1. Trang chủ
  2. » Luận Văn - Báo Cáo

Tài liệu Báo cáo Y học: Role of electrostatics in the interaction between plastocyanin and photosystem I of the cyanobacterium Phormidium laminosum ppt

10 674 0
Tài liệu đã được kiểm tra trùng lặp

Đang tải... (xem toàn văn)

THÔNG TIN TÀI LIỆU

Thông tin cơ bản

Tiêu đề Role of electrostatics in the interaction between plastocyanin and photosystem I of the cyanobacterium Phormidium laminosum
Tác giả Beatrix G. Schlarb-Ridley, José A. Navarro, Matthew Spencer, Derek S. Bendall, Manuel Hervás, Christopher J. Howe, Miguel A. De la Rosa
Trường học University of Cambridge
Chuyên ngành Biochemistry
Thể loại Research paper
Năm xuất bản 2002
Thành phố Cambridge
Định dạng
Số trang 10
Dung lượng 523,3 KB

Các công cụ chuyển đổi và chỉnh sửa cho tài liệu này

Nội dung

De la Rosa2 1 Department of Biochemistry and Cambridge Centre for Molecular Recognition, University of Cambridge, UK;2Instituto de Bioquı´mica Vegetal y Fotosı´ntesis, Centro de Investig

Trang 1

Role of electrostatics in the interaction between plastocyanin

Beatrix G Schlarb-Ridley1, Jose´ A Navarro2, Matthew Spencer1, Derek S Bendall1, Manuel Herva´s2, Christopher J Howe1and Miguel A De la Rosa2

1

Department of Biochemistry and Cambridge Centre for Molecular Recognition, University of Cambridge, UK;2Instituto de Bioquı´mica Vegetal y Fotosı´ntesis, Centro de Investigaciones Cientı´ficas Isla de la Cartuja, Universidad de Sevilla y CSIC, Spain

The interactions between photosystem I and five charge

mutants of plastocyanin from the cyanobacterium

Phormi-dium laminosumwere investigated in vitro The dependence

of the overall rate constant of reaction, k2, on ionic strength

was investigated using laser flash photolysis The rate

con-stant of the wild-type reaction increased with ionic strength,

indicating repulsion between the reaction partners

Remov-ing a negative charge on plastocyanin (D44A) accelerated the

reaction and made it independent of ionic strength; removing

a positive charge adjacent to D44 (K53A) had little effect

Neutralizing and inverting the charge on R93 slowed the

reaction down and increased the repulsion Specific effects of

MgCl2were observed for mutants K53A, R93Q and R93E

Thermodynamic analysis of the transition state revealed

positive activation entropies, suggesting partial desolvation

of the interface in the transition state In comparison with plants, plastocyanin and photosystem I of Phormidium laminosumreact slowly at low ionic strength, whereas the two systems have similar rates in the range of physiological salt concentrations We conclude that in P laminosum, in con-trast with plants in vitro, hydrophobic interactions are more important than electrostatics for the reactions of plastocya-nin, both with photosystem I (this paper) and with cyto-chrome f [Schlarb-Ridley, B.G., Bendall, D.S & Howe, C.J (2002) Biochemistry 41, 3279–3285] We discuss the impli-cations of this conclusion for the divergent evolution of cyanobacterial and plant plastocyanins

Keywords: cyanobacteria; electron transfer; photosystem I; plastocyanin; weak interaction

Electron-transfer chains like that of oxygenic

photosyn-thesis impose special restraints on the proteins involved

Reactions must be fast to allow rapid turnover of the

chain Binding between the reaction partners must be

transient, while at the same time sufficient specificity needs

to be retained Surface properties of proteinaceous

reac-tion partners play a crucial role in meeting these criteria

The aim of our research was to increase our

understand-ing of how one property of the protein surface, the charge

pattern, influences the rate constant of the overall reaction

and how it may have evolved Our model protein is

plastocyanin, a soluble photosynthetic redox protein

which accepts an electron from cytochrome f in the

cytochrome bf complex and passes it on to P+700 in

photosystem I In a previous study [1], we mutated

negatively and positively charged residues on the proposed interaction site of plastocyanin with cytochrome f and analysed the reaction of these mutants with the soluble redox-active domain of cytochrome f (Cyt f) in vitro This paper presents results on the interaction in vitro between a representative subset of these charge mutants with the physiological electron acceptor of plastocyanin, photosys-tem I Hence, we can compare two sets of protein–protein interactive surfaces operating in the same compartment with similar functional selection pressures, with the aim of identifying common features

The organism from which plastocyanin and both its reaction partners, Cyt f [1] and photosystem I (this paper), were taken is a moderately thermophilic cyanobacterium, Phormidium laminosum Studying these photosynthetic electron-transfer reactions of cyanobacteria is of evolu-tionary interest: whereas the overall three-dimensional structure of plastocyanin is highly conserved among plants and cyanobacteria, the surface charge pattern varies greatly [1] Comparing cyanobacterial data with the wealth

of information available for the higher plant reaction [2–5] reveals which functional aspects are variable Further-more, the type I copper protein plastocyanin can be replaced by cytochrome c6, a redox protein of similar size but entirely different folding, in a number of eukaryotic algae and cyanobacteria including P laminosum [6,7] Hence two more sets of protein–protein interactive surfaces with the same function as Cyt f – plastocyanin and plastocyanin–photosystem I – are available for identi-fication of features common to interprotein electron-transfer reactions [4,7] To our knowledge, this is the first

Correspondence toB G Schlarb-Ridley, Department of Biochemistry,

University of Cambridge, Building O, The Downing Site,

Cambridge CB2 1QW, UK.

Fax: + 44 1223 333345, Tel.: + 44 1223 333684,

E-mail: bgs9@mole.bio.cam.ac.uk

Abbreviations: Cyt f, soluble redox-active domain of cytochrome f;

k obs , observed first-order rate constant; k on , rate constant of protein

association; k off , rate constant of complex dissociation before electron

transfer has taken place; k et , rate constant of intracomplex electron

transfer; k 2 , bimolecular rate constant of the overall reaction; k ¥ , k 2 at

infinite ionic strength.

(Received 10 June 2002, revised 5 September 2002,

accepted 15 October 2002)

Trang 2

case in which kinetic data of the interaction of

plastocya-nin with both Cyt f and photosystem I have been

collected in a homologous cyanobacterial system This is

essential for informed discussion of evolutionary

relation-ships

The structure and charge properties of plastocyanin have

been described previously in detail (Introduction in [1]) Its

primary electron acceptor is P+700 of photosystem I, a

photo-oxidized chlorophyll a-dimer The crystal structure

of a cyanobacterial photosystem I has been solved at a

resolution of 2.5 A˚ [8] In higher plants, the positively

charged N-terminal lumenal helix of PsaF has been shown

to be involved in binding of plastocyanin [9,10] In

cyanobacteria, deletion of PsaF did not change the kinetics

of photosystem I reduction by either plastocyanin or

cytochrome c6[10,11] Schubert et al [12] suggest that, in

cyanobacteria, subunits PsaA and PsaB are largely

respon-sible for binding plastocyanin or cytochrome c6in a shallow

pocket

In the reaction between photosystem I and

plastocya-nin from different organisms, three different types of

kinetics have been observed, which may represent

vari-ations on a single reaction scheme [13,14] Type I kinetics

are characterized by monophasic decay of the absorbance

of photo-oxidized P+

700 at 820 nm on reduction by plastocyanin, and linear dependence of the observed

pseudo-first-order rate constant kobs on the plastocyanin

concentration This type is observed for weak

interac-tions: in a range of experimentally reasonable

plastocy-anin concentrations, no sign of saturation is apparent

Type II also exhibits monophasic kinetics; however, kobs

approaches a saturating value at high plastocyanin

concentrations, which provides explicit evidence for

complex formation followed by intracomplex electron

transfer Type III shows biphasic kinetics, which provides

evidence for the formation of an additional reaction

complex (compared to Type II) so that rearrangement

must occur before intracomplex electron transfer The

reaction between plastocyanin and photosystem I of

P laminosum is of Type I [7]

Determination of the ionic strength dependence of rates is

an important method of studying electrostatic interactions

[1] The salt commonly added to increase ionic strength is

NaCl However, it has been reported that bivalent cations

can play a specific role in the reaction in vitro between

photosystem I and both plastocyanin [15–17] and

cyto-chrome c6 [13,18–21] by forming electrostatic bridges

between negative charges on the interacting surfaces In

this study, we investigated the dependence of the

second-order rate constant of the overall reaction, k2, on both NaCl

and MgCl2concentration

Information about the thermodynamic parameters of

the transition state can be obtained by measuring the

temperature dependence of k2 This analysis has been

performed for the interactions of plastocyanin and/or

cytochrome c6 with their respective homologous

photo-system I from various plants, green algae and

cyanobac-teria [14,15,19,22] (including P laminosum wild-type [7])

We determined the activation parameters and their

dependence on NaCl and MgCl2 concentration for the

reaction of P laminosum photosystem I with P

lamino-sum plastocyanin wild-type as well as five charge

mutants

M A T E R I A L S A N D M E T H O D S

Molecular biology and mutagenesis Molecular biological methods were essentially as described

by Schlarb-Ridley et al [1]

Protein methods Expression, purification and characterization of wild-type and mutant plastocyanins were carried out essentially as in Schlarb et al [23]

Photosystem I preparations

P laminosum photosystem I particles were obtained by solubilization with b-dodecyl maltoside as described by Ro¨gner et al [24] and Herva´s et al [21] The chlorophyll/

P700ratio of the resulting photosystem I preparation was

150 : 1 The P700 content in photosystem I samples was calculated from the photoinduced absorbance increase at

820 nm using an absorption coefficient of 6.5 mM )1Æcm)1 [25] Chlorophyll concentration was determined by the method of Arnon [26]

Kinetic analysis The second-order rate constant, k2, and its ionic strength dependence were measured using laser-flash-induced absorbance changes of photosystem I at 820 nm Unless stated otherwise, the experimental setup and programmes used in the analysis were as in Herva´s et al [13] The standard experimental conditions were as described by

De la Cerda et al [27] Measurements of the dependence of

kobson the concentration of plastocyanin were carried out in the following buffer: 20 mMtricine/KOH (pH 7.5), 10 mM

MgCl2, 100 lMmethyl viologen and 0.03% (w/v) b-dodecyl maltoside to which photosystem I-enriched particles (0.39 mg chlorophyll per ml) were added The same reaction mixture but without the 10 mM MgCl2 was used for measuring the dependence of k2on ionic strength The ionic strength was adjusted with small aliquots of concentrated solutions of NaCl or MgCl2, and correction was made for the resulting dilution of the reaction mixture All experi-ments were carried out at 278, 283, 288, 293 and 298 K Thermodynamic activation parameters DH, DSand DG were obtained according to the transition state theory by fitting plots of k2/T vs T to the Eyring equation:

k2

T ¼kB

h expðDGz=RTÞ

¼kB

h expðDHz=RTÞ expðDSz=RÞ ð1Þ where kBis the Boltzmann constant, h is the Planck constant, and R is the gas constant Nonlinear regression by the least-squares method gave the standard error of DG To obtain an independent error estimate for each of the correlated parameters DH and DS, the Exhaustive Search Method [28,29] was applied Plots of rate constants, k2, against ionic strength were fitted to the monopole–monopole version of the Watkins equation (Eqn 2) by a nonlinear least-squares method ( TMversion 3.51; Synergy Software):

Trang 3

k2¼ k1exp½Viiexpð0:3295q ffiffi

I

p Þ=ð1 þ 0:3295q ffiffi

I p ð2Þ where q is the radius of the interactive site (in A˚), and the

factor 0.3295 ffiffi

I

p

is the Debye-Hu¨ckel parameter j at 298 K

[30] The allowable error was set to 10)4% For the criteria

used to determine the data range, see the Discussion

Overall errors in the experimental determination of kinetic

constants were estimated to be 10%

Electrostatic potentials

Electrostatic potentials of wild-type and mutant

plastocy-anins in the reduced form were calculated by a finite

difference solution of the linear Poisson–Boltzmann

equa-tion withDELPHIII [31] TheSWISS-PDBVIEWERwas used to

add polar and aromatic ring hydrogens to chain A of pdb

file 1baw, and was also used to introduce mutations Atomic

radii and partial charges were assigned from the PARSE list

of Sitkoff et al [32]

R E S U L T S

Concentration dependence ofkobs

and standard thermodynamic analysis

Five charge mutants of plastocyanin from P laminosum

were chosen for analysis with wild-type photosystem I

isolated from the same organism (Fig 1) All of them were

in a surface patch shown to interact with photosystem I in

the plant case [33] One mutant neutralized a negative

charge (D44A), one neutralized an adjacent positive charge

(K53A), and three neutralized or inverted the charge on R93

(R93A, R93Q, R93E), a residue situated close to the charge

cluster that includes D44 and K53 and at the edge of the

hydrophobic flat end of the protein surrounding the copper

ligand H92 R93 has been shown to be essential for the interaction of plastocyanin with photosystem I in Anabaena [15], and is highly conserved in cyanobacterial plastocya-nins Mutagenesis, expression, purification and character-ization of the plastocyanins has been described [1] Representations of the electrostatic surfaces showing the changes introduced by the mutations are displayed in Fig 1 The decay of the flash-induced absorbance of P+700 at

820 nm was monoexponential for all proteins at each of the five temperatures (278, 283, 288, 293 and 298 K) In the range of concentrations and temperatures used in this study,

kobsshowed no sign of rate saturation The best interpret-ation of the results as a whole was a linear response to plastocyanin concentration through the origin Examples at

293 K and 298 K are shown in Fig 2 Thus wild-type and

Fig 1 Representations of the electrostatic surface potentials of

wild-type and mutant P laminosum plastocyanin drawn with GRASP[50].

The molecular surface (probe radius 1.4 A˚) is coloured according to

electrostatic potential on a scale of red (acidic) to blue (basic) The

orientation is similar to that of Fig 2 of [1].

Fig 2 Dependence of k obs on plastocyanin concentration: wild-type and mutant P laminosum plastocyanin reacting with wild-type P laminosum photosystem I at (A) 293 K and (B) 298 K The data were fitted to the equation k ¼ k [plastocyanin].

Trang 4

all mutants were treated as following kinetic Type I Balme

et al [7] have already reported Type I behaviour for the

wild-type protein From the slopes of the linear regressions

in Fig 2A the bimolecular rate constants for the overall

reaction, k2, were determined (Table 1) The rate constant

increased when a negatively charged residue was neutralized

(D44A), hardly changed when an adjacent positively

charged residue was neutralized (K53A), but decreased

markedly when the charge of R93 was neutralized (R93A,

R93Q), and even more so when it was inverted (R93E) The

results are summarized in Table 1 and are qualitatively

similar to those obtained in the reaction with Cyt f [1]

Balme et al [7] have previously reported a slightly higher

value for k2of the wild-type reaction, and we attribute this

to the use of different photosystem I preparations

The thermodynamic parameters obtained from

tempera-ture-dependence measurements of kobs at 10 mM MgCl2

show that DGdecreases slightly for D44A compared with

wild-type, remains essentially unchanged for K53A, and

increases for all three R93 mutants, most markedly for

R93E (Table 1) Owing to the correlation between DHand

DS, their independent errors, determined by the Exhaustive

Search Method, are large Hence in all but one case (DHof

R93E), DSand DHlie within the 67% confidence interval

of the wild-type values However, the trends parallel those

seen for DG: a decrease relative to wild-type for D44A, no

change for K53A, and an increase for all three R93 mutants,

again most pronounced in R93E It is noteworthy that, with

67% confidence, all DS values are positive under these

conditions Implications for the structure of the transition

state are described in the Discussion

Ionic strength dependence

Response to NaCl The dependence of the second-order

rate constant, k2, on the concentration of NaCl was

investigated at five different temperatures (278, 283, 288,

293 and 298 K) Figure 3 shows the result for all proteins at

298 K; the other temperatures gave analogous results For

wild-type plastocyanin, the rate increased with increasing

salt concentration, as observed by Balme et al [7] This is in

clear contrast with the reaction of wild-type plastocyanin

with Cyt f, where the rate decreases with increasing ionic

strength [1] The mutant D44A showed no dependence on

ionic strength, but K53A reacted slightly more slowly than

wild-type and exhibited a shallower dependence on NaCl concentration R93A and R93Q were slower still with a similar steepness, and again R93E showed the most pronounced effect Experimental results were fitted to the Watkins equation (see Materials and methods), as shown in Fig 3, to obtain estimates of k2at infinite ionic strength (k¥) (Table 1) Modification of charge at positions 44 and 53 had

no significant effect on k¥, but values were significantly lower for mutants of R93

Response to MgCl2 In some systems, enhancement effects have been reported when bivalent rather than univalent cations were used in measurements of ionic strength dependence (see the Introduction) Hence, the dependence

of k2of wild-type and all mutants on the concentration of MgCl2 was investigated at 278, 283, 288, 293 and 298 K

Table 1 Kinetic and thermodynamic parameters of the reaction between wild-type and mutant P laminosum plastocyanin with wild-type P laminosum photosystem I Errors given are either standard errors obtained from curve fitting by least squares (k 2 , k ¥ , DG) or 67% confidence limits derived by the Exhaustive Search Method (DH  , DS  ).

Plastocyanin

k 2 at 298 K a

(l M )1 Æs)1)

k¥at 298 K b

(l M )1 Æs)1)

k¥at 298K c

(l M )1 Æs)1)

DG a

(kJÆmol)1)

DH a

(kJÆmol)1)

DS a

(JÆmol)1ÆK)1) Wild-type 7.1 ± 0.5 10.6 ± 0.5 10.0 ± 0.7 34.06 ± 0.08 40.2 (34.2–46.5) 20.8 (0.3–42.5) D44A 12.1 ± 0.5 10.9 ± 2.1 11.7 ± 0.2 32.66 ± 0.06 37.9 (34.8–41.1) 17.9 (7.3–28.8) K53A 7.8 ± 0.1 12.4 ± 0.7 12.3 ± 1.0 33.74 ± 0.08 39.8 (34.5–45.4) 20.7 (2.5–39.8) R93A 3.3 ± 0.2 5.9 ± 0.5 7.6 ± 1.3 36.00 ± 0.12 47.5 (42.7–52.5) 39.2 (22.9–56.4) R93Q 4.1 ± 0.1 7.0 ± 0.6 6.7 ± 0.5 35.45 ± 0.11 46.2 (44.4–48.0) 36.6 (30.5–42.9) R93E 1.3 ± 0.1 8.5 ± 3.4 3.4 ± 0.6 38.37 ± 0.12 50.4 (47.9–52.9) 40.9 (32.4–49.6)

a Buffer used contained 10 m M MgCl 2 b Buffer contained no MgCl 2 ; ionic strength was adjusted with NaCl The first datapoint was not included in the fit (see Discussion) c Buffer contained no NaCl; ionic strength was adjusted with MgCl 2 The first datapoint was not included

in the fit (see Discussion).

Fig 3 Ionic strength dependence (NaCl) of k 2 : wild-type and mutant

P laminosum plastocyanin reacting with wild-type P laminosum photo-system I at 298 K All measured data points are shown; for the fits to the Watkins equation the first data point was excluded (see Discus-sion) Values for k obtained from the fit are given in Table 1.

Trang 5

Wild-type plastocyanin showed little or no significant

difference in its ionic strength dependence whether NaCl

or MgCl2was used (Fig 4A; this has also been reported in

[7]) The same was the case for the mutants D44A and

R93A For mutants K53A, R93Q and R93E, however, the

rate constant increased faster with ionic strength when

MgCl2rather than NaCl was added (Fig 4B,C) Figure 4

shows the results obtained at 298 K, and analogous effects

were observed at the other temperatures

Activation parameters

Nonlinear Eyring plots of the effect of temperature on k2at

each salt concentration were used to determine the effect of

ionic strength on the activation enthalpy, entropy and free

energy No significant difference was observed between the

thermodynamics of the NaCl and MgCl2 dependencies

Figure 5A–C shows DH and –TDS at 298 K plotted

against the square root of ionic strength (using MgCl2) for

wild-type, K53A and R93E The noise in the wild-type data

buries any trend, if there is one Although there is still

considerable noise in the K53A data, a trend in both DH

(increasing with ionic strength) and –TDS(decreasing with

increasing ionic strength) is emerging For R93E, this trend

is clear and considerably larger than any noise These trends

have also been observed for DHand –TDSof plastocyanin

and cytochrome c6from Synechocystis sp PCC 6803, and a

trend of opposite sign has been reported for plastocyanin

from Anabaena (each reacting with their respective

homo-logous photosystem I), whereas Anabaena sp PCC 7119

cytochrome c6showed an increase for both DHand –TDS

[14]

Comparison betweenP laminosum and spinach

The response to ionic strength of the reaction between

P laminosum plastocyanin and photosystem I was in

marked contrast with the behaviour of the homologous

system in spinach A direct comparison of the two systems

at 298 K is shown in Fig 6 (spinach data taken from [14])

Below 100 mMNaCl, the plant system reacted at least one

order of magnitude faster than that of the cyanobacterium,

but with increasing NaCl concentration the difference

diminished; the point of intersection of the two curves can

be extrapolated to 270 mM NaCl Eyring plots can be

used to extrapolate k2 to 318 K [7], the temperature at

which P laminosum is cultured When the resulting data

were plotted together with the spinach data at 298 K (an

acceptable growth temperature for spinach), the point of

intersection moved to 150 mMNaCl To our knowledge,

the ionic strength of the thylakoid lumen has not been

determined Published values of the ionic strength in the

stroma of chloroplasts vary from 130 mM to 200 mM

[34,35], and it seems reasonable to assume that the lumenal

ionic strength lies within a similar range Hence, at

physiological ion concentrations and temperatures, the

plant and cyanobacterial systems show similar rates

D I S C U S S I O N

To our knowledge, the work described here and in the

related publications [1,7] is the first kinetic analysis of the

in vitrointeractions Cyt f–plastocyanin and plastocyanin–

Fig 4 Comparison of ionic strength curves obtained by using NaCl or MgCl 2 : wild-type and mutant P laminosum plastocyanin reacting with wild-type P laminosum photosystem I at 298 K (A) wild-type; (B) K53A; (C) R93Q and R93E.

Trang 6

photosystem I from the same cyanobacterium Here, we will discuss the new results presented in the context of these two directly related publications [1,7]

Concentration and ionic strength dependence Comparison with the reaction Cyt f–plastocyanin [1] In all cases but one, the concentration dependence of kobsfor the mutant plastocyanins reacting with photosystem I showed qualitatively the same effect as that observed in the reaction with Cyt f, i.e neutralizing acidic residues speeded the reaction up, and neutralizing or inverting the charge on basic residues slowed it down (Fig 3 in [1], Fig 2

in this paper) This indicates that the interacting sites on plastocyanin used for the two reactions were similar The exception is the mutant K53A, which was slower than wild-type plastocyanin in reaction with Cyt f, but appeared to be virtually identical with wild-type plastocyanin in reaction with photosystem I This difference may be due to specific effects of Mg2+(see below)

Response to NaCl In the NaCl-based ionic strength dependence of k2, the effects of the mutations relative to wild-type plastocyanin resembled those observed in reaction with Cyt f, confirming that similar interactive sites were used ([1] and Fig 3) However, the wild-type curve differed dramatically between the reaction with photosystem I and with Cyt f Whereas the reaction with Cyt f showed an overall attraction between the reaction partners, the reaction with photosystem I exhibited a repulsion The attraction between wild-type plastocyanin and Cyt f could be virtually abolished by neutralizing a single positive charge (K53A), whereas the repulsion between wild-type plastocyanin and

Fig 5 Ionic strength dependence (MgCl 2 ) of DH  and –TDS  at 298 K.

(A) wild-type; (B) K53A: (C) R93E Error bars indicate 67%

confid-ence limits obtained by the Exhaustive Search Method [28,29].

Fig 6 Comparison between P laminosum and a plant: ionic strength dependence (NaCl) of k 2 for wild-type P laminosum plastocyanin reacting with wild-type P laminosum photosystem I and wild-type spinach plastocyanin reacting with wild-type spinach photosystem I Experimental data were collected at 298 K; for P laminosum the rate constants were also extrapolated to 318 K, the temperature at which

P laminosum is cultured The lines represent an interpolation between the data points.

Trang 7

photosystem I was abolished by neutralizing a single

negative charge (D44A) The plastocyanin-interaction sites

of Cyt f and photosystem I therefore appear to have a

difference in charge of 2

These conclusions can be drawn from the shape of the

data curve without using curve fitting To extrapolate to

infinite ionic strength and so obtain k¥, the monopole–

monopole version of the Watkins equation was applied [30]

The Watkins equation is based on a model that assumes that

only charges of the protein at the active site are relevant, and

these can be represented by a disc of fixed radius and

uniform charge The same analysis has been used for the

interaction between Cyt f and plastocyanin [1], and for an

in-depth discussion of the advantages and limitations of the

Watkins model the reader is referred to [1] As described in

[1], deviations from the overall curvature are observed at low

ionic strength, probably due to changes in Debye length (the

distance over which the electrostatic field around a charge is

reduced to a value of 1/e of what it would have been in the

absence of electrostatic screening, and thus a measure of the

radius of effective electrostatic influence of a charge), which

the Watkins model does not accommodate Hence in both

analyses (Fig 4 in [1], Fig 3 in this paper), omission of

datapoints at low ionic strength led to better fits and more

reliable k¥values (Table 1 in [1] and in this paper) Although

the curves are shallow, making extrapolation more difficult,

and the number of datapoints is smaller than in the case of

the interaction between Cyt f and plastocyanin [1], the

values shown in Table 1 confirm the qualitative conclusion

that changes in position 93 have a more pronounced and

specific effect For all three R93 mutants, k¥is significantly

slower than that of wild-type or the other mutants,

indicating that in addition to the electrostatic effect, which

leads to low rates at low ionic strength, another

nonelectro-static factor, e.g altered structure of the complex, reduces

the rate at infinite ionic strength

Response to MgCl2 For three mutants, the ionic strength

dependence using MgCl2was markedly different from that

using NaCl (Fig 4B,C) For macromolecular systems,

electrostatic theory, such as the Gouy–Chapman theory

applied to a model membrane [36], can predict stronger

effects for bivalent ions compared with univalent ions at

equivalent ionic strength However, the fact that not all

plastocyanins in this study show an enhancement effect

suggests a different cause, e.g binding A bivalent cation

such as Mg2+ can function as a bridge between two

negative charges on two interaction sites more effectively

than a univalent cation, and may thus speed up a reaction

by neutralizing repelling acidic groups Figure 1 shows that,

in comparison with wild-type plastocyanin, where little or

no enhancement effect occurs (Fig 4A), K53A has gained a

strongly acidic region, as one of the two basic residues

counteracting D44 and D45 has been lost Binding of Mg2+

to this region would counteract its repulsive effect, leading

to the enhancement observed The reaction buffer for the

concentration dependence with photosystem I contained

10 mM MgCl2, whereas the buffer for the analogous

reaction with Cyt f contained only NaCl (90 mM[1]) This

may be the explanation for the fact that K53A reacted as

fast as type with photosystem I, but slower than

wild-type with Cyt f in the concentration dependence of kobs The

enhancement seen for R93E (Fig 4C) can be explained in

an analogous way It is less obvious why R93Q exhibits an effect (Fig 4C) whereas R93A does not, especially as the representations of the electrostatic surface of both mutants (Fig 1) show very little difference However, Gln is polar and also protrudes further into the solvent than the hydrophobic Ala Mg2+may bind to the partially negat-ively charged oxygen of Gln, leading to the observed effect Analysis of activation parameters

This analysis was based on the transition state theory of Eyring [37] The interpretation of the activation parameters

in Table 1 depends on whether the reaction is diffusion-limited or activation-diffusion-limited In the former case, the transition state would be that of association (kon) One would then expect to see a fast phase with a rate constant independent of plastocyanin concentration in the experi-mental traces, which was not observed in this study Hence the reaction is likely to be activation controlled, and DG,

DH and DS listed in Table 1 represent more than one transition state, i.e that of binding (konand koff) and that of electron transfer (ket) With the information to hand, the magnitude or even sign of each contribution to the measured parameters cannot be precisely determined It has to be remembered that transition state theory was developed for elementary chemical reaction steps, not for the interaction of macromolecules in solution Furthermore, the magnitude (but not the sign) of the thermodynamic parameters of activation measured with the same experimental setup can vary with different photosystem I preparations (compare Table 2 in [7] and Table 1 in this paper) In what follows we summarize the expected contributions of binding and electron transfer to DHand DSand estimate their relative importance in the light of the data in Table 1

The contributions to DHfrom binding are expected to

be positive as repulsion between the reaction partners has to

be overcome The contribution of solvent effects on DHis determined by the molecular structure of the protein– protein interface and the degree of desolvation in the transition state If this contribution were negative, it would

be expected to be small The nuclear factor of ket also contributes positively to DH(equation 35 in [38,39]) Hence there is no definite source of negative DH, and it is not surprising that the measured values for DHwere positive for wild-type and all mutants (Table 1) However, the situation is different for DS The loss of translational and rotational degrees of freedom on complex formation and the electronic factor make a negative contribution to DS [38] The only source of positive DS is solvent exclusion from the complex interface, as water molecules gain degrees

of freedom when leaving the ordered protein solvation shell and joining the bulk solvent The fact that DSwas positive for wild-type and all mutants (within at least 67% confid-ence; Table 1) indicates that solvent effects play an import-ant role in the interaction between plastocyanin and photosystem I When copper proteins react with small inorganic reagents where little solvent exclusion occurs, DS

is negative [40], but when plastocyanin reacts with Cyt f, involving a relatively large interface [3], it is positive [41] We can conclude that the transition state for binding in the reaction between plastocyanin and photosystem I is parti-ally desolvated The importance of desolvation of the encounter complex in protein–protein association has been

Trang 8

stressed by others [42,43] The increase in both DHand DS

with increasing ionic strength, which was most clearly seen

for R93E (Fig 5C), has also been observed for the

cyanobacterium Synechocystis sp PCC 6803 [14] and the

green alga Monoraphidium braunii These three systems are

characterized by repulsion between plastocyanin and

phot-osystem I

In contrast, both DHand DSdecreased for Anabaena

sp PCC 7119 plastocyanin [14], where plastocyanin and

photosystem I attract each other All reactions mentioned

show no fast phase in their laser flash kinetics; hence they

are expected to be activation controlled and should have

activation parameters that represent transition states of

binding (konand koff) and electron transfer (ket) as discussed

above The qualitative analysis applied above to the values

in Table 1 is, by itself, inadequate to explain the direction of

the changes in DHand DS Most notably, for repelling

reaction partners, an increase in ionic strength would be

expected to decrease the positive DH of binding; any

solvent effect would go in the same direction The inverse is

the case for the attraction in the case of Anabaena Also, if

the structure of the final complex in which electron transfer

takes place remained unaltered with increasing ionic

strength, no significant changes would be expected for the

activation parameters of electron transfer for both P

lami-nosumand Anabaena Hence, in both cases, the measured

trend is the opposite of what one would expect

The observed effect could be explained, however, if one

assumed that increasing ionic strength modifies the structure

of the final complex [17,19] In the case of P laminosum, the

final complex may be tighter at high ionic strength, as

repelling electrostatic forces between the reaction partners

have been screened out In this case, the activation

parameters associated with koffand ket would experience

additional changes For koff, a tighter final binding complex

would mean that both DHand DSbecome more negative,

as a greater degree of re-solvation implies liberation of extra

solvation energy and loss in degrees of freedom for the

additional molecules of bulk solvent re-solvating the

protein The overall effect on the measured parameters

would be an increase of both DHand DS This would have

to overcompensate the decrease in DH and DS with

increasing ionic strength predicted for kon For Anabaena

plastocyanin, the converse of the above would lead to the

measured decrease for DHand DS

Modifications in the structure of the final complex may

also influence the transition state of electron transfer The

most important effects are likely to be on the electronic

factor If the complex became tighter with increasing ionic

strength, this contribution to DS would become less

negative The inverse would apply for Anabaena Hence

for both organisms, the changes in DS of the electronic

factor would have the same sign as the measured trend

Further discussion of the effects of ionic strength on

thermodynamic parameters can be found in Dı´az et al [19]

and Herva´s et al [14]

Evolutionary implications

Comparison of Cyt f–photosystem I In the case of the

reaction of plastocyanin with Cyt f, it is clear that the acidic

residues D44 and D45 slow the reaction down and that R93

has a more specific role than K46 or K53 The experiments

reported here were intended to clarify if R93 has the same specific role in reaction with photosystem I (as is the case for the analogous residue in Anabaena sp PCC 7119, R88 [15]) and if the interaction with photosystem I is the reason for having acidic residues in the interface (especially the better conserved D44; see alignment in [44]) The results of this study indicate that similar interaction sites are being used for both the reactions with Cyt f and photosystem I R93 has a specific role in both interactions, and D44 slows the reaction down in both cases Electrostatics seem to play a minor role in both reactions

Whereas it is not surprising that R93 is being used in both interactions, one may ask why the acidic residues (especially D44, see above) have been conserved Reasons for conser-ving surface residues fall into two broad categories The first category comprises external reasons such as modification

of interactions with other proteins (enhancing the interac-tion with reacinterac-tion partners, discouraging unfavourable contacts) or with the membrane (avoiding sticking to it while enabling two-dimensional diffusion) D44 clearly does not enhance the reaction with either Cyt f or photosystem I, but may do so with cytochrome oxidase, another potential reaction partner of plastocyanin in cyanobacteria It remains to be clarified if the acidic residues serve to discourage futile reactions (e.g with photosystem II) and/or modify interactions with the membrane The second category of causes for conserving surface residues comprises

internal reasons, for example to enhance stability of the protein Networks of surface salt bridges have been reported

to enhance protein stability [45]; this may be one role of the negative charges on D44 and D45 It is interesting to note that the D72K mutant of cytochrome c6from Anabaena sp PCC 7119 shows increased activity (faster rates with its endogenous reaction partners), but decreased stability (C Lange, personal communication)

Comparison of cyanobacterial and plant systems.Figure 6 reveals that, although the reaction between plant plastocy-anin and photosystem I is faster than that of P laminosum

at low ionic strength values, the difference disappears in the region of physiological salt concentrations This is in accordance with the results obtained for the reaction Cyt f–plastocyanin We have argued previously [1] that

in vitrothe plant and cyanobacterial systems reach similar rates in different ways: the plant system uses mainly electrostatic interactions, as indicated by the steep decrease

in the rate constant with increasing ionic strength and its low k8 P laminosum relies on hydrophobic interactions, as indicated by its shallow ionic strength dependence and the high k8 The importance of hydrophobic interactions has also been shown in the plastocyanin–photosystem I system

of Synechocystis [14] and Prochlorothrix [46] It seems reasonable to conclude from Fig 6 by extrapolation that k8

of the plant plastocyanin–photosystem I reaction is lower than that of P laminosum Qualitatively this would lead to the same conclusion as that drawn for the Cyt f–plastocy-anin interaction in vitro The behaviour in vivo, however, which is of key evolutionary relevance, remains to be investigated Studies of both reactions in the green alga Chlamydomonas reinhardtiihave led to the conclusion that, under favourable growth conditions the charge properties of plastocyanin seem to be without kinetic consequence, but may become so under some conditions of stress, thus

Trang 9

accounting for their conservation [10,47–49] The question

remains why and when such a shift between hydrophobic

and electrostatic enhancement of the rate has happened

during evolution One possibility is that primitive

cyano-bacteria at the time of chloroplast origin had poorly defined

interactions between Cyt f and plastocyanin and/or

plasto-cyanin and photosystem I, and subsequently the nature of

the interactions in the two lineages (hydrophobic in

cyanobacteria, electrostatic in chloroplasts) diverged as a

result of different environmental conditions In

cyanobac-teria there may have been a greater degree of exposure to

environmental fluctuations in ionic strength compared with

chloroplasts This is highlighted in the extant

cyanobacte-rium Gleobacter, where the photosynthetic apparatus is

located in the cytoplasmic membrane However, selective

pressures for formation of acidic patches on plastocyanin

have still to be identified

A C K N O W L E D G E M E N T S

We are extremely grateful to Alexis Balme for his expert help and

advice, to O¨rjan Hansson, Wolfgang Haehnel, Hualing Mi, William

Teale and Ju¨rgen Wastl for fruitful discussions, and to Barry Honig for

making available the programs DelPhi and Grasp This work was

supported by the Biotechnology and Biological Sciences Research

Council, UK, the Oppenheimer Fund, University of Cambridge, UK,

Corpus Christi College Cambridge, UK, Ministerio de Ciencia y

Tecnologı´a, Junta de Andalucı´a, Spain and the Research Training

Network TRANSIENT in the Programme Human Potential and

Mobility of Researchers of the European Commission

(HPRN-CT-1999-00095).

R E F E R E N C E S

1 Schlarb-Ridley, B.G., Bendall, D.S & Howe, C.J (2002) The role

of electrostatics in the interaction between cytochrome f and

plastocyanin of the cyanobacterium Phormidium laminosum

Bio-chemistry 41, 3279–3285.

2 Kannt, A., Young, S & Bendall, D.S (1996) The role of acidic

residues of plastocyanin in its interaction with cytochrome f.

Biochim Biophys Acta 1277, 115–126.

3 Ubbink, M., Ejdeba¨ck, M., Karlsson, B.G & Bendall, D.S (1998)

The structure of the complex of plastocyanin and cytochrome f,

determined by paramagnetic NMR and restrained rigid-body

molecular dynamics Structure 6, 323–335.

4 Navarro, J.A., Herva´s, M & De la Rosa, M.A (1997)

Co-evo-lution of cytochrome c 6 and plastocyanin, mobile proteins

trans-ferring electrons from cytochrome b 6 f to photosystem I J Biol.

Inorg Chem 2, 11–22.

5 Hope, A.B (2000) Electron transfers amongst cytochrome f,

plastocyanin and photosystem I Kinetics and mechanisms

Bio-chim Biophys Acta 1456, 5–26.

6 Ho, K.K & Krogmann, D.W (1984) Electron-donors to P700 in

cyanobacteria and algae: an instance of unusual

genetic-vari-ability Biochim Biophys Acta 766, 310–316.

7 Balme, A., Herva´s, M., Campos, L.A., Sancho, J., De la Rosa,

M.A & Navarro, J.A (2002) A comparative study of the thermal

stability of plastocyanin, cytochrome c 6 and photosystem I in

thermophilic and mesophilic cyanobacteria Photosynth Res 70,

281–298.

8 Jordan, P., Fromme, P., Witt, H.T., Klukas, O., Saenger, W &

Krauss, N (2001) Three-dimensional structure of cyanobacterial

photosystem I at 2.5 A˚ resolution Nature (London) 411, 909–917.

9 Wynn, R.M & Malkin, R (1988) Interaction of plastocyanin with

photosystem I: a chemical cross-linking study of the polypeptide

that binds plastocyanin Biochemistry 27, 5863–5869.

10 Hippler, M., Reichert, J., Sutter, M., Zak, E., Altschmied, L., Schro¨er, U., Herrmann, R.G & Haehnel, W (1996) The plasto-cyanin binding domain of photosystem I EMBO J 15, 6374– 6384.

11 Xu, Q., Yu, L., Chitnis, V.P & Chitnis, P.R (1994) Function and organization of photosystem I in a cyanobacterial mutant strain that lacks PsaF and PsaJ subunits J Biol Chem 269, 3205– 3211.

12 Schubert, W.D., Klukas, O., Krauss, N., Saenger, W., Fromme, P.

& Witt, H.T (1997) Photosystem I of Synechococcus elongatus at

4 A˚ resolution: comprehensive structure analysis J Mol Biol 272, 741–769.

13 Herva´s, M., Navarro, J.A., Dı´az, A., Bottin, H & De la Rosa, M.A (1995) Laser-flash kinetic analysis of the fast electron transfer from plastocyanin and cytochrome c 6 to photosystem I Experimental evidence on the evolution of the reaction mechan-ism Biochemistry 34, 11321–11326.

14 Herva´s, M., Navarro, J.A., Dı´az, A & De la Rosa, M.A (1996) A comparative thermodynamic analysis by laser-flash absorption spectroscopy of photosystem I reduction by plastocyanin and cytochrome c 6 in Anabaena PCC 7119, Synechocystis PCC 6803, and spinach Biochemistry 35, 2693–2698.

15 Molina-Heredia, F.P., Herva´s, M., Navarro, J.A & De la Rosa, M.A (2001) A single arginyl residue in plastocyanin and cyto-chrome c 6 from the cyanobacterium Anabaena sp PCC 7119 is required for efficient reduction of photosystem I J Biol Chem.

276, 601–605.

16 Haehnel, W., Hesse, V & Pro¨pper, A (1980) Electron transfer from plastocyanin to P700 Function of a subunit of photosystem

I reaction center FEBS Let 111, 79–82.

17 Takabe, T., Ishikawa, H., Niwa, S & Itoh, S (1983) Electron-transfer between plastocyanin and P700 in highly-purified photosystem-I reaction center complex: effects of pH, cations, and subunit peptide composition J Biochem (Tokyo) 94, 1901–1911.

18 Medina, M., Dı´az, A., Herva´s, M., Navarro, J.A., Go´mez-Moreno, C., De la Rosa, M.A & Tollin, G (1993) A comparative laser-flash absorption spectroscopy study of Anabaena PCC 7119 plastocyanin and cytochrome c 6 photooxidation by photosystem I particles Eur J Biochem 213, 1133–1138.

19 Dı´az, A., Herva´s, M., Navarro, J.A., De la Rosa, M.A & Tollin,

G (1994) A thermodynamic study by laser-flash photolysis of plastocyanin and cytochrome c 6 oxidation by photosystem I from the green alga Monoraphidium braunii Eur J Biochem 222, 1001– 1007.

20 Herva´s, M., De la Rosa, M.A & Tollin, G (1992) A comparative laser flash absorption spectroscopy study of algal plastocyanin and cytochrome c-552 photooxidation by photosystem I particles from spinach Eur J Biochem 203, 115–120.

21 Herva´s, M., Ortega, J.M., Navarro, J.A., De la Rosa, M.A & Bottin, H (1994) Laser flash kinetic analysis of Synechocystis PCC

6803 cytochrome c 6 and plastocyanin oxidation by photosystem I Biochim Biophys Acta 1184, 235–241.

22 De la Cerda, B., Dı´az-Quintana, A., Navarro, J.A., Herva´s, M &

De la Rosa, M.A (1999) Site-directed mutagenesis of cytochrome

c 6 from Synechocystis sp PCC 6803: the heme protein possesses a negatively charged area that may be isofunctional with the acidic patch of plastocyanin J Biol Chem 274, 13292–13297.

23 Schlarb, B.G., Wagner, M.J., Vijgenboom, E., Ubbink, M., Bendall, D.S & Howe, C.J (1999) Expression of plastocyanin and cytochrome f of the cyanobacterium Phormidium laminosum in Escherichia coli and Paracoccus denitrificans and the role of leader peptides Gene 234, 275–283.

24 Ro¨gner, M., Nixon, P.J & Diner, B.A (1990) Purification and characterization of photosystem-I and photosystem-II core com-plexes from wild-type and phycocyanin-deficient strains of the cyanobacterium Synechocystis PCC-6803 J Biol Chem 265, 6189–6196.

Trang 10

25 Mathis, P & Se´tif, P (1981) Near-infrared absorption-spectra of

the chlorophyll-a cations and triplet-state in vitro and in vivo Isr J.

Chem 21, 316–320.

26 Arnon, D.I (1949) Copper enzymes in isolated chloroplasts Plant

Physiol 24, 1–15.

27 De la Cerda, B., Navarro, J.A., Herva´s, M & De la Rosa, M.A.

(1997) Changes in the reaction mechanism of electron transfer

from plastocyanin to photosystem I in the cyanobacterium

Synechocystis sp PCC 6803 as induced by site-directed

mutagenesis of the copper protein Biochemistry 36, 10125–10130.

28 Roelofs, T.A., Lee, C.H & Holzwarth, A.R (1992) Global target

analysis of picosecond chlorophyll fluorescence kinetics from

pea-chloroplasts: a new approach to the characterization of the

pri-mary processes in photosystem-II alpha-units and beta-units.

Biophys J 61, 1147–1163.

29 Holzwarth, A.R (1995) Data analysis of time-resolved

measure-ments In Biophysical Techniques in Photosynthesis (Amesz, J &

Hoff, A.J., eds), pp 75–92 Kluwer Academic Publishers,

Dord-recht.

30 Watkins, J.A., Cusanovich, M.A., Meyer, T.E & Tollin, G (1994)

A parallel plate electrostatic model for bimolecular rate constants

applied to electron transfer proteins Protein Sci 3, 2104–2114.

31 Gilson, M., Sharp, K & Honig, B (1988) Calculating electrostatic

interactions in bio-molecules: Method and error assessment.

J Comput Chem 9, 327–335.

32 Sitkoff, D., Sharp, K.A & Honig, B (1994) Accurate calculation

of hydration free energies using macroscopic solvent models.

J Phys Chem 98, 1978–1988.

33 Young, S., Sigfridsson, K., Olesen, K & Hansson, O¨ (1997) The

involvement of the two acidic patches of spinach plastocyanin in

the reaction with photosystem I Biochim Biophys Acta

Bio-Energetics 1322, 106–114.

34 Hall, D.O (1976) The coupling of phosphorylation to electron

transport in isolated chloroplasts In The Intact Chloroplast

(Barber, J., ed.), pp 135–170 Elsevier North-Holland Biomedical

Press, Amsterdam.

35 Kaiser, W.M., Weber, H & Sauer, M (1983) Photosynthetic

capacity, osmotic response and solute content of leaves and

chloroplasts from Sinacea oleracea under salt stress Z

Pflanzen-physiol 113, 15–27.

36 Barber, J (1982) Influence of surface charges on thylakoid

struc-ture and function Annu Rev Plant Physiol 33, 261–295.

37 Eyring, H (1935) The activated complex in chemical reactions.

J.Chem Phys 3, 107–115.

38 Marcus, R.A & Sutin, N (1985) Electron transfers in chemistry

and biology Biochim Biophys Acta 811, 265–322.

39 DeVault, D (1984) Quantum-Mechanical Tunnelling in Biological Systems, 2nd edn Cambridge University Press, Cambridge.

40 Rosen, P & Pecht, I (1976) Conformational equilibria accom-panying the electron transfer between cytochrome c (P551) and azurin from Pseudomonas aeruginosa Biochemistry 15, 775– 786.

41 Wood, P.M (1974) Rate of electron transfer between plastocya-nin, cytochrome f, related proteins and artificial redox reagents in solution Biochim Biophys Acta 357, 370–379.

42 Camacho, C.J., Weng, Z.P., Vajda, S & Delisi, C (1999) Free energy landscapes of encounter complexes in protein–protein association Biophys J 76, 1166–1178.

43 Gabdoulline, R.R & Wade, R.C (1999) On the protein– protein diffusional encounter complex J Mol Recognit 12, 226–234.

44 Bendall, D.S., Wagner, M.J., Schlarb, B.G., Ubbink, M & Howe, C.J (1999) Electron transfer between cytochrome f and plastocyanin in Phormidium laminosum In The Phototrophic Prokaryotes (Peschek, G.A., ed.), pp 315–328 Plenum Press, New York.

45 Xiao, L & Honig, B (1999) Electrostatic contributions to the stability of hyperthermophilic proteins J Mol Biol 289, 1435– 1444.

46 Navarro, J.A., Herva´s, M., Babu, C.R., Molina-Heredia, F.P., Bullerjahn, G.S & De la Rosa, M.A (1998) Kinetic mechan-isms of PSI reduction by plastocyanin and cytochrome c 6 in the ancient cyanobacteria Pseudoanabaena sp PCC 6903 and Pro-chlorothrix hollandica In Photosynthesis: Mechanisms and Effects (Garab, G., ed.), pp 1605–1608 Kluwer Academic Publishers, Dordrecht.

47 Soriano, G.M., Ponamarev, M.V., Tae, G.S & Cramer, W.A (1996) Effect of the interdomain basic region of cytochrome f on its redox reactions in vivo Biochemistry 35, 14590–14598.

48 Soriano, G.M., Ponamarev, M.V., Piskorowski, R.A & Cramer, W.A (1998) Identification of the basic residues of cytochrome f responsible for electrostatic docking interactions with plastocya-nin in vivo: relevance to the electron transfer reaction in vivo Biochemistry 37, 15120–15128.

49 Farah, J., Rappaport, F., Choquet, Y., Joliot, P & Rochaix, J.D (1995) Isolation of a PsaF-deficient mutant of Chlamydomonas reinhardtii: efficient interaction of plastocyanin with the photo-system I reaction center is mediated by the PsaF subunit EMBO

J 14, 4976–4984.

50 Nicholls, A., Sharp, K.A & Honig, B (1991) Protein folding and association: insights from the interfacial and thermodynamic properties of hydrocarbons Proteins 11, 281–296.

Ngày đăng: 21/02/2014, 01:21

TỪ KHÓA LIÊN QUAN

TÀI LIỆU CÙNG NGƯỜI DÙNG

TÀI LIỆU LIÊN QUAN