Essential dynamics of lipase binding sites: the effect of inhibitors of different...

10
peng$$1213 Protein Engineering vol.10 no.2 pp.149–158, 1997 Essential dynamics of lipase binding sites: the effect of inhibitors of different chain length Gu ¨ nther H.Peters 1,2,4 , D.M.F.van Aalten 3,5 , A.Svendsen 2 (Brady et al., 1990; Derewenda et al., 1992), human and R.Bywater 2 pancreatic (Winkler et al., 1990), Geotrichum candidum (Schrag and Cygler, 1993), Candida rugosa (Grochulski et al., 1 Chemistry Department III, H.C. Ørsted Institutet, University of 1993), Rhizopus delemar (Derewenda et al., 1994a), Copenhagen, Universitetsparken 5, DK-2100 Copenhagen Ø, Denmark, Pseudomonas glumae (Noble et al., 1993), Penicillium 3 Department of Biochemistry and Molecular Biology, University of Leeds, Leeds LS2 9JT, UK and 2 Novo-Nordisk A/S, Novo All ´ e 1, DK-2880 camembertii (Derewenda et al., 1994b), Humicola lanuginosa Bagsvaerd, Denmark, 5 Present address: Keck Structural Biology, (Lawson et al., 1994) and Cutinase (Martinez et al., 1992) Cold Spring Harbor Laboratory, 1 Bungtown Rd, NY 11724, USA have been crystallographically resolved. The crystal structures 4 To whom correspondence should be addressed at University of Copenhagen revealed that the catalytic serine is part of a triad equivalent to that found in the serine protease (Brady et al., 1990; Winkler The biochemical activity of enzymes, such as lipases, is often et al., 1990; Derewenda and Derewenda, 1991). The active associated with structural changes in the enzyme resulting site is shielded from the surrounding environment by loops in selective and stereospecific reactions with the substrate. and/or helices. The most striking difference between the amino To investigate the effect of a substrate and its chain length acid sequences is their length. For instance, human pancreatic on the dynamics of the enzyme, we have performed molecular lipase contains 449 amino acids (Lowe et al., 1989), while the dynamics simulations of the native Rhizomucor miehei lipase fungal enzyme from R.miehei contains 269 residues (Boel (Rml) and lipase–dialkylphosophate complexes, where the et al., 1988). Other related lipases exhibit the same kind of length of the alkyl chain ranges from two to 10 carbon atoms. variation. The degrees of similarity between the different Simulations were performed in water and trajectories of 400 lipase groups is very low, apart from that in the immediate ps were used to analyse the essential motions in these systems. environment of the catalytic site. There is a common motif Our results indicate that the internal motions of the Rml and present in most of the lipases: Gly-X 1 -Ser-X 2 -Gly, where X 1 Rml complexes occur in a subspace of only a few degrees of and X 2 are variable residues. The upstream Gly is not conserved freedom. A high flexibility is observed in solvent-exposed in all lipases, e.g. the motifs in Candida antartica lipase B segments, which connect β-sheets and helices. In particular, loop regions Gly35–Lys50 and Thr57–Asn63 fluctuate (Uppenberg et al., 1994) and Bacillus subtilis lipase (Dartois extensively in the native enzyme. Upon activation and et al., 1992) are Thr-X 1 -Ser-X 2 -Gly and Ala-X 1 -Ser-X 2 -Gly binding of the inhibitor, involving the displacement of the respectively. The consensus sequence, Gly-X 1 -Ser-X 2 -Gly, is active site loop, these motions are considerably suppressed. also observed in other esterases and is part of the substrate- With increasing chain length of the inhibitor, the fluctu- binding site (Svendsen, 1994a). The mechanism of activation ations in the essential subspace increase, levelling off at a for R.miehei lipase (Rml) has been revealed by its crystal chain length of 10, which corresponds to the size of the structure, in which a helical loop is positioned on top of the active-site groove. catalytic site, effectively acting as a lid, which buries non- Keywords: concerted motion/enzyme activation/lipase/molecu- polar residues underneath. The helix is supported by two lar dynamics/protein dynamics extended and flexible segments. The 3-D crystal structure of the Rml enzyme complexed to the substrate analogue diethylphosphate (Brzozowski et al., 1991; Derewanda et al., Introduction 1992) shows that the conformational changes observed during activation involves a rigid body hinge-type motion of a single Lipases (acylglycerol acylhydrolases, EC 3.1.1.3) are efficient helix (‘lid’). The resultant complex is thought to be reminiscent catalysts for lipolytic reactions initiating the catabolism of of the tetrahedral transition state which forms on binding fats and oils by hydrolysing the fatty acyl ester bonds of the triglyceride substrate (Dodson and Lawson, 1992). The acylglycerols (Wulfson, 1994). Their capability of catalysing displacement of the helix lid exposes previously buried hydro- a wide variety of reactions allows a widespread application in phobic residues, which favour interaction with a lipid interface. industry: the removal of oils and fats from fabrics, machinery At the same time, a tryptophan residue located on the underside and waste water, the production of mono- and diglycerides for of the lid, which in the uncomplexed enzyme blocks access food emulsifiers and stereospecific synthesis of compounds to the active serine, faces outward in the activated conformation incuding precursors for biologically active therapeutics, herbi- and presumably interacts favourably with the lipid interface cides or pesticides (Macrae, 1983; Macrae and Hammond, (Holmquist et al., 1993). As the lid rolls over, polar residues 1985; Boland et al., 1991; Haas et al., 1992; van Kuiken and on the lid helix and on the surface adjacent to it become buried Behnke, 1994). Lipases differ from one another in their and water molecules, which in the inactive enzyme form a physical properties and biochemical features and are selective well-defined network on the polar surface, are displaced. This in respect of the length and the level of saturation of the fatty adds a favourable entropic term to the stability of the activated acid chains (Haas et al., 1992). enzyme (Ashbaugh and Paulaitis, 1992). A series of X-ray studies on lipases and related enzymes The phenomenon of ‘interfacial activation’ seen in lipases has revealed the structural basis of their catalytic mechanism. The 3-D structures of lipases from Rhizomucor miehei has for a long time given rise to discussion about what © Oxford University Press 149

Transcript of Essential dynamics of lipase binding sites: the effect of inhibitors of different...

Page 1: Essential dynamics of lipase binding sites: the effect of inhibitors of different ...davapc1.bioch.dundee.ac.uk/pdf/lipII.pdf · 1997. 9. 5. · Essential dynamics of lipase binding

peng$$1213

Protein Engineering vol.10 no.2 pp.149–158, 1997

Essential dynamics of lipase binding sites: the effect of inhibitorsof different chain length

Gunther H.Peters1,2,4, D.M.F.van Aalten3,5, A.Svendsen2 (Brady et al., 1990; Derewendaet al., 1992), humanand R.Bywater2 pancreatic (Winkler et al., 1990), Geotrichum candidum

(Schrag and Cygler, 1993),Candida rugosa(Grochulskiet al.,1Chemistry Department III, H.C. Ørsted Institutet, University of1993), Rhizopus delemar (Derewenda et al., 1994a),Copenhagen, Universitetsparken 5, DK-2100 Copenhagen Ø, Denmark,Pseudomonas glumae(Noble et al., 1993), Penicillium3Department of Biochemistry and Molecular Biology, University of Leeds,

Leeds LS2 9JT, UK and2Novo-Nordisk A/S, Novo All´e 1, DK-2880 camembertii(Derewendaet al., 1994b),Humicola lanuginosaBagsvaerd, Denmark,5Present address: Keck Structural Biology, (Lawson et al., 1994) andCutinase(Martinez et al., 1992)Cold Spring Harbor Laboratory, 1 Bungtown Rd, NY 11724, USA have been crystallographically resolved. The crystal structures4To whom correspondence should be addressed at University of Copenhagenrevealed that the catalytic serine is part of a triad equivalent

to that found in the serine protease (Bradyet al., 1990; WinklerThe biochemical activity of enzymes, such as lipases, is oftenet al., 1990; Derewenda and Derewenda, 1991). The activeassociated with structural changes in the enzyme resultingsite is shielded from the surrounding environment by loopsin selective and stereospecific reactions with the substrate.and/or helices. The most striking difference between the aminoTo investigate the effect of a substrate and its chain lengthacid sequences is their length. For instance, human pancreaticonthedynamicsof theenzyme,wehaveperformedmolecularlipase contains 449 amino acids (Loweet al., 1989), while thedynamics simulations of the nativeRhizomucor mieheilipasefungal enzyme fromR.mieheicontains 269 residues (Boel(Rml) and lipase–dialkylphosophate complexes, where theet al., 1988). Other related lipases exhibit the same kind oflength of the alkyl chain ranges from two to 10 carbon atoms.variation. The degrees of similarity between the differentSimulations were performed in water and trajectories of 400lipase groups is very low, apart from that in the immediateps were used to analyse the essential motions in these systems.environment of the catalytic site. There is a common motifOur results indicate that the internal motions of the Rml andpresent in most of the lipases: Gly-X1-Ser-X2-Gly, where X1Rml complexes occur in a subspace of only a few degrees ofand X2 are variable residues. The upstream Gly is not conservedfreedom. A high flexibility is observed in solvent-exposedin all lipases, e.g. the motifs inCandida antarticalipase Bsegments, which connectβ-sheets and helices. In particular,

loop regions Gly35–Lys50 and Thr57–Asn63 fluctuate (Uppenberget al., 1994) andBacillus subtilislipase (Dartoisextensively in the native enzyme. Upon activation and et al., 1992) are Thr-X1-Ser-X2-Gly and Ala-X1-Ser-X2-Glybinding of the inhibitor, involving the displacement of the respectively. The consensus sequence, Gly-X1-Ser-X2-Gly, isactive site loop, these motions are considerably suppressed. also observed in other esterases and is part of the substrate-With increasing chain length of the inhibitor, the fluctu- binding site (Svendsen, 1994a). The mechanism of activationations in the essential subspace increase, levelling off at a for R.miehei lipase (Rml) has been revealed by its crystalchain length of 10, which corresponds to the size of the structure, in which a helical loop is positioned on top of theactive-site groove. catalytic site, effectively acting as a lid, which buries non-Keywords:concerted motion/enzyme activation/lipase/molecu-polar residues underneath. The helix is supported by twolar dynamics/protein dynamics extended and flexible segments. The 3-D crystal structure

of the Rml enzyme complexed to the substrate analoguediethylphosphate (Brzozowskiet al., 1991; Derewandaet al.,

Introduction 1992) shows that the conformational changes observed duringactivation involves a rigid body hinge-type motion of a singleLipases (acylglycerol acylhydrolases, EC 3.1.1.3) are efficienthelix (‘lid’). The resultant complex is thought to be reminiscentcatalysts for lipolytic reactions initiating the catabolism ofof the tetrahedral transition state which forms on bindingfats and oils by hydrolysing the fatty acyl ester bonds ofthe triglyceride substrate (Dodson and Lawson, 1992). Theacylglycerols (Wulfson, 1994). Their capability of catalysingdisplacement of the helix lid exposes previously buried hydro-a wide variety of reactions allows a widespread application inphobic residues, which favour interaction with a lipid interface.industry: the removal of oils and fats from fabrics, machineryAt the same time, a tryptophan residue located on the undersideand waste water, the production of mono- and diglycerides forof the lid, which in the uncomplexed enzyme blocks accessfood emulsifiers and stereospecific synthesis of compoundsto the active serine, faces outward in the activated conformationincuding precursors for biologically active therapeutics, herbi-and presumably interacts favourably with the lipid interfacecides or pesticides (Macrae, 1983; Macrae and Hammond,(Holmquistet al., 1993). As the lid rolls over, polar residues1985; Bolandet al., 1991; Haaset al., 1992; van Kuiken andon the lid helix and on the surface adjacent to it become buriedBehnke, 1994). Lipases differ from one another in theirand water molecules, which in the inactive enzyme form aphysical properties and biochemical features and are selectivewell-defined network on the polar surface, are displaced. Thisin respect of the length and the level of saturation of the fattyadds a favourable entropic term to the stability of the activatedacid chains (Haaset al., 1992).enzyme (Ashbaugh and Paulaitis, 1992).A series of X-ray studies on lipases and related enzymes

The phenomenon of ‘interfacial activation’ seen in lipaseshas revealed the structural basis of their catalytic mechanism.The 3-D structures of lipases fromRhizomucor miehei has for a long time given rise to discussion about what

© Oxford University Press 149

Page 2: Essential dynamics of lipase binding sites: the effect of inhibitors of different ...davapc1.bioch.dundee.ac.uk/pdf/lipII.pdf · 1997. 9. 5. · Essential dynamics of lipase binding

G.H.Peterset al.

determines selectivity. Effective hydrolysis probably dependson structural changes in the enzyme upon activation and thephysical state and structure of the lipid substrate. Many lipasesare sensitive to surface pressures (Pie´roni et al., 1990), pH(Huge-Jensenet al., 1987), ionic strength (Laironet al.,1980) and lipid composition (Muderhwa and Brockman, 1992;G.H.Peters, U.Dahmen, K.de Meijere, S.Toxvaerd, H.Mo¨hwaldand G.Brezesinski (unpublished). A higher activity is measuredwith increasing pH and surface pressures and, for many lipases,optima are observed. The catalytic reaction in the lipid–waterinterphase involves at least four processes: (i) binding to thelipid surface, (ii) penetration into the lipid phase, (iii) activationof the enzyme (i.e. displacement of the lid) and (iv) catalytichydrolysis (Svendsen, 1994b). In our previous studies, wehave independently investigated the phases and interfacialproperties of the lipid substrate (Peterset al., 1994, 1995a)and the activation process of Rml (Peterset al., 1995b, 1996a).Our findings showed that electrostatic interactions have a keyrole in the activation of lipases and, hence, the hydrophilicityof the lipid surface is an important factor in the activation step(Peterset al., 1995a).

The objective of this study is to examine the essentialmotions and structural changes in Rml when the transitionstate between the enzyme and substrate is formed. The crystalstructures of the transition state itself are not available, sincesuch a state is by its nature unstable. Therefore, we have usedthe crystal structure of Rml complexed to the substrate analoguediethylphosophate, where we have varied the length of one ofthe alkyl chains. We have applied the essential dynamicsanalysis technique (Amadeiet al., 1993) to extract informationabout concerted atomic motions in the protein. This techniquehas been used to correlate essential motions in lysozyme(Amadei et al., 1993) and thermolysin (van Aaltenet al.,1995) to their biological function. Similarly, this study focuseson investigating the effect of the chain length of the substrate(analogue) on the essential motions in the protein, which mayincrease our understanding of the selectivity of lipases towardsthe length of the fatty acid chains.

Model and simulation detailsThe high resolution crystal structures of the native Rml solvedto 1.9 Å resolution (Derewendaet al., 1992) and a lipaseinhibitor complex solved to 2.6 Å resolution (Brzozowskiet al., 1991) were used as models for the inactive and activestructures respectively. The structures were obtained from theProtein Data Bank at Brookhaven (Bernsteinet al., 1977)(entry codes 3tgl and 4tgl). The inhibitor is diethylphosophate,which is covalently bonded to the active site serine. Visualinspection of the structure shows the hydrophobic groove,where the alkyl chain of the substrate is placed during theFig. 1. Secondary structures of (a) inactive and (b) active (inhibitor-bound)catalytic reaction (Derewenda, 1994). We extended the alkylconformations of theR.mieheilipase. The lid region is indicated by the red

colour. In the active form (B), ethyl-decanoylphosophate is covalentlychain of the diethylphosophate inhibitor by four and eightbound to the active serine (Ser144). Green, red and yellow represent themethylene groups. The alkyl chains were placed in an all-transmethyl(ene), oxygen and phosphor atoms respectively.conformation using the QUANTA software package

(CHARMM, 1992). Figure 1(a) and (b) displays the inactiveand active conformations of the Rml. The lid region is indicated Examination of the molecular structures and analysis of the

trajectories were carried out using the WHAT IF modellingby the red colour. In the active form, ethyl-decanoyl-phosophateis covalently bonded to the active site serine (Ser144). The program (Vriend, 1990) and the ESSDYN essential dynamics

menu supplied therein. Simulations were performed at 300 Kmain-chain structure of the Rml is composed of a centralβ-sheet system of nine strands. The structure is stabilized by with a time step of 2 fs. Periodic boundary conditions were

applied using a truncated octahedral simulation cell, filled withthree cysteine bridges.Molecular dynamics simulations were performed using the water molecules taken from an equilibrated liquid configura-

tion. Water was described by the simple point charge (SPC)GROMOS program (van Gunsteren and Berendsen, 1987).

150

Page 3: Essential dynamics of lipase binding sites: the effect of inhibitors of different ...davapc1.bioch.dundee.ac.uk/pdf/lipII.pdf · 1997. 9. 5. · Essential dynamics of lipase binding

Essential dynamics of lipase binding sites

Fig. 2. Correlation coefficient as a function of the eigenvector indexdetermined from the stimulation of the Rml in an aqueous environmentcompared to an ideal Gaussian distribution derived from the eigenvalue.

Fig. 3. Cumulative normalized eigenvectors determined from the Cαcoordinates covariance matrix of the trajectory obtained from the simulationof the Rml in water.

model (Berendsenet al., 1981). The non-bonded pair list wasupdated every 10 fs and the non-bonded and electrostaticinteractions were truncated at 8 and 10 Å respectively. TheSHAKE algorithm (Ryckaertet al., 1977) was applied toconstrain the bond lengths to their equilibrium positions andthe equations of motion were solved using the Verlet algorithm(Allen and Tildesley, 1989).

Simulations were started from the crystallographic structureof Rml and subjected to a steepest descents energy minimiza-tion for 2500 steps. The final energy derivative was,0.01kJ/mol in all simulations. The minimization was followed bya 5 ps molecular dynamics (MD) simulation, where theinitial velocities of the atoms were taken from a Maxwelliandistribution. Simulations were run for 400 ps and the last 300Fig. 4. The root mean square displacement, number of hydrogen bonds and

radius of gyration (from the top) computed during the simulation of theps of the trajectories were used for the essential dynamicsdifferent systems.analysis. The stability of the simulations was checked by

monitoring several geometrical properties (the radius of gyra-tion, number of hydrogen bonds, root mean square displace-ment, accessibility, number of residues in random coilconformation and number of strained dihedrals) and the last

151

Page 4: Essential dynamics of lipase binding sites: the effect of inhibitors of different ...davapc1.bioch.dundee.ac.uk/pdf/lipII.pdf · 1997. 9. 5. · Essential dynamics of lipase binding

G.H.Peterset al.

300 ps trajectories were used for the essential dynamicsanalyses. The essential dynamics method (Amadeiet al., 1993)is based on the diagonalization of the covariance matrix builtfrom atomic fluctuations in a trajectory from which overallthe translation and rotations have been removed:

M 5 ⟨(Xi 2 Xi,0)(Xj 2 Xj,0)⟩ (1)

in which X is the separatex, y, z coordinates of the atoms,fluctuating around their averageX0. The translational androtational motion were eliminated by fitting on a referencestructure. The centres of mass were superposed and sub-sequently a least-quarter fit procedure on the Cα coordinateswas performed. The diagonalization of Equation 1 yields a setof eigenvectors and corresponding eigenvalues. The eigenvec-tors describe directions in the 3N configurational space, repres-enting correlated positional changes of groups of coordinates.The eigenvalues indicate the total mean square fluctuation alongthese directions. As a matter of convention, the eigenvectors are

Fig. 5. Average values of the projections of the separate trajectories onto ordered by the size of their corresponding eigenvalues, i.e. thethe eigenvectors extracted from the Cα coordinates covariance matrix of the first eigenvector is the eigenvector with the largest eigenvalue.concatenated trajectories as a function of the eigenvector index. The central hypothesis of essential dynamics is that only

the eigenvectors with large corresponding eigenvalues areimportant in describing the overall motion of the protein. Ingeneral, it has been shown for several proteins (Amadeiet al.,1993; van Aaltenet al., 1995, 1996a) that taking the first feweigenvectors represents ~80–95% of the total motion of theprotein. For instance, as shown in Figure 2, the first 50eigenvectors describe ~85% of the motion in Rml whensimulated in an aqueous solution (Peterset al., 1996b).The motions described by higher eigenvectors are essentiallyGaussian (thermal) motions. Figure 3 shows the correlationcoefficient between an ideal Gaussian distribution and thedistributions calculated from the eigenvectors. The coefficientis ~0.95 after 20 eigenvectors.

A useful method for comparing the essential dynamicsproperties of two simulations on similar systems is the so-called ‘combined’ analysis (van Aaltenet al., 1995). In thismethod, two or more trajectories, fitted on the same referencestructure, are concatenated and a covariance matrix is con-structed. The separate trajectories are then projected onto the

Fig. 6. Mean square fluctuations of projections of the separate trajectories resulting eigenvectors and the properties of these projectionsonto the eigenvectors extracted from the Cα coordinates covariance matrix are compared for all simulations. There are two main effectsof the concatenated trajectories as a function of the eigenvector index. to be studied.

Table I. Summary of the residues involved in the changes of average projection and mean square fluctuations along different eigenvectors.

Eigenvector Residues involved Interpretation

1 Mainly lid Static shift, mainly describing the difference between the active and inactive structures2 (A) 20–29, (B) 37–48, (C) 97–106, Static shift and small changes in dynamics, which correlate with the ligand size.

(D) 208–216, (E) 227–230, (F) 236–252, Conformational changes involve the folding back of (D) and (F) onto the protein(G) 256–265, (H) N-terminal surface. Conformational changes in (E) and (F) (van der Waals contacts) are due to the

folding back of (D). (E) and (G) push against the N-terminal (displacement)3 (A) 88–95, (B) 225–238, (C) N-terminal Static shift. The C2 inhibitor aligns anti-parallel into the groove4 (A) 39–50, (B) 89–97, (C) 98–105, Fluctuations. Most of the fluctuations are around the active-site groove. The difference

(D) 209–215 in the fluctuation of the lid is correlated to the size of the inhibitor. Inhibitors withlonger alkyl chains reduce the lid motion

5 (A) 35–50, (B) 68–75, (C) 131–137, Fluctuations, which are mainly found in loops (A)–(D). Large fluctuations in loop (A),(D) 160–168, (E) 209–215 which due to van der Waals contacts, cause the motion in (B)–(D)

6 (A) 35–50, (B) 57–63, (C) 86–94, (D) 93–101, Fluctuations, which are mainly observed in the native Rml. Opening of lid (C)(D) 189–194 corresponds to folding back of loop (B) onto the protein surface.

7 (A) 57–63, (B) 84–90, (C) 92–94, (D) 97–105, Fluctuations, which are mainly observed for the C10 inhibitor. Folding back of loop(E) 209–215, (F) 224–232 (E) due to the size of the C10 inhibitor

8 (A) 35–50, (B) 57–63, (C) 80–87, (D) 91–95, Fluctuations, which are mainly observed in the native Rml. Fluctuations in active-site(E) 226–232 lid regions (C) and (D)

152

Page 5: Essential dynamics of lipase binding sites: the effect of inhibitors of different ...davapc1.bioch.dundee.ac.uk/pdf/lipII.pdf · 1997. 9. 5. · Essential dynamics of lipase binding

Essential dynamics of lipase binding sites

(i) The average projection, indicating that simulations have in solvent. The stability of the simulations was checked bydifferent average displacements along eigenvectors, i.e. havecomputing the root mean square displacement (r.m.s.d.) witha different equilibrium structure in that direction. respect to the starting structure, the number of hydrogen bonds(ii) The mean square fluctuation in the projection, revealing aand the radius of gyration. As shown in Figure 4, a slightdifference in dynamics along this direction. increase in the r.m.s.d. is observed, but the other geometricalHere, this method is applied to the concatenated trajectoriesquantities are stable and fluctuating around a constant value,of the inactive enzyme and those with the different inhibitorsindicating that we have obtained stable trajectories which maybound. be used for essential dynamics analyses.

The basic results of an essential dynamics analysis on lipasesResults and discussions has been described elsewhere (Peterset al.1996b). We showed

that there is a pronounced motion in the lid and that, as alsoMolecular dynamics simulations of the native Rml and Rml–found for other proteins, there are only a few ‘essential’dialkylphosophate complexes, where the length of the alkyl

chain ranges from two to 10 carbon atoms, were performed eigenvectors. To investigate the effect of the inhibitors on the

Fig. 7. Continued on next page.

153

Page 6: Essential dynamics of lipase binding sites: the effect of inhibitors of different ...davapc1.bioch.dundee.ac.uk/pdf/lipII.pdf · 1997. 9. 5. · Essential dynamics of lipase binding

G.H.Peterset al.

dynamics of the Rml, we have performed a ‘combined’ and the superpositions of 50 sequential projections of the Cαmotion onto selected eigenvectors are displayed in Figure 7.essential dynamics analysis of the trajectories of the native

Rml and the Rml–inhibitor complexes. The covariance matrix In the Cα traces, the catalytic residue Ser144 is indicated bythe orange colour and the glycines known to be flexible linkswas constructed from the Cα coordinates from the different

systems and subsequently diagonalized. in proteins are coloured purple. In all projections, it is noticeablethat the active serine located at the end of a helix is rigid,Motion within the subspace can be studied by projecting

the trajectory onto the individual eigenvectors. This dot product whereas fluctuations of the active-site lid (coloured red)are observed along several eigenvectors. The C-terminal isprovides information about the time dependence of the con-

formational changes. The average and mean square fluctuation relatively rigid due to disulphide bridge Cys268–Cys29, whichanchors the terminus to the N-terminal helix. It is generallyof these projections as a function of eigenvector indices are

shown in Figures 5 and 6 respectively. For several eigenvectors, observed thatβ-sheets are stiff due to the formation of hydrogenbonds between the sheets, whereasα-helices, when exposedthe changes in the average projection (structure) and mean

square fluctuations (dynamics) correlate with the inhibitor size. to the solvent in particular, are flexible.Motions along eigenvectors 1–3 describe mainly conforma-A summary of the involved residues is provided in Table I

Fig. 7. Continued on next page.

154

Page 7: Essential dynamics of lipase binding sites: the effect of inhibitors of different ...davapc1.bioch.dundee.ac.uk/pdf/lipII.pdf · 1997. 9. 5. · Essential dynamics of lipase binding

Essential dynamics of lipase binding sites

Fig. 7. (a)–(g) Superposition of 50 configurations extracted from the concatenated trajectories and obtained by projecting the C motion onto eigenvectors2–8. (h) Initial configuration displaying selected residues and residue numbers. Eigenvector 1 is not shown since it corrresponds to the different position ofthe active site loop in the native and active forms. In the Cα traces, the catalytic residue Ser144 is indicated by the orange colour and the glycines known tobe flexible links in proteins are coloured purple.

eigenvectors 6 and 8 are predominantly observed in the nativeRml. For eigenvectors 2, 4 and 6, the magnitude of themean square fluctuations correlates to the inhibitor size, i.e.combinations of 2–6–10 or 10–6–2 are found.

The motions in the subspace spanned by eigenvector 1correspond to the difference in the equilibrium structures ofthe native Rml and the Rml–inhibitor complexes and nosignificant differences are observed in the mean square fluctu-ations (Figure 6). The configurational changes and fluctuationsalong eigenvector 2 involve the folding back of loop Leu208–Leu216 and the displacement of hinge region Val97–Lys106.The folding back of loop Leu208–Leu216 widens the groove,but the C10 inhibitor is too short to fill the whole groove and,consequently, segment Val97–Lys106 (a turn which is locatedat the end of the groove) moves into the free space. Increasingthe chain length of the inhibitor causes van der Waals contactsbetween the methyl group of the inhibitor and residue Pro209.The C6 inhibitor fits well in the active-site groove as indicated

Fig. 8. Output from the moving window superposition method using theby the all-trans state of the alkyl chain. Gauche defects areminimum and maximum structures observed in the subspaces spanned byobserved in the alkyl chain of the C10 inhibitor.eigenvectors 4, 5, 7 and 8. The positions of the glycines are indicated by

the open and filled circles. See the text for more details. Similar conformational changes are seen in the subspacespanned by eigenvector 3 (Figure 7b). Large displacements

tional (static) changes in the Rml structure due to the inhibitor.are observed in lid region Trp88–Val95, which are larger inThere are differences in the fluctuations (dynamics) in thethe simulation with the C6 inhibitor (Figure 6). It is the contactmotions along eigenvectors.3. The mean square fluctuations between the Asp91 and the end of the alkyl chains whichfor eigenvectors 4, 5 and 7 are the highest for the C2, C6 andcauses the displacement of the lid. This interaction is not

present in the simulations with the C10 inhibitor, since theC10 inhibitors respectively (Figure 6). The motions along

155

Page 8: Essential dynamics of lipase binding sites: the effect of inhibitors of different ...davapc1.bioch.dundee.ac.uk/pdf/lipII.pdf · 1997. 9. 5. · Essential dynamics of lipase binding

G.H.Peterset al.

Fig. 9. Structural sequence alignment and the structures of theR.mieheilipase (Rml),H.lanuginosalipase (Hll),R.delemar(Rhi) andPenicillium camembertii(Pen). The active site loop is indicated by the red colour. Loops Gly35–Lys50 and Thr57–Asn63 are displayed in yellow and blue respectively. The colourcoding indicates the number of identical residues in the sequence of the different lipases (green–blue–yellow–red corresponds to 4–3–2–0).

and its hinge–bend region (Thr93–Pro101). Fluctuations areobserved for loops Gly35–Lys50, Thr57–Asn63 and Pro194–Val189. The instability of the last turn is probably due toGly192 causing a flexible link between aβ-sheet and a shorthelix. The other two turns are exposed to solvent and are notstabilized by hydrogen bonds. It is noticeable that fluctuationsalong loop Thr57–Asn63 are not found in the simulations withthe Rml–inhibitor complexes. The flexibility is drasticallyreduced by the movement of the lid going from the inactive(closed) to the active (open) conformation. van der Waalsinteractions between residues in lid region Ser83–Asn87 andloop Thr57–Asn63 cause the turn to be locked between thelid region and link Tyr115–Val118. The motions described byeigenvector 6 are also observed in the Rml–inhibitor com-plexes. The motion becomes more significant with an increasingalkyl chain length of the inhibitor (Figure 6) and corresponds

Fig. 10. Total fluctuation in the essential space defined by the first 20 to fluctuations of loop Thr57–Asn63, which causes a displace-eigenvectors as a function of the inhibitor size. ment of the lid in the direction normal to the helix. The

distortion is damped by Gly104, which forms a flexible linkaccommodation of C10 requires that the active-site lid opensbetween active-site hinge region Thr93–Pro101 and a helix.further causing the roll-over of the lid and resulting in the This finding is contrary to the result observed in the directionexposure of Trp84 to the solvent. Shortening the alkyl chainof eigenvector 4, where the fluctuations were mainly observedlength of the inhibitor causes a larger range of lid fluctuationsin the C2 inhibitor and the amount of fluctuations decreases(Ile89–Val97) as shown in the motion along eigenvector 4with increasing chain length. Comparing the motions along(Figure 7c). Fluctuations in active-site loop Arg80–Val95 areeigenvectors 4 and 6, it shows that motions in loop Thr57–correlated with the motion in loop Leu208–216, which is Asn63 are absent in the subspace spanned by eigenvector 4.located opposite active-site hinge region Phe94–Pro96. As theVisualization of several configurations shows that the alkyllength of the alkyl chain increases the interaction decreaseschain of the C2 inhibitor is too short and is aligned perpendic-and C10 is long enough to cause the folding back of loopular to the groove, whereas the C6 and C10 inhibitors are longLeu208–Leu216 as seen in eigenvector 2. The motions in theenough to favour a parallel alignment of the alkyl chains. Thesubspaces spanned by eigenvectors 6 and 8 are predominantlyperpendicular conformation of the C2 inhibitor causes hinge–observed in native Rml. In particular, the motions occurringbend region Lys73–Ser84 to be further pushed against turnin the latter subspace are only minor in the Rml–inhibitorThr57–Asn63 than is observed for the other inhibitors. This

effect diminishes with increasing chain length mainly due tocomplexes. Fluctuations are observed in the active-site helix

156

Page 9: Essential dynamics of lipase binding sites: the effect of inhibitors of different ...davapc1.bioch.dundee.ac.uk/pdf/lipII.pdf · 1997. 9. 5. · Essential dynamics of lipase binding

Essential dynamics of lipase binding sites

the parallel orientation of the alkyl chain with respect to can utilize substrate binding energy directly to lower the freeenergy of activation of the catalysed reaction. To analyse furtherthe groove.

Differences between the C10 and C6 inhibitors seen in eigen- the motion in the enzyme, we have calculated the total fluctuationin the essential space defined by the first 20 eigenvectors. Thevectors 4 and 6, reflect the extent of the lid opening. The groove

can accommodate a 12-carbon fatty acid chain (Derewenda, total fluctuation as a function of inhibitor size is displayed inFigure 10. The fluctuations increase with increasing inhibitor1994). For shorter chains, the groove is partially closed causing

a distortion of the lid, i.e. for the C6 inhibitor the lid is open length and seem to level off at an alkyl chain length of 10carbons. This agrees well with structural considerations. It hasaround Arg86–Ile89 and further closed at Leu92–Pro96. This

twist, which is less pronounced for the C10 inhibitor may gener- been estimated that the groove formed during the activationis long enough to accommodate a 12-carbon fatty acid chainateunfavourable contactsbetween the lidand loop Thr57–Asn63

influencing the motion in this turn. The influence of the inhibitor (Derewenda, 1994). According to the induced fit model (Kosh-land, 1958), in the inactive enzyme the residues at the active siteis also observed in the motions along eigenvectors 4, 5 and 7,

where fluctuations at region Ser82–Ile89 are increased with that are required for catalysis are not located in the correctposition to interact with the substrate. The binding of the sub-increasing chain length supporting the view that for shorter

chains the lid is partially closed. Also motions along loop Gly69– strate creates interactions with the active site that lead to a changein enzyme conformation and shift the catalytic groups into theThr74 are only found in the C2 and C6 inhibitor complexes. In

the C10 complex, this motion is present in loop Thr57–Asn63. correct positions relative to the reacting groups of the substrateso that the catalysed reaction may take place. Essentially, bindingBoth loops are separated by aβ-sheet and, as discussed earlier,

the distortion of the active-site lid, which is observed for the of the substrate causes an entropy variation. The formation ofthe complex (enzyme–substrate) causes a decrease in entropy,shorter inhibitors, suppresses the fluctuations in loop Thr57–

Asn63. whereas the displacement of water located in the binding pocketincreases the configurational entropy of the solvent (AshbaughThere are several glycines in the Rml structure, which provide

flexible links and give rise to the internal motions in Rml. A and Paulaitis, 1992). The increase in fluctuations (‘entropy’)observed in Figure 10 is due to the presence of an aqueousmoving window superposition method (van Aaltenet al., 1996b)

was used to locate the hinge–bend regions in Rml. The root environment. As described earlier, the catalytic reaction takesplace in the interfacial plane of the lipid substrate. Activationmean square deviations (r.m.s.d.) calculated from the minimum

and maximum structures of the eigenvector motion as a function (i.e. displacement of the lid) occurs in a low dielectric constantmedium. As the lid rolls over, polar residues on the active-of residue number are shown in Figure 8. Significant r.m.s.d.

values are observed close to glycines, whose positions are indi- site loop and protein surface are buried. At the same time,hydrophobic residues are exposed to the lipid interface stabiliz-cated by the filled and open circles. The motions described by

eigenvectors 4, 5, 7 and 8 are predominately found in the C2 ing the active conformation. In the simulations, these residuesare exposed to an aqueous environment. The unfavourable inter-inhibitor complex, C6 inhibitor complex, C10 inhibitor complex

and water respectively (see Figure 6). In a residue range.150, actions between the hydrophobic residues and water cause adestabilization of the active conformation and hence larger fluc-the r.m.s.d. decreases with increasing inhibitor size. This is

caused by van der Waals contacts between strings Gly175– tuations in the essential space (Figure 10).Ala183 and Glu201–Phe215, which are in contact with the inhib-

Conclusionitor. Several glycines (indicated by the filled circles) are con-served in the structures ofR.mieheilipase,H.lanuginosalipase Analysis of the essential motions in lipases is important for a

detailed understanding of the relationship between structure(Hll), R.delemar(Rhi) and P.camembertii(Pen). The resultsof the structural sequence alignment and the structures of the and biological function, which could provide information about

possible changes in the lipase function. Progress has been madedifferent lipases are shown in Figure 9. In general, Hll, Rhi andPen are structurally similar to the Rml. The active site loop is by several crystallographic studies providing information about

the location and nature of the catalytic centre, substrate bindingcoloured red, whereas loops Gly35–Lys50 and Thr57–Asn63are coloured yellow and blue respectively. Pairwise comparisons pocket and the structural basis for interfacial activation. How-

ever, no attempt has been made to describe in detail the largeshow that Pen and Hll share 41% of identical amino acids, versus55% for the Rml–Rhi pair (Derewendaet al., 1994b). It is concerted motions in the Rml and their relation to the biological

function. In the present study, we have investigated the concertednoticeable, that loops Gly35–Lys50 and Thr57–Asn63, whichare most affected by the inhibitor, show low homology between motions in the native Rml and the effect of inhibitors on these

motions by performing molecular dynamics simulations andthe lipases.The present results indicate that the conformational changes essential dynamics analyses. Fluctuations of the active-site loop

and its hinge–bend region (Thr93–Pro101) are observed in sev-and internal motions in the protein are significantly influencedby the inhibitor. The formation of an enzyme–substrate inter- eral eigenvectors. The other hinge–bend region (Lys73–Ser84)

shows much less fluctuations, since it is part of the centralβ-mediate is complex and the favourable binding interactionenergy between the active site and substrates and the unfavour- sheet system.

The combined analysis of the trajectories shows that the fluc-able binding or exclusion of non-specific substrates may explainthe selectivity and specificity of the lipases. Four mechanisms tuations in the essential space initially increase with increasing

chain length of the inhibitor. The activation step involves thethrough which binding energy may be utilized to bring about anincrease in the rate of reaction of an enzyme–substrate complex displacement (roll-over) of the active-site helix, by which polar

residues on the lid and protein surface are buried and hydro-are (i) induced fit, (ii) destabilization of the substrate relative tothe transition state, (iii) productive binding and (iv) entropy phobic residues are exposed. The increase in fluctuations corre-

sponds to unfavourable interactions between the non-polarloss (Koshland, 1958; Jencks, 1975). Induced fit and (non-)productive binding provide mechanisms for specificity and con- residues and water molecules. As observed in many proteins

(Amadei et al., 1993; van Aaltenet al., 1995, 1996a,b), thetrol, whereas substrate destabilization and the loss of entropy

157

Page 10: Essential dynamics of lipase binding sites: the effect of inhibitors of different ...davapc1.bioch.dundee.ac.uk/pdf/lipII.pdf · 1997. 9. 5. · Essential dynamics of lipase binding

G.H.Peterset al.

Brzozowski,A.M., Derewenda,U., Derewenda,Z.S., Dodson,G.G.,catalytic site (Ser144, at the end of a helix) is rigid. HelicesLawson,D.M., Turkenburg,J.P., Bjorkling,F., Huge-Jensen,B., Patkar,S.A.are relatively rigid due to interactions with the centralβ-sheet and Thim,L. (1991)Nature, 351,491–494.

system. The motions are mainly observed in the loops. In particu-CHARMM (1992)CHARMM Molecular Modeling Software Package, Release3.3. Molecular Simulations Inc., Waltham, MA.lar, loops Gly35–Lys50 and Thr57–Asn63 show extensive flex-

Dartois,V., Baulard,A., Schanck,K. and Colson,C. (1992)Biochim. Biophys.ibility in the native enzyme. In the Rml–inhibitor complexes,Acta, 1131,253–260.motions in loop Thr57–Asn63 are significantly suppressed. ForDerewenda,Z.S. (1994)Adv. Protein Chem., 45, 1–52.

shorter alkyl chains, the groove partially closes causing a twistDerewenda,Z.S. and Derewenda,U. (1991)Biochem. Cell Biol., 69, 842–851.Derewenda,Z.S., Derewenda,U. and Dodson,G.G. (1992)J. Mol. Biol., 227,in the lid. This creates unfavourable van der Waals interactions

818–839.between hinge–bend region Lys73–Ser84 of the active-site lidDerewenda,U., Swenson,L., Wei,Y., Green,R., Kobos,P.M., Joerger,R.,and turn Thr57–Asn63 essentially locking this turn. For the C10 Haas,M.J. and Derewenda,Z.S. (1994a)J. Lipid Res., 35, 524–534.

inhibitor complex, the lid is open further releasing the stressDerewenda,U., Swenson,L., Green,R., Wei,Y., Dodson,G.G., Yamaguchi,S.,Haas,M.J. and Derewenda,Z.S. (1994b)Nature Struct. Biol., 1, 36–47.between the hinge region and loop Thr57–Asn63. Extensive

Dodson,G.G. and Lawson,D.M. (1992)Faraday Discuss., 93, 95–105.fluctuations are observed in loop Leu208–Leu216 in the simula-Grochulski,P., Yunge,L., Schrag,J.D., Bouthillier,F., Smith,P., Harrison,D.,

tions with the C2 inhibitor, which have a pronounced effect on Rubin,B. and Cygler,M. (1993)J. Biol. Chem., 268,12843–12847.the protein dynamics. For the C10 inhibitor complex, this loopHaas,M.J., Cichowicz,D.J. and Bailey,D.G. (1992)Lipids, 27, 571–576.

Holmquist,M., Martinelle,M., Berglund,P., Clausen,I.G., Patkar,S.,folds back causing a large displacement of the N-terminus andSvendsen,A. and Hult,K. (1993)J. Protein Chem., 12, 749–757.loop Gly35–Lys50. Motions in the loop are observed in severalHuge-Jensen,B., Galluzzo,D.R. and Jensen,R.G. (1987)Lipids, 22, 559–565.

eigenvectors indicating that this loop is rather flexible. In fact,Jencks,W.P. (1975) In Meister,A. (ed.),Advances in Enzymology. J. Wiley,New York, pp. 219–410.it is only stabilized by a few hydrogen bonds with residues in

Koshland,D.E.,Jr (1958)Proc. Natl Acad. Sci. USA, 44, 98–106.the N-terminal helix.Lairon,D., Charbonnier-Augeire,M., Nalbone,G., Leonardi,J., Hauton,J.C.,Structural sequence alignment ofR.miehei lipase, Pieroni,G., Ferrato,F. and Verger,R. (1980)Biochim. Biophys. Acta, 618,

H.lanuginosalipase,R.delemarandP.camembertii(see Figure9) 106–118.Lawson,D.M., Brzozowski,A.M., Dodson,G.G., Hubbard,R.E., Huge-indicates that several glycines are conserved in the lipase struc-

Jensen,B., Boel,E. and Derewenda,Z.S. (1994) In Woolley,P. andtures, which strongly suggests that these residues are importantPetersen,S.B. (eds),Lipases—Their Structure, Biochemistry and

for the flexibility (i.e. internal motions) and biological function Applications. Cambridge University Press, Cambridge, pp. 77–94.of these enzymes. Structural similarity is also observed for loopsLowe,M.E., Rosemblum,J.L. and Strauss,A.W. (1989)J. Biol. Chem., 264,

20042–20048.Gly35–Lys50 and Thr57–Asn63. Both loops are found inMacrae,A.R. (1983)J. Am. Oil Chem. Soc., 60, 291–294.H.lanuginosalipase,R.delemarandP.camembertii. The loops Macrae,A.R. and Hammond,R.C. (1985)Biotech. Gen. Engng Rev.,3,193–217.

are coloured yellow (Gly35–Lys50) and blue (Thr57–Asn63) inMartinez,C., de Geus,P., Lauwereys,M., Matthyssens,G. and Cambillau,C.(1992)Nature, 356,615–618.Figure 9. Our previous study, where we investigated the effect

Muderhwa,J.M. and Brockman,H.L. (1992)J. Biol. Chem., 267,24184–24192.of hydrophobic or hydrophilic solvents on the concerted motionsNoble,M.E.M., Cleasby,A., Johnson,L.N., Egmond,M.R. and Frenken,L.G.J.in R.mieheilipase (Peterset al., 1996), indicated that in a hydro- (1993)FEBS Lett., 331,123–128.

phobic solvent loop Gly35–Lys50 folds back on the proteinPieroni,G., Gargouri,Y., Sarda,L. and Verger,R. (1990)Adv. Coll. Int. Sci., 32,341–378.surface. This change is accompanied by a displacement of loop

Peters,G.H., Toxvaerd,S., Svendsen,A. and Olsen,O.H. (1994)J. Chem. Phys.,Thr57–Asn63, which pushes against the active site hinge region100,5996–6010.

and causes, as a consequence of van der Waals contacts, a slightPeters,G.H., Toxvaerd,S., Larsen,N.B., Bjørnholm,T., Schaumburg,K. andopeningof theactivesite lid.Thismaysuggest that improving the Kjaer,K. (1995a)Nature Struct. Biol., 2, 401–409.

Peters,G.H., Toxvaerd,S., Olsen,O.H. and Svendsen,A. (1995b)Protein Engng,interactions (e.g. additional salt bridges) between loop Gly35–71, 119–129..Lys50 and residues on the protein surface may enhance the

Peters,G.H., Olsen,O.H., Svendsen,A. and Wade,R.C. (1996a)Biophys. J., 71,activation of lipases, which could be investigated in site-directed 119–129.

Peters,G.H., van Aalten,D.M.F., Edholm,O., Toxvaerd,S. and Bywater,R.mutagenesis experiments.(1996b)Biophys. J., in press.

Ryckaert,J.-P., Ciccotti,G. and Berendsen,H.J.C. (1977)J. Comput. Phys., 23,327–341.Acknowledgements

Schrag,J.D. and Cygler,M. (1993)J. Mol. Biol., 230,575–591.Computations were performed at Novo-Nordisk A/S. G.H.P. would like to Svendsen,A. (1994a) In Woolley,P. and Petersen,S.B. (eds),Lipases—Theiracknowledge financial support from European grant BIOZ-CT93-5507 and Structure, Biochemistry and Applications. Cambridge University Press,the University of Copenhagen. Cambridge, pp. 1–21.

Svendsen,A. (1994b)Inform, 351,619–623.Uppenberg,J., Hansen,M.T., Patkar,S. and Jones,T.A. (1994)Structure, 2,

References 293–308.Van Aalten,D.M.F., Amadei,A., Linssen,A.B.M., Eijsink,V.G.H. and Vriend,G.Allen,M.P. and Tildesley,D.J. (1989)Computer Simulation of Liquids.

(1995)Proteins: Struct. Funct. Genet., 22, 45–54.Clarendon, Oxford.Van Aalten,D.M.F., Amadei,A., Bywater,R., Findlay,J.B.C., Berendsen,H.J.C.,Amadei,A., Linssen,A.B.M. and Berendsen,H.J.C. (1993)Proteins: Struct.

Sander,C. and Stouten,P.F.W. (1996a)Biophys. J., 70, 684–692.Funct. Genet., 17, 412–425.Van Aalten,D.M.F., Findlay,J.B.C., Amadei,A. and Berendsen,H.J.C. (1996b)Ashbaugh,H.S. and Paulaitis,M.E. (1992)J. Phys. Chem., 100,1900–1913. Protein Engng, 8, 1129–1135.

Berendsen,H.J.C., Postma,J.P.M., van Gunsteren,W.F. and Hermans,J. (1981)Van Gunsteren,W.F. and Berendsen,H.J.C. (1987)GROMOS: GroningenIn Pullman,A.et al. (eds)Membrane Proteins: Structures, Interactions and Molecular Simulation Computer Program Package. University ofModels. Kluwer Academic Publishers, Leiden University, The Netherlands, Groningen, The Netherlands.pp. 457–470. Van Kuiken,B.A. and Behnke,W.D. (1994)Biochimica et Biophysica Acta,

Bernstein,F.C., Koetzle,T.F., Williams,G.J.B., Meyer,E.F., Brice,M.D., 1214, 148–160.Rogers,J.R., Kennard,O., Shimanouchi,T. and Tasumi,M. (1977)J. Mol. Vriend,G. (1990)J. Mol. Graph., 8, 52–56.Biol., 112,535–542. Winkler,F.K., D’Arcy,A. and Hunziker,W. (1990)Nature, 343, 771–774.

Boel,E., Huge-Jensen,B., Christensen,M., Thim,L. and Fiil,N. (1988)Lipids, Wulfson,E.J. (1994) In Woolley,P. and Petersen,S.B. (eds),Lipases—Their23, 701–706. Structure, Biochemistry and Applications. Cambridge University Press,

Boland,W., Frobel,C. and Lorentz,M. (1991)J. Synth. Org. Chem., 12, Cambridge, pp. 271–288.1049–1072.

Received May 22, 1996; revised August 21, 1996; accepted August 23, 1996Brady,L. et al. (1990)Nature, 343,767–770.

158