• 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!
19
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 doctoral thesis in the

Institutional Repository of the University of Leiden

Downloaded from:

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

(2)

affiliations of the authors

Introduction to electron transfer in a

biological context

(3)

8 C h a p te r 1

Abstract

(4)

9 C h a p te r 1

1.1 E lectron 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 (VD A). The rate of electron transfer (kE T) is directly related to the square of VD A 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 M arcus equation:[1]

Eq. 2

(5)

10 C h a p te r 1

parameter O 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 < O,

kETincreases with increasing driving force (-'Go) . This behaviour falls into what

is referred to as the ‘normal’ region. Somewhat counter-intuitively though, when -'G0 > O, 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 O 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 m atrix

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

O. 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=O. Rates then decrease at higher driving forces (inverted effect). Figure made

(6)

11 C h a p te r 1

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 E and the distance between donor and acceptor sites d (Eq. 3):

Eq. 3 Values for E 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 U.[9] The empirically derived dependence of k

ET on the edge-to-edge DA

(7)

12 C h a p te r 1 Eq. 4 The distance parameter 3.6 Å in this equation represents the distance at van der W aals interaction at which kET,max§13 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 H.[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 W aals contact. An important prediction from the TP model has been that strongly hydrogen bonded E-sheet structures form more effective mediators of long-range electronic coupling than D-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

(8)

13 C h a p te r 1 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 E-strands (E=1.0 Å-1) and

D-helices (E=1.3 Å-1) respectively, whilst

(9)

14 C h a p te r 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. M olecular w ires

1.3.1. ET through molecular w ires

(10)

15 C h a p te r 1

Table 1.1: Dependence of different ET mechanisms on temperature and DA distance mechanism Temperature (T) DA distance (d) coherent tunnelling, ‘superexchange’ none exp(-Ed) 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 McC reery 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 S-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

(11)

16 C h a p te r 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 a25 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]

(12)

17 C h a p te r 1

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

(13)

18 C h a p te r 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,

(14)

19 C h a p te r 1

References

[1.] R. A. Marcus, N. Sutin, Biochimica et Biophysica Acta 1985, 811 265-322. [2.] H. B. Gray, J. R. Winkler, Q uarterly 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 N ational Academy of Sciences of the U nited States of America 1997, 94 4246-4249.

[5.] H. B. Gray, J. R. Winkler, Proceedings of the N ational Academy of Sciences of the U nited 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, N ature 1998, 396 60-63.

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

(15)

20 C h a p te r 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

(16)

21 C h a p te r 1

[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 O pinion 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 O rganic 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 L etters 1998, 1229-1230.

[43.] T. Hayashi, Journal of Synthetic O rganic 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 O rganic 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.

(17)

22 C h a p te r 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.

(18)
(19)

2

Referenties

GERELATEERDE DOCUMENTEN

The autocorrelation of bright times (Fig. 4G, blue) shows a characteristic decay time of hundreds of seconds while the autocorrelation of dark times (Fig. 4G, red) is closer to that

License: Licence agreement concerning inclusion of doctoral thesis in the Institutional Repository of the University of Leiden Downloaded.

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

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