tecnico.ulisboa.pt/jgmartinho/pdf/2008...nitrophenyl...

9
Biochemical and structural characterisation of cutinase mutants in the presence of the anionic surfactant AOT V. Brissos a, , E.P. Melo a,b , J.M.G. Martinho c , J.M.S. Cabral a a IBB Institute for Biotechnology and Bioengineering, Centre for Biological and Chemical Engineering, Instituto Superior Técnico, Av. Rovisco Pais, 1049-001 Lisboa, Portugal b Institute for Biotechnology and Bioengineering, Centre for Molecular and Structural Biomedicine, Campus de Gambelas, 8005-139 Faro, Portugal c Centro de Química-Física Molecular, Instituto Superior Técnico, Av. Rovisco Pais, 1049-001 Lisboa, Portugal abstract article info Article history: Received 12 February 2008 Received in revised form 8 April 2008 Accepted 22 April 2008 Available online 4 May 2008 Keywords: Cutinase AOT Kinetics Stability Protein folding The reactivity, stability and unfolding of wild-type (WT) Fusarium solani pisi cutinase and L153Q, S54D and T179C variants were studied in the absence and presence of the dioctyl sulfosuccinate sodium salt (AOT) surfactant. In the absence of surfactant the S54D variant catalytic activity is similar to that of the WT cutinase, whereas L153Q and T179C variants show a lower activity. AOTaddition induces an activity reduction for WT cutinase and its variants, although for low AOT concentrations a small increase of activity was observed for S54D and T179C. The enzyme deactivation in the presence of 0.5 mM AOT is relatively slow for the S54D and T179C variants when compared to wild-type cutinase and L153Q variant. These results were correlated with secondary and tertiary structure changes assessed by the CD spectrum and uorescence of the single tryptophan and the six tyrosine residues. The WT cutinase and S54D variant have similar secondary and tertiary structures that differ from those of T179C and L153Q variants. L153Q, S54D and T179C mutations prevent the formation of hydrophobic crevices responsible for the unfolding by anionic surfactants, with the consequent decrease of the AOTcutinase interactions. © 2008 Elsevier B.V. All rights reserved. 1. Introduction Cutinases are enzymes whose name derives from their ability to degrade cutin polymers, but they additionally hydrolyse a wide variety of both water-soluble esters and emulsied triacylglycerols [1,2]. Cutinase has the enzymatic capabilities of true esterases and true lipases, being an efcient catalyst both in solution and at waterlipid interfaces. In addition, cutinase is stable under alkaline conditions and its structural integrity does not depend on calcium [3]. These properties make cutinases good candidates to improve the removal of fat-based stains in detergent formulations. However, anionic surfactants present in detergent formulations rapidly inactivate wild-type (WT) Fusarium solani pisi cutinase, which limits their use [1,4,5]. Electrostatic interactions were already identied as being important for the stabilization of cutinase in the presence of the anionic surfactants. For instance, in the presence of lithium dodecyl sulphate, a cutinase variant with a more positively charged surface in position 172, (N172K), was found to be less stable than the WT cutinase, while the introduction of negative charged residues in position 17 and 196 (R17E and R196E) increased the stability of the mutants [6]. This work studies three mutants of the WT F. solani pisi cutinase found by directed evolution to improve the stability of cutinase in the presence of AOT (S54D, L153Q and T179C), schematically represented in Fig. 1 [7,8]. S54D mutant was chosen because in the high-throughput screen- ing it showed the highest stability against dioctyl sulfosuccinate sodium salt (AOT) without a signicant decrease in activity. This residue is located in a α-helix (5163) far from the active site (Fig. 1). The substitution of the serine residue by an aspartic acid is thought to prevent the creation of a large hydrophobic crevice that is observed in the course of unfolding of cutinase due to the separation of the helices encompassing residues 5163 and 191211 [4]. L153Q mutant was selected to study the importance of changing the hydrophobic residue (leucine) by a less hydrophobic amino acid (glutamine) in the loop 151166 (Fig. 1), that was previously identied to be a weak spot regarding the stability of cutinase [4]. However, several mutants were found in this region in the high-throughput screening with a slight stability improvement in the presence of AOT [7,8]. Finally, T179C mutant, located in one of the binding loops (171191) (Fig. 1), was chosen to understand the effect of introducing a cysteine (more space lling residue), near the active centre and next to another cysteine involved in a disulphide bridge (Cys171Cys178), on the protein structure and catalytic activity. The enzymatic activity and the stability of the WT cutinase and the three mutants were measured using the hydrolysis reaction of p- nitrophenylbutyrate (p-NPB) in the absence and presence of several amounts of the anionic surfactant, AOT. The hydrolysis reaction of the WT cutinase and its variants follows MichaelisMenten kinetics, in the Biochimica et Biophysica Acta 1784 (2008) 13261334 Corresponding author. Fax: +351218 419 062. E-mail addresses: [email protected], [email protected] (V. Brissos). 1570-9639/$ see front matter © 2008 Elsevier B.V. All rights reserved. doi:10.1016/j.bbapap.2008.04.017 Contents lists available at ScienceDirect Biochimica et Biophysica Acta journal homepage: www.elsevier.com/locate/bbapap

Upload: others

Post on 17-Jul-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Biochimica et Biophysica Acta - ULisboaweb.tecnico.ulisboa.pt/jgmartinho/PDF/2008...nitrophenyl butyrate (p-NPB) (98%) were purchased from Sigma. 2.3. Enzymatic hydrolysis kinetics

Biochimica et Biophysica Acta 1784 (2008) 1326–1334

Contents lists available at ScienceDirect

Biochimica et Biophysica Acta

j ourna l homepage: www.e lsev ie r.com/ locate /bbapap

Biochemical and structural characterisation of cutinase mutants in the presence ofthe anionic surfactant AOT

V. Brissos a,⁎, E.P. Melo a,b, J.M.G. Martinho c, J.M.S. Cabral a

a IBB — Institute for Biotechnology and Bioengineering, Centre for Biological and Chemical Engineering, Instituto Superior Técnico, Av. Rovisco Pais, 1049-001 Lisboa, Portugalb Institute for Biotechnology and Bioengineering, Centre for Molecular and Structural Biomedicine, Campus de Gambelas, 8005-139 Faro, Portugalc Centro de Química-Física Molecular, Instituto Superior Técnico, Av. Rovisco Pais, 1049-001 Lisboa, Portugal

⁎ Corresponding author. Fax: +351 218 419 062.E-mail addresses: [email protected], vaniabrisso

1570-9639/$ – see front matter © 2008 Elsevier B.V. Aldoi:10.1016/j.bbapap.2008.04.017

a b s t r a c t

a r t i c l e i n f o

Article history:

The reactivity, stability and Received 12 February 2008Received in revised form 8 April 2008Accepted 22 April 2008Available online 4 May 2008

Keywords:CutinaseAOTKineticsStabilityProtein folding

unfolding of wild-type (WT) Fusarium solani pisi cutinase and L153Q, S54D andT179C variants were studied in the absence and presence of the dioctyl sulfosuccinate sodium salt (AOT)surfactant. In the absence of surfactant the S54D variant catalytic activity is similar to that of the WT cutinase,whereas L153Q and T179C variants show a lower activity. AOT addition induces an activity reduction for WTcutinase and its variants, although for low AOT concentrations a small increase of activity was observed forS54D and T179C. The enzyme deactivation in the presence of 0.5 mM AOT is relatively slow for the S54D andT179C variants when compared to wild-type cutinase and L153Q variant. These results were correlated withsecondary and tertiary structure changes assessed by the CD spectrum and fluorescence of the singletryptophan and the six tyrosine residues. The WT cutinase and S54D variant have similar secondary andtertiary structures that differ from those of T179C and L153Q variants. L153Q, S54D and T179C mutationsprevent the formation of hydrophobic crevices responsible for the unfolding by anionic surfactants, with theconsequent decrease of the AOT–cutinase interactions.

© 2008 Elsevier B.V. All rights reserved.

1. Introduction

Cutinases are enzymes whose name derives from their ability todegrade cutin polymers, but they additionally hydrolyse awide varietyof both water-soluble esters and emulsified triacylglycerols [1,2].Cutinase has the enzymatic capabilities of true esterases and truelipases, being an efficient catalyst both in solution and at water–lipidinterfaces. In addition, cutinase is stable under alkaline conditions andits structural integrity does not depend on calcium [3]. Theseproperties make cutinases good candidates to improve the removalof fat-based stains in detergent formulations. However, anionicsurfactants present in detergent formulations rapidly inactivatewild-type (WT) Fusarium solani pisi cutinase, which limits their use[1,4,5]. Electrostatic interactions were already identified as beingimportant for the stabilization of cutinase in the presence of theanionic surfactants. For instance, in the presence of lithium dodecylsulphate, a cutinase variant with a more positively charged surface inposition 172, (N172K), was found to be less stable than the WTcutinase, while the introduction of negative charged residues inposition 17 and 196 (R17E and R196E) increased the stability of themutants [6].

This work studies three mutants of the WT F. solani pisi cutinasefound by directed evolution to improve the stability of cutinase in the

[email protected] (V. Brissos).

l rights reserved.

presence of AOT (S54D, L153Q and T179C), schematically representedin Fig. 1 [7,8].

S54D mutant was chosen because in the high-throughput screen-ing it showed the highest stability against dioctyl sulfosuccinatesodium salt (AOT) without a significant decrease in activity. Thisresidue is located in a α-helix (51–63) far from the active site (Fig. 1).The substitution of the serine residue by an aspartic acid is thought toprevent the creation of a large hydrophobic crevice that is observed inthe course of unfolding of cutinase due to the separation of the helicesencompassing residues 51–63 and 191–211 [4]. L153Q mutant wasselected to study the importance of changing the hydrophobic residue(leucine) by a less hydrophobic amino acid (glutamine) in the loop151–166 (Fig. 1), that was previously identified to be a weak spotregarding the stability of cutinase [4]. However, several mutants werefound in this region in the high-throughput screening with a slightstability improvement in the presence of AOT [7,8]. Finally, T179Cmutant, located in one of the binding loops (171–191) (Fig. 1), waschosen to understand the effect of introducing a cysteine (more spacefilling residue), near the active centre and next to another cysteineinvolved in a disulphide bridge (Cys171–Cys178), on the proteinstructure and catalytic activity.

The enzymatic activity and the stability of theWTcutinase and thethree mutants were measured using the hydrolysis reaction of p-nitrophenylbutyrate (p-NPB) in the absence and presence of severalamounts of the anionic surfactant, AOT. The hydrolysis reaction of theWTcutinase and its variants followsMichaelis–Menten kinetics, in the

Page 2: Biochimica et Biophysica Acta - ULisboaweb.tecnico.ulisboa.pt/jgmartinho/PDF/2008...nitrophenyl butyrate (p-NPB) (98%) were purchased from Sigma. 2.3. Enzymatic hydrolysis kinetics

Fig. 1. Ribbon structure of F. solani pisi cutinase [42] displayed using Swiss pdb viewer.The residues of the catalytic triad (Ser120, yellow; His188, blue; Asp175, red), arerepresented in CPK. The residues S54, W69, L153 and T179 are in black.

1327V. Brissos et al. / Biochimica et Biophysica Acta 1784 (2008) 1326–1334

absence and presence of AOT. The catalytic activity of the enzymes andthe inhibition effect of AOT were correlated with structural changes ofthe protein, assessed by both the intrinsic fluorescence of the singletryptophan (Trp69) and the six tyrosine residues (Tyr38, 77, 119, 149,162 and 191) distributed along the backbone and by far-UV circulardichroism (CD) spectra. The creation of hydrophobic patches inducedby the addition of AOT was probed by the fluorescence of 8-anilino-1-naphtalene-sulfonic acid (1,8-ANS). The unfolding induced by guani-dinium hydrochloride (GdnHCl) is identical for theWTcutinase and itsvariants and does not vary with the amount of AOT. The concentra-tions of AOT used are below the critical micellar concentration (CMC)in water (2.5 mM) [9] and the results were interpreted in terms of theAOT/Cutinase molar ratio (MR).

2. Materials and methods

2.1. Protein production and preparation of samples

The over-expression of recombinant cutinase cloned in pMac5-8 (a gift from CorvasIntl.) and its variants (S54D, L153Q and T179C) was performed with Escherichia coliWK6 strain. The variants were obtained by complete site mutagenesis [7,8]. Proteinpurification was carried out through an osmotic shock, acid precipitation, dialysis andtwo sequential anion exchange chromatographic steps, followed by a final dialysis andsubsequently freeze-dried [2]. Cutinase purity was confirmed by 12.5% SDS-PAGEelectrophoresis [10].

Stock solutions of cutinase were prepared by weight and the final proteinconcentration was determined by UV–Vis absorption at 280 nm using the extinctioncoefficient ε280 nm=10,668 M−1 cm−1. Samples for analysis were prepared by dilution ofcutinases and AOT stock solutions. The solutions were prepared in 20 mM Tris–HClbuffer, pH 8.0, previously filtered with 0.2 μm filter discs.

2.2. Chemicals

Tris-(hydroxymethyl)-aminomethane, and acetonitrile (p.a.) were purchased fromMerck. Guanidinium hydrochloride (GdnHCl), 8-anilino-1-naphtalene-sulfonic acid(1,8-ANS) (99%), dioctyl sulfosuccinate sodium salt (AOT) (99%) and the substrate p-nitrophenyl butyrate (p-NPB) (98%) were purchased from Sigma.

2.3. Enzymatic hydrolysis kinetics

The hydrolysis of p-NPB (0.1–0.8 mM) catalysed by WT cutinase and its variantswas studied in the absence and presence of several amounts of AOT. The amount of p-nitrophenol produced during the reaction was calculated by UV absorption at 400 nm(ε=15,400 M−1 cm−1), using a Hitachi U-2000 spectrophotometer.

Using the Michaelis–Menten kinetics scheme with competitive inhibition,

where the reversible complex (ES) is formed between the substrate (S) and enzyme(E) with equilibrium constant, Km. The product (P) results from the dissociation ofthe complex with rate constant kcat. In the presence of the surfactant (I) a competitivebinding was considered by the formation of an inactive complex (EI) withequilibrium constant, KI = [E][I] / [EI].

The kinetic scheme gives for the initial rate the Lineweaver–Burk equation:

v−10 ¼ 1kcat E½ �T

þ K Vmkcat E½ �T S½ �T

ð1Þ

where [E]T and [S]T are the total concentrations of enzyme and substrate, respectively.From the linear plot, kcat and K′m=Km(1+[I] /KI) were calculated from the intercept andslope, respectively. The inhibition constant KI was finally obtained from the slope ratiosof the Lineweaver–Burk plots in the presence and absence of surfactant.

2.4. Stability measurements in the presence of AOT

The inactivation of WT cutinase and its variants by AOT was determined using300 nM cutinase in assay buffer at 40 °C containing 0.5mMof AOT. At several times,15 μlaliquots were removed and diluted 100 fold in assay buffer containing 0.5 mM of sur-factant. The residual activity of the samples was determined by adding 0.7 mM p-NPBand measuring the absorbance of p-nitrophenol at 400 nm over 1 min periods at 30 °C.The assays were made in triplicate. The stability plots were fitted with the Henley andSadana model [11,12] that incorporates two consecutive first-order deactivation steps,

EYk1

E1β1ð Þ

Yk2

E2β2ð Þ

where k1 and k2 are the deactivation rate coefficients and β1 and β2 the relative ratiosof the specific activities of enzyme conformational states, E1 and E2, respectively. Theintegration of the rate equations, considering that each species is weighted by its ownactivity, gives the following expression for the relative activity,

Relative acitivity ¼ 1þ β1k1k2−k1

−β2k2k2−k1

� �e−k1 t−

β1k1k2−k1

−β2k1k2−k1

� �e−k2 t þ β2 ð2Þ

2.5. Circular dichroism

The CD spectra were acquired with a Jasco 720-spectropolarimeter at roomtemperature. Cutinase solutions (5×10−6 M) at pH 8.0 containing several amounts ofAOT were incubated overnight. The far-UV CD spectra (190–300 nm) were recordedunder nitrogen flow in a 1×1 mm square quartz cuvette, setting the experimentalconditions to: 0.5 mm bandwidth; 100 mdeg sensitivity; 0.2 mm resolution; 4 sresponse; 20 nm/min scanning speed. Themean residue ellipticity [Θ] (deg cm2 dmol−1)was calculated from the CD spectra considering 114.8 g mol−1 as the mean residueweight for cutinase. The ellipticity at 222 nm is inversely proportional to the α-helixcontent and was used to follow the loss of secondary structure.

2.6. Fluorescence measurements

The fluorescence spectra were recorded in a SLM-AMINCO spectrofluorimeter(450 W ozone-free Xenon lamp), using a right angle geometry. The excitation light wasselected by a double monochromator (4 nm bandwidth), being the fluorescenceselected by a double monochromator (4 nm bandwidth) and detected by a cooledphotomultiplier (Hammamatsu, R928). Protein solutions (5×10−6 M) were excited at280 nm (excitation of both tryptophan and tyrosine residues) and at 296 nm (forselective excitation of tryptophan residue), and emission was collected between 290–450 nm and 305–450 nm, respectively. The emission of the six tyrosine residues ofcutinase was obtained from the fluorescence spectrum by excitation at 280 nm, aftersubtracting the emission due to the single tryptophan. The fluorescence spectra wererecorded after overnight incubation of the samples. The spectra were corrected for thebackground using the spectrum of the solvent recorded in identical conditions.

Page 3: Biochimica et Biophysica Acta - ULisboaweb.tecnico.ulisboa.pt/jgmartinho/PDF/2008...nitrophenyl butyrate (p-NPB) (98%) were purchased from Sigma. 2.3. Enzymatic hydrolysis kinetics

Table 1Effect of AOT on kinetic parameter values for cutinase and its mutants L153Q, S54D andT179C in aqueous solutions

Protein [AOT] (mM) kcat (104 s−1) K′m(mM) KI (mM)

WT 0 1.56±0.29 0.35±0.02 0.72±0.060.5 1.55±0.10 0.48±0.051 1.74±0.13 0.85±0.011.5 1.52±0.14 1.12±0.122 1.38±0.17 1.08±0.16

S54D 0 1.47±0.13 0.33±0.03 1.38±0.140.5 2.47±0.15 0.74±0.071 2.42±0.27 0.72±0.011.5 2.52±0.15 1.14±0.112 3.21±0.29 1.74±0.75

L153Q 0 1.10±0.77 0.31±0.03 0.44±0.040.5 1.60±0.84 0.69±0.071 1.79±0.22 1.31±0.131.5 2.58±0.18 3.57±0.212 2.85±0.25 4.33±0.41

T179C 0 0.79±0.09 0.20±0.02 0.46±0.050.5 0.96±0.05 0.49±0.051 1.11±0.07 0.66±0.061.5 1.10±0.12 1.09±0.112 1.15±0.12 1.56±0.16

1328 V. Brissos et al. / Biochimica et Biophysica Acta 1784 (2008) 1326–1334

2.7. 1,8-ANS fluorescence

Solutions ofWTcutinase and its variants (5×10−6 M) containing several amounts ofAOT were incubated with 1.25×10−4 M of 1,8-ANS. Fluorescence spectrum of ANS wasrecorded between 420–650 nm, by excitation at 410 nm.

2.8. Unfolding equilibrium studies

Guanidinium hydrochloride unfolding curves were obtained from both theellipticity at 222 nm and the tryptophan fluorescence intensity at the maximum byexcitation at 296 nm. The stock solution containing the surfactant was let to incubatewith the required concentration of AOT for 1.5 h. By dilution, several solutions with aconstant concentration of cutinase (5×10−6 M) and several concentrations (0–3 M) ofGdnHCl and AOT (0–10 MR) were prepared. Samples were allowed to equilibrateovernight prior to the measurements.

The unfolding curves were fitted with a two-state model. The apparent equilibriumconstant of unfolding (Keq) and the free-energy change (ΔG0), were calculated by:

Keq ¼ yf−yð Þ= y−yuð Þ ð3Þ

ΔG0 ¼ −RTln Keq� � ð4Þ

where y is the spectroscopic property (ellipticity or fluorescence intensity) at the givendenaturant concentration and yf and yu are the corresponding values for the folded andunfolded protein states, respectively. These values were extrapolated from the pre- andpost-transition baselines obtained by a linear fit. The ΔG0 varies linearly with theconcentration of denaturant,

ΔG0 ¼ ΔG0 H2Oð Þ−m denaturant½ � ð5Þ

where ΔG0(H2O) is the free-energy change in the absence of denaturant and m is ameasure of the ΔG0 dependence with the denaturant concentration. The midpoint ofdenaturation concentration (Cm) corresponds to the denaturant concentration at whichΔG0=0.

The unfolding curves were fitted with Eq. (6) [13], using a nonlinear least-squaresprocedure,

F ¼ aD þ bDcþ aN þ bNcð Þ10mD−N c−c50kð Þ1þ 10mD–N c−c50kð Þ ð6Þ

where F is the spectroscopic property, c the denaturant concentration, mD–N thesensitivity of the equilibrium constant to unfolding, c50% the denaturant concentrationcorresponding to half unfolded protein, aD, aN the intercepts and bD, bN the slopes ofthe fluorescence baselines at low (N) and high (D) denaturant concentrations beforeand after the transition region, respectively.

3. Results

3.1. Catalysed hydrolysis of p-NPB

After purification, enzymes were analysed by SDS-PAGE and themolar mass found for the WT cutinase and its mutants was identical(~22 KDa).

Fig. 2. Variation of the initial rate (v0) of the p-NPB hydrolysis with the concentration ofAOT: WT (□) L153Q (●), S54D (▲) and T179C (⋄). The reaction was performed in 20 mMTris–HCl (pH 8) containing 3 nM cutinase, 0.7 mM p-NPB and several concentrations ofAOT.

The initial rate of p-NPB hydrolysis catalysed byWTcutinase and itsmutantswas studied as a function of AOTconcentration. The activity inthe absence of AOT is identical for the WT cutinase and S54D variant,butmuch lower for the L153Q and T179Cmutants (Fig. 2). The shape ofthe curves is also different, decreasing with the AOT concentration fortheWTcutinase and T179C variant, while the other twomutants (S54Dand L153Q) displayamaximumresulting froman initial increase at lowconcentrations followed by a continuous decrease. The initial ratesfollow Michaelis–Menten kinetics, being the parameters, in theabsence and presence of AOT (below the CMC), summarized in Table1. In the absence of AOT, the kinetic parameters of the S54Dmutant arepractically equal to those of the WT cutinase, while the catalyticefficiency of the L153Q and T179C mutants are lower. The addition ofAOT leads to Km values higher than Km(1+[I] /KI). The L153Q andT179C mutants (KI=0.44±0.04 mM and KI =0.46±0.05 mM, respec-tively) show higher and the S54D mutant (KI=1.38±0.14 mM) lowerinhibition by AOT than the WT cutinase (KI =0.72±0.06 mM). Thedecrease of the initial rate with the increase of AOT concentrationobserved for the WT cutinase and T179C variant is essentially due tothe strong inhibition of the surfactant because kcat is practicallyinvariant with the AOT concentrations. The shapes of the S54D andL153Q curves are due to two competing effects: the increase of kcat andthe inhibition with increasing AOT concentrations.

Fig. 3. Deactivation profile of cutinase WT (□), L153Q (●), S54D (▲), T179C (⋄) in thepresence of 0.5 mMAOT. The experimental results were fitted (full line) with the Sadanamodel (Eq. (2).

Page 4: Biochimica et Biophysica Acta - ULisboaweb.tecnico.ulisboa.pt/jgmartinho/PDF/2008...nitrophenyl butyrate (p-NPB) (98%) were purchased from Sigma. 2.3. Enzymatic hydrolysis kinetics

Table 2Deactivation rate coefficients (k1 and k2), ratio of specific activities (β1 and β2) andaverage activity over the course of 1 h of the wild-type cutinase and the L153Q, S54Dand T179C mutants in 20 mM Tris–HCl pH 8 containing 0.5 mM AOT

Protein k1 (min−1) k2 (min−1) β1 β2 Average activity(mM s−1)

WT 0.83 0.026 0.82 0 37S54D 0.083 0.004 0.81 0.006 86L153Q 0.025 0 0 0 37T179C 6.85 0.014 1.16 0.011 46

1329V. Brissos et al. / Biochimica et Biophysica Acta 1784 (2008) 1326–1334

3.2. Stability measurements in the presence of AOT

The loss of initial activity of the WT cutinase and its variants uponincubation with 0.5 mM AOT was monitored over the course of 1 h at40 °C (Fig. 3). The stability curves of theWTcutinase and theirmutantsin the absence of AOT are identical, and were used as a control. In thepresence of AOT, the three mutants show different stability profiles.The L153Q mutant stability curve can be fitted with a singleexponential (pseudo-first-order kinetics) with an inactivation rateconstant of k1=0.025 min−1 (Table 2). The deactivation curves of theother variants are more complex and were fitted with the Henley andSadana model [11,12]. A good fit was obtained, as judged by the plot ofthe residuals and the correlation coefficients. The S54D variant decaysto an intermediate state E1 with rate constant k1=0.083 min−1. Thisintermediate has lower activity than the initial enzyme (β1=0.81) anddeactivates with rate constant k2=0.004 min−1 to a final state E2 withresidual specific activity (β2=0.006). The T179C mutant decays withrate constant k1=6.85 min−1 to E1 with specific activity higher thanthe initial enzyme (β1=1.16), followed by the deactivation with rate

Fig. 4. Fluorescence spectra (λexc=280 nm) of 5 μM cutinase WT (A), S54D (B) L153Q (C), andline); MR=100 (dashed line).

constant k2=0.014 min−1 to the final state E2 with very low activity(β2=0.011).

The WT cutinase and L153Q mutant display the same averageactivity of 37mM s−1, while the S54Dmutant has a significantly higheraverage activity of 86 mM s−1 (Table 2). The T179C variant has anaverage activity (46 mM s−1) higher than the WT cutinase, despite thelow initial activity due to the high activity of intermediate E1 and slowdeactivation paths that compensate for the low initial activity of theenzyme.

3.3. Fluorescence and CD measurements

The fluorescence spectra of the WT cutinase and its variants wererecorded in buffer solutions at pH 8.0 (Fig. 4). The fluorescence ofcutinase by excitation at 280 nm is due to both the single tryptophanand the six tyrosine residues. In proteins having both tyrosine andtryptophan aromatic residues the fluorescence of tryptophan gen-erally dominates [14]. This does not occur in the WT cutinase becausethe single tryptophan (Trp69) fluorescence is strongly quenched(Φ=0.01, [15]) by the adjacent disulfide bond between Cys31 andCys109, located approximately 4 Å away from tryptophan [16–19].Thus, the emission of native cutinase is dominated by six tyrosines(Tyr38, 77, 119, 149, 162 and 191) as was observed for the WT cutinaseand S54D, L153Q variants in buffer solution. The fluorescence of thenative T179Cmutant has amajor contribution from tryptophan, whichsuggests that for this mutation the tryptophan is not in the vicinity ofthe disulfide bridge.

The emission of the unfolded proteins (WT and variants) isdominated by the tryptophan fluorescence because upon denatura-tion by GdnHCl, the tryptophan residue is released from the vicinity of

T179C (D) dissolved in 20 mM Tris–HCl, pH 8.0. Native (full line); 3.0 M GdnHCl (dotted

Page 5: Biochimica et Biophysica Acta - ULisboaweb.tecnico.ulisboa.pt/jgmartinho/PDF/2008...nitrophenyl butyrate (p-NPB) (98%) were purchased from Sigma. 2.3. Enzymatic hydrolysis kinetics

Fig. 5. Fluorescence intensity at 335 nm (A) by excitation at λexc=296 nm and 305 nm(B) as a function of MR. Cutinase WT (□), L153Q (●), S54D (▲) and T179C (⋄).

1330 V. Brissos et al. / Biochimica et Biophysica Acta 1784 (2008) 1326–1334

the disulfide bridge and in turn the strong tryptophan quenchingdisappears. By addition of AOT the fluorescence spectra approachthose of the unfolded proteins. At MR ([AOT]/ [protein])=100 the

Fig. 6. Far-UV CD spectra for 5 μM cutinase WT (A), S54D (B), L153Q (C) and T179C (D) solut(dashed line).

spectra of the WT cutinase and L153Q variant are identical (intensityand shape), suggesting that unfolding is occurring. The shape of thespectrum of the T179C variant does not change significantly with AOTaddition, although a large increase of the fluorescence intensity at340 nm was observed. The S54D variant fluorescence spectrum iscomposed of both the tryptophan and tyrosine emissions. Thisindicates that at MR=100 the enzymes still exist in the folded state(fluorescence dominated by tyrosines) however, a fraction of proteinsshow already structural changes (fluorescence dominated by thetryptophan).

In order to further study the influence of AOT, the fluorescenceintensity at the maximum of tryptophan (λexc=296 nm) and tyrosineresidues (obtained from the spectrum by excitation at λexc=280 nmafter subtracting the fluorescence of tryptophan) as function of MRwas studied (Fig. 5A and B respectively). The tryptophan fluorescenceintensity of the WT cutinase and S54D, L153Q variants are identicaluntil MR~40 increasing substantially for higher MR values (Fig. 5A).This indicates that the enzymes begin to denature at this point. Afterthis value the tertiary structure of all the enzymes changes, with theS54D variant being the most resistant to unfolding by AOT. The T179Cvariant has much higher tryptophan fluorescence intensity and variesin a non-monotonic way with the addition of AOT, showing amaximum at MR~10 and a minimum at MR~40, increasing after-wards until reaching a plateau.

The tyrosine fluorescence of the WT cutinase and its variants as afunction of MR is shown in Fig. 5B. In the absence of surfactant theWTcutinase and S54D mutant have the same intensity, while the L153Qand T179C mutants show much smaller values. The tyrosinefluorescence intensity of the WT cutinase slightly increases at lowMR and then decrease. The S54D variant has a similar profile to theWT cutinase at low MR but for higher MR values remains almostconstant with increasing amounts of surfactant. The tyrosine

ions in 20 mM Tris–HCl, pH 8.0. Native (full line); 3.0 M GdnHCl (dotted line); MR=100

Page 6: Biochimica et Biophysica Acta - ULisboaweb.tecnico.ulisboa.pt/jgmartinho/PDF/2008...nitrophenyl butyrate (p-NPB) (98%) were purchased from Sigma. 2.3. Enzymatic hydrolysis kinetics

Fig. 7. Mean residue ellipticity at 222 nm versus the MR: cutinase WT (□), L153Q (●),S54D (▲), T179C (⋄).

Fig. 9. Equilibrium unfolding of WT cutinase by GdnHCl followed by the tryptophanfluorescence intensity (closed symbols) and CD residual ellipticity (open symbols):MR=0 (▪, □), MR=5 (●, ○) and MR=10 (▲, ▵).

Table 3Thermodynamics parameters of wild-type cutinase and its mutants' unfolding inducedby GdnHCl and at several values of MR

Protein MR ΔG0(H2O)(kcal mol−1)

m(kcal mol−1 M−1)

Cm (M)

1331V. Brissos et al. / Biochimica et Biophysica Acta 1784 (2008) 1326–1334

fluorescence of the L153Q and T179C variants does not change withMR, being the intensity of the T179C mutant much lower than that ofL153Q.

The ellipticity of theWTcutinase and its variants is shown in Fig. 6.The far-UV absorption spectra of proteins (typically 240 nm to190 nm) are mainly due to the peptide bond. There is a weak butbroad n→π⁎ transition centred around 210 nm and an intense π→π⁎

transition around 190 nm. The CD spectra in this spectral region can beused to quantify the overall secondary structure content of proteinsnamely by following the 222 nm signal, which is assigned to andconsidered inversely proportional to the α-helix content [20–22]. Theunfolding by GdnHCl leads to the loss of the secondary structure,while by addition of surfactant the secondary structure is almostconserved. This is consistent with earlier studies showing thatproteins denatured by detergents contained a large degree of orderstructure [23–25]. The WT cutinase and S54D variant have muchhigher α-helix content than the L153Q and T179C variants.

In order to further study the influence of AOT in the secondarystructure, the ellipticity at 222 nm of all the proteins was plotted as afunction of MR (Fig. 7). The ellipticity of the WT cutinase and S54Dvariant is identical in the absence of AOT. Upon AOT addition up toMR~40 there is a decrease of the S54Dmutant ellipticity (correspond-ing to an α-helix content increase), to increase afterwards (α-helixcontent decrease) slightly for the S54D variant and more pronounc-edly for the WT cutinase. The α-helix content of the L153Q and T179Cvariants is much lower than that of WT cutinase and S54D variant

Fig. 8. Fluorescence intensity of 1.25×10−4 M solution of 1,8-ANS at 500 nm byexcitation at λexc=410 nm as a function of MR: cutinase WT (□), L153Q (●), S54D (▲),T179C (⋄).

(reduction of 38% and 45%, respectively) and practically does not varywith MR.

3.4. 1,8-ANS binding studies

The fluorescence of 1,8-ANS is very sensitive to its own environ-ment. In water the ANS fluorescence is very weak, while inhydrophobic media it is relatively strong [26–29]. This property wasexplored to monitor the development of hydrophobic patches oncutinase surface induced by the binding/interactions of AOT mole-cules. AOT is expected to promote less rigid packing of hydrophobicclusters upon binding thus allowing greater interactions between theANS and the protein to occur.

The ANS fluorescence intensity at 500 nm is practically constant orslightly increases for all the enzymes until MR=30 (Fig. 8). Byincreasing the AOT concentration, the fluorescence intensity increasesrapidly for the WT cutinase and moderately for its variants. The S54Dmutant shows the smaller increase, followed by the T179C and L153Qvariants.

WT Steady state fluorescence 0 6.7±0.1 4.05±0.07 1.64±0.005 7.7±0.4 4.80±0.23 1.61±0.13

10 7.9±0.9 4.92±0.30 1.61±0.13Far-UV CD 0 10.3±0.4 6.08±0.22 1.69±0.13

5 6.0±0.2 3.60±0.10 1.68±0.0910 7.9±0.1 4.64±0.08 1.71±0.06

S54D Steady state fluorescence 0 8.0±0.2 4.72±0.10 1.69±0.075 7.0±1.3 4.17±0.70 1.67±0.02

10 7.8±0.0 4.76±0.03 1.64±0.00Far-UV CD 0 9.3±0.5 5.51±0.33 1.68±0.13

5 9.1±0.8 5.56±0.48 1.63±0.0910 9.8±0.5 5.95±0.32 1.65±0.06

L153Q Steady state fluorescence 0 6.9±0.1 4.25±0.08 1.62±0.065 7.0±0.2 4.35±0.14 1.61±0.11

10 6.1±0.1 4.05±0.09 1.51±0.07Far-UV CD 0 6.9± 1.6 4.27±0.94 1.62±0.03

5 8.4±0.2 5.40±0.16 1.57±0.0910 5.8±1.3 3.56±0.84 1.63±0.02

T179C Steady state fluorescence 0 7.0±1.5 4.00±0.96 1.75±0.045 6.8±0.1 3.81±0.03 1.77±0.02

10 5.9±0.3 3.50±0.07 1.68±0.03Far-UV CD 0 8.5±0.1 4.82±0.05 1.77±0.04

5 9.2±1.2 5.10±0.71 1.80±0.1310 9.8±0.4 5.54±0.20 1.77±0.13

Page 7: Biochimica et Biophysica Acta - ULisboaweb.tecnico.ulisboa.pt/jgmartinho/PDF/2008...nitrophenyl butyrate (p-NPB) (98%) were purchased from Sigma. 2.3. Enzymatic hydrolysis kinetics

1332 V. Brissos et al. / Biochimica et Biophysica Acta 1784 (2008) 1326–1334

3.5. Unfolding equilibrium studies in GdnHCl

The WT cutinase unfolding induced by GdnHCl, obtained from theellipticity at 222 nm and the tryptophan fluorescence intensity in theabsence of surfactant and for MR=5 and MR=10, is shown in Fig. 9. Thecurves calculated from the CD and fluorescence signals are identicaland the influence of AOT is minor. Similar curves were obtained for thethree variants (results not shown). The shape of the curves suggeststhat unfolding proceeds without intermediates, involving only thefolded and unfolded states. However, ΔG0 (H2O) and m values in theabsence and in the presence of AOT are consistently higher for CD thanfor the fluorescence intensity (Table 3). This can indicate the presenceof an intermediate state, in accordance with a previous work where anunfolding intermediate (molten globule) was identified in the unfold-ing of cutinase by GdnHCl [30].

4. Discussion

In the absence of surfactant, the catalytic performance of the S54Dvariant, whose mutation is located far from the active site, is identicalto that of the WT cutinase. The other variants, L153Q and T179C,whose mutations are near the active centre, have much lower activity(Table 1). The KI values of the L153Q and T179C variants are muchlower (higher inhibition effect) than those of the WT cutinase andS54D mutant indicating that these mutants have a higher affinity toAOT. In the presence of AOT, the S54D variant shows the highest KI anda substantial increase of kcat when compared to WT cutinase. Theeffect of AOT on the catalytic activity is complex and a simplemechanism of inhibition of catalytic activity may not be sufficient tointerpret what was observed experimentally as both activation andinhibition seem to occur. The maximum changes in the activity occurbefore the CMC, suggesting that these changes result from the bindingof monomeric AOT to regions other than the substrate binding site. Itwas also observed that the anionic surfactant sodium dodecylsulphate (SDS) increases the reactivity of active site residuespresumably as a result of localised unfolding [1,31]. Some authors[5] suggested that cutinase interactionswith SDS result in a hyperbolicmixed inhibition behaviour, for surfactant concentrations above CMC.This anionic surfactant has also been shown to have an activationeffect on Aspergillus niger catalase [32], on the tetrametric enzyme β-galactosidase from E. coli [33], on BSA [34] and on Sulfolobussolfataricus β-glycosidase [35]. The catalytic performance of L153Qand T179C decreases in the presence of AOT as was observed for theWT cutinase. The variation of the initial rates with the amount of AOTfor the L153Q variant is similar to that of the WT cutinase, except forlow AOT concentrations where some activation seems to occur, whilefor the T179C variant the partial unfolding occurs at higher AOTconcentrations than those observed for the WT cutinase. Theinactivation evolution with time of the enzymes in the presence of0.5 mM of AOT is faster for theWTcutinase and L153Q variant than forthe S54D and T179Cmutants. This suggests that the interaction of AOTis stronger with the WT cutinase and L153Q than with the S54D andT179C variants. The initial activity of the WT cutinase and S54Dmutant is much higher than that of L153Q and T179C variants and thisis attributed to their different secondary structures. Indeed, the α-helix content of the S54D is similar to that of the WT but much higherthan that observed for the L153Q and T179C variants (Fig. 6). Thepresence of AOT did not have a significant effect on the secondarystructure of L153Q and T179C variants (Fig. 7). For the S54D mutantlow AOTconcentrations seem to induce a gradual increase inα-helicalcontent up to MR=40, which is constant for WT cutinase. For MRN40the α-helical content decreases for the WT cutinase and remainspractically constant for the S54D variant. It was observed that SDS alsopromoted an increase in α-helix content of ovalbumin, carbonicanydrase [36] and of β-glycosidase from the thermophilic archaeon S.solfataricus [37].

The tryptophan fluorescence intensity (Fig. 5A) of the WTcutinase and S54D, L153Q mutants does not change until MR=40.This indicates that the tertiary structure around the singletryptophan located in the opposite site of the active centre doesnot change. For MRN40 the intensity increases substantially,suggesting the beginning of denaturation. The increase is lessprominent for the S54D mutant indicating a stronger resistance ofthis mutant against AOT. Tryptophan fluorescence intensity of theT179C variant is much higher and tyrosine fluorescence lower thanthat of the WT cutinase and the other mutants, i.e. the tryptophansuffers less quenching from the disulphide bridge, without asignificant variation with AOT concentration change. As thetryptophan is located in the opposite side of the mutation (Fig. 1),this suggests that large structural changes occurred. A possibility isthat Cys179, introduced by this mutation, forms an intramoleculardisulphide bridge with Cys171 or Cys178, that are involved in adisulfide bridge in WT cutinase. This could imply a distortion on thestructure and lead to the displacement of the tryptophan from thedisulfide bridge (Cys31–Cys109), and/or the approach of a tyrosinewith the consequent increase in tyrosine–tryptophan energy transferefficiency. The structural conformation changes that probably occurnear the active site correlate well with the decrease of activity of thismutant.

The β-strand (residues 67–73) in which Trp69 is located liesbetween two α-helices (residues 51–63 and 91–108) that have beenidentified, by molecular dynamics simulations, as the lowest mobilesecondary structure elements [4]. It is believed that once the bindingsites are saturated with AOT, the interaction proceeds to other regionswith higher mobility, allowing the access of surfactant tails tohydrophobic regions. From 1,8-ANS fluorescence (Fig. 8) it is clearthat all the three mutations prevent the development of hydrophobicsolvent accessible patches. The serine 54 is located in the α-helix 51–63 and an aspartic acid in this position probably promotes a hydrogenbond between carboxylate D54 and the main carbonyl chain of A195[7,8]. This would prevent the separation of helices encompassingresidues 51–63 and 191–211, thus avoiding the formation of thelarger hydrophobic crevice that is thought to be the most importantweak region for unfolding by anionics [4]. This mutation is the onethat leads to the lowest fluorescence of the probe ANS for the wholerange of AOT concentrations indicating the development of fewerhydrophobic solvent patches when compared to the WT cutinase.Since no significant changes in the tertiary structure near thetryptophan region or near the active centre were observed, it isreasonable to assume that these hydrophobic crevices that are lessexposed in the structure of this mutant are involved in hydrophobicinteractions with the surfactant, inducing the unfolding of cutinase.The replacement of a hydrophobic amino acid (L153Q) also reducedthe development of hydrophobic solvent accessible patches (therelative fluorescence intensity of ANS was two orders of magnitudelower than that of the WT cutinase). However, the effect of AOT in thevicinity of the tryptophan is similar to the observed for the WTcutinase and probably this is also true near the active centre. Thismay suggest the existence of stronger electrostatic interactionsbetween AOT and this mutant, responsible for a high destabilisationof the tertiary structure. The T179C mutation in one of the bindingloops leads to a decrease in the fluorescence intensity of the probeANS by 2.5 orders of magnitude when compared to the WT cutinase.This suggests that this mutation also avoids the exposition ofhydrophobic crevices since both regions near the tryptophan andthe active centre exhibit a weaker destabilisation in the presence ofthe surfactant. Generally, the ANS fluorescence increases beforechanges in the tryptophan and tyrosine fluorescence occur, indicatingthat the access of ANS to the protein hydrophobic regions occursbefore any significant changes in the tertiary structure can bedetected by the tryptophan and tyrosine fluorescence. The unfoldingcurves of the mutants suggest the presence of an intermediate, that

Page 8: Biochimica et Biophysica Acta - ULisboaweb.tecnico.ulisboa.pt/jgmartinho/PDF/2008...nitrophenyl butyrate (p-NPB) (98%) were purchased from Sigma. 2.3. Enzymatic hydrolysis kinetics

1333V. Brissos et al. / Biochimica et Biophysica Acta 1784 (2008) 1326–1334

clearly shares some of the characteristics found in other proteins,that is molten globe in character, e.g., native-like secondary structure,with disordered tertiary structure and higher ANS fluorescence[38–41].

It is worth to note that the susceptibility to unfolding of cutinasescannot be explained solely by electrostatic interactions. The degree ofstabilisation of the native state, which in turn might reveal thevulnerability of surfactant association and the stabilisation oftransition states and intermediates in the unfolding pathway shouldalso be considered. To clarify this point the equilibrium unfolding ofthe WT cutinase and mutants induced by GdnHCl was studied.However, the unfolding does not allow a better understanding sinceWT cutinase and variants show the same unfolding profiles, not beinginfluenced by the addition of AOT until MR=10.

5. Concluding remarks

The results show that the S54D mutant of cutinase is significantlymore resistant to AOT denaturation than theWT. This mutant shows asecondary and tertiary structure similar to that of the WT and theincreased resistance to AOT may result from the establishment of ahydrogen bond between the carboxylate of the aspartic acid and thecarbonyl chain of the residue A195. This prevents the separation ofhelices encompassing residues 51–63 and 191–211, thus avoiding theformation of large hydrophobic crevices, responsible for the unfoldingby anionic surfactants.

The L153Qmutation also reduced the development of hydrophobicsolvent accessible patches. However, stronger electrostatic interac-tions may occur between the anionic surfactant and this mutant, asthe effect of AOT in the tertiary structure is the same as that observedfor the WT cutinase.

The T179C mutation located close to the active centre and todisulfide bond Cys171–Cys178 introduced changes in the cutinasestructure that were observed even in the cutinase region around thetryptophan residue. Nevertheless, this mutation also reduced thedevelopment of hydrophobic solvent accessible patches.

After the saturation of the binding sites by AOT, this surfactant caninteract with cutinase regions of higher mobility, allowing access ofthe surfactant tails to hydrophobic regions. As cutinase has certainlymore than one independent binding site, the interaction of thesurfactant with all sites at the protein surface cannot be avoided.However, if the L153Q, S54D and T179C mutations prevent thedevelopment of hydrophobic crevices formed upon initial unfoldingor even by the initial binding of surfactant to specific sites, theinteraction between surfactant and protein may be reduced. Theseregions can be detergent sensitive zones at the cutinase surface andfurnish the key for the design of biocatalysts with sufficient stabilityfor practical applications in detergent industry.

Acknowledgements

V. Brissos gratefully acknowledges PhD grant SFRH/BD/9019/2002from Fundação para a Ciência e Tecnologia. J. M. G. Martinhoacknowledges the financial support from the IN — Instituto deNanociências e Nanotecnologias. The authors are indebted to Prof.João Pessoa for Circular Dichroism facilities.

References

[1] P.E. Kolattukudy, Cutinases from fungi and pollen, in: B. Borgstrom, H. Brockman(Eds.), Lipases, Amsterdam, Elsevier, 1984, pp. 471–504.

[2] M. Lauwereys, P. de Geus, J. de Meutter, P. Stanssen, G. Matthyssens, Cloning,expression and characterisation of cutinase, a fungal lipolytic enzyme, in: L.Alberghina, R.D. Schmid, R. Verger (Eds.), Lipases, Structure, Mechanism andGenetic Engineering, GBF Monograph, 16, VCH, Weinheim, 1991, pp. 243–251.

[3] J.A.C. Flipsen, A.C.M. Appel, H.T.W.M. van der Hijden, C.T. Verrips, Mechanism ofremoval of immobilized triacylglycerol by lipolytic enzymes in a sequentiallaundry wash process, Enzyme Microb. Technol. 23 (1998) 274–280.

[4] L.D. Creveld, A. Amadei, R.C. van Schaik, H.A.M. Pepermans, J. de Vlieg, H.J.C.Berendsen, Identification of functional and unfolding motions of cutinase asobtained from molecular dynamics computer simulations, Proteins, StructureFunction and Genetics 33 (1998) 253–264.

[5] D.J. Pocalyko, M. Tallman, Effects of amphipaths on the activity and stability ofFusarium solani pisi cutinase, Enzyme Microb. Technol. 22 (1998) 647–651.

[6] M.R. Egmond, J. de Vlieg, H.M. Verheij, G.H. de Haas, Strategies and design ofmutations in lipases, in: F.X. Malcata (Ed.), Engineering of/with Lipases, KluwerAcademic Publishers, Dordrecht, The Netherlands, 1996, pp. 193–202.

[7] V. Brissos, T. Eggert, J.M.S. Cabral, K.E. Jaeger, Improving activity and stability ofcutinase towards the anionic detergent AOT by complete saturation mutagenesis,Protein Eng. Des. Sel. 21 (2008) 387–393.

[8] V. Brissos, Instituto Superior Técnico, Universidade Técnica de Lisboa, Lisbon,2006.

[9] J.M. Neugebauer, Detergents — an overview, Methods in Enzymology 182 (1990)239–253.

[10] G.S. Makowski, M.L. Ramsby, Protein molecular weight determination by sodiumdodecyl sulfate polyacrylamide gel electrophoresis, in: T.E. Creighton (Ed.),Protein Structure — a Practical Approach, Oxford University Press, Oxford, 1997,pp. 1–27.

[11] J.P. Henley, A. Sadana, Categorization of enzyme deactivations using a series-typemechanism, Enzyme Microb. Technol. 7 (1985) 50–60.

[12] A. Sadana, Biocatalysis: Fundamentals in Enzyme Deactivation Kinetics, PrenticeHall, Englewwod Cliffs, NY, 1991.

[13] J. Clarke, A.R. Fersht, Engineered disulfide bonds as probes of the folding pathwayof barnase — increasing the stability of proteins against the rate of denaturation,Biochemistry 32 (1993) 4322–4329.

[14] J.R. Lakowicz, Principles of Fluorescence Spectoscopy, Second Edition, KluwerAcademic/Plenum Publishers, N.Y., 1999.

[15] E.P. Melo, S.M.B. Costa, J.M.S. Cabral, Denaturation of a recombinant cutinase fromFusarium solani in AOT-iso-octane reverse micelles: a steady-state fluorescencestudy, Photochem. Photobiol. 63 (1996) 169–175.

[16] J.J. Prompers, C.W. Hilbers, H.A.M. Pepermans, Tryptophan mediated photoreduc-tion of disulfide bond causes unusual fluorescence behaviour of Fusarium solanipisi cutinase, Febs Lett. 456 (1999) 409–416.

[17] J.M.G. Martinho, A.M. Santos, A. Fedorov, R.P. Baptista, M.A. Taipa, J.M.S. Cabral,Fluorescence of the single tryptophan of cutinase: temperature and pH effect onprotein conformation and dynamics, Photochem. Photobiol. 78 (2003) 15–22.

[18] P.C.M. Weisenborn, H. Meder, M.R. Egmond, T.J.W.G. Visser, A. vanHoek,Photophysics of the single tryptophan residue in Fusarium solani cutinase:Evidence for the occurrence of conformational substrates with unusualfluorescence behaviour, Biophys. Chem. 58 (1996) 281–288.

[19] M.T. Neves-Petersen, Z. Gryczynski, J. Lakowicz, P. Fojan, S. Pedersen, E.Petersen, S.B. Petersen, High probability of disrupting a disulphide bridgemediated by an endogenous excited tryptophan residue, Protein Sci. 11 (2002)588–600.

[20] W.C. Johnson, Protein secondary structure and circular-dichroism — a practicalguide, Proteins, Structure Function and Genetics 7 (1990) 205–214.

[21] S.M. Kelly, N.C. Price, The application of circular dichroism to studies of proteinfolding and unfolding, Biochim. Biophys. Acta, Prot. Struct. Mol. Enzymol. 1338(1997) 161–185.

[22] S.M. Kelly, N.C. Price, The use of circular dichroism in the investigation of proteinstructure and function, Curr. Protein Pept. Sci. 1 (2000) 349–384.

[23] T. Ternstrom, A. Svendsen, M. Akke, P. Adlercreutz, Unfolding and inactivation ofcutinases by AOT and guanidine hydrochloride, Biochim. Biophys. Acta, Prot.Proteomic. 1748 (2005) 74–83.

[24] P. Sehgal, S. Bang Nielsen, S. Pedersen, R. Wimmer, D.E. Otzen, Modulation ofcutinase stability and structure by phospholipid detergents, Biochim. Biophys.Acta 1774 (2007) 1544–1554.

[25] T.K. Das, S. Mazumdar, S. Mitra, Characterization of a partially unfolded structureof cytochrome c induced by sodium dodecyl sulphate and the kinetics of itsrefolding, Eur. J. Biochem. 254 (1998) 662–670.

[26] H. Roder, W. Colon, Kinetic role of early intermediates in protein folding, Curr.Opin. Struct. Biol. 7 (1997) 15–28.

[27] L. Shi, D.R. Palleros, A.L. Fink, Protein conformational changes induced by 1,1′-bis(4-anilino-5-naphthalenesulfonic acid): preferential binding to the moltenglobule of DnaK, Biochemistry 33 (1994) 7536–7546.

[28] V.N. Uversky, S.Winter, G. Lober, Use of fluorescence decay times of 8-ANS-proteincomplexes to study the conformational transitions in proteins which unfoldthrough the molten globule state, Biophys. Chem. 60 (1996) 79–88.

[29] V.N. Uversky, S. Winter, G. Lober, Self-association of 8-anilino-1-naphthalene-sulfonate molecules: spectroscopic characterization and application to theinvestigation of protein folding, Biochim. Biophys. Acta, Prot. Struct. Mol. Enzymol.1388 (1998) 133–142.

[30] E.P. Melo, T.Q. Faria, L.O. Martins, A.M. Goncalves, J.M.S. Cabral, Cutinase unfoldingand stabilization by trehalose andmannosylglycerate, Proteins, Structure Functionand Genetics 42 (2001) 542–552.

[31] W. Koller, P.E. Kolattukudy, Mechanism of action of cutinase — chemicalmodification of the catalytic triad characteristic for serine hydrolases, Biochem-istry 21 (1982) 3083–3090.

[32] M.N. Jones, A. Finn, A. Mosavimovahedi, B.J. Waller, The Activation of Aspergillusniger catalase by sodium N-dodecyl-sulfate, Biochim. Biophys. Acta 913 (1987)395–398.

[33] A. Muga, J.L.R. Arrondo, T. Bellon, J. Sancho, C. Bernabeu, Structural and functionalstudies on the interaction of sodium dodecyl-sulfate with beta-galactosidase,Arch. Biochem. Biophys. 300 (1993) 451–457.

Page 9: Biochimica et Biophysica Acta - ULisboaweb.tecnico.ulisboa.pt/jgmartinho/PDF/2008...nitrophenyl butyrate (p-NPB) (98%) were purchased from Sigma. 2.3. Enzymatic hydrolysis kinetics

1334 V. Brissos et al. / Biochimica et Biophysica Acta 1784 (2008) 1326–1334

[34] C. Giancola, C. DeSena, D. Fessas, G. Graziano, G. Barone, DSC studies on bovineserum albumin denaturation — effects of ionic strength and SDS concentration,Int. J. Biol. Macromol. 20 (1997) 193–204.

[35] S. DAuria, M. Rossi, R. Nucci, G. Irace, E. Bismuto, Perturbation of conformationaldynamics, enzymatic activity, and thermostability of beta-glycosidase fromarchaeon Sulfolobus solfataricus by pH and sodium dodecyl sulfate detergent,Proteins, Structure Function and Genetics 27 (1997) 71–79.

[36] Y. Nozaki, J.A. Reynolds, C. Tanford, Interaction of a cationic detergent with bovineserum-albumin and other proteins, J. Biol. Chem. 249 (1974) 4452–4459.

[37] S. DAuria, R. Barone, M. Rossi, R. Nucci, G. Barone, D. Fessas, E. Bertoli, F.Tanfani, Effects of temperature and SDS on the structure of beta-glycosidasefrom the thermophilic archaeon Sulfolobus solfataricus, Biochem. J. 323 (1997)833–840.

[38] G.V. Semisotnov, N.A. Rodionova, O.I. Razgulyaev, V.N. Uversky, A.F. Gripas, R.I.Gilmanshin, Study of the “molten globule” intermediate state in protein folding bya hydrophobic fluorescent probe, Biopolymers 31 (1991) 119–128.

[39] O.B. Ptitsyn, Protein folding — hypotheses and experiments, J. Protein Chem. 6(1987) 273–293.

[40] M. Ohgushi, A. Wada, Molten-globule state— a compact form of globular-proteinswith mobile side-chains, Febs Lett. 164 (1983) 21–24.

[41] G.V. Semisotnov, N.A. Rodionova, V.P. Kutyshenko, B. Ebert, J. Blanck, O.B. Ptitsyn,Sequential mechanism of refolding of carbonic anhydrase-B, Febs Lett. 224 (1987)9–13.

[42] S. Longhi, M. Czjzek, V. Lamzin, A. Nicolas, C. Cambillau, Atomic resolution (1.0 A)crystal structure of Fusarium solani cutinase: stereochemical analysis, J. Mol. Biol.268 (1997) 779–799.