• No results found

Dimers of Azurin as model systems for electron transfer Jongh, Thyra Estrid de

N/A
N/A
Protected

Academic year: 2021

Share "Dimers of Azurin as model systems for electron transfer Jongh, Thyra Estrid de"

Copied!
193
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Dimers of Azurin as model systems for electron transfer

Jongh, Thyra Estrid de

Citation

Jongh, T. E. de. (2006, September 12). Dimers of Azurin as model systems for

electron transfer. Retrieved from https://hdl.handle.net/1887/4554

Version: Corrected Publisher’s Version

License: Licence agreement concerning inclusion of doctoralthesis in the Institutional Repository of the University of Leiden

Downloaded from: https://hdl.handle.net/1887/4554

(2)

Proefschrift

ter verkrijging van

de graad van Doctor aan de Universiteit Leiden, op gezag van de Rector Magnificus Dr.D.D.Breimer,

hoogleraar in de faculteit der Wiskunde en Natuurwetenschappen en die der Geneeskunde,

volgens besluit van het College voor Promoties te verdedigen op dinsdag 12 september 2006

klokke 13.45 uur

door

Thyra Estrid de Jongh

geboren te Amersfoort in 1978

(3)

Promotiecommissie

Promotor: Prof. Dr. G.W. Canters Co-promotor: Dr. M. Ubbink

Referent: Prof. Dr. O. Farver (DFU Kopenhagen) Overige leden:

Prof. Dr. O. Einsle (Georg-August Universität Göttingen) Prof. Dr. M.T.M. Koper

Prof. Dr. J. Brouwer

Thyra de Jongh, Amsterdam 2006

Printed by: PrintPartners Ipskamp B.V., Enschede

Cover: Cartoon representation of a close-up of the interface between two molecules of wild type azurin (P. aeruginosa). The copper atoms are depicted as dark blue spheres. The sidechains of residues His117 are shown in stick representation. The ordered water molecules in the crystal lattice are indicated by semi-transparent blue spheres. (This picture was generated using PyMol 0.98).

(4)

Horizon

Een immer zinvol streven

Een doel dat nooit benaderd worden kon Mijn horizon

Een lijn zo recht en vast getrokken Geen enkel punt dat hem bepalen kon Mijn horizon

Zo vaak verborgen achter duist’re wolken Nooit een die hem verdrijven kon Mijn horizon

Altijd aanwezig en niet te bereiken Geen mens die hem begroeten kon Mijn horizon

Geen vaster punt dat steeds beweegt Geen beter doel dat steeds weer vlucht Ik ga er heen

Uit: Gedichten van een Kleine Zeeman op een Grote Zee Ricor de Jongh

Now, bring me that horizon!

Captain Jack Sparrow, Pirates of the Caribbean

(5)

Abbreviations

1,5-dip 1,5-di(imidazol-1-yl)pentane 1,6-dih 1,6-di(imidazol-1-yl)hexane azu azurin BMME bis-maleimidomethylether DA donor-acceptor pair DTT dithiotreitol ε extinction coefficient

E. coli Escherichia coli

EPR electron paramagnetic resonance

e.s.e. electron self-exchange

ET electron transfer

HSQC heteronuclear single quantum correlation

IPTG isopropyl-b-thiogalactopyranoside

Kd dissociation constant

kese rate of electron self-exchange

kET rate of electron transfer

KPi potassium phosphate buffer

MES 2-morpholinoethanesulfonic acid

NMP N-methylpyrrolidone

NMR nuclear magnetic resonance

ox oxidised

P. aeruginosa Pseudomonas aeruginosa

red reduced

RT room temperature

TCEP tris(carboxyethyl)phosphine

UV-Vis ultraviolet and visible spectroscopy

WT wild type

Amino acids

A Ala Alanine M Met Methione

C Cys Cysteine N Asn Asparagine

D Asp Aspartate P Pro Proline

E Glu Glutamate Q Gln Glutamine

F Phe Phenylalanine R Arg Arginine

G Gly Glycine S Ser Serine

H His Histidine T Thr Threonine

I Ile Isoleucine V Val Valine

K Lys Lysine W Trp Tryptophan

(6)

Table of contents

Chapter 1:

Introduction to electron transfer in a biological context 7

Chapter 2:

Modelling of interprotein electron transfer by covalent and non-covalent

crosslinking 25

Chapter 3:

Electron self-exchange in N42C-BMME azurin dimers as a function of

temperature 47

Chapter 4:

Electron transfer and complex formation in a pH sensitive dimer of

azurin

59

Chapter 5:

Electron transfer in a crosslinked protein dimer mediated by a hydrogen- bonded network across the interface 85

Chapter 6:

Crystal structure of a non-covalent dimer of H117G azurin 111

Chapter 7:

Hotwiring redox proteins by ligand reconstitution using conducting

molecular wires 131

Chapter 8:

General discussion, conclusions and perspectives 159

Summary

170

Samenvatting

174

Appendices:

179

Full colour images 180

Curriculum Vitæ 190

(7)



(8)

affiliations of the authors

Introduction to electron transfer in a

biological context

Chapter

1

(9)



Chapter 1

Abstract

(10)



Chapter 1

Introduction to electron transfer in a biological context

1.1 Electron transfer in a biological context

The generation and subsequent transfer of electrons is a crucial component of a wide variety of biological processes. Once a flow of electrons is generated, for instance by light induced excitation of photoreceptive molecules, it will proceed along a series of acceptors and ultimately results in the formation of energy storage molecules such as ATP. In biology, electron acceptors or donors often are specific organic or inorganic cofactors incorporated into a protein scaffold. By encapsulating the redox centres in insulating protein matrices, nature assures that electrons are not passed on to wanton acceptors but are specifically and efficiently channelled along designated electron transfer (ET) pathways formed by a series of successive redox proteins. Because of the fact that the interacting centres are buried inside the proteins, electron transfer may have to proceed over distances as large as 20 Å, resulting in a relatively weak electronic coupling between the donor and acceptor sites. This long-range electron transfer (LRET) usually relies on electron tunnelling and is governed by the overlap between the wavefunctions of donor (D) and acceptor (A), which in turn is dependent on the DA distance separation as well as on the respective orbital orientations. As the tail of a wavefunction decays exponentially with distance, so does the electronic coupling strength (VDA). The rate of electron transfer (kET) is

directly related to the square of VDA according to Fermi’s rule:

Eq. 1

Here, FC denotes the Franck-Condon factor, a measure of the nuclear motion involved in the ET process. The classical theory for non-adiabatic electron transfer, pioneered in the early seventies by Levich and Dogonadze, allows this term to be approximated to the following expression, commonly referred to as the Marcus equation: [1]

Eq. 2

(11)

10

Chapter 1

parameter λ is the energy that is required to reorganise the ligands and the solvent environment of the protein during ET, while ∆Go is the standard free energy of

the reaction. As is apparent from equation 2, under conditions where -∆G0 < λ, kET increases with increasing driving force (-∆Go) . This behaviour falls into what

is referred to as the ‘normal’ region. Somewhat counter-intuitively though, when -∆G0 > λ, the so-called 'inverted' region, a further increase of the driving force

results in a decrease of kET. The different situations are visualised in Figure 1.1. It is clear that fastest ET is achieved when -∆G0 and λ are perfectly matched so that the

activationless ET is purely under control of the electronic coupling. This parameter is strongly dependent on the intervening medium between the redox centres.

1.2. The protein matrix

Although the actual uptake and donation of electrons is usually restricted to the organic or inorganic cofactors embedded in the protein, the protein matrix itself plays various essential roles in the ET process. As mentioned before, insulating the redox centres inside proteins helps to prevent spurious ET whilst determining the specificity and affinity with which appropriate electron acceptors are recognized by organizing the polypeptide chain into a 3-dimensional structure with electronic,

Figure 1.1: Left: Potential energy diagram for ET between the reactants R resulting

in the formation of products P, indicating the relationship between ET, -∆G0 and

λ. Right: Driving-force dependence of ET rates predicted by semi-classical theory (Eq. 2). Rates increase with driving force until they reach a maximum value (kETo) at

-∆Go=λ. Rates then decrease at higher driving forces (inverted effect). Figure made

(12)

11

Chapter 1

Introduction to electron transfer in a biological context

steric and/or hydrophobic determinants. Another important modulating function of the protein structure concerns the tuning of the redox potential of the reactive centres by adjusting the dielectric properties of the protein medium in the proximity of the redox centre. In this way, similar redox centres can be tuned to produce large differences in redox potentials. Furthermore, by shielding of the redox centre, outer-sphere solvent reorganization effects are greatly reduced. Additional reduction of the inner-sphere reorganization energy can be accomplished by stabilization of the geometry of the redox centre.[3] The lowering of reorganizational energy by

protein encapsulation is clearly illustrated in blue copper proteins, which effectively attenuate the large structural changes associated with reduction of copper complexes in aqueous solution.[4]

1.2.1. Intramolecular ET through protein media

The ways in which the protein matrix participates in mediating LRET has been a topic of interest for many years and several useful theoretical descriptions have arisen, the most widely employed of which will be discussed in more detail here. In electron tunnelling processes, kET is exponentially dependent on the efficiency with which different media are capable of mediating ET, expressed by the electronic distance decay constant b and the distance between donor and acceptor sites d (Eq. 3):

Eq. 3 Values for b can be as high as 3-5 Å-1 in vacuum and range from about 1.7 Å-1 in

water to as little as < 0.1 Å-1 for highly conductive phenylenevinylene molecular

wires [Figure 1.2].[5-7] For proteins, Dutton et al. have determined an averaged

decay factor of 1.4 Å-1 from a series of different protein/cofactor complexes, a

value approximately in between that of covalent systems and vacuum.[8] Although

the approach taken by these authors is often quoted as considering the protein a homogeneous ‘organic glass’-like medium with a single averaged electronic decay constant, it has since been refined to explicitly recognize that protein structure can play a modulating role in the form of an additional parameter termed the packing density ρ.[9] The empirically derived dependence of k

ET on the edge-to-edge DA

(13)

12

Chapter 1

Eq. 4 The distance parameter 3.6 Å in this equation represents the distance at van der Waals interaction at which kET,max ≈ 1013 s-1. Even though the inclusion of a volume

occupancy parameter allows for some deviations from averaged behaviour, this uniform barrier (UB) model assumes that nature does not act by optimising specific ET pathways within a protein structure and does not attribute great importance to individual amino acid residues. In the approach developed by Beratan and co-workers on the other hand, ET is viewed as proceeding along a specific bridging pathway comprising three types of interatomic interactions: covalent bonds (C), hydrogen bonds (H) and through-space contacts (S), each with a distinct decay factor ε.[10-12] The overall electronic coupling for a given tunnelling pathway (TP)

connecting the redox centres is taken as the product of its individual components:

where Eq. 5

V0 is the electronic coupling strength when the donor and acceptor sites would be in direct van der Waals contact. An important prediction from the TP model has been that strongly hydrogen bonded b-sheet structures form more effective mediators of long-range electronic coupling than α-helices do [Figure 1.2]. Based on the pathway analysis approach powerful search algorithms have been developed to identify electronically coupled pathways in known protein structures.[13;14]

These were shown to be very efficient in explaining why in certain ruthenated complexes of azurin differences in ET rate by orders of magnitude were found despite their similar DA distances.[15] More experimental support for the pathway

model has been provided by Farver et al. who observed LRET from the pulse radiolytically reduced intramolecular Cys3,26 disulfide bond to the copper centre in wild type and mutants of azurin (P. aeruginosa) and were thus able to identify distinct intramolecular pathways.[16-18] Although pathway calculations have proven

(14)

13

Chapter 1

Introduction to electron transfer in a biological context

1.2.2. Interprotein ET

All models discussed so far have been developed to treat mostly intraprotein ET and complex dynamics or solvent effects are of relatively little consequence. These factors will however become of major importance when considering intermolecular ET. One problem plaguing the identification of tunnelling pathways between protein complexes is that often the exact structure of the complex is not known or the complex is highly dynamic, not maintaining a single orientation for prolonged times, and the respective orientation of donor and acceptor is not clearly defined. Furthermore, true ET rates may become obscured when diffusional, gating or other regulatory effects become rate limiting. When no direct structural information on the complex is available, either from co-crystallization or from solution NMR studies, possible pathways can sometimes be identified by molecular dynamics simulations.[26] As

will be discussed in more detail in Chapter 2, chemical crosslinking can also be used as a means to render protein complexes amenable to structural analysis. In the pathway approach, intermolecular ET is treated by breaking down the pathway into two intramolecular steps, connecting the donor and acceptor sites to their respective protein edges, and one intermolecular through-space jump which may involve additional solvent interactions. The explicit involvement of solvent molecules in

Figure 1.2: Tunneling timetable for ET

through different media as a function of the edge-to-edge DA distance.[5] The

distance decay for ET through vacuum is shown as a black wedge, that for water as a grey wedge. The dashed lines illustrate the tunneling-pathway predictions for coupling along b-strands (b=1.0 Å-1) and

α-helices (b=1.3 Å-1) respectively, whilst

(15)

14

Chapter 1

mediating interprotein ET is gaining recognition. In this light, the electron self-exchange complex of wild type azurin, which reveals a water-based hydrogen bond network across the protein interface, is often mentioned.[27-29] Analysis of the ET

in crystals of zinc-doped cytochrome c, as well as in encounter complexes of the photosynthetic reaction centre with cytochrome c2 (R. sphaeroides) similarly suggests the involvement of water molecules in electronic coupling between redox centres. [30-32]

1.3. Molecular wires

1.3.1. ET through molecular wires

(16)

15

Chapter 1

Introduction to electron transfer in a biological context

Table 1.1: Dependence of different ET mechanisms on temperature and DA distance

mechanism Temperature (T) DA distance (d)

coherent tunnelling, ‘superexchange’ none exp(-bd)

incoherent, diffusive tunnelling none exp(-bd)

hopping exp(-a/T) d-1

* a and b denote constants independent of temperature or DA distance respectively (table after McCreery et al.[33])

1.3.2. DBA system properties

Molecular wires cover a broad spectrum of molecules and materials, including well-known examples such as carbon nanotubes, DNA and conjugated oligomers. Common oligomer building blocks include ethene, ethyne, thiophene, benzene, pyridine and aniline modules, all of which can be linked together to form extended conjugated chains [Figure 1.3].

Connecting multiple π-systems results in additional electron delocalization and increased conjugation, facilitating the ET process. Although the conventional linear representation of chemical structures may create the illusion of extensive conjugation, in reality the conjugation of a system may become severely distorted by geometric deviations from planarity. Functionalising the individual moieties by incorporating electron withdrawing or donating side-groups can also have a profound effect on the electronic bridge structure, allowing for a certain degree of tunability. It is clear that both geometric and electronic factors are important determinants of the effectiveness of molecular wires. Another important consideration in the design of DBA systems is the contact between the electrodes and the wire.[34] For

metallic electrodes, wires are usually connected either by formation of a thiol bond, chemiadsorption onto the surface -potentially mediated by a self-assembled monolayer (SAM)- or hydrophobic interaction. In order to achieve optimal contact, a good orbital overlap between electrode and bridge is required such that electrons

Figure 1.3: Chemical groups that are frequently encountered as building blocks for

(17)

1

Chapter 1

can move freely into and out of the bridge, avoiding build up of charge at the interface. In general, strong chemical bonds lead to better electronic connection.

1.3.3. Applications of molecular wires in biosensors

The obvious appeal of molecular wires lies in the miniaturisation of electronic circuitry, eventually down to the single molecule level. Particularly interesting applications of molecular wires rest in their potential to effectively transmit protein mediated biosensory signals to a connected electrode or to drive enzymatic reactions in bionanotechnological devices.[35] By definition a biosensor is a biomacromolecule

based system able to detect the presence of a certain substrate and convert the stimulus into an electric signal which can then be relayed to an electrode. Enzymes are exquisitely suited for applications in biosensors due to their high natural sensitivity for specific compounds and their generally fast rate of catalytic turn-over. When binding and the subsequent conversion of a substrate leads to the generation of an electric current, a signal can be transduced to an electrode. However, as touched upon earlier, the protein shell itself is an essentially insulating medium that will impair efficient electronic communication. For relatively small proteins where the distance between the donor and acceptor sites does not exceed ∼25 Å, direct electrooxidation or reduction is still possible but larger molecules may be unable to directly interact with the electrode. The large glucose oxidase (GOx) molecule, which forms the basis of commercial glucose sensors, is a case in point. In order to enhance its efficiency as a biosensor, ferrocene derivatives are attached to the protein that act as electron relay stations.[36] These relay stations serve to reduce the

DA distance between subsequent redox centres and can be anchored to the protein surface by reaction with residues such as histidines or lysines.[37]

(18)

17

Chapter 1

Introduction to electron transfer in a biological context

This strategy has been put to the test in a number of enzyme/cofactor systems. Since the biomedically very interesting GOx enzyme is an FAD containing protein and these prosthetic groups can readily be extracted from the protein by acid precipitation whilst maintaining the native protein fold, reconstituted forms of GOx have been amongst the foremost to be investigated for their ability to transmit an electron flow to a connected electrode. Willner and co-workers employed amine functionalized FAD to tether apo-GOx to gold electrodes modified with a PQQ (pyrroloquinoline quinone) monolayer by treatment with the amine reactive agent EDC (1-ethyl-3-[(dimethylamino)propyl]-carbodiimide) so that the PQQ moiety serves as the mediator for the directional ET from the FAD to the electrode [Figure 1.4A].[48-50] The PQQ anchoring method has similarly been applied to reconstitute

apo-lactic dehydrogenase with modified NAD+ enabling the conversion of lactate

to pyruvate.[51;52]

Although the PQQ molecule exhibits substantial electron delocalization, the overall chain extending between the cofactor and the connected electrode is not strictly speaking considered a molecular wire. In fact, the usage of true molecular wires in conjunction with protein based biosensors is still in its infancy but is beginning to gather some momentum. Recently, Willner and co-workers have created a variation on their GOx sensor by substituting the PQQ relay with a highly conducting polyaniline based wire wrapped around a DNA template [Figure 1.4B].[53] It is to

Figure 1.4: GOx based sensors for the detection of glucose. The electrons generated

(19)

1

Chapter 1

be anticipated that further exploits in this direction are not far behind as molecular wires may well represent the most efficient means for electron transduction in biosensors, provided that the wires can be connected to both the donor and acceptor sites with proper matching of their corresponding energy levels.

Although metal ion ligands are not normally viewed as cofactors, ligand substitution can have many of the same applications as cofactor reconstitution. The H117G mutant of azurin, which forms the basis of the work presented in Chapters 6 and 7, is a prime example. Mutagenesis of the copper ligand His117 for a much smaller glycine residue, creates a solvent exposed aperture in the protein which can then be filled by exogenous molecules such as imidazole or pyridine that ligate to the copper centre.[55-58] If these molecules are connected to a suitable wire structure,

(20)

1

Chapter 1

Introduction to electron transfer in a biological context

References

[1.] R. A. Marcus, N. Sutin, Biochimica et Biophysica Acta 1985, 811 265-322. [2.] H. B. Gray, J. R. Winkler, Quarterly Reviews of Biophysics 2003, 36 341-372. [3.] E. Sigfridsson, M. H. M. Olsson, U. Ryde, Journal of Physical Chemistry

2001, 105 5546-5552.

[4.] J. R. Winkler, P. Wittung-Stafshede, J. Leckner, B. G. Malmstrom, H. B. Gray, Proceedings of the National Academy of Sciences of the United States of

America 1997, 94 4246-4249.

[5.] H. B. Gray, J. R. Winkler, Proceedings of the National Academy of Sciences of

the United States of America 2005, 102 3534-3539.

[6.] A. Ponce, H. B. Gray, J. R. Winkler, Journal of the American Chemical Society

2000, 122 8187-8191.

[7.] W. B. Davis, W. A. Svec, M. A. Ratner, M. R. Wasielewski, Nature 1998,

396 60-63.

[8.] C. C. Moser, J. M. Keske, K. Warncke, R. S. Farid, P. L. Dutton, Nature

1992, 355 796-802.

[9.] C. C. Moser, C. C. Page, X. Chen, P. L. Dutton, Journal of Bioinorganic

Chemistry 1997, 2 393-398.

[10.] D. N. Beratan, J. N. Onuchic, J. N. Winkler, H. B. Gray, Science 1992, 258 1740-1741.

[11.] D. N. Beratan, J. N. Onuchic, J. J. Hopfield, Journal of Chemical Physics

1987, 86 4488-4498.

[12.] D. N. Beratan, J. N. Betts, J. N. Onuchic, Science 1991, 252 1285-1288. [13.] J. N. Betts, D. N. Beratan, J. N. Onuchic, Journal of the American Chemical

Society 1992, 114 4043-4046.

[14.] I. V. Kurnikov, D. N. Beratan, Journal of Chemical Physics 1996, 105 9561. [15.] M. J. Bjerrum, D. R. Casimiro, I.-J. Chang, A. J. Di Bilio, H. B. Gray, M.

G. Hill, R. Langen, G. A. Mines, L. K. Skov, J. R. Winkler, D. S. Wuttke,

Journal of Bioenergetics and Biomembranes 1995, 27 295-302.

[16.] O. Farver, L. J. C. Jeuken, G. W. Canters, I. Pecht, European Journal of

Biochemistry 2000, 267 3123-3129.

(21)

20

Chapter 1

Wherland, I. Pecht, Chemical Physics 1996, 204 271-277.

[18.] O. Farver, I. Pecht, Journal of Biological Inorganic Chemistry 1997, 2 387-392.

[19.] S. S. Skourtis, J. N. Onuchic, D. N. Beratan, Inorganica Chimica Acta 1996,

243 167-175.

[20.] W. B. Curry, M. D. Grabe, I. V. Kurnikov, S. S. Skourtis, D. N. Beratan, J. J. Regan, A. J. A. Aquino, P. Beroza, J. N. Onuchic, Journal of Bioenergetics

and Biomembranes 1995, 27 285-293.

[21.] T. Kawatsu, T. Kakitani, T. Yamato, Journal of Physical Chemistry B 2001,

105 4424-4435.

[22.] S. S. Skourtis, G. Archontis, Q. Xie, Journal of Chemical Physics 2001, 115 9444-9462.

[23.] S. S. Skourtis, I. A. Balabin, T. Kawatsu, D. N. Beratan, Proceedings of the

National Academy of Sciences of the United States of America 2005, 102

3552-3557.

[24.] C. Kobayashi, K. Baldridge, J. N. Onuchic, Journal of Chemical Physics 2003,

119 3550-3558.

[25.] H. Nishioka, A. Kimura, T. Yamato, T. Kawatsu, T. Kakitani, Journal of

Physical Chemistry B 2005, 109 15621-15635.

[26.] F. Musiani, A. Dikiy, A. Y. Semenov, S. Ciurli, Journal of Biological Chemistry

2005, 280 18833-18841.

[27.] K. V. Mikkelsen, L. K. Skov, H. Nar, O. Farver, Proceedings of the National

Academy of Sciences of the United States of America 1993, 90 5443-5445.

[28.] H. Nar, A. Messerschmidt, R. Huber, M. van de Kamp, G. W. Canters,

Journal of Molecular Biology 1991, 221 765-772.

[29.] I. M. C. van Amsterdam, M. Ubbink, O. Einsle, A. Messerschmidt, A. Merli, D. Cavazzini, G. L. Rossi, G. W. Canters, Nature Structural Biology

2002, 9 48-52.

[30.] F. A. Tezcan, B. R. Crane, J. R. Winkler, H. B. Gray, Proceedings of the

National Academy of Sciences of the United States of America 2001, 98

5002-5006.

[31.] O. Miyashita, M. Y. Okamura, J. N. Onuchic, Proceedings of the National

Academy of Sciences of the United States of America 2005, 102 3558-3563.

[32.] O. Miyashita, H. L. Axelrod, J. N. Onuchic, Journal of Biological Physics

(22)

21

Chapter 1

Introduction to electron transfer in a biological context

[33.] R. L. McCreery, Chemistry of Materials 2004, 16 4477-4496.

[34.] S. N. Yaliraki, M. Kemp, M. A. Ratner, Journal of the American Chemical

Society 1999, 121 3428-3434.

[35.] G. Gilardi, A. Fantuzzi, S. J. Sadeghi, Current Opinion in Structural Biology

2001, 11 491-499.

[36.] A. E. G. Cass, G. Davis, G. D. Francis, H. A. O. Hill, W. J. Aston, I. J. Higgins, E. V. Plotkin, L. D. L. Scott, A. P. F. Turner, Analytical Chemistry

1984, 56 667-671.

[37.] A. Heller, Accounts of Chemical Research 1990, 23 128-134.

[38.] I. Hamachi, T. Nagase, Y. Tajiri, S. Shinkai, Bioconjugate Chemistry 1997, 8 862-868.

[39.] I. Hamachi, Y. Tajiri, T. Nagase, S. Shinkai, Chemistry-A European Journal

1997, 3 1025-1031.

[40.] I. Hamachi, S. Shinkai, European Journal of Organic Chemistry 1999, 539-549.

[41.] I. Hamachi, H. Takashima, Y. Z. Hu, S. Shinkai, S. Oishi, Chemical

Communications 2000, 1127-1128.

[42.] T. Hayashi, A. Tomokuni, T. Mizutani, Y. Hisaeda, H. Ogoshi, Chemistry

Letters 1998, 1229-1230.

[43.] T. Hayashi, Journal of Synthetic Organic Chemistry Japan 1998, 56 745-754. [44.] T. Hayashi, Y. Hitomi, H. Ogoshi, Journal of the American Chemical Society

1998, 120 4910-4915.

[45.] T. Hayashi, Y. Hitomi, T. Takimura, A. Tomokuni, T. Mizutani, Y. Hisaeda, H. Ogoshi, Coordination Chemistry Reviews 1999, 192 961-974.

[46.] T. Hayashi, H. Dejima, T. Matsuo, H. Sato, D. Murata, Y. Hisaeda, Journal

of the American Chemical Society 2002, 124 11226-11227.

[47.] T. Hayashi, Journal of Synthetic Organic Chemistry Japan 2002, 60 573-580. [48.] E. Katz, A. Riklin, V. Heleg-Shabtai, I. Willner, A. F. Buckmann, Analytica

Chimica Acta 1999, 385 45-58.

[49.] I. Willner, V. Heleg-Shabtai, R. Blonder, E. Katz, G. L. Tao, Journal of the

American Chemical Society 1996, 118 10321-10322.

[50.] M. Zayats, E. Katz, I. Willner, Journal of the American Chemical Society 2002,

(23)

22

Chapter 1

[51.] E. Katz, V. Heleg-Shabtai, A. Bardea, I. Willner, H. K. Rau, W. Haehnel,

Biosensors & Bioelectronics 1998, 13 741-756.

[52.] A. Bardea, E. Katz, A. F. Buckmann, I. Willner, Journal of the American

Chemical Society 1997, 119 9114-9119.

[53.] L. X. Shi, Y. Xiao, I. Willner, Electrochemistry Communications 2004, 6 1057-1060.

[54.] E. Katz, O. Lioubashevski, I. Willner, Journal of the American Chemical

Society 2005, 127 3979-3988.

[55.] T. den Blaauwen, G. W. Canters, Journal of the American Chemical Society

1993, 115 1121-1129.

[56.] T. den Blaauwen, C. W. G. Hoitink, G. W. Canters, J. Han, T. M. Loehr, J. Sanders-Loehr, Biochemistry 1993, 32 12455-12464.

[57.] T. den Blaauwen, M. van de Kamp, G. W. Canters, Journal of the American

Chemical Society 1991, 113 5050-5052.

[58.] G. van Pouderoyen, T. den Blaauwen, J. Reedijk, G. W. Canters, Biochemistry

(24)

23

Chapter 1

(25)

24

(26)

Chapter

2

Modelling of interprotein electron

transfer by covalent and non-covalent

crosslinking

(27)

2

Chapter 2

Abstract

(28)

27

Chapter 2

Modelling of interprotein electron transfer by covalent and non-covalent crosslinking

Introduction

Many biological processes involve multiple partner proteins. The specificity as well as the affinity with which these form a complex is governed by the requirements to be met. In some cases, such as in antibody-antigen or signal transduction complexes, it will be necessary to form reasonably long lived and stable complexes. In contrast, there are many instances where rather than tight binding, a high rate of turnover is the main prerequisite and complexes will be formed with much lower binding affinities and significantly reduced lifetimes. Clear examples of the latter are found most notably amongst protein complexes involved in electron transfer (ET). The transport of electrons in cells proceeds along a chain comprised of multiple redox proteins, some of which may be membrane-bound, and shuttling of electrons between them is mediated by small soluble electron carrier proteins. The strong demand for fast and efficient ET, as well as the necessity for carrier proteins to be able to specifically interact with more than one partner protein, requires that these complexes are transient and their formation is governed by an interplay between the conflicting demands of sufficient specificity of recognition and that of rapid dissociation. Forces that help balance these needs include geometric surface compatibility, electrostatics or hydrophobic interactions.

To fully appreciate how exactly proteins interact, a detailed description of the surfaces involved in complex formation is needed. The study of ET protein complexes by conventional techniques such as X-ray diffraction is unfortunately hampered by their transient nature. To date very few of these short-lived and weakly binding complexes have been successfully co-crystallized. Most information on redox complex formation has come from NMR studies in which chemical shift perturbation mapping was used to identify protein binding sites and to specify the nature of the interaction.[1-5]

(29)

2

Chapter 2

cytochrome f and plastocyanin[9], cytochrome c and flavodoxin[10], cytochrome c

peroxidase and yeast cytochrome c[11] and several others. This chapter will focus

on the application of crosslinking in the formation of dimers of the redox protein azurin.

2.1. Azurin

2.1.1. Structure and function of azurin (P.aeruginosa)

Azurin is a small, 14 kDa, soluble redox protein belonging to the type I blue copper proteins, also known as cupredoxins. It is found in the periplasmic space of many gram negative bacteria including that of Pseudomonas aeruginosa, the source of the azurin discussed in this thesis. Its in vivo biological function has remained somewhat unclear, having been disproven as an obligatory partner of nitrite reductase (NIR) and as an alternative electron donor to cytochrome c551. It has been suggested that the physiological role of azurin is related to the cellular response to redox stress as increased azurin expression is observed under conditions of oxidative stress.[12] In vitro azurin is capable of interacting with nitrite reductase as well as with various

dehydrogenases.[13-17]

The overall structure of P. aeruginosa azurin consists of two b-sheets connected by an α-helix.[18] Buried about 7 Å beneath the protein surface, a single copper atom is

located that is coordinated by three strong ligands (His117, His46 and Cys112) that form a near-equatorial plane. On either side of the plane the ligation sphere of the copper atom is completed by a weak axial ligand (Gly45 and Met121). His117 is protruding towards the protein surface of an area known as the ‘northern face’ of the protein. Here it is surrounded by a cluster of hydrophobic residues, collectively referred to as the hydrophobic patch (HP) [Figure 2.1].

The presence of a hydrophobic region near the copper site is a conserved feature among all members of the cupredoxin family, although the exact size and amino acid composition of the region may vary.[19] The azurins are, however, unique in that

in all wild type crystal structures two molecules are found with their hydrophobic patches packed closely together [Figure 2.2].[18;20-22] As will be discussed in more

(30)

2

Chapter 2

Modelling of interprotein electron transfer by covalent and non-covalent crosslinking

within the HP have conclusively confirmed its involvement in the e.s.e. reaction of azurin. The introduction of both positive (M44K) or negative charges (M64E) in the HP was shown to lead to a marked pH dependent decrease of kese.[23-25]

In addition to its involvement in e.s.e., the HP has also been identified as the main site of interaction for ET between azurin and cytochrome c551, nitrite reductase and zinc cytochrome c.[24;26] Initially also a second site, located around histidine35, was

proposed as entry and/or exit point for electrons but mutagenesis studies in which the histidine residue was substituted for leucine, glutamine or phenylalanine have shown that this patch is not involved in self-exchange.[15]

Figure 2.1: Representation of P. aeruginosa

azurin seen from the northern face. Residues that make up the hydrophobic patch surrounding the copper ligand histidine117 (dark grey) are shaded and numbered. At the rim of the hydrophobic patch residue 42 (black) is situated. [Full colour image shown on page 180].

Figure 2.2: Cartoon depiction of

(31)

30

Chapter 2

2.1.2 Electron transfer and electron self-exchange in azurin

The relatively small size of azurin, combined with its distinctive spectroscopic properties, stability and high yield of expression have made azurin a classic model protein in the study of intramolecular and intermolecular biological ET. Much information on intramolecular long-range ET (LRET) has been obtained from studies on ruthenium-modified azurins in which excitable ruthenium complexes are anchored to residues on the protein surface.[27-32] Photochemical reduction of

the inorganic complex is succeeded by intramolecular ET to the Cu(II) centre. Alternatively, a unique intramolecular disulfide bond (Cys3-Cys26) that is located at the opposite end of the molecule with respect to the copper centre is pulse radiolytically reduced by generated CO2- radicals.[33-41] The free electron is then

intramolecularly transferred from the reduced RSSR- radical anion to the Cu(II)

ion. By mutagenesis of residues in the intervening protein medium, pathways for ET have been identified and the relative importance of specific residues could be determined. Azurin has also been used in a number of studies of intermolecular ET with various acceptors or donors.

A particularly attractive model system for the study of intermolecular ET in biological systems is the electron self-exchange reaction in which electrons are shuttled between the oxidized and reduced forms of a protein. Although this reaction is presumably non-physiological, its intrinsic simplicity has made it the focus of several studies centred on ET between redox proteins, including azurin. For WT azurin, kese is on the order of 106 M-1s-1 and is only weakly dependent on pH

or ionic strength.[42] When compared with e.s.e. rates found for other blue copper

proteins, that observed for azurin is relatively high.[43] This has been attributed to

(32)

31

Chapter 2

Modelling of interprotein electron transfer by covalent and non-covalent crosslinking

of a strongly coupled electronic pathway connecting the copper atoms.[44] Inspired

by the apparent agreement between the e.s.e. behaviour of azurin in solution and its distinctive protein packing observed in the crystal structure of azurin, several crosslinked dimers of azurin have been created to help further the understanding of some of the factors involved in interprotein ET.

2.2 Covalent crosslinking

The most commonly applicable method of covalent crosslinking of proteins is through use of bifunctional chemical crosslinking agents. The reactive groups targeted by such agents typically belong to either one of three distinct groups: 1) amine groups which can be recognized by agents such as glutaraldehyde or 1-ethyl-3-[(dimethylamino)propyl]carbodiimide (EDC), 2) thiol groups that are reactive towards, for instance, maleimides or iodoacetamides or 3) alcoholic groups which can be oxidized into more chemically reactive aldehyde groups by treatment with periodate. The natural abundance of alcoholic or amine containing amino acids, however, makes these less selective targets for crosslinking. In contrast, the low abundance of surface accessible thiols in most biomolecules combined with their high chemical reactivity, makes thiol groups ideal targets for controlled and selective chemical crosslinking. Unique sites for thiol based crosslinking can easily be introduced on a protein surface through site-directed mutagenesis. For azurin, a series of single and double cysteine mutants have been designed and expressed for various purposes [Table 2.1].

Table 2.1: Single and double cysteine mutants of azurin (P. aeruginosa)

Mutant Purpose

Q12C crosslinking with flavodoxinup, attachment of paramagnetic spin label (EPR)[45]

K27C Immobilization on gold surface (STM, AFM)[46], crosslinking[47], attachment of spin labels (EPR)[45]

N42C crosslinking[48;49], attachment of spin label s(EPR)[45], pulse radiolysis[50]

S118C crosslinking[47], attachment of spin labels (EPR)[45]

Q12C/K27C attachment of spin labels (EPR)mip

(33)

32

Chapter 2

2.2.1 Direct crosslinking of azurin through the introduction of surface exposed cysteines

2.2.1.1 N42C azurin

The relationship between the rate of interprotein ET and the structure of the association complex was explored by constructing covalent dimers of azurin that closely mimic the relative protein orientation of the proposed e.s.e. complex. Solvent exposed cysteines were engineered at position 42 on the protein surface enabling crosslinkage by intermolecular disulfide bond formation[48;49]. The crystal structure

of WT azurin shows Asn42 is located at the rim of the HP and formation of an intermolecular disulfide bond between cysteines at this position was expected to bring together the proteins with their hydrophobic patches facing each other and the copper centres positioned sufficiently close for fast e.s.e. [Figure 2.3, Figure 2.4].

Rates of e.s.e. were determined from NMR line broadening of the His46 proton resonance upon addition of small amounts of oxidized protein into a solution of reduced protein, as well as from simulation of the lineshape of the redox sensitive Val31 1Hγ2 resonance using the MEX/MEXICO software package.[51-53] Quite

unexpectedly, it was found that the disulfide crosslinked N42C azurin dimer displayed a very low kese within the dimer (keseintra < 10 s-1). In contrast, a high value

Figure 2.3: Proposed structure of the

e.s.e. complex of WT azurin based on the crystal packing orientation, showing the target site (N42) for covalent crosslinking (top right) by intermolecular disulfide bond formation between opposing monomers (bottom right).

Figure 2.4: Cartoon representation of

(34)

33

Chapter 2

Modelling of interprotein electron transfer by covalent and non-covalent crosslinking

for kese between dimers was determined (keseinter = 4.2 x 105 M-1 s-1), which is less

than an order of magnitude slower than for WT azurin. Elucidation of the crystal structure of the N42C disulfide bridged dimer offered a clear explanation for this initially surprising behaviour. It shows a dimer in which the two subunits are forced apart and the observed intramolecular Cu-to-Cu distance of 25.9 Å significantly exceeds the range of commonly observed donor-acceptor distances in biological systems with high ET rates [Figure 2.5].[54-56] The large distance was confirmed

in solution by pulsed electron-electron-double resonance (DEER) experiments indicating a distance of 26 Å.[57] Apparently, the introduced disulfide bond is too

short and rigid to allow the hydrophobic surfaces of the two monomers to come together to form a productive ET complex. The resulting ‘open’ structure, however, leaves the HP of one of the protein subunits exposed enough for interaction with a second dimer. The crystal structure shows two separate dimers in the asymmetric unit with an intermolecular Cu-to-Cu distance of 15 Å for the two nearest subunits, comparable to the Cu-to-Cu distance for WT azurin, which accounts for the high keseinter.

Figure 2.5: Cartoon depiction of the

(35)

34

Chapter 2

2.2.1.2 S118C azurin

An alternative site for direct crosslinking of azurin was created by replacement of a solvent exposed serine at position 118 by cysteine.[47] This residue is located close

to the HP and is adjacent to the copper coordinating ligand H117 [Figure 2.4]. Through formation of an intermolecular disulfide bond between the introduced cysteines, a short covalent path (H117-C118-C118-H117) connecting the copper centres is created [Figure 2.6].

As for the N42C disulfide linked dimer, the e.s.e. behaviour of S118C disulfide crosslinked azurin was studied by analysis of the redox sensitive Val31 proton resonance. Intriguingly, the S118C dimer was found to display anti-cooperativity in the two-step oxidation reaction of the dimer with a difference in redox potential between the first and second oxidation step of 33 mV, suggesting that the semi-reduced form of the dimer is stabilised over the fully oxidized form. Unlike for the N42C dimer, the position of the Val31 proton resonance of the semi-reduced form of the S118C dimer is slightly shifted compared to that of the fully reduced form suggestive of a small change in the geometry of the reduced copper centre upon oxidation of the opposing monomer. Together these observations imply that oxidation of the first copper centre induces strain in the molecule, disfavouring further oxidation. More indications for a structural rearrangement of the S118C dimer stem from the fact that, compared to other azurin dimers, several changes in resonance positions are detected and a number of NMR peaks are relatively

Figure 2.6: Schematic representation

(36)

35

Chapter 2

Modelling of interprotein electron transfer by covalent and non-covalent crosslinking

broadened. Finally, additional support for the presence of strain arises from the observation that the absorption maximum of the absorption around 630 nm is similarly dependent on the overall oxidation state of the dimer. The e.s.e. within, as well as between, dimers of S118C azurin was found to be very slow. Pathway calculations on the covalent path connecting the copper centres predict a keseintra

of approximately 60 s-1. From the experimental data, however, an upper limit of

only 3 s-1 was determined. This difference by more than an order of magnitude was

accounted for by assuming that the additional strain introduced by the disulfide bond has led to an increase in the reorganization energy of the copper site. The low keseinter is in turn explained in terms of a shielding of the HP as a result of

dimerization.

2.2.1.3 K27C azurin

To further explore the importance of properly orienting the association complex for efficient ET, dimers of azurin were created in which, rather than in a face-on orientation, the proteins were crosslinked at the “southern face” of the protein by replacement of Lys27 for a cysteine. Since position 27 is located at the far end of the molecule with respect to the copper centre, formation of a disulfide bond at this position is expected to result in the formation of a dimer in which the intramolecular Cu-to-Cu distance is maximized to about 60 Å [Figure 2.4]. Consistent with non-adiabatic ET theory, it was indeed found that the intramolecular e.s.e. in this complex is essentially absent (keseintra <10 s-1). The hydrophobic patches of each of the protein

monomers are, however, still accessible for interaction with other dimers, reflected by the fact that the experimentally observed keseinter is comparable to that of WT

azurin (2.6 x 106 M-1s-1) (unpublished results).

2.2.2 Covalent crosslinking through bifunctional thiol reactive spacers

The dimers formed by direct disulfide bond formation of N42C azurin, as well as those of S118C azurin, are examples of cases in which so-called ‘zero-length’ crosslinking leads to the formation of complexes that are not optimised for efficient ET. The use of molecular spacers that selectively react with groups on the protein surface may then help to alleviate the steric strain.

2.2.2.1 N42C-BMME azurin

(37)

bis-3

Chapter 2

maleimidomethylether (BMME) was introduced [Figure 2.8A].[49] Each

thiol-reactive moiety of this bifunctional linker was reacted with a single cysteine on the protein surface. The homodimer thus created was expected to have a higher degree of rotational flexibility around the site of crosslinking, thereby facilitating the formation of a productive intramolecular ET complex. It was found that the BMME crosslinked dimer of N42C azurin indeed displayed a high keseintra ( ≥ 5 x 104 s-1). The

crystal structure of the dimer confirmed the formation of an association complex in which the connected protein monomers are oriented in a fashion reminiscent of the WT crystal packing structure with a Cu-to-Cu distance of 14.6 Å [Figure 2.7]. The experimentally observed keseintra is in good agreement with the theoretically predicted

ET rate (kET = 4 x 105 s-1) for this donor-acceptor distance according to the empirical

model proposed by Dutton which describes kET as a function of the DA separation and the packing density ρ of the intervening medium [Eq. 1]:[56]

Eq. 1

In the dimer interface two hydrogen bonded water molecules were observed connecting to the copper coordinating H117 residues. As mentioned before, the presence of such ordered water molecules was first noted in the crystal packing of WT azurin. Through the formation of a hydrogen bond network these water molecules are thought to create a strong electronic coupling between the copper atoms. The importance of water molecules in the coupling between the redox centres is underscored by calculations of the electronic coupling element which predict a dramatic decrease of kese in the absence of this structurally ordered water.[44;58] Unlike

the direct disulfide crosslinked dimer of N42C azurin, the BMME crosslinked dimer displays a very low keseinter, which can be explained by the shielding of the HP in the

(38)

37

Chapter 2

Modelling of interprotein electron transfer by covalent and non-covalent crosslinking

2.3 Non-covalent crosslinking by cofactor or ligand reconstitution

An intriguing alternative to covalent crosslinking is presented by the fact that many proteins require non-protein based cofactors for their functioning. For many proteins their natural cofactor can be removed through chemical means and the resulting apo-protein can be reconstituted with a suitable artificial cofactor. By replacement of the native cofactor with a functionalized derivative, the spectroscopic and functional properties of a protein can be altered and non-native properties conferred onto the complex.

Hamachi and Hayashi have demonstrated some of the possibilities offered by this method by reconstitution of myoglobins with synthetic protohemin derivatives, thus creating proteins with artificial functionalities such as photoactivation, selective partner recognition, substrate sensing or membrane incorporation.[59-65] Of particular

interest is also the work of Willner et al. who utilised cofactor reconstitution of glucose oxidase and lactic dehydrogenase as a means of establishing electrical contact between the enzyme and an electrode. Furthermore, cofactor reconstitution enables highly specific non-covalent crosslinking of proteins through use of

Figure 2.7: Crystal structure of the N42C-BMME

(39)

3

Chapter 2

bifunctional linkers that are composed of cofactor or ligand-like moieties connected by a flexible spacer.

2.3.1 Dimerization of H117G azurin

The principle of ligand reconstitution has been extended to azurin. It has been shown that it is possible to replace one of the copper ligating histidines, H117, by a glycine without severely disrupting the overall protein structure.[66;67] The removal

of the coordinating imidazole side chain leaves a gap in the coordination sphere of the copper ion that, since H117 normally protrudes towards the protein surface, can be filled by exogenous copper coordinating ligands.[67] Depending on the nature of

the external ligand, the spectroscopic features of the mutant protein resemble either that of the native Type-1 site or that of a Type-2 copper site. In the absence of any external ligands, the coordination sphere of the copper ion is completed by water, resulting in the formation of a Type-2 like copper site.[68] Addition of monodentate

ligands like Cl-, Br-, imidazole (derivatives) or pyridine (derivatives) was shown to

restore the spectroscopic properties of the native Type-1 site.[69;70] Reconstitution

with derivatives of imidazole, an analogue of the natural histidine, allows direct ‘hotwiring’ of the copper site.

Van Pouderoyen et al. have demonstrated how hotwiring can be used to achieve non-covalent crosslinking of azurin.[71;72] Bifunctional molecules were synthesized

in which two imidazole moieties are connected by a flexible alkane chain, the length of which was varied [Figure 2.8B-D]. It was found that addition of these

bis-imidazolyl(CH2)n linkers with either n=5 or n=6 led to dimerization of

Cu(II)-Figure 2.8: Bifunctional linkers

(40)

3

Chapter 2

Modelling of interprotein electron transfer by covalent and non-covalent crosslinking

H117G azurin and to restoration of the spectroscopic WT features. In contrast, the linker with n=4 proved to be too short for efficient dimer formation. EPR spectra of H117G azurin dimerized with 1,5-di(imidazol-1-yl)pentane or 1,6-di(imidazol-1-yl)hexane respectively revealed small, yet distinct differences in the geometry of the copper site. The H117G azurin-1,5-di(imidazol-1-yl)pentane complex exhibited a spectrum composed of two slightly different Type-1 copper sites. Most likely, the linker is unable to bind to two proteins in exactly the same fashion as a result of steric constraints and the two imidazole rings are forced into slightly different orientations with respect to the copper sites. For the somewhat longer 1,6-di(imidazol-1-yl)hexane linker this steric constraint was relieved and the EPR spectrum showed only a single Type-1 species. The work presented in Chapter 6 discusses a crystal structure of Zn-H117G azurin dimerized with 1,6-di(imidazol-1-yl)hexane, the first crystal structure of a metal-bound and ligand-coordinated form of H117G azurin. Such structural information on crosslinked or reconstituted systems is of vital importance for their potential application in biotechnological devices such as those discussed in Chapter 1.

Conclusions and prospects

(41)

40

Chapter 2

The various examples presented in this review stress that care must be taken in the design of efficient crosslinked ET complexes. It has long been thought that the use of short crosslinkers is preferable in the study of redox complexes in order to minimize the distance between the respective redox centres. Often, however, this strategy has resulted in the formation of complexes that proved to be inefficient. The structure and e.s.e. properties of the disulfide crosslinked dimer of N42C azurin illustrate how short linkers can impede the formation of properly oriented ET complexes by restricting the freedom of motion of the partner proteins. Zero-distance crosslinking may even cause a slight distortion of the redox centre itself, as demonstrated by the redox behaviour of S118C azurin. These findings clearly show how even relatively small changes in energy or geometry can have a profound effect on the rate of ET. In such cases, the use of longer and more flexible spacers can relieve the steric constraints and allow the formation of more favourable complexes that exhibit fast ET. The N42C-BMME azurin dimer serves as a good example of this principle. The crystal structures of both wild type and N42C-BMME azurin reveal the presence of ordered water molecules in the protein interface. The formation of a hydrogen bond network connecting the copper centres may be of vital improtance in the efficient electronic coupling of the redox centres as shown by pathway calculations of the azurin dimer that highlight that solvent molecules should be explicitly taken into account when analyzing ET pathways in redox complexes.

(42)

41

Chapter 2

Modelling of interprotein electron transfer by covalent and non-covalent crosslinking

of the H117G-imidazole complex is possible, provided that the rate of reoxidation is faster than the rate of dissociation of imidazole.[73] In principle this implies that,

as long as linkers with sufficiently strong electronic coupling are used, the formation of redox active complexes of H117G azurin is possible.

References

[1.] A. Bergkvist, M. Ejdeback, M. Ubbink, G. Karlsson, Protein Science 2001,

10 2623-2626.

[2.] P. B. Crowley, G. Otting, B. G. Schlarb-Ridley, G. W. Canters, M. Ubbink,

Journal of the American Chemical Society 2001, 123 10444-10453.

[3.] M. Ubbink, M. Ejdeback, B. G. Karlsson, D. S. Bendall, Structure 1998, 6 323-335.

[4.] E. R. P. Zuiderweg, Biochemistry 2002, 41 1-7.

[5.] J. A. R. Worrall, Y. J. Liu, P. B. Crowley, J. M. Nocek, B. M. Hoffman, M. Ubbink, Biochemistry 2002, 41 11721-11730.

[6.] T. A. Alleyne, M. T. Wilson, G. Antonini, F. Malatesta, B. Vallone, P. Sarti, M. Brunori, Biochemical Journal 1992, 287 951-956.

[7.] L. M. Peerey, H. M. Brothers, J. T. Hazzard, G. Tollin, N. M. Kostic,

Biochemistry 1991, 30 9297-9304.

[8.] J. S. Zhou, H. M. Brothers, J. P. Neddersen, L. M. Peerey, T. M. Cotton, N. M. Kostic, Bioconjugate Chemistry 1992, 3 382-390.

[9.] L. Qin, N. M. Kostic, Biochemistry 1993, 32 6073-6080.

[10.] J. L. Dickerson, J. J. Kornuc, D. C. Rees, Journal of Biological Chemistry

1985, 260 5175-5178.

[11.] H. S. Pappa, T. L. Poulos, Biochemistry 1995, 34 6573-6580.

[12.] E. Vijgenboom, J. E. Busch, G. W. Canters, Microbiology-Uk 1997, 143 2853-2863.

[13.] P. Rosen, M. Segal, I. Pecht, European Journal of Biochemistry 1981, 120 339-344.

[14.] M. C. Silvestrini, F. Cutruzzola, R. Dalessandro, M. Brunori, N. Fochesato, E. Zennaro, Biochemical Journal 1992, 285 661-666.

(43)

42

Chapter 2

[16.] S. L. Edwards, V. L. Davidson, Y. L. Hyun, P. T. Wingfield, Journal of

Biological Chemistry 1995, 270 4293-4298.

[17.] Y. L. Hyun, V. L. Davidson, Biochemistry 1995, 34 12249-12254. [18.] H. Nar, A. Messerschmidt, R. Huber, M. van de Kamp, G. W. Canters,

Journal of Molecular Biology 1991, 221 765-772.

[19.] R. B. King, J. K. Burdett, R. H. Crabtree, C. M. Lukehart, R. A. Scott, R. L. Wells, in Copper proteins in Type 1 sitesEd.: E. N. Baker), Wiley Interscience, Chichester, UK 1994, p. pp. 883-923.

[20.] Z. W. Chen, M. J. Barber, W. S. McIntire, F. S. Mathews, Acta

Crystallographica Section D-Biological Crystallography 1998, 54 253-268.

[21.] E. N. Baker, Journal of Molecular Biology 1988, 203 1071-1095.

[22.] F. E. Dodd, S. S. Hasnain, Z. H. L. Abraham, R. R. Eady, B. E. Smith,

Acta Crystallographica Section D-Biological Crystallography 1995, 51

1052-1064.

[23.] M. van de Kamp, R. Floris, F. C. Hali, G. W. Canters, Journal of the

American Chemical Society 1990, 112 907-908.

[24.] M. van de Kamp, G. W. Canters, C. R. Andrew, J. Sanders-Loehr, C. J. Bender, J. Peisach, European Journal of Biochemistry 1993, 218 229-238. [25.] G. van Pouderoyen, S. Mazumdar, N. I. Hunt, H. A. O. Hill, G. W.

Canters, European Journal of Biochemistry 1994, 222 583-588.

[26.] E. V. Sokerina, G. M. Ullmann, G. van Pouderoyen, G. W. Canters, N. M. Kostic, Journal of Biological Inorganic Chemistry 1999, 4 111-121. [27.] M. J. Bjerrum, D. R. Casimiro, I.-J. Chang, A. J. Di Bilio, H. B. Gray, M.

G. Hill, R. Langen, G. A. Mines, L. K. Skov, J. R. Winkler, D. S. Wuttke,

Journal of Bioenergetics and Biomembranes 1995, 27 295-302.

[28.] C. M. Che, R. Margalit, H. J. Chiang, H. B. Gray, Inorganica Chimica

Acta-Bioinorganic Chemistry 1987, 135 33-35.

[29.] A. J. Di Bilio, M. G. Hill, N. Bonander, B. G. Karlsson, R. M.

Villahermosa, B. G. Malmstrom, J. R. Winkler, H. B. Gray, Journal of the

American Chemical Society 1997, 119 9921-9922.

[30.] A. J. Di Bilio, M. G. Hill, L. K. Skov, I. J. Chang, F. M. Tam, J. R. Winkler, H. B. Gray, Abstracts of Papers of the American Chemical Society

1994, 207 498-INOR.

[31.] N. M. Kostic, R. Margalit, C. M. Che, H. B. Gray, Journal of the American

(44)

43

Chapter 2

Modelling of interprotein electron transfer by covalent and non-covalent crosslinking

[32.] J. R. Winkler, H. B. Gray, Chemical Reviews 1992, 92 369-379.

[33.] O. Farver, L. K. Skov, M. van de Kamp, G. W. Canters, I. Pecht, European

Journal of Biochemistry 1992, 210 399-403.

[34.] O. Farver, L. K. Skov, T. Pascher, B. G. Karlsson, M. Nordling, L. G. Lundberg, T. Vänngård, I. Pecht, Biochemistry 1993, 32 7317-7322. [35.] O. Farver, I. Pecht, Biophysical Chemistry 1994, 50 203-216.

[36.] O. Farver, L. K. Skov, G. Gilardi, G. van Pouderoyen, G. W. Canters, S. Wherland, I. Pecht, Chemical Physics 1996, 204 271-277.

[37.] O. Farver, N. Bonander, L. K. Skov, I. Pecht, Inorganica Chimica Acta

1996, 243 127-133.

[38.] O. Farver, L. J. C. Jeuken, G. W. Canters, I. Pecht, European Journal of

Biochemistry 2000, 267 3123-3129.

[39.] O. Farver, I. Pecht, Proceedings of the National Academy of Sciences of the

United States of America 1989, 86 6968-6972.

[40.] O. Farver, I. Pecht, Molecular Crystals and Liquid Crystals 1991, 194 215-224.

[41.] O. Farver, I. Pecht, Journal of Biological Inorganic Chemistry 1997, 2 387-392.

[42.] C. M. Groeneveld, G. W. Canters, European Journal of Biochemistry 1985,

153 559-564.

[43.] P. Kyritsis, C. Dennison, W. J. Ingledew, W. McFarlane, A. G. Sykes,

Inorganic Chemistry 1995, 34 5370-5374.

[44.] K. V. Mikkelsen, L. K. Skov, H. Nar, O. Farver, Proceedings of the National

Academy of Sciences of the United States of America 1993, 90 5443-5445.

[45.] M. G. Finiguerra, I. M. C. van Amsterdam, S. Alagaratnam, M. Ubbink, M. Huber, Chemical Physics Letters 2003, 382 528-533.

[46.] C. M. Halliwell, J. A. Davies, J. C. Gallop, R. Josephs-Franks,

Bioelectrochemistry 2004, 63 225-228.

[47.] I. M. C. van Amsterdam, M. Ubbink, G. W. Canters, Inorganica Chimica

Acta 2002, 331 296-302.

[48.] I. M. C. van Amsterdam, M. Ubbink, L. J. C. Jeuken, M. P. Verbeet, O. Einsle, A. Messerschmidt, G. W. Canters, Chemistry-A European Journal

(45)

44

Chapter 2

[49.] I. M. C. van Amsterdam, M. Ubbink, O. Einsle, A. Messerschmidt, A. Merli, D. Cavazzini, G. L. Rossi, G. W. Canters, Nature Structural Biology

2002, 9 48-52.

[50.] O. Farver, G. W. Canters, I. van Amsterdam, I. Pecht, Journal of Physical

Chemistry A 2003, 107 6757-6760.

[51.] A. D. Bain, G. J. Duns, Journal of Magnetic Resonance Series A 1995, 112 258-260.

[52.] A. D. Bain, G. J. Duns, Canadian Journal of Chemistry-Revue Canadienne de

Chimie 1996, 74 819-824.

[53.] A. D. Bain, Progress in Nuclear Magnetic Resonance Spectroscopy 2003, 43 63-103.

[54.] R. A. Marcus, N. Sutin, Biochimica et Biophysica Acta 1985, 811 265-322. [55.] C. C. Page, C. C. Moser, P. L. Dutton, Current Opinion in Chemical Biology

2003, 7 551-556.

[56.] C. C. Page, C. C. Moser, X. Chen, P. L. Dutton, Nature 1999, 42 47-52. [57.] M. Huber, I. M. C. van Amsterdam, M. Ubbink, G. W. Canters,

Biophysical Journal 2002, 82 477A.

[58.] W. B. Curry, M. D. Grabe, I. V. Kurnikov, S. S. Skourtis, D. N. Beratan, J. J. Regan, A. J. A. Aquino, P. Beroza, J. N. Onuchic, Journal of Bioenergetics

and Biomembranes 1995, 27 285-293.

[59.] I. Hamachi, Y. Tajiri, T. Nagase, S. Shinkai, Chemistry-A European Journal

1997, 3 1025-1031.

[60.] I. Hamachi, S. Shinkai, European Journal of Organic Chemistry 1999, 539-549.

[61.] T. Hayashi, A. Tomokuni, T. Mizutani, Y. Hisaeda, H. Ogoshi, Chemistry

Letters 1998, 1229-1230.

[62.] T. Hayashi, Y. Hitomi, H. Ogoshi, Journal of the American Chemical Society

1998, 120 4910-4915.

[63.] T. Hayashi, T. Ando, T. Matsuda, H. Yonemura, S. Yamada, Y. Hisaeda,

Journal of Inorganic Biochemistry 2000, 82 133-139.

[64.] T. Hayashi, H. Dejima, T. Matsuo, H. Sato, D. Murata, Y. Hisaeda,

Journal of the American Chemical Society 2002, 124 11226-11227.

(46)

45

Chapter 2

Modelling of interprotein electron transfer by covalent and non-covalent crosslinking

[66.] L. J. C. Jeuken, M. Ubbink, J. H. Bitter, P. van Vliet, W. Meyer-Klaucke, G. W. Canters, Journal of Molecular Biology 2000, 299 737-755.

[67.] T. den Blaauwen, M. van de Kamp, G. W. Canters, Journal of the American

Chemical Society 1991, 113 5050-5052.

[68.] S. J. Kroes, J. Salgado, G. Parigi, C. Luchinat, G. W. Canters, Journal of

Biological Inorganic Chemistry 1996, 1 551-559.

[69.] T. den Blaauwen, G. W. Canters, Journal of the American Chemical Society

1993, 115 1121-1129.

[70.] W. M. Albers, J. O. Lekkala, L. Jeuken, G. W. Canters, A. P. F. Turner,

Bioelectrochemistry and Bioenergetics 1997, 42 25-33.

[71.] G. van Pouderoyen, T. den Blaauwen, J. Reedijk, G. W. Canters,

Biochemistry 1996, 35 13205-13211.

[72.] G. van Pouderoyen, T. den Blaauwen, J. Reedijk, G. W. Canters,

Biochemistry 1998, 37 7656.

[73.] L. J. C. Jeuken, P. van Vliet, M. P. Verbeet, R. Camba, J. P. McEvoy, F. K. Armstrong, G. W. Canters, Journal of the American Chemical Society 2000,

(47)

4

(48)

Chapter

3

Electron self-exchange in N42C-BMME

azurin dimers as a function of

(49)

4

Chapter 3

Abstract

The electron self-exchange (e.s.e.) reaction in a covalent homodimer of N42C-BMME azurin was investigated as a function of temperature in an attempt to determine the reorganization energy λ. Rates of intramolecular e.s.e. were obtained by simulation of the redox sensitive Val31 1Hγ2 resonance in NMR spectra recorded

at temperatures between 312 and 272K. At T ≥ 303K a lower limit for keseintra was

found of ≥ 2x104 s-1 which is consistent with previous reports. At lower temperatures

Referenties

GERELATEERDE DOCUMENTEN

Although the PQQ molecule exhibits substantial electron delocalization, the overall chain extending between the cofactor and the connected electrode is not strictly

the formation of an association complex in which the hydrophobic patches of the partner proteins are positioned in a way that is similar to the situation observed in

in which 'Q represents the chemical shift difference between the positions of the analysed resonances in the fully oxidized and fully reduced protein, the line

In the case of the N42C azurin dimer it was found that intermolecular disulfide bond formation resulted in a sterically hindered complex incapable of intramolecular

Artificial protein complexes can serve as controlled model systems for biological electron transfer and provide valuable information for the design of efficient

Partial reduction of the dimer, however, can result in either one of two possibilities: 1) the reduced protein dissociates from the complex leaving behind the oxidized form

In contrast, the insertion of a spacing molecule (BMME) in between the cysteine residues allowed rearrangement of the N42C-BMME azurin complex and formation of a dimer

Hoofdstuk 5 beschrijft de analyse van een azurine dimeer met kenmerken van zowel de in hoofdstuk 3 als in hoofdstuk 4 beschreven systemen: een dimeer van N42C/M64E-azurine