You are on page 1of 6

Long-range electron transfer

Harry B. Gray† and Jay R. Winkler


Beckman Institute, California Institute of Technology, Pasadena, CA 91125

Edited by Jack Halpern, University of Chicago, Chicago, IL and accepted January 28, 2005 (received for review January 5, 2005)

Recent investigations have shed much light on the nuclear and electronic factors that control the rates of long-range electron tunnel-
ing through molecules in aqueous and organic glasses as well as through bonds in donor– bridge–acceptor complexes. Couplings
through covalent and hydrogen bonds are much stronger than those across van der Waals gaps, and these differences in coupling
between bonded and nonbonded atoms account for the dependence of tunneling rates on the structure of the media between re-
dox sites in Ru-modified proteins and protein–protein complexes.

electron tunneling 兩 hopping 兩 glass 兩 protein

distance range (⬍9 Å for ␤ ⫽ 1.0 Å⫺1).

I
ndependent efforts by Gamow (1) tunneling energy gap (⌬␧), defined as the
in late 1928 and Gurney and Con- vertical energy required to remove an Longer tunneling distances can be exam-
don (2) in early 1929 to explain electron from the donor, or a hole from ined if D and A are immobilized. In a typ-
radioactive decay using the ‘‘new the acceptor, and place it on the bridge. If ical experiment, a small concentration of
quantum mechanics’’ produced an ex- the donor-acceptor separation (d) is a lin- electron or hole donors is embedded in a
pression for the probability that a particle ear function of n (d ⫽ do ⫹ n␦), then glassed solvent amid a higher concentra-

冉 冊
will tunnel through a square potential tion of randomly distributed acceptors.
n⫺1
energy barrier. The transmission coeffi- h Ab h bb The donor is a photoexcited chromophore
cient (␬) decays exponentially with in- HA B ⫽ h bB
⌬␧ ⌬␧ or a radiolytically generated radical. The
creasing barrier width (r) and the decay time-dependent survival probability of the
constant varies as the square root of the
product of the barrier height (⌬E) times ⬀ e ⫺冉 ␦冊 ln
1
冉 冊.
⌬⑀
h bb d [3] donor depends on the concentration of
acceptors, the rate constant for electron兾
the mass (m) of the particle (Eq. 1). HAB will decay exponentially with dis- hole transfer when D and A are in van der
⫺共2兾 –h兲 冑2m⌬Er tance (hbb ⬍⬍ ⌬␧; Eq. 3). Using the result Waals contact (ko), and the distance de-
␬⬀e . [1] from semiclassical theory that ET rates cay factor ␤. Extracting reliable values for
are proportional to HAB 2
(Eq. 2), it is pos- ko and ␤ from time-resolved spectroscopic
This theory rationalized the microsec- measurements, however, can be rather
ond tunneling of an ␣-particle through a sible to define effective tunneling barriers
(⌬Eeff) in terms of superexchange param- difficult because the two parameters are
3 ⫻ 10⫺4-Å, 6-MeV barrier (2) and pre-
eters, as well as the exponential decay highly correlated (19, 22). In the case of
dicted that in the same time period an
electron could tunnel 19 Å through a constant (␤) describing the variation of photoinitiated ET in glasses, measure-
rates with distance (Eq. 4). ments of luminescence decay kinetics and
1-eV barrier. In the intervening years,
luminescence quantum yields at several

冉 冊冋冉 冊 冉 冊册
electron tunneling has been found to –h2 2
play pivotal roles in solid-state physics 2 ⌬␧ different quencher concentrations provide
⌬Eeff ⫽ ln enough information to decouple ko and ␤,
(3), chemistry (4, 5), and biology (5–7). 8m e ␦ h bb
permitting reliable values to be deter-

冉 冊
A Landau-Zener treatment of the re-
actant-product transition probability –h2 mined for each parameter (19).
produces the familiar semiclassical ex- ⫽ ␤2 Our experimental investigation of
8m e Ru(tpy)2⫹
pression for the rate of nonadiabatic 2 luminescence quenching by
Fe(OH2)3⫹ 6 in aqueous acidic glasses
electron transfer (ET) between a donor ⫽ 共0.952 eV A2兲 ␤ 2. [4]
(D) and acceptor (A) held at fixed dis- placed rigorous limits on the distance de-
tance (Eq. 2) (5). Saturated hydrocarbon spacers typically cay constant for tunneling through water
exhibit ␤ ⬇1.0 Å⫺1 (⌬Eeff ⫽ 0.95 eV) (23). The luminescence lifetime of

kET ⫽ 冑 4␲3
H2
h ␭ RT AB
2
(10–12). The decay constant for phe-
nylene bridges depends on the dihedral
angle between adjacent aromatic rings
*Ru(tpy)2⫹2 in aqueous glasses is long
enough to allow a significant distance
range (⬇25 Å) to be probed. A distance

䡠exp ⫺再 共⌬G⬚ ⫹ ␭ 兲 2
4 ␭ RT
. 冎 [2]
[0.4–0.8 Å⫺1 (13, 14); ⌬Eeff ⫽ 0.15–0.61
eV]. Polyene and phenylenevinylene
bridges exhibit remarkably efficient ET
decay constant of 1.58(5) Å⫺1 was ob-
tained for H2SO4兾H2O, HSO3F兾H2O, and
D2SO4兾D2O glasses (⌬Eeff ⫽ 2.4 eV) (14).
The electronic coupling matrix element over very long distances: ␤-values as We also have determined ␤ and ⌬Eeff
(HAB) reflects the strength of the interac- low as 0.04 Å⫺1 (⌬Eeff ⫽ 0.002 eV) have values for electron tunneling from
tion between reactants and products at been reported (15). electronically excited [Ir(␮-pyrazolyl)(1,5-
the nuclear configuration of the transition cyclooctadiene)]2 to 2,6-dichloro-1,4-
state. McConnell (8), building on prior Aqueous and Organic Glasses
work by Halpern and Orgel (9), argued Several experimental investigations have
This paper was submitted directly (Track II) to the PNAS
that the interaction energy (2HAB) be- demonstrated that solvent hole and elec- office.
tween two redox centers separated by a tron states can mediate long-range elec-
Abbreviations: ET, electron transfer; cyt, cytochrome; ccp,
covalent bridge composed of n identical tron tunneling (16–21). In fluid solution, cyt c peroxidase; MTHF, 2-methyltetrahydrofuran; Hb,
repeat units depends on the coupling when the positions of D and A are not hemoglobin.
strength between the redox sites and the constrained by a covalent bridge, diffusion †Towhom correspondence should be addressed. E-mail:
bridge (hAb, hbB), the coupling between places an upper limit on the time scale hbgray@caltech.edu.
adjacent bridge elements (hbb), and the (⬍10⫺9 s) and, therefore, the tunneling © 2005 by The National Academy of Sciences of the USA

3534 –3539 兩 PNAS 兩 March 8, 2005 兩 vol. 102 兩 no. 10 www.pnas.org兾cgi兾doi兾10.1073兾pnas.0408029102


SPECIAL FEA TURE: PERSPECTIVE
biological ET reaction (24, 25). Contin- ing distances (Ru-Cu) in these five
uum models suggest that embedding a Ru(bpy)2(im)(HisX)2⫹-azurins and
redox center inside a low dielectric cavity Ru(bpy)2(im)(His-83)2⫹-azurin are
can lower the outer-sphere ␭ by as much shown in Fig. 2.
as 50% (26). Moreover, by constraining Measurements of Cu(I) 3 Ru(III) ET
the coordination environment around (⫺⌬G° ⫽ 0.7 eV) in the set of Ru-azurins
metal centers, the inner-sphere ␭ can be established the distance dependence of
reduced as well (25). Copper is a case in ET along ␤-strands (7, 31, 32). The driv-
point. The reorganization energy for elec- ing force-optimized azurin tunneling time-
tron self-exchange in Cu(phen)2⫹/⫹2 is table reveals a nearly perfect exponential
⬇2.4 eV; the value for Cu(II兾I) in distance dependence, with a decay con-
Pseudomonas aeruginosa azurin is 0.7 eV stant (␤) of 1.1 Å⫺1, and an intercept at
(24). The 1.7-eV reduction in ␭ reflects close contact (ro ⫽ 3 Å) of 1013 s⫺1. This
the transition-state stabilization imposed decay constant is quite similar to that
by the azurin fold. found for superexchange-mediated tunnel-
In 1982, we demonstrated long-range ing across saturated alkane bridges
electron tunneling through Ru-modified (␤ ⬇1.0 Å⫺1) (12, 35), strongly indicating
cytochrome (cyt) c (27). Subsequent work that a similar coupling mechanism is oper-
in our laboratory has focused on the elu- ative in the polypeptide (Fig. 1). Impor-
cidation of distant electronic couplings tantly, kinetics data obtained by Farver
Fig. 1. Timetable for activationless electron tun- between redox sites in several Ru-proteins and Pecht (36) in their studies of long-
neling through various media: vacuum (black, ␤ ⫽ (7, 28–32). In particular, work on Ru- range ET from radiolytically generated
2.9 – 4.0 Å⫺1), MTHF glass (violet, ␤ ⫽ 1.57–1.67 azurin has provided a reference point for disulfide radical anion to the blue copper
Å⫺1), aqueous glass (cyan, ␤ ⫽ 1.55–1.65 Å⫺1), and electron tunneling through folded center in azurin also have been inter-
toluene glass (green, ␤ ⫽ 1.18 –1.28 Å⫺1). Investi- polypeptide structures (7, 31, 32). The preted successfully in terms of this cou-
gations of ET rates in D-(bridge)-A complexes have copper center in azurin is situated at one pling model.
produced exponential distance dependences: xylyl
end of an eight-stranded ␤-barrel, ligated Our studies have shown that Cu(I) to
bridges, ␤ ⫽ 0.76 Å⫺1 (red) (12); alkane bridges, ␤ ⫽
1.0 Å⫺1 (orange) (10); and ␤-strand bridges in
in a trigonal plane by two imidazoles Ru(III) or Os(III) ET rates in labeled
ruthenium-modified azurin, ␤ ⫽ 1.1 Å⫺1 (yellow). (His-46 and His-117) and a thiolate (Cys- azurin crystals are nearly identical with
112); in addition, there are weak axial solution values for each donor–acceptor
interactions (Met-121 thioether sulfur and pair (34). The energy gap between the
benzoquinone in 2-methyltetrahydrofuran Gly-45 carbonyl oxygen) (33, 34). The donor–acceptor redox levels and those of
(MTHF) and toluene glasses at 77 K (14). azurin from P. aeruginosa has two addi- oxidized or reduced intermediate states is
The effective barrier height for electron tional His residues, one of which (His-83) the primary criterion in determining when
tunneling through toluene (⌬Eeff ⫽ 1.4 reacts readily with Ru-labeling reagents. hole or electron hopping becomes impor-
eV) is substantially lower than the barrier A H83Q base mutant was prepared and tant. In Ru-azurin, photogenerated
in MTHF (⌬Eeff ⫽ 2.6 eV); and the bar- individual mutant His residues were intro- Ru(bpy)2(im)(His)3⫹ (E° ⫽ 1.0 V vs.
rier for tunneling in aqueous sulfuric acid duced at five locations on ␤-strands ex- NHE) (37) potentially could oxidize Trp
glasses (⌬Eeff ⫽ 2.4 eV) is very near that tending from the Cys-112 and Met-121 or Tyr residues (7). If the Cu(I) center is
for MTHF. Distance decay parameters ligands (K122H, T124H, T126H, replaced by redox-inert Zn(II) in the pro-
are 1.62 (MTHF), 1.58 (H2O), and 1.23 Q107H, and M109H) (31, 32). Tunnel- tein, however, we find that photogener-
Å⫺1 (toluene) (14). In toluene and
MTHF, coupling between bridge units is
mediated by van der Waals contacts,
whereas the aqueous glass is interlaced
with strong hydrogen bonds. It is possible
that the hydrogen bonds between mole-
cules in the aqueous glass compensate for
the large molecular orbital (H2O) energy
gap to produce a tunneling barrier on par
with that of MTHF.
The 1.62-Å⫺1 distance decay constant
for MTHF confirms that there is a signifi-
cant coupling penalty associated with tun-
neling across the van der Waals gaps be-
tween solvent molecules. Taking 20 Å as a
reference distance, we find that tunneling
across an alkane bridge (␤ ⬇ 1.0 Å⫺1) is
⬇40,000 times faster than tunneling
through MTHF; and, to underscore the
point, recent experiments on D-oligoxy-
lene-A complexes have shown that 20-Å
tunneling across covalently linked xylenes
is almost 3,000 times faster than tunneling
through a toluene glass (Fig. 1) (14).
Ru-Proteins Fig. 2. Tunneling distances and backbone structure models showing locations of the Cu active site (blue)
The protein fold plays a central role in and the Ru(bpy)2(im)(HisX)2⫹ label (orange) in Ru-azurins: Ru-Cu (Å), HisX (X ⫽ 122, 15.9; 83, 16.9; 109,
lowering the reorganization energy of a 17.9; 124, 20.6; 107, 25.7; 126, 26.0).

Gray and Winkler PNAS 兩 March 8, 2005 兩 vol. 102 兩 no. 10 兩 3535
smaller subunits linked by covalent bonds, time for protein ET (具␶典) is on the order
hydrogen bonds, or through-space jumps. of 200 ns. We emphasize, however, that
More elaborate computational protocols eight tunneling times measured for four
also have shed light on the factors that different Ru-proteins are ⬍200 ns
determine distant couplings in proteins (Fig. 3).
(42–46). Indeed, in this issue of PNAS, Bixon and Jortner (53, 54) have noted
Beratan and coworkers (47) present a de- that there are several examples of ET
tailed analysis, including the effects of rates exceeding the solvent-controlled adi-
protein dynamics on the distant couplings abatic limit; model calculations suggest a
of six Ru-azurins (Fig. 2). possible explanation. Reactions at low
driving force (⫺⌬G° ⬍ ␭) require sub-
Medium Dynamics stantial reorganization along solvent coor-
The nonadiabatic ET model embodied in dinates and rates are predicted to exhibit
Eq. 2 rests on the assumption that the a pronounced dependence on relaxation
electronic transition from the reactant dynamics. The calculations suggest, how-
potential energy surface (D ⫹ A) to the ever, that the rates of activationless
product surface (D⫹ ⫹ A⫺) is much (⫺⌬G° ⬇ ␭) and inverted (⫺⌬G° ⬎ ␭)
slower than the frequency of nuclear mo- reactions will be nearly independent of ␬
tion on these surfaces. Both theoreticians and, hence, the dynamics of medium re-
and experimentalists have long been inter- laxation, as we have found in the case of
Fig. 3. Tunneling timetable for intraprotein ET in ested in charge-transfer processes that are Ru(diimine)-protein ET reactions (7).
Ru-modified azurin (blue circles), cyt c (red circles), not well described by this model (48–54).
myoglobin (yellow triangles), cyt b562 (green Protein–Protein Reactions
squares), HiPIP (orange diamonds), and for inter- ␬NA At least three elementary steps are re-
protein ET Fe:Zn-cyt c crystals (fuchsia triangles). ␬⫽ [5] quired to complete a redox reaction be-
The solid lines illustrate the tunneling-pathway
1⫹␬
tween soluble proteins: (i) formation of an
predictions for coupling along ␤-strands (␤ ⫽ 1.0
Å⫺1) and ␣-helices (␤ ⫽ 1.3 Å⫺1); the dashed line 4␲HA2
B具 ␶ 典
active donor–acceptor complex; (ii) elec-
␬⫽ . [6] tron tunneling within the donor–acceptor
illustrates a 1.1-Å⫺1 distance decay. Distance decay –h␭
for electron tunneling through water is shown as a
o complex; and (iii) dissociation of the oxi-
cyan wedge. Estimated distance dependence for dized and reduced products. Because the
Theory suggests that, under certain cir-
tunneling through vacuum is shown as the black dynamics of the first and third steps ob-
cumstances, the time scales for reorienta-
wedge. scure the electron tunneling reaction,
tion of solvent molecules can be slower experimental studies often focus on ET
than the reactant-product transition fre- reactions within protein–protein com-
ated holes in Ru(bpy)2(im)(HisX)3⫹ com- quency. In this solvent-controlled adia- plexes that form at low ionic strength. It
plexes remain localized on the metal batic limit, reactions are limited by the has been difficult to interpret the results,
center. The energy gap between the dynamics of solvent relaxation. Bixon and however, as neither the donor–acceptor
Ru(III) hole and oxidized bridge states Jortner (53, 54) developed a generalized docking geometries nor the conformations
therefore must be ⬎75 meV (3 kBT at expression for ET rates that explicitly ac- of these complexes are known with cer-
295 K). Our finding that the Cu(I) 3 counts for solvent relaxation dynamics tainty. With the aid of rapid triggering
Ru(III) ET rate in Ru(bpy)2(im)(HisX)- (Eqs. 5 and 6). The factor ␬ is a solvent methods, it has been possible to measure
azurin does not decrease in going from adiabaticity factor and kNA is the nonadia- rates of long-range ET between redox
300 to 240 K and actually increases batic ET rate given by Eq. 2. Small values sites in protein–protein complexes. In
slightly at 160 K demonstrates that hop- of ␬ correspond to the nonadiabatic limit; many complexes, there are multiple bind-
ping does not occur in this case, as a reac- large values result in solvent-controlled ing sites and it is not uncommon to find
tion with an endergonic step would be adiabatic processes. Typical solvent relax- that the ET kinetics often are regulated
highly disfavored at low temperature. We ation times (具␶典) are ⱕ10⫺11 s, so that sol- by the dynamics of conformational
conclude that the Ru-azurin timetable vent relaxation dynamics are expected to changes in the complex (60, 61). The
(Fig. 1) provides a benchmark for super- become important only in relatively usual interpretation is that surface diffu-
exchange-mediated electron tunneling strongly coupled systems. An adiabaticity sion of the two proteins produces a tran-
through proteins. factor of 1 in a solvent with a 1-ps relax- sient complex with enhanced electronic
The rates of high driving-force ET reac- ation time, for example, corresponds to a coupling and faster ET. Consequently,
tions have been measured for ⬎30 Ru(dii- coupling matrix element of ⬇50 cm⫺1 rates depend strongly on solvent viscosity
mine)-labeled metalloproteins (7, 29, 30, (␭o ⫽ 0.5 eV). rather than intrinsic ET parameters. A
38). Driving-force-optimized values are Our studies of ET in Ru-proteins indi- further complication associated with stud-
scattered around the Ru-azurin 1.1-Å⫺1 cate that a large part of the contribution ies of protein–protein ET in solution is
exponential distance decay. Rates at a to ␭o comes from reorientation of the that binding sites and, hence, locations of
single distance can differ by as much as a polypeptide matrix (34). The dynamics of redox cofactors, often are unknown.
factor of 103, and D兾A distances that large-scale nuclear motions in polypep- Crystals containing photoactivatable
differ by as much as 5 Å can produce tides are expected to be substantially donors and acceptors at specific lattice
identical rates (Fig. 3). In seminal work, slower than those of most solvents. Relax- sites are ideal media for investigating tun-
Beratan, Onuchic, and coworkers (39–41) ation times ranging from picoseconds to neling between proteins. In crystal lattices
developed a generalization of the McCon- microseconds have been reported for the of tuna cyt c, chains of protein molecules
nell superexchange coupling model that heme pocket of myoglobin (55–58). In- form helices with a 24.1-Å separation be-
accounts for rate scatter attributable to deed, electrochemical measurements by tween neighboring metal centers (62).
protein structural complexity. In this tun- Waldeck and coworkers (59) using cyt c By doping Zn-cyt c into this lattice, inter-
neling-pathway model, the medium be- adsorbed onto self-assembled monolayers protein ET between triplet-excited Zn-
tween D and A is decomposed into suggest that the characteristic relaxation porphyrin and a neighboring Fe(III)-cyt c

3536 兩 www.pnas.org兾cgi兾doi兾10.1073兾pnas.0408029102 Gray and Winkler


could be investigated; the rate constant a minor contribution to protein ET reor- complex of the two yeast proteins. Inter-
was found to be 4(1) ⫻ 102 s⫺1, and ganization energies (34). estingly, the complex between yeast ccp
charge recombination was about four The ET reaction between cyt c and cyt and horse cyt c exhibited a slightly differ-
times faster [2.0(5) ⫻ 103 s⫺1] (62). b5 has been the subject of experimental ent structure. Analysis of the yeast兾yeast
Rapid relay of electrons involving at and theoretical investigations for ⬎40 complex suggested an electronic coupling
least one soluble redox enzyme requires years (70, 71). The structural model pro- pathway from the cyt c heme to the ccp
the formation of short-lived, weakly posed by Salemme in 1976 (72) for the heme via Trp-191. On the basis of these
bound protein–protein complexes. The precursor complex of this protein pair crystallographic results, Pelletier and
recognition sites between proteins in such stimulated a great deal of experimental Kraut argued that ccp and cyt c form a
complexes tend to be smaller (⬍1,200 Å2) work. Careful spectroscopic studies re- highly specific 1:1 ET complex.
and include more water molecules than vealed that these oppositely charged pro- Hoffman and coworkers (77) have used
the interfaces between subunits in oligo- teins form a stable 1:1 complex at low metal-substituted ccp and cyt c to explore
meric proteins (63). The interprotein in- ionic strength [KA ⫽ 8(3) ⫻ 104 M⫺1, pH the ET kinetics between these two pro-
teractions in crystals of tuna cyt c involve 7, ␮ 10 mM; KA ⫽ 4(3) ⫻ 106 M⫺1, pH 7, teins. Results from quenching studies,
relatively few contacts: 760 Å2 of surface ␮ 1 mM] (73). temperature and ionic strength depen-
area is buried in an interface with 31 van McLendon and Miller (74) used a com- dences, species variations, and electro-
der Waals contacts (3.2 ⱕ d ⱕ 3.9 Å) and bination of photochemical and pulse- static modeling provide compelling
16 water molecules (3 of which form radiolytic methods to probe the driving- evidence for two distinct cyt c binding
bridging hydrogen bonds across the inter- force dependence of heme-heme ET in sites on ccp. The higher-affinity binding
face) but only one direct hydrogen bond the 1:1 complex. The ET rates exhibited a site is the locus for Trp-191 radical reduc-
bridging the two proteins. Indeed, the cyt near-Gaussian free energy dependence, in tion by cyt c. Heme (ccp) reduction by cyt
c–cyt c interface is reminiscent of that excellent agreement with a 0.8-eV reorga- c can occur from either the high- or low-
between natural redox partners, e.g., cyt c nization energy. Evidence for more com- affinity binding site but, when exchange
and cyt c peroxidase (ccp) (770 Å2) (64), plex ET processes came from studies in between the two is rapid, reduction from
or cyt c2 and the photosynthetic reaction which photochemically generated reduc- the low-affinity site dominates (77). These
center, discussed in this issue of PNAS tants injected electrons into preformed studies of ccp兾cyt c and cyt b5兾c ET, in-
by Onuchic and coworkers (65). Our find- Fe-cyt b5兾Fe-cyt c complexes. In one cluding an important contribution by
ing that ET rates in Zn-doped tuna cyt c study, the rate of b5 3 c ET (1.7 ⫻ 103 Hoffman and coworkers (79) in this issue
crystals fall well within the protein range s⫺1) was reported to depend on viscosity of PNAS, have shed much light on the
in the Ru-protein tunneling timetable and surface mutations (75). Later laser mechanisms of protein–protein ET
(Fig. 3) (7, 62) demonstrates that small flash photolysis experiments revealed a processes.
interaction zones of low density are quite rate-limiting second-order reduction of In the terminal reaction of the respira-
effective in mediating interprotein redox Fe-cyt b5兾Fe-cyt c and no sign of satura- tory chain, cyt c oxidase (CcO) removes
reactions. tion, suggesting that the intracomplex ET electrons from cyt c and passes them on
Kinetics measurements on crystallo- rate was ⬎104 s⫺1 (76). to O2 (80). CcO is a multisubunit mem-
graphically characterized metal-substituted Durham, Millett, and coworkers (71) brane-bound enzyme with four redox co-
hemoglobin (Hb) hybrids provided some and Meyer et al. (76) used Ru-modified factors (CuA, cyt a, cyt a3, and CuB). The
of the earliest insights into interprotein cyt b5 and photochemical triggering meth- locations of these metal complexes in CcO
ET rates (66, 67). Because Hb is a very ods to examine the kinetics of ET in cyt were revealed in the 1990s by the x-ray
strongly bound complex of four polypep- b5兾c complexes. Rapid intraprotein reduc- crystal structures of bacterial (81) and
tide subunits, ET measurements are not tion (⬍100 ns) of Fe(III)-cyt b5 by elec- bovine enzymes (82, 83). CuA, a binuclear
complicated by the dynamical problems tronically excited Ru(bpy)2⫹3 made it site with bridging S(Cys) atoms, is the pri-
that plague interpretation of rates in more possible to probe b5 3 c ET kinetics. Two mary electron acceptor from cyt c. Studies
weakly bound assemblies. Replacement of concentration-independent ET rates (4 ⫻ with Ru-modified cyt c reveal rapid (6 ⫻
the native Fe center in the ␤-subunits of 105 s⫺1 and 3.4 ⫻ 104 s⫺1) were observed, 104 s⫺1) (84) electron injection from
Hb with Zn or Mg creates the opportu- suggesting that two cyt b5兾c species are Fe(II) into CuA at low driving force (⌬G°
nity for photoinitiated ET reactions. The present in solution. Studies of ionic ⫽ ⫺0.03 eV) (85). Modeling suggests that
reacting metal centers in the Hb hybrids strength dependences and the effects of cyt c binds to the enzyme at an acidic
are separated by 25 Å so that rates are mutations indicated that the slower patch on subunit II (86, 87). The cyt c
relatively slow even at high driving forces. Fe(III)-cyt c reduction phase may be lim- heme is very near the Trp-104 (subunit II)
The time constant for ET from a triplet- ited by conformational changes within one indole ring, a residue that appears from
excited Zn-porphyrin in the ␤-subunit to of the complexes (71). mutagenesis experiments to be critical for
an Fe(III) center in the ␣-subunit is ⬇16 Ccp catalyzes the two-electron reduc- rapid cyt c 3 CuA ET (88–91). Solomon
ms (66). Extensive studies of temperature tion of H2O2 by ferrocyt c. Peroxide re- and coworkers (92) have identified a pos-
dependence of hybrid Hb ET rates led to acts rapidly with the resting ferric form of sible electron tunneling path from this cyt
the conclusion that the reorganization ccp to produce a species referred to as c binding site through Trp-104 to the
energy for these reactions (␭ ⬇1 eV) is compound I, which contains a ferryl bridging S(Cys-200) ligand on CuA.
dominated by outer-sphere contributions [Fe(IV)O2⫹] heme and a protein radical The 19.6-Å ET from CuA to cyt a pro-
(68, 69). Measurements of ET rates in located on Trp-191. The ET reactions in- ceeds rapidly at low driving force (⬇104
cryogenic glasses suggest that the polypep- volving these physiological redox partners s⫺1; ⌬G° ⬇⫺0.05 eV) (84, 93). Ramirez et
tide is the primary outer-sphere medium have been studied in great detail (77). At al. (80), Regan et al. (94), and Solomon
for the reaction and that bulk solvent re- low ionic strength, acidic ccp and basic cyt and coworkers (92, 95) have identified a
organization does not play an important c form a stable complex. A model of a 1:1 coupling route that proceeds from CuA
role. Moreover, it was suggested that even complex, based on the crystal structures ligand His-204 (subunit II) across one hy-
at room temperature, the protein medium of the two independent proteins, was pro- drogen bond to Arg-438 (subunit I)
in Hb acts like a frozen glass. Results posed by Poulos and Kraut (78) in 1980. [H204(N␧)-R438(O), 3.36 Å], and another
from measurements on Ru-azurin crystals Twelve years later, Pelletier and Kraut H-bond (2.95 Å) from the Arg-438 N-
also indicate that bulk solvent makes only (64) reported the crystal structure of a 1:1 amide to the cyt a heme propionate.

Gray and Winkler PNAS 兩 March 8, 2005 兩 vol. 102 兩 no. 10 兩 3537
ping rates for any set of driving-force,
temperature, and distance parameters.
Consider the two-step tunneling reaction
defined in Eq. 7 (reactants, R ⫽ D-I-A;
redox intermediate, H ⫽ D⫹-I⫺-A or
D-I⫹-A⫺; products, P ⫽ D⫹-I-A⫺) The
general solution to the rate law for this
process calls for biexponential production
of P, although under some circumstances
the appearance of P can be approximated
by a single exponential function. Taking a
value of ␭ ⫽ 0.8 eV for both tunneling
reactions (i.e., R 3 H and H 3 P) and a
distance decay constant of 1.1 Å⫺1, we
can calculate the time dependence of the
populations of all three reacting species
for various values of ⌬GRH ° , ⌬GHP
° , rRH,
and rHP (7). Results for the particular case
in which ⌬GRH ° ⫽ ⫺⌬GHP ° and rRH ⫽ rHP
are illustrated in Fig. 5. This model ap-
proximates a biological electron-transport
Fig. 4. Tunneling-time (1兾kobs) contours as functions of donor-acceptor distance (␤ ⫽ 1.1 Å⫺1) and driving
chain (⌬GRP ° ⫽ 0) reaction with a single
force [in units of ␭; kBT兾␭ ⫽ kB(295 K)兾(0.8 eV) ⫽ 0.318].
endergonic step. Transport across 20 Å is
104 times faster than a single tunneling
Based on a tunneling currents analysis, cases deliver electrons or holes rapidly step at this distance and submillisecond
Stuchebrukhov and colleagues (96) sug- to very distant sites (7, 30, 99). Require- transfers can be realized. For D-A separa-
gested a slightly different CuA-to-cyt a ments for functional hopping include tions ⬍20 Å, endergonic intermediate
coupling route through His-204. It is likely optimal positioning of redox centers and steps as large as 0.4–0.5 eV will afford
that, owing to strong Cu-S(Cys) electronic fine-tuning of reaction driving forces. submillisecond transport times. An impor-
interactions, pathways involving the bridg- Modeling the kinetics of electron hop- tant conclusion is that hopping can facili-
ing Cys residues are important for mediat- ping is a straightforward problem that tate electron flow over distances ⬎20 Å in
ing coupling even though they involve can be solved analytically without using cases where the free-energy changes for
more bonds than the His-204 route. The simplifying approximations (7). endergonic intermediate steps are no
current view is that the sequence Cys-200兾 more than 0.2 eV.
Ile-199兾Arg-439兾heme-propionate (cyt a) k RH k HP
is the dominant CuA to cyt a electron tun- -0 H |
R| -0 P [7] Concluding Remarks
neling pathway (92, 96). k HR k PH More than 75 years have passed since the
Both Regan et al. (94) and Stuchebruk- Gamow (1) and Gurney-Condon (2) pa-
hov and colleagues (96) have identified Using the well defined properties of ET pers appeared. Activity in the electron
pathways between cyt a and cyt a3. In- reactions (Eq. 2), and the average dis- tunneling field over the last 20 years has
cluded among these routes is a direct tance dependence defined by Ru-protein been intense, most especially on the ex-
covalent pathway from the heme-a axial tunneling timetables, we can predict hop- perimental side, where investigators have
ligand His-378 through Phe-377 to the a3
His-376. Importantly, although the CuA-
cyt a3 distance (22.4 Å) is similar to that
of CuA to cyt a, neither Regan et al. (94)
nor Stuchebrukhov and colleagues (96)
found a coupling pathway that would fa-
cilitate electron flow to a3 in a single step
from the binuclear copper center.

Hopping
Electron tunneling times must be in the
millisecond to microsecond range for bio-
logical redox machines to function properly.
As a result, the maximum center-to-center
distance for single-step tunneling through
proteins can be no more than ⬇20 Å (Fig.
4). The structures of several redox enzyme
assemblies, however, suggest that charge
transport may occur over distances that
far exceed this single-step limit (7, 97, 98).
How can charge transport in proteins
cover distances well over 20 Å? One pos- Fig. 5. Distance dependences of the rates of single-step and two-step electron tunneling reactions. Solid
sibility is by hopping, as it can be shown line indicates theoretical distance dependence for a single-step, ergoneutral (⌬GRP ⫽ 0) tunneling process
that coupled tunneling reactions, particularly (␤ ⫽ 1.1 Å⫺1). Dashed lines indicate distance dependence calculated for two-step ergoneutral tunneling
with endergonic steps, can in favorable (R%H%P) with the indicated free-energy changes for the R%H step.

3538 兩 www.pnas.org兾cgi兾doi兾10.1073兾pnas.0408029102 Gray and Winkler


elucidated many of the factors that con- is another hot topic these days; and two version of all forms of energy. It also is
trol reaction rates through bonded as well articles in this issue of PNAS report excit- essential for successful operation of mo-
as nonbonded atoms in molecules and ing results in this area (47, 79). Distant lecular-scale electronic devices. We have
molecular assemblies, including the impor- (⬎20 Å) charge transport in DNA is still laid a firm foundation for these applica-
tant case of folded polypeptide structures. another area of great interest. Much work tions, but we must greatly ramp up both
In 2005, the field is booming, as tunnel- in this area has been done by Barton and theoretical and experimental investigations
ing-related solar cell, sensor, and other coworkers; and, in this issue of PNAS, a of multielectron and other coupled redox
device technologies are being developed report from her laboratory, in collaboration processes if we are to realize the full
at a rapid pace. A critical issue here is with the David group at Utah, suggests potential of this simplest of chemical
reactions.
understanding bridge energy effects on that guanine radicals, which facilitate iron-
charge transport through molecular mate- sulfur cluster oxidation in a DNA兾MutY Our work is supported by the National
rials, as discussed in this issue of PNAS by complex, may stimulate DNA repair (101). Institutes of Health, the National Science
Ratner, Wasielewski, and coworkers Controlled electron flow is an absolute Foundation, BP, and the Arnold and Mabel
(100). The role of dynamics in protein ET requirement for efficient storage and con- Beckman Foundation.

1. Gamow, G. (1928) Z. Phys. 51, 204–212. 37. Chang, I.-J., Gray, H. B. & Winkler, J. R. (1991) J. Am. 73. Mauk, M. R., Reid, L. S. & Mauk, A. G. (1982) Bio-
2. Gurney, R. W. & Condon, E. U. (1929) Phys. Rev. 33, Chem. Soc. 113, 7056–7057. chemistry 21, 1843–1846.
127–140. 38. Babini, E., Bertini, I., Borsari, M., Capozzi, F., Luchinat, 74. McLendon, G. & Miller, J. R. (1985) J. Am. Chem. Soc.
3. Esaki, L. (1974) Science 183, 1149–1155. C., Zhang, X., Moura, G. L. C., Kurnikov, I. V., Beratan, 107, 7811–7816.
4. Miller, J. R. (1975) Science 189, 221–222. D. N., Ponce, A., et al. (2000) J. Am. Chem. Soc. 122, 75. Qin, L., Rodgers, K. K. & Sligar, S. G. (1991) Mol. Cryst.
5. Marcus, R. A. & Sutin, N. (1985) Biochim. Biophys. Acta 4532–4533. Liq. Cryst. 194, 311–316.
811, 265–322. 39. Beratan, D. N. & Onuchic, J. N. (1989) Photosynth. Res. 76. Meyer, T. E., Rivera, M., Walker, F. A., Mauk, M. R.,
6. De Vault, D., Parkes, J. H. & Chance, B. (1967) Nature 22, 173–186. Mauk, A. G., Cusanovich, M. A. & Tollin, G. (1993)
215, 642–644. 40. Onuchic, J. N., Beratan, D. N., Winkler, J. R. & Gray, H. B. Biochemistry 32, 622–627.
7. Gray, H. B. & Winkler, J. R. (2003) Q. Rev. Biophys. 36, (1992) Annu. Rev. Biophys. Biomol. Struct. 21, 349–377. 77. Nocek, J. M., Zhou, J. S., DeForest, S., Priyadarshy, S.,
341–372. 41. Beratan, D. N., Betts, J. N. & Onuchic, J. N. (1991) Beratan, D. N., Onuchic, J. N. & Hoffman, B. M. (1996)
8. McConnell, H. M. (1961) J. Chem. Phys. 35, 508–515. Science 252, 1285–1288. Chem. Rev. 96, 2459–2489.
9. Halpern, J. & Orgel, L. E. (1960) Disc. Faraday Soc. 42. Regan, J. J. & Onuchic, J. N. (1999) Adv. Chem. Phys. 107, 78. Poulos, T. L. & Kraut, J. (1980) J. Biol. Chem. 255,
32–41. 497–553. 10322–10330.
10. Oevering, H., Paddon-Row, M. N., Heppener, M., Oliver, 43. Balabin, I. A. & Onuchic, J. N. (1998) J. Phys. Chem. B 79. Hoffman, B. M., Celis, L. M., Cull, D. A., Patel, A. D.,
A. M., Cotsaris, E., Verhoeven, J. W. & Hush, N. S. 102, 7497–7505. Seifert, J. L., Wheeler, K. E., Wang, J., Yao, J., Kurnikov,
(1987) J. Am. Chem. Soc. 109, 3258–3269. 44. Skourtis, S. S. & Beratan, D. N. (1997) J. Biol. Inorg. I. V. & Nocek, J. M. (2005) Proc. Natl. Acad. Sci. USA
11. Johnson, M. D., Miller, J. R., Green, N. S. & Closs, G. L. Chem. 2, 378–386. 102, 3564–3569.
(1989) J. Phys. Chem. 93, 1173–1176. 45. Kumar, K., Kurnikov, I. V., Beratan, D. N., Waldeck, D. H. 80. Ramirez, B. E., Malmström, B. G., Winkler, J. R. & Gray,
12. Smalley, J. F., Finklea, H. O., Chidsey, C. E. D., Linford, & Zimmt, M. B. (1998) J. Phys. Chem. A 102, 5529–5541. H. B. (1995) Proc. Natl. Acad. Sci. USA 92, 11949–11951.
M. R., Creager, S. E., Ferraris, J. P., Chalfant, K., 46. Stuchebrukhov, A. A. (1996) J. Chem. Phys. 105, 10819– 81. Iwata, S., Ostermeier, C., Ludwig, B. & Michel, H. (1995)
Zawodzinsk, T., Feldberg, S. W. & Newton, M. D. (2003) 10829. Nature 376, 660–669.
J. Am. Chem. Soc. 125, 2004–2013. 47. Skourtis, S. S., Balabin, I. A., Kawatsu, T. & Beratan, 82. Tsukihara, T., Aoyama, H., Yamashita, E., Tomizaki, T.,
13. Helms, A., Heiler, D. & McLendon, G. (1992) J. Am. D. N. (2005) Proc. Natl. Acad. Sci. USA 102, 3552–3557. Yamaguchi, H., Shinzawa-Itoh, K., Nakashima, R., Yaono,
Chem. Soc. 114, 6227–6238. 48. Calef, D. F. & Wolynes, P. G. (1983) J. Phys. Chem. 87, R. & Yoshikawa, S. (1995) Science 269, 1071–1074.
14. Wenger, O. S., Leigh, B. S., Villahermosa, R. M., Gray, 3387–3400. 83. Yoshikawa, S., Shinzawa-Itoh, K., Nakashima, R.,
H. B. & Winkler, J. R. (2005) Science 307, 99–102. 49. Calef, D. F. & Wolynes, P. G. (1983) J. Chem. Phys. 78, Yaono, R., Yamashita, E., Inoue, N., Yao, M., Fei, J. M.,
15. Davis, W. B., Svec, W. A., Ratner, M. A. & Wasielewski, 470–482. Libeu, C. P., Mizushima, T., et al. (1998) Science 280,
M. R. (1998) Nature 396, 60–63. 50. Maroncelli, M., MacInnis, J. & Fleming, G. (1989) Sci- 1723–1729.
16. Sakata, Y., Tsue, H., O’Neil, M. P., Wiederrecht, G. P. & ence 243, 1674–1681. 84. Geren, L. M., Beasley, J. R., Fine, B. R., Saunders, A. J.,
51. Hynes, J. T. (1986) J. Phys. Chem. 90, 3701–3706. Hibdon, S., Pielak, G. J., Durham, B. & Millett, F. (1995)
Wasielewski, M. R. (1994) J. Am. Chem. Soc. 116, 6904–
52. Sumi, H. & Marcus, R. A. (1986) J. Chem. Phys. 84, J. Biol. Chem. 270, 2466–2472.
6909.
4894–4914. 85. Pan, L. P., Hibdon, S., Liu, R.-Q., Durham, B. & Millett,
17. Newton, M. D. (1997) J. Electroanal. Chem. 438, 3–10.
53. Bixon, M. & Jortner, J. (1993) Chem. Phys. 176, 467–481. F. (1993) Biochemistry 32, 8492–8498.
18. Hayashi, S. & Kato, S. (1998) J. Phys. Chem. A 102,
54. Bixon, M. & Jortner, J. (1999) Adv. Chem. Phys. 106, 35–202. 86. Roberts, V. A. & Pique, M. E. (1999) J. Biol. Chem. 274,
3333–3342.
55. Genberg, L., Richard, L., McLendon, G. & Miller, 38051–38060.
19. Weidemaier, K., Tavernier, H. L., Swallen, S. F. & Fayer,
R. J. D. (1991) Science 251, 1051–1054. 87. Flöck, D. & Helms, V. (2002) Proteins Struct. Funct.
M. D. (1997) J. Phys. Chem. A 101, 1887–1902.
56. Bashkin, J. S., McLendon, G., Mukamel, S. & Marohn, J. Genet. 47, 75–85.
20. Miller, J. R. (1975) J. Phys. Chem. 79, 1070–1078.
(1990) J. Phys. Chem. 94, 4757–4761. 88. Witt, H., Malatesta, F., Nicoletti, F., Brunori, M. &
21. Miller, J. R., Beitz, J. V. & Huddleston, R. K. (1984)
57. Pierce, D. W. & Boxer, S. G. (1992) J. Phys. Chem. 96, Ludwig, B. (1998) J. Biol. Chem. 273, 5132–5136.
J. Am. Chem. Soc. 106, 5057–5068.
5560–5566. 89. Zhen, Y. J., Hoganson, C. W., Babcock, G. T. & Fergu-
22. Swallen, S. W., Weidemaier, K., Tavernier, H. L. & Fayer, 58. Hagen, S. J. & Eaton, W. A. (1996) J. Chem. Phys. 104, son-Miller, S. (1999) J. Biol. Chem. 274, 38032–38041.
M. D. (1996) J. Phys. Chem. 100, 8106–8117. 3395–3398. 90. Wang, K. F., Zhen, Y. J., Sadoski, R., Grinnell, S., Geren,
23. Ponce, A., Gray, H. B. & Winkler, J. R. (2000) J. Am. 59. Khoshtariya, D. E., Wei, J., Liu, H., Yue, H. & Waldeck, L., Ferguson-Miller, S., Durham, B. & Millett, F. (1999)
Chem. Soc. 122, 8187–8191. D. H. (2003) J. Am. Chem. Soc. 125, 7704–7714. J. Biol. Chem. 274, 38042–38050.
24. Winkler, J. R., Wittung-Stafshede, P., Leckner, J., Malm- 60. Stemp, E. D. A. & Hoffman, B. M. (1993) Biochemistry 91. Drosou, V., Malatesta, F. & Ludwig, B. (2002) Eur.
ström, B. G. & Gray, H. B. (1997) Proc. Natl. Acad. Sci. 32, 10848–10865. J. Biochem. 269, 2980–2988.
USA 94, 4246–4249. 61. Pletneva, E. V., Fulton, D. B., Kohzuma, T. & Kostic, 92. George, S. D., Metz, M., Szilagyi, R. K., Wang, H.,
25. Gray, H. B., Malmström, B. G. & Williams, R. J. P. (2000) N. M. (2000) J. Am. Chem. Soc. 122, 1034–1046. Cramer, S. P., Lu, Y., Tolman, W. B., Hedman, B.,
J. Biol. Inorg. Chem. 5, 551–559. 62. Tezcan, F. A., Crane, B. R., Winkler, J. R. & Gray, H. B. Hodgson, K. O. & Solomon, E. I. (2001) J. Am. Chem.
26. Simonson, T. (2002) Proc. Natl. Acad. Sci. USA 99, (2001) Proc. Natl. Acad. Sci. USA 98, 5002–5006. Soc. 123, 5757–5767.
6544–6549. 63. Lo Conte, L., Chothia, C. & Janin, J. (1999) J. Mol. Biol. 93. Farver, O., Einarsdóttir, O. & Pecht, I. (2000) Eur.
27. Winkler, J. R., Nocera, D. G., Yocom, K. M., Bordignon, E. 285, 2177–2198. J. Biochem. 267, 950–954.
& Gray, H. B. (1982) J. Am. Chem. Soc. 104, 5798–5800. 64. Pelletier, H. & Kraut, J. (1992) Science 258, 1748–1755. 94. Regan, J. J., Ramirez, B. E., Winkler, J. R., Gray, H. B. &
28. Winkler, J. R. & Gray, H. B. (1992) Chem. Rev. 92, 369–379. 65. Miyashita, O., Okamura, M. Y. & Onuchic, J. N. (2005) Malmström, B. G. (1998) J. Bioenerg. Biomembr. 30, 35–39.
29. Gray, H. B. & Winkler, J. R. (1996) Annu. Rev. Biochem. Proc. Natl. Acad. Sci. USA 102, 3558–3563. 95. Gamelin, D. R., Randall, D. W., Hay, M. T., Houser,
65, 537–561. 66. McGourty, J. L., Blough, N. V. & Hoffman, B. M. (1983) R. T., Mulder, T. C., Canters, G. W., de Vries, S., Tolman,
30. Winkler, J. R., Di Bilio, A., Farrow, N. A., Richards, J. H. J. Am. Chem. Soc. 105, 4470–4472. W. B., Lu, Y. & Solomon, E. I. (1998) J. Am. Chem. Soc.
& Gray, H. B. (1999) Pure Appl. Chem. 71, 1753–1764. 67. Kuila, D., Baxter, W. W., Natan, M. J. & Hoffman, B. M. 120, 5246–5263.
31. Langen, R., Chang, I.-J., Germanas, J. P., Richards, J. H., (1991) J. Phys. Chem. 95, 1–3. 96. Medvedev, D. M., Daizadeh, I. & Stuchebrukhov, A. A.
Winkler, J. R. & Gray, H. B. (1995) Science 268, 1733– 68. Peterson-Kennedy, S. E., McGourty, J. L., Kalweit, J. A. (2000) J. Am. Chem. Soc. 122, 6571–6582.
1735. & Hoffman, B. M. (1986) J. Am. Chem. Soc. 108, 1739– 97. Sjöberg, B. M. (1997) Struct. Bonding 88, 139–173.
32. Regan, J. J., Di Bilio, A. J., Langen, R., Skov, L. K., 1746. 98. Stubbe, J., Nocera, D. G., Yee, C. S. & Chang, M. C. Y.
Winkler, J. R., Gray, H. B. & Onuchic, J. N. (1995) Chem. 69. Dick, L. A., Malfant, I., Kuila, D., Nebolsky, S., Nocek, (2003) Chem. Rev. 103, 2167–2201.
Biol. 2, 489–496. J. M., Hoffman, B. M. & Ratner, M. A. (1998) J. Am. 99. Page, C. C., Moser, C. C., Chen, X. & Dutton, P. L. (1999)
33. Adman, E. T. (1991) Adv. Protein Chem. 42, 145–197. Chem. Soc. 120, 11401–11407. Nature 402, 47–52.
34. Crane, B. R., Di Bilio, A. J., Winkler, J. R. & Gray, H. B. 70. Mauk, A. G., Mauk, M. R., Moore, G. R. & Northrup, 100. Goldsmith, R. H., Sinks, L. E., Kelley, R. F., Betzen, L. J.,
(2001) J. Am. Chem. Soc. 123, 11623–11631. S. H. (1995) J. Bioenerg. Biomemb. 27, 311–330. Liu, W., Weiss, E. A., Ratner, M. A. & Wasielewski,
35. Smalley, J. F., Feldberg, S. W., Chidsey, C. E. D., Linford, 71. Durham, B., Fairris, J. L., McLean, M., Millett, F., Scott, M. R. (2005) Proc. Natl. Acad. Sci. USA 102, 3540–3545.
M. R., Newton, M. D. & Liu, Y.-P. (1995) J. Phys. Chem. J. R., Sligar, S. G. & Willie, A. (1995) J. Bioenerg. 101. Yavin, E., Boal, A. K., Stemp, E. D. A., Boon, E. M.,
99, 13141–13149. Biomemb. 27, 331–340. Livingston, A. L., O’Shea, V. L., David, S. S. & Barton,
36. Farver, O. & Pecht, I. (1999) Adv. Chem. Phys. 107, 555–589. 72. Salemme, F. R. (1976) J. Mol. Biol. 102, 563–568. J. K. (2005) Proc. Natl. Acad. Sci. USA 102, 3546–3551.

Gray and Winkler PNAS 兩 March 8, 2005 兩 vol. 102 兩 no. 10 兩 3539

You might also like