+ All Categories
Home > Documents > Stochastic theory of interfacial enzyme kinetics: A kinetic Monte Carlo study

Stochastic theory of interfacial enzyme kinetics: A kinetic Monte Carlo study

Date post: 05-Sep-2016
Category:
Upload: biswajit-das
View: 214 times
Download: 1 times
Share this document with a friend
7
Stochastic theory of interfacial enzyme kinetics: A kinetic Monte Carlo study Biswajit Das, Gautam Gangopadhyay S.N. Bose National Centre For Basic Sciences, Block-JD, Sector-III, Salt Lake, Kolkata 700098, India article info Article history: Received 3 June 2011 In final form 15 November 2011 Available online 3 December 2011 Keywords: Stochastic processes Interfacial enzyme kinetics Monte Carlo simulation abstract In the spirit of Gillespie’s stochastic approach we have formulated a theory to explore the advancement of the interfacial enzyme kinetics at the single enzyme level which is ultimately utilized to obtain the ensemble average macroscopic feature, lag-burst kinetics. We have provided a theory of the transition from the lag phase to the burst phase kinetics by considering the gradual development of electrostatic interaction among the positively charged enzyme and negatively charged product molecules deposited on the phospholipid surface. It is shown that the different diffusion time scales of the enzyme over the fluid and product regions are responsible for the memory effect in the correlation of successive turnover events of the hopping mode in the single trajectory analysis which again is reflected on the non-Gaussian distribution of turnover times on the macroscopic kinetics in the lag phase unlike the burst phase kinetics. Ó 2011 Elsevier B.V. All rights reserved. 1. Introduction The study of interfacial enzymatic reaction is gaining increasing importance in biological science as enzyme plays a crucial role as catalyst of lipid metabolism on the membrane and as mediator of cell signaling processes [1]. It is a heterogeneous enzymatic reaction where the rate of the reaction depends on both the mechanical and chemical steps involved. From the experimental observation it is well known that the activity of an interfacial en- zyme is maximum where both the fluid and gel state phospholipid molecules coexist [2,3]. Due to different packing pattern, gel state phospholipid molecules are tightly packed than the fluid state molecules whereby an enzyme adsorbs exclusively in the fluid region and gradually diffuses to the fluid–gel boundary [2,4]. With the progress of the reaction the product molecules i.e., lyso-phos- pholipids and fatty acids are accumulated in the surface and forms a product domain in between the gel and fluid domain [3,5–10]. Usually the product molecules i.e., fatty acids are negatively charged and with increase in size of the product domain the elec- trostatic interaction between the positively charged enzyme and the negatively charged surface also increases. It is observed that the formation of an appreciable size of product domain is respon- sible for the lag-burst kinetics which is characterized by initial slow hydrolysis in the lag phase followed by a sudden increase in activity of the enzyme by two or three orders of magnitude, the burst phase [11–16]. Lag-burst kinetics is the most important macroscopic features of interfacial enzyme kinetics and previously various kinetic analysis had been performed by considering the interactions among the enzyme-phospholipid molecules [12,13]. However, no microscopic theoretical study is found in terms of the dynamical processes by considering the single molecule activ- ity on the phospholipid monolayer. In this work we have studied the macroscopic feature of inter- facial enzyme kinetics starting from the single enzyme activity. At the single molecule level this reaction kinetics is a stochastic pro- cess and the analysis involves single molecule trajectory [17–19]. It is well known that due to thermal hopping an enzyme can come out from the gel domain or it can also keep on doing hydrolysis of successive substrate molecules in the scooting mode as well due to electrostatic binding of the enzyme molecule on the gel sur- face. We have simulated the stochastic processes for the hopping and scooting mode of motion as both the modes are operating at the same time probabilistically depending on the amount of prod- uct formed in the trajectory of a single enzyme to calculate the suc- cessive turnover events which on ensemble averaging gives the macroscopic rate of the reaction by which the lag-burst kinetics can be described. In the spirit of Gillespie’s method [20,21] we have studied the stochastic turnover events due to mechanical and chemical steps of the single enzyme activity. Finally, we have searched for any dynamic correlation which can develop due to the motion of enzyme over various time scales of motion in the differ- ent heterogeneous phases. In what follows we have introduced a model for the hopping mode and scooting mode of motion of the interfacial enzyme kinetics in Section 2. A stochastic formulation and simulation tech- nique is provided in Section 3. Numerical results are discussed in Section 4 by first providing single enzyme activity for some exper- imental parameters and then the ensemble average kinetics in the bulk. Finally the paper is concluded in Section 5. 0301-0104/$ - see front matter Ó 2011 Elsevier B.V. All rights reserved. doi:10.1016/j.chemphys.2011.11.024 Corresponding author. E-mail address: [email protected] (G. Gangopadhyay). Chemical Physics 393 (2012) 58–64 Contents lists available at SciVerse ScienceDirect Chemical Physics journal homepage: www.elsevier.com/locate/chemphys
Transcript

Chemical Physics 393 (2012) 58–64

Contents lists available at SciVerse ScienceDirect

Chemical Physics

journal homepage: www.elsevier .com/locate /chemphys

Stochastic theory of interfacial enzyme kinetics: A kinetic Monte Carlo study

Biswajit Das, Gautam Gangopadhyay ⇑S.N. Bose National Centre For Basic Sciences, Block-JD, Sector-III, Salt Lake, Kolkata 700098, India

a r t i c l e i n f o a b s t r a c t

Article history:Received 3 June 2011In final form 15 November 2011Available online 3 December 2011

Keywords:Stochastic processesInterfacial enzyme kineticsMonte Carlo simulation

0301-0104/$ - see front matter � 2011 Elsevier B.V. Adoi:10.1016/j.chemphys.2011.11.024

⇑ Corresponding author.E-mail address: [email protected] (G. Gangopad

In the spirit of Gillespie’s stochastic approach we have formulated a theory to explore the advancement ofthe interfacial enzyme kinetics at the single enzyme level which is ultimately utilized to obtain theensemble average macroscopic feature, lag-burst kinetics. We have provided a theory of the transitionfrom the lag phase to the burst phase kinetics by considering the gradual development of electrostaticinteraction among the positively charged enzyme and negatively charged product molecules depositedon the phospholipid surface. It is shown that the different diffusion time scales of the enzyme over thefluid and product regions are responsible for the memory effect in the correlation of successive turnoverevents of the hopping mode in the single trajectory analysis which again is reflected on the non-Gaussiandistribution of turnover times on the macroscopic kinetics in the lag phase unlike the burst phasekinetics.

� 2011 Elsevier B.V. All rights reserved.

1. Introduction

The study of interfacial enzymatic reaction is gaining increasingimportance in biological science as enzyme plays a crucial role ascatalyst of lipid metabolism on the membrane and as mediatorof cell signaling processes [1]. It is a heterogeneous enzymaticreaction where the rate of the reaction depends on both themechanical and chemical steps involved. From the experimentalobservation it is well known that the activity of an interfacial en-zyme is maximum where both the fluid and gel state phospholipidmolecules coexist [2,3]. Due to different packing pattern, gel statephospholipid molecules are tightly packed than the fluid statemolecules whereby an enzyme adsorbs exclusively in the fluidregion and gradually diffuses to the fluid–gel boundary [2,4]. Withthe progress of the reaction the product molecules i.e., lyso-phos-pholipids and fatty acids are accumulated in the surface and formsa product domain in between the gel and fluid domain [3,5–10].Usually the product molecules i.e., fatty acids are negativelycharged and with increase in size of the product domain the elec-trostatic interaction between the positively charged enzyme andthe negatively charged surface also increases. It is observed thatthe formation of an appreciable size of product domain is respon-sible for the lag-burst kinetics which is characterized by initialslow hydrolysis in the lag phase followed by a sudden increase inactivity of the enzyme by two or three orders of magnitude, theburst phase [11–16]. Lag-burst kinetics is the most importantmacroscopic features of interfacial enzyme kinetics and previouslyvarious kinetic analysis had been performed by considering the

ll rights reserved.

hyay).

interactions among the enzyme-phospholipid molecules [12,13].However, no microscopic theoretical study is found in terms ofthe dynamical processes by considering the single molecule activ-ity on the phospholipid monolayer.

In this work we have studied the macroscopic feature of inter-facial enzyme kinetics starting from the single enzyme activity. Atthe single molecule level this reaction kinetics is a stochastic pro-cess and the analysis involves single molecule trajectory [17–19]. Itis well known that due to thermal hopping an enzyme can comeout from the gel domain or it can also keep on doing hydrolysisof successive substrate molecules in the scooting mode as welldue to electrostatic binding of the enzyme molecule on the gel sur-face. We have simulated the stochastic processes for the hoppingand scooting mode of motion as both the modes are operating atthe same time probabilistically depending on the amount of prod-uct formed in the trajectory of a single enzyme to calculate the suc-cessive turnover events which on ensemble averaging gives themacroscopic rate of the reaction by which the lag-burst kineticscan be described. In the spirit of Gillespie’s method [20,21] wehave studied the stochastic turnover events due to mechanicaland chemical steps of the single enzyme activity. Finally, we havesearched for any dynamic correlation which can develop due to themotion of enzyme over various time scales of motion in the differ-ent heterogeneous phases.

In what follows we have introduced a model for the hoppingmode and scooting mode of motion of the interfacial enzymekinetics in Section 2. A stochastic formulation and simulation tech-nique is provided in Section 3. Numerical results are discussed inSection 4 by first providing single enzyme activity for some exper-imental parameters and then the ensemble average kinetics in thebulk. Finally the paper is concluded in Section 5.

lag phase

burst phase

Phospholipid

Enzyme

LysophospholipidFatty acid

hopping

diffusive motion chemical reaction

(gel state)

(fluid state)

Phospholipid

(b)(a)

(c)

(d)

scooting

(product)

Enzyme substrate

complex

Fig. 1. (a) An enzyme is attached with the fluid state phospholipid molecules on theLangmuir monolayer. (b) Through the diffusion, the enzyme molecule reaches thegel–fluid interface. (c) After hydrolysing a phospholipid molecule it predominantlyleaves the surface in the lag phase. (d) The enzyme is strictly attached to the surfacewith scooting mode of motion and the burst phase appears.

B. Das, G. Gangopadhyay / Chemical Physics 393 (2012) 58–64 59

2. Movement of the enzyme on the interface: hopping andscooting motion

In this section we have given a probabilistic description of twofamiliar interfacial enzymatic reaction schemes, namely hoppingand scooting mode of motion in terms of the desorption andadsorption probability of a surface bound enzyme as shown inFig. 1. The significant difference between these two mechanismsis that an enzyme moves out from the gel state after completinga Michaelis–Menten cycle along with the diffusion in fluid andproduct region in the hopping mechanism whereas, in scootingmode an enzyme is strictly attached with the phospholipid mono-layer and gradually hydrolyzes the phospholipid molecules. There-fore, the reaction scheme of the hopping mode for one turnovercan be written as

E�f !Kd1 E�P!

Kd2 E�g þ S �k1

k�1

E�gS!k2 E�gP!k3 E�0g þ P! EðbulkÞ ð1Þ

Here, E�f and E�P are designated as the conformations of an interfacialenzyme in the fluid and the product region, respectively. Actuallythese two conformations are mainly responsible for the diffusivemotion of the enzyme along the fluid and the product region. Tohydrolyze a substrate molecule i.e., a gel state phospholipid mole-cule, an enzyme is first converted into the conformation E�g fromthe conformation E�P so that a substrate molecule binds to the inter-facial enzyme. The other conformations, like E�gS, E�gP, and E�0g aredesignated as the substrate-bound enzyme, the product-bound en-zyme, and the enzyme after just releasing a product molecule,respectively. In the hopping mode an enzyme leaves the surfaceof the monolayer and goes to the bulk. The conformation E repre-sents the free enzyme in the bulk.

Similarly the reaction scheme for the scooting mode can bewritten as,

E�g þ S �k1

k�1

E�gS!k2 E�gP!k3 E�0g þ P ð2Þ

and

E�0g !d

E�g:

The terms E�g;E�gS;E�gP, and E�0g carry similar meaning as described in

the above case. In this mechanism, the conformation E�0g changesquickly to the conformation E�g and the enzyme starts another turn-over cycle.

From the above two mechanisms it is observed that a surfacebound enzyme may desorb or reside at the monolayer after com-pletion of a Michaelis–Menten turnover cycle. Therefore, at any in-stant these two modes compete each other and this competitionstrictly depends on the electrostatic interaction between the posi-tively charged enzyme and the negatively charged surface. Here,the desorption probability of an enzyme is designated as pd andthe adsorption probability is pa, where pd = (1 � pa). With increasein the fraction of negatively charged product molecules, h, the elec-trostatic energy as well as the adsorption probability, pa, of an en-zyme increases with time. If we consider that D(h) is theelectrostatic binding energy developed due to the fraction of prod-uct formed, h at time t, then one can find

dDðhÞ ¼ kDðhÞdh

where k is the proportionality constant and it is unitless. It dependson the polarity of the substrate molecules. Integrating up toh = hburst, at which the interaction energy D(h) reaches a saturationvalue, D(h)burst, we get

DðhÞ ¼ DðhÞburst ½expð�kðhburst � hÞÞ� ð3ÞHere hburst is designated as the fraction of product molecules respon-sible for the burst kinetics. Actually beyond the value of hburst, enzymeis strictly attached to the gel surface and gradually hydrolyzes the gelstate phospholipid molecules in the so called ‘scooting mode’.

Furthermore, the above equation can be written in terms of theadsorption probability as,

pa ¼DðhÞ

DðhÞburst¼ ½expð�kðhburst � hÞÞ� ð4Þ

In the initial stage, when h is very small, pd dominates over pa

and the enzyme follows the hopping mode motion but whenh� hburst, enzyme strictly follows the scooting mode motion aspa dominates over pd. However, at any intermediate time boththe hopping and scooting mode mechanism will be operative withthe respective probabilities pd and pa. Basically it is observed thatpd� pa during the lag period and after the burst the reverse phe-nomenon is occurred. Therefore, the magnitude of the probabilitiespd and pa determine the different enzymatic motion which ulti-mately dictates the macroscopic lag-burst kinetics.

3. Stochastic formulation of interfacial enzyme kinetics andsimulation technique

In this section we have provided a stochastic simulation kineticsfollowing the approach of Gillespie for the chemical and mechani-cal steps of single enzyme activity which upon ensemble averagingwill give the macroscopic rate.

To simulate the stochastic turnover time of a single enzyme, wehave considered the kinetics along its own trajectory. The move-ment of an enzyme along the fluid region is modelled by a twodimensional Brownian motion. The Monte–Carlo method has beenused for simulating two-dimensional Brownian motion in a squareplane of side L as a random walk model in which each displace-ment is of equal length, say l, but in random direction [22]. By thistechnique the mean square displacement, hd2i is calculated by thefollowing relation,

hd2i ¼ 4DMtc ¼ Ml2 ð5Þ

60 B. Das, G. Gangopadhyay / Chemical Physics 393 (2012) 58–64

where D is the diffusion coefficient of the particle, M is the suffi-ciently large Monte–Carlo (MC) steps, l is the length covered bythe particle per MC step and tc is the time interval between two suc-cessive MC steps. The total diffusion time, t can be calculated by therelation t = Mtc, where tc ¼ l2

4D.Here our main interest is to calculate the time, sfluid, required to

cross the fluid region of an arbitrary finite area by an enzyme i.e.,the residence time of the enzyme in that region. For this purposewe have considered that nf number of phospholipid molecules arepresent in the fluid region. If be the area of the head group of aphospholipid molecule is a, then the total area of the fluid regionis (a � nf). Here we assume that the distance between two adjacentmolecules is negligible compared to the dimension of the enzyme.The number of fluid state phospholipid molecules covered by an en-zyme along its trajectory in a turnover is say, n where n is an integerrandom number and that ranges from 1 to nf. Here n is equivalent tothe number of MC steps in Eq. (5). However, to calculate hd2i, theMonte–Carlo steps, M should be large enough but in our case n canbe small. Therefore, the above relation in Eq. (5) can not be directlyuseful in this context. So to calculate the residence time in the fluidregion, sfluid, we have first calculated the area covered by an enzyme,(n � a) and then divided it by the diffusion coefficient of the enzymein the fluid region. Hence sfluid can be written as,

sfluid ¼ðn� aÞ

Dfluid: ð6Þ

Similarly, if we consider that at any time t, m number of productmolecules be present in the product region and the area of thehead group of a product molecule, lyso-phospholipid, be b thenthe residence time of the enzyme in the product region can be ex-pressed as

sprod ¼ðm� bÞ

Dprodð7Þ

where Dprod is the diffusion coefficient of the enzyme in the productregion. The above two steps are mechanical in nature but when theenzyme molecule reaches the gel–fluid boundary it starts perform-ing chemical reaction which is modelled here as Michaelis–Mentensteps. It is true that ideally the unit of rate constant k1 should besecond order, however, we consider the activities in terms of singleenzyme and when the enzyme molecule reaches the gel–fluidboundary after crossing the fluid and product regions it starts per-forming chemical reaction which is modelled here as a Michaelis–Menten type. During a turnover cycle a surface bound enzymecan hydrolyze only one substrate molecule among all the substratemolecules that E�g ‘‘sees’’ on the surface which are closer to its path.If for an enzyme at time t, S(t) number of substrate molecules are

closer to it’s path, then the rate of the reaction, Eþ S!k01 ES will be

k1 ¼ k01 � SðtÞ. Therefore, the rate constant of the reaction becomessecond order as the value of S(t) can vary with time but change inthis number is very small compared to the total substrate moleculespresent in the surface. Therefore, we have taken this number S(t) asa constant value. As a result the rate constant k1 of the above reac-tion is considered as pseudo first order. Except at the very end of thereaction when the accumulation of the product molecule is veryhigh and well dispersed, this approximation can be physically ten-able. In the single enzyme level, the enzyme kinetics becomes sto-chastic in nature. Hence the rate equation can be written in terms ofthe probabilities. To calculate the residence time in the gel state,sgel, for a single enzyme molecule we have considered the followingprobabilistic rate equations [17,19],

dPE�g ðsÞds

¼ �k1PE�g ðsÞ þ k�1PE�gSðsÞ ð8Þ

dPE�gSðsÞds

¼ k1PE�g ðsÞ � ðk�1 þ k2ÞPE�gSðsÞ ð9Þ

and

dPE�gPðsÞds

¼ k2PE�gSðsÞ: ð10Þ

Hence the residence time distribution of an enzyme in the gel stateis,

fgelðsÞ ¼dPE�PðsÞ

ds¼ k2PE�SðsÞ ð11Þ

or,

fgelðsÞ ¼k1k2

2Ae�ðB�AÞs � e�ðBþAÞs� �

ð12Þ

where B = (k1 + k�1 + k2)/2 and A ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiðk1 þ k�1 þ k2Þ2=4� ðk1k2Þ

q.

The average value of sgel is,

hsgeli ¼k1 þ k�1 þ k2

k1k2¼ 1þ KM

k2ð13Þ

where KM is the Michaelis–Menten constant of the correspondingenzyme.

In the next reaction step, E�gP!k3 E�0g , a gel state substrate mole-cule is converted into a product molecule which is valid both forthe hopping and the scooting modes of motion. The time takento complete the reaction is designated here as sconvert, which is arandom quantity at the single enzyme level. As the reaction isuni-molecular, its probability density function must be an expo-nential distribution which can be written as

fconvertðsÞ ¼ k3e�ðk3sÞ: ð14Þ

The average time required for an enzyme to complete the reactionstep is

hsconverti ¼1k3: ð15Þ

To get the random time, sgel from the corresponding residence timedistribution, fgel, we use the inversion generating method and calcu-late the corresponding time in terms of the uniform random num-ber [23]. The value of sgel is,

sgel ¼1

ðB� AÞ lnk1k2

2AðB� AÞr

� �ð16Þ

and sconvert is,

sconvert ¼1k3

ln1r

� �ð17Þ

where r is a uniform random number. (For detailed calculation seeAppendix A.)

From the above discussion, now we can calculate the corre-sponding turnover time i.e., total time required to hydrolyze agel state phospholipid molecule during the hopping or scootingmode which will obviously be stochastic in nature. The turnovertime in the hopping mode is,

shop ¼ sfluid þ sprod þ sgel þ sconvert: ð18Þ

Similarly for the scooting mode the turnover time can be written as,

sscot ¼ sgel þ sconvert: ð19Þ

As the time required to hydrolyze a phospholipid molecule isgreater in the hopping mode, the product formation rate must behigher in the scooting mode. After obtaining the correspondingturnover times for the hopping and scooting modes with theirprobabilistic occurrence with pd and pa, respectively, one can getthe average turnover time of a trajectory for a single enzyme. Thenwe can easily calculate the ensemble average rate of productformation corresponding to the bulk kinetics. If X be the numberof enzymes present in the monolayer and in a particular time

Ef* + S Eg S Eg

* P E*0 PEgEp**

+

Fluidregion

Productregion

Gel region

+ S E S E + PEg* * * 0PE g

*g

Gel region

*

(a)

(b)

kd 1k

k k

d2 3

−1

2

1 2 k3

kk k

k−1

k1E

Bulk

τ τ τgel convert

τgel τ convert

g

Hydrolysis of one phospholipid molecule

molecule

τfluid

Hydrolysis of one phospholipidDiffusion of enzyme

g

Scooting mechanism for a single turnover

Hopping mechanism for a single turnover

prod

Fig. 2. Schematic representation of the interfacial enzyme kinetics reaction in the(a) hopping and (b) scooting mode mechanism for a single turnover. Here kd1

andkd2 are the inverse of sfluid and sprod which represent the diffusion rates in the fluidand product regions, respectively.

B. Das, G. Gangopadhyay / Chemical Physics 393 (2012) 58–64 61

interval, sint the number of product molecules formed by these en-zymes are Q1,Q2, . . . ,QX, respectively then the average macroscopicrate of the reaction per unit enzyme is

vnet ¼hQisint

ð20Þ

where hQi ¼P

iQ i

X , is the average number of product moleculesformed per unit enzyme over the monolayer.

From the earlier theoretical and experimental [17,18,24–28]studies on the kinetics of single enzyme molecule due to conforma-tional fluctuation, the memory effect is explored through auto-cor-relation function among the random turnover times,

CðmÞ ¼ hDsð0ÞDsðmÞihDs2i : ð21Þ

If C(m) = 0 for (m > 0), it is considered that the dynamic correlationis absent in the reaction system and in the presence of dynamic cor-relation, C(m) decays from the initial (m = 1) value.

The stochastic analysis of the turnover time can be performedfrom the detailed calculation of the correlation function of theturnover time for the hopping and scooting modes by using theformula [29],

CðmÞ ¼n2

n�m

Pn�mi¼1 sisiþm �

Pni¼1si

2

nPn

i¼1s2i �

Pni¼1si

2 ð22Þ

where n is the total number of turnovers, si is the time of the ithturnover and m is the number of turnovers separating si and si+m

in the time sequence.

4. Numerical result: from single enzyme stochastic turnovertime to the bulk interfacial kinetics

In this section we have applied the simulation approach to ob-tain the stochastic turnover time of a single enzyme activity forPLA2 enzyme. Then we have calculated the ensemble average rateprofile for the description of the bulk kinetics of lag burst phenom-enon. We have considered the diffusion parameters of the systemfor the stochastic features of the single trajectory of PLA2 enzymewhich is studied recently through wide-field fluorescence micros-copy [2] whereas the chemical rate parameters are taken from theexperimental papers of Berg et al. [3]. An arbitrary set of reactionrate parameters of single enzyme activity [17] are also taken toshow the general validity of our theory.

4.1. Description of the single enzyme kinetics

To describe the kinetic schemes in the hopping and scootingmodes, we have shown the various mechanical and chemical reac-tion steps which is shown in Fig. 2. According to the reactionscheme in the hopping mode, the steps E�f ! E�p, and E�p ! E�g, areoccurred in the fluid and product region. These two reaction stepsare completely mechanical in nature because in these steps onlythe diffusive movement of the enzyme is occurred. The other con-formational states e.g., E�g; E�gS and E�gP must be in the gel region.Therefore, the reaction steps E�g ! E�gS;E�gS! E�g, and E�gS! E�gPare occurred in the gel region. The other more nontrivial reactionstep is, E�g P! E�0g because in this chemical step, a phospholipidmolecule is totally converted into the product molecules, lyso-phospholipid and fatty acid. Therefore, a turnover time can be writ-ten as,

shop ¼ smech þ schem ð23Þ

where smech and schem are designated as the time required to com-plete the mechanical and the chemical steps, respectively.

In the scooting mode, the turnover time can be written as,

sscoot ¼ schem ð24Þ

thereby the time required to hydrolyze a phospholipid molecule isgreater for the hopping mode than the scooting mode. We haveconsidered the diffusion coefficient values along the fluid regionand product region are 3 and 0.2 lm2/sec, respectively. The exper-imental values of the average residence time along the fluid andproduct region is 30 and 220 ms, respectively. For simulation wehave taken the value of k in Eq. (4) is 300, which is unit less. Wehave suitably chosen the distribution of phospholipid moleculesin the fluid region to obtain such average residence times in thefluid and product regions.

When a PLA2 enzyme molecule starts hydrolyzing a substratemolecule, it first forms an intermediate complex E�gS followed byanother intermediate complex, E�gP which finally gives a productmolecule. Fig. 3(a) and (b) are displayed by taking the first set ofparameters values, k1 = 1350.0 s�1, k�1 = 35.0 s�1, k2 = 400.0 s�1

and k3 = 450.0 s�1 [3], whereas Fig. 3(c) and (d) are drawn by con-sidering the second set of rate parameter values, k1 = 120.0 s�1,k�1 = 30.0 s�1, k2 = 40.0 s�1 and k3 = 50.0 s�1. The first set of param-eters values are considered experimentally [3] whereas the secondset of parameters are arbitrarily considered which corresponds tosingle enzyme processes [17]. From Fig. 3(a) and (c), we observethat the probability of remaining in the gel phase i.e., fgel(s), first in-creases then after a certain time interval it decreases which indi-cates that first an intermediate complex, E�gS is formed and thenit starts to convert into another intermediate complex, E�gP. If theassociation rate constant, k1 is large then E�gS complex is formedquickly and it is converted into E�gP complex very soon which isshown in the Fig. 3(a) and (c). As the probability density functionof converting a product molecule i.e., fconvert(s), is an exponentialdistribution, so fconvert(s), decreases with sconvert exponentiallywhich indicates that the intermediate state E�gP is converted intoE�0g state exponentially according to the value of k3.

In what follows, we have calculated the correlation coefficient,C(m) among the turnover times, si and si+m where m is the numberof turnovers separating si and si+m in the time sequence. Fig. 4(a)and (b) are displayed by taking the first set of parameter valueswhereas Fig. 4(c) and (d) are plotted by considering the second

0 0.007 0.0140

100

200

300

0 0.007 0.0140

200

400

0 0.1 0.20

10

20

0 0.05 0.10

30

60

τgel

fge

l

τ

f

ττgel

f f con

vert

conv

ert

convert

convert

gel

(a) (b)

(c) (d))s ni()s ni(

(in s) (in s)

Fig. 3. (a) Plot of fgel versus sgel is displayed with the rate parameters,k1 = 1350.0 s�1, k�1 = 35.0 s�1, k2 = 400.0 s�1 and k3 = 450.0 s�1 (b) fconvert is plottedwith sconvert for the same rate parameters as in Fig. 3(a). Here fconvert is anexponential distribution where k3 is average rate constant of the product formationstep. Similar plots, fgel versus sgel and fconvert versus sconvert are plotted in Fig. 3(c) and(d) by considering the rate parameters k1 = 120.0 s�1, k�1 = 30.0 s�1, k2 = 40.0 s�1

and k3 = 50.0 s�1, respectively.

0 40 80m

0.3

0.6

0.9

C(m

)

0 40 80m

-0.2

0

0.2

0 40 80

m

-0.2

0

0.2

C(m

)

0 40 80

m

-0.02

0

0.02

C(m

)C

(m)

(a) (b)

(c) (d)

Fig. 4. Plot of autocorrelation function, C(m) versus m is given for the turnovertimes, si in hopping mode of motion in Fig. 4(a) and scooting mode of motion inFig. 4(b). Here m is the number of turnovers separating si and si+m in the successiveturnover sequence and the rate parameters k1 = 1350.0 s�1, k�1 = 35.0 s�1,k2 = 400.0 s�1 and k3 = 450.0 s�1, respectively. Similar plots are given in Fig. 4(c)and (d) for hopping and scooting mode of motion by considering the rateparameters k1 = 120.0 s�1, k�1 = 30.0 s�1, k2 = 40.0 s�1 and k3 = 50.0 s�1,respectively.

0 20 40 60 80time (in s)

0

100

200

0 100 200 300time (in s)

0

10

20

v net

net

v

(a)

(b)

Fig. 5. (a) Plot of rate of product formation versus time (in sec) is given with therate parameters, k1 = 1350.0 s�1, k�1 = 35.0 s�1, k2 = 400.0 s�1 and k3 = 450.0 s�1.These rate parameters are taken from the experimental paper of Berg et al. [3]. (b)Same as (a) by considering the rate parameters k1 = 120.0 s�1, k�1 = 30.0 s�1,k2 = 40.0 s�1 and k3 = 50.0 s�1. The value of hburst = 0.05. The value of hburst is takenthe same in both the curves. We observe that the rate of product formation isincreased suddenly after the burst.

62 B. Das, G. Gangopadhyay / Chemical Physics 393 (2012) 58–64

set of parameter values. From Fig. 4(b) and (d), we have observedthat the correlation coefficient, C(m) among the turnover times inthe scooting mode, sscoot, fluctuates around zero. Therefore, onecan conclude that through this mechanism no memory effect isdeveloped in the system. In the scooting mode only chemical reac-tion steps are involved. The chemical reactions are simple Michae-lis–Menten types in which no conformational fluctuation of theenzyme is considered. Such reactions are simply a renewal typeof processes in which no memory effect is observed [25,30]. How-ever, turnover times in hopping mode, shop, are strongly correlatedwhich is observed in Fig. 4(a) and (c). The correlation indicates that

during the hopping mode mechanism a memory effect is devel-oped in the system. In the hopping mode both the mechanicaland chemical steps are involved. As we have observed that nomemory effect is developed due to the chemical reactions, wecan conclude that the mechanical movements i.e., variable diffu-sion coefficients of the enzyme along the fluid and product regionscreate this memory effect. Recently, Cao et al. had shown that if aBrownian particle travels through two distinct diffusive areas ofvarious sizes and geometrical arrangements i.e., diffusion alongthe heterogeneous environments, then a memory effect is devel-oped for such processes [31]. A similar quantitative model can bebuilt up here to understand the more detailed statistics of thememory effect which can in principle be observed on the single en-zyme trajectory.

4.2. Description of bulk properties of interfacial kinetics

We have calculated the rate of product formation vnet and ob-serve a certain enhancement of product formation rate after theburst which is shown in the Fig. 5(a) and (b). As the time requiredto hydrolyze a phospholipid molecule is greater during the lagphase, the product formation rate must be higher at the burstphase. If the lag phase is absent then the curve of product forma-tion rate versus time be a traditional hyperbolic one and that isusually observed from the ensemble average kinetics experimentscarried out by Berg et al. [3].

As the turnover time is random in nature the statistical featurescan be understood from its distribution. Here we have calculatedthe probability distribution function of such random quantity,P(sturnover) and this distribution is non-Gaussian in nature whichis expected for general non-Markovian process. From Fig. 6(a)and (b) we see that with increase in the value of hburst, the maxi-mum height of the distribution curve decreases which indicatesthat more substrate molecules are hydrolyzed by the hoppingmode of motion. From the correlation of successive turnover timesin single trajectory analysis in Fig. 4, it is found that the correlationarises in the hopping mode which again is predominantly presentin the lag phase in macroscopic reactions. In the lag phase turnovertimes are larger and the values of the turnover times are sparsely

0 0.01 0.020

0.05

0.1

0 0.1 0.20

0.03

0.06θ

burst= 0.05

θθ

burst

burst

= 0.17

= 0.3

P(τ

turn

over

)

P(τ

turn

over

turnoverτ

turnover

(a) (b)θ

burst= 0.05

burstθ = 0.17

burstθ = 0.3

(in s) (in s)

Fig. 6. Probability distribution of sturnover is plotted in Fig. 6(a) for three differentvalues of hburst with the rate parameters k1 = 1350.0 s�1, k�1 = 35.0 s�1,k2 = 400.0 s�1 and k3 = 450.0 s�1, respectively. Similar plot is provided by takingthe rate parameters k1 = 120.0 s�1, k�1 = 30.0 s�1, k2 = 40.0 s�1 and k3 = 50.0 s�1,respectively. The memory effect is reflected on the skewed non-Gaussian distribu-tion towards the range of larger values of turnover characteristic of the lag-phase.

B. Das, G. Gangopadhyay / Chemical Physics 393 (2012) 58–64 63

distributed over a large range and thereby the distribution be-comes skewed towards the higher values of turnover times. How-ever, in absence of various diffusion time scales of enzymaticmotion over the fluid and product phase the distribution wouldtend to be a Gaussian one.

5. Conclusion

In the spirit of Gillespie’s stochastic approach we have provideda kinetic Monte Carlo simulation technique for the study of inter-facial enzyme kinetics which interpolates between the single en-zyme trajectory to that of bulk surface. This trajectory basedanalysis is essential to describe the microscopic details and the sta-tistical features at the single enzyme level which can in principlebe observed by the single molecule fluorescence techniques. Byaveraging over many trajectories we can get the ensemble averageproperties like the lag-burst phenomenon for bulk interfacial en-zyme kinetics.

Our model is based on the experimental observation that thepresence of negatively charged hydrolysis product makes the elec-trostatic binding between the enzyme and the product molecules.We have defined the probabilities of occurrence of both the com-peting processes, namely the thermal hopping and scooting modeof motion which ultimately dictates the preference of an enzyme tochoose one of them at a time. We have observed that after the for-mation of some critical number of product molecules, the enzymegets strictly attached to the surface of the phospholipid and followsthe scooting mode of motion. We have also applied the simulationto get the macroscopic results on the overall kinetic rate to showthe burst followed by a lag period.

From the single trajectory analysis it is found that the varioustime scales of diffusion of the enzyme over the fluid and productregions develop a dynamic correlation among the turnover times.The source of this correlation is very different from the dynamiccorrelation usually observed in single molecule enzymology dueto the conformational fluctuations [24,25,30]. However, it corrobo-rates the fact that if a Brownian particle travels through two dis-tinct diffusive areas of various sizes i.e., diffusion along theheterogeneous environments, then a memory effect is developed[31]. This memory effect is also studied in terms of the distributionof turnover times over the average of many trajectories to obtain amacroscopic impact of this correlation. The memory effect is iden-tified with the range of the lag-phase in the overall rate profilewhich can again be characterized by the non-Gaussian distributionof random time steps in the hopping mode motion. In the lag phaseturnover times are larger than in the burst phase and the values ofthe turnover times are distributed over a larger range and thereby

the distribution becomes skewed towards the higher values ofturnover times. However, in the absence of various diffusion timescales of enzymatic motion over the fluid and product phase thethe distribution would tend to be a Gaussian one characteristicof chemical steps of the process.

This simulation technique can also be applicable to many com-plex biological processes where various mechanical steps are in-volved along with the chemical steps in the overall rate process,e.g., kinetics of the restriction enzyme on a DNA molecule [32–34]. By generalizing the model of thermal hopping we can also ob-tain the effect of temperature and pressure on the turnover rateand lag-burst feature of interfacial enzyme kinetics which is underinvestigation.

Appendix A. Calculation of sgel and sconvert

Here we have calculated sgel and sconvert from the correspondingprobability density function, fgelðsÞ ¼ k1k2

2A ½e�ðB�AÞs � e�ðBþAÞs� andfconvertðsÞ ¼ k3e�ðk3sÞ by the inversion generating method. In theinversion generating method the random numbers from the uni-form distribution in the unit interval is used to construct randomnumbers distributed according to any prescribed probability den-sity function. In this method for generating a random number x,according to a given density function P(x) is simply to draw a ran-dom number ‘r’ from the uniform distribution in the unit intervaland take the value x which satisfies F(x) = r, where FðxÞ ¼

R x�1

Pðx0Þdx0. In other words, we take x = F�1(r), where F(x) is the cumu-lative distribution function of the given probability density func-tion. For calculating the reaction time sgel, we consider theprobability density function is, fgel(s). Hence the cumulativedistribution function, Fgel(s) corresponding to the residence timedistribution function fgel(s) is,

FgelðsÞ ¼ �K

B� A½e�ðB�AÞsgel � 1� þ K

Bþ A½e�ðBþAÞsgel � 1� ð25Þ

where K ¼ k1k22A .

Now according to the inversion method, one can write Fgel(s) = rwhere r is a uniform random number. Substituting the value of rinstead of the term Fgel(s) in the above equation and after simplify-ing we get,

rK� 2A

ðB2 � A2Þ¼ � e�ðB�AÞsgel

ðB� AÞ 1� B� ABþ A

e�2Asgel

� �: ð26Þ

As the value of ‘A’ is large, therefore one can write e�2Asgel ! 0 andwe get the value of sgel as,

sgel ¼1

ðB� AÞ lnk1k2

2AðB� AÞr

� �ð27Þ

where r is a uniform random number. Similarly, the random time,sconvert can be calculated from the probability density functionfconvert(s). The residence time distribution, fconvertðsÞ ¼ k3e�ðk3sconvertÞ

of this event gives the cumulative distribution function,

Fconvert ¼ 1� exp½�k3sconvert�: ð28Þ

According to the inversion method, we can write Fconvert = r, where‘r’ is a uniform random number. Hence r = 1 � exp (�k3sconvert)gives,

sconvert ¼1k3

ln1r

� �: ð29Þ

References

[1] I. Sitkiewicz, K.E. Stockbauer, J.M. Musser, Trends Microbiol. 15 (2007) 63.[2] M. Gudmand, S. Rocha, N.S. Hatzakis, K. Peneva, K. Mullen, D. Stamou, H. Uji-i,

J. Hofkens, T. Bjornholm, T. Heimburg, Biophys. J. 98 (2010) 1873.

64 B. Das, G. Gangopadhyay / Chemical Physics 393 (2012) 58–64

[3] O.G. Berg, B.-Z. Yu, J. Rogers, M.K. Jain, Biochemistry 30 (1991) 7283.[4] S. Rocha, J.A. Hutchison, K. Peneva, A. Herrmann, K. Mullen, M. Skjot, C.I.

Jorgensen, A. Svendsen, F.C. De Schryver, J. Hofkens, H. Uji-i, Chem. Phys. Chem.10 (2009) 151.

[5] O.G. Berg, M.K. Jain, Interfacial Enzyme Kinetics, First ed., Wiley, New-York,2002.

[6] O.G. Berg, M.H. Gelb, M.-D. Tsai, M.K. Jain, Chem. Rev. 101 (2001) 2613.[7] F. Ghomashchi, T. O’Hare, D. Clary, M.H. Gelb, Biochemistry 30 (1991) 7298.[8] F. Ghomashchi, B.-Z. Yu, O. Berg, M.K. Jain, M.H. Gelb, Biochemistry 30 (1991)

7318.[9] M.K. Jain, J. Rogers, O. Berg, M.H. Gelb, Biochemistry 30 (1991) 7340.

[10] M.H. Gelb, M.K. Jain, A.M. Hanel, O.G. Berg, Ann. Rev. Biochem. 64 (1995) 653.[11] T. Bayburt, B.-Z. Yu, H.-K. Lin, J. Browning, M.K. Jain, M.H. Gelb, Biochemistry

32 (1993) 573.[12] T. Honger, K. Jorgensen, R.L. Biltonen, O.G. Mouritsen, Biochemistry 35 (1996)

9003.[13] A. Halperin, O.G. Mouritsen, Eur. Biophys. J. 34 (2005) 967.[14] D.P. Cherney, G.A. Myers, R.A. Horton, J.M. Harris, Anal. Chem. 78 (2006) 6928.[15] C. Leidy, L. Linderoth, T.L. Andresen, O.G. Mouritsen, K. Jorgensen, G.H. Peters,

Biophys. J. 90 (2006) 3165.[16] R. Morales, E. Zavala, M.S. Fernandez, Biochem. Molecular Biol. Int. 44 (1998)

1111.[17] S.C. Kou, B.J. Cherayil, W. Min, B.P. English, X.S. Xie, J. Phys. Chem. B 109 (2005)

19068.

[18] H.P. Lu, L. Xun, X.S. Xie, Science 282 (1998) 1877.[19] W. Xu, J.S. Kong, P. Chen, J. Phys. Chem. C 113 (2009) 2393.[20] D.T. Gillespie, J. Comput. Phys. 22 (1976) 403.[21] D.T. Gillespie, J. Phys. Chem. 81 (1977) 2340.[22] J. Martins, K.R. Naqvi, E. Melo, J. Phys. Chem. B 104 (2000) 4986.[23] W.H. Press, S.A. Teukolsky, W.T. Vetterling, B.P. Flannery, Numerical Recipes in

Fortran, Second ed., Cambridge University Press, 1992.[24] J. Cao, R.J. Silbey, J. Phys. Chem. B 112 (2008) 12867.[25] J. Cao, Chem. Phys. Lett. 327 (2000) 38.[26] X.S. Xie, H.P. Lu, J. Biol. Chem. 274 (1999) 15967.[27] C. Bai, C. Wang, X.S. Xie, P.G. Wolynes, Proc. Natl. Acad. Sci. USA 96 (1999)

11075.[28] B.P. English, W. Min, A.M. van Oijen, K.T. Lee, G. Luo, H. Sun, B.J. Cherayil, S.C.

Kou, X.S. Xie, Nat. Chem. Biol. 2 (2006) 87.[29] A. Molski, Chem. Phys. 352 (2008) 276.[30] S. Yang, J. Cao, J. Chem. Phys. 117 (2002) 10996.[31] J. Cao, Phys. Rev. E. 63 (2001) 041101.[32] V. Namasivayam, R.G. Larson, D.T. Burke, M.A. Burns, Anal. Chem. 75 (2003)

4188.[33] J. Lin, R. Qi, C. Aston, J. Jing, T.S. Anantharaman, B. Mishra, O. White, M.J. Daly,

K.W. Minton, J.C. Venter, D.C. Schwartz, Science 285 (1999) 1558.[34] H.-W. Li, E.S. Yeung, Anal. Chem. 77 (2005) 4374.


Recommended