You are on page 1of 19

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/12079747

Electron Spin Resonance in Studies of Membranes and Proteins

Article  in  Science · February 2001


DOI: 10.1126/science.291.5502.266 · Source: PubMed

CITATIONS READS
313 1,180

5 authors, including:

Antonio J Costa-Filho Keith A. Earle


University of São Paulo University at Albany, The State University of New York
150 PUBLICATIONS   2,295 CITATIONS    59 PUBLICATIONS   1,516 CITATIONS   

SEE PROFILE SEE PROFILE

Jozef K. Moscicki
Jagiellonian University
84 PUBLICATIONS   1,499 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Unraveling the molecular bases of the early protein secretory pathway in humans using biophysical techniques View project

Hsp70s View project

All content following this page was uploaded by Antonio J Costa-Filho on 19 May 2014.

The user has requested enhancement of the downloaded file.


Electron Spin Resonance in Studies of Membranes and
Proteins
P. P. Borbat,1 A. J. Costa-Filho,1 K. A. Earle,1 J. K. Moscicki,1, 2 J. H. Freed1*

We provide a review of current electron spin resonance (ESR) techniques for studying
basic molecular mechanisms in membranes and proteins by using nitroxide spin labels. In
particular, nitroxide spin label studies with high-field/high-frequency ESR and two-
dimensional Fourier transform ESR enable one to accurately determine distances in
biomolecules, unravel the details of the complex dynamics in proteins, characterize the
dynamic structure of membrane domains, and discriminate between bulk lipids and
boundary lipids that coat transmembrane peptides or proteins; these studies can also
provide time resolution to studies of functional dynamics of proteins. We illustrate these
capabilities with recent examples.
1
Baker Laboratory of Chemistry and Chemical Biology, Cornell University, Ithaca, NY
14853-1301, USA.
2
Smoluchowski Institute of Physics, Jagellonian University, Ulica Reymonta 4, 30-059
Krakow, Poland.
*
To whom correspondence should be addressed. E-mail: jhf@ccmr.cornell.edu

ESR, a sister technique to nuclear magnetic resonance (NMR), has some key advantages
and disadvantages in comparison to NMR. Because the electron has a much greater
magnetic moment than a nucleus, ESR is much more sensitive per spin. In time domain
experiments [e.g., free induction decays, spin echoes, or two-dimensional (2D)
spectroscopy], ESR's time scale is nanoseconds versus NMR's time scale of milliseconds.
Thus, ESR is naturally suited for studying faster dynamics. The absence of unpaired
electrons in most chemical and biological materials would appear to be a major
impediment for the applicability of ESR, but it can be used as an important virtue. Spin
labeling of selected molecules with a small spin-bearing moiety that has a simple ESR
signal, as well as the limited number of spins contributing to the ESR experiment, enables
a broad range of problem-targeted studies. Most biological studies have been performed
with conventional ESR (1-5), but we consider newer developments here.

ESR spectroscopy, based on nitroxide spin labeling, is largely driven by the sensitivity of
the nitroxide label to its surroundings. This comes about in a number of ways. The
interaction of the unpaired electron spin of the nitroxide with the 14N magnetic nucleus
(I = 1) leads to an electron-nuclear dipolar (or hyperfine, hf) tensor, which, for a rapidly
tumbling molecule, averages to a nonzero value (the hf splitting). This is also the case for
the g tensor of the electron spin (analogous to the chemical shift tensor in NMR), which,
when averaged, yields the g shift. If the investigated system (e.g., a membrane) is
macroscopically aligned, then one can observe the different "single-crystal-like" spectra
obtained for each orientation of the system with respect to the constant magnetic field.
The homogeneous line broadening (hb) of the spectra would then reflect the motional
dynamics.
ESR spectra are known to change dramatically as the tumbling motion of the probe
slows, thus providing great sensitivity to fluidity in the neighborhood of the spin probe
(1). This is different from NMR, where the molecular motions in liquids lead to nearly
complete averaging of the motion-dependent terms in the spin Hamiltonian. Their
residual effects are then reflected only in the relaxation times, T1 and T2, and may be
accounted for by Redfield perturbation theory. However, the equivalent terms for ESR are
much larger, frequently leading to effects that are too dramatic to be addressed by
perturbation theory. Instead, an approach based on the stochastic Liouville equation,
known as slow-motional theory, has been developed (6-10), which shows that the
dramatic line shape changes are particularly sensitive to the microscopic details of the
dynamics.

These methods of analysis enable a quantitative assessment of ESR spectra in terms of


motional rates of spinning of the nitroxide moiety, the end-over-end tumbling of the
tagged molecule, and local environmental constraints to the motion. For a spin-labeled
protein, for example, the motions could be the spinning about the tether attaching the
nitroxide and the overall tumbling of the protein. In a bilayer, typical of a cell membrane,
reorientation of the lipids is constrained by the surrounding lipids and other molecules.
The motional rates lead to a rotational diffusion tensor, whereas the motional constraints
lead to an orientational order parameter.

The changes in magnitude of the hf splitting and g shift can be used to monitor features
of the local surroundings, such as its polarity (11, 12). In addition, unpaired electron spins
from different nitroxides (on the same or different molecules) interact weakly through
long-range magnetic dipolar interactions or strongly through short-range Heisenberg spin
exchange. In fluid media, these latter interactions can be used to monitor microscopic
translational dynamics (13, 14). In frozen or very viscous media, the dipolar interactions
can be used to measure distances (4), either by continuous wave (cw) ESR (4, 5, 10, 15,
16) or pulsed ESR (4, 17).

Often, one must deal with the complications of samples that are MOMD [i.e.,
microscopically ordered but macroscopically disordered (7, 9)]. Membrane vesicles offer
a physical example of this property. Spin-labeled moieties in the different vesicle regions
may then be oriented at all angles with respect to the magnetic field, thus providing a
"powder-like" spectrum with inhomogeneous line broadening (ib) superimposed on the
hb. The degree of the ib is determined by the extent of local (microscopic) ordering. This
ib masks the hb, resulting in reduced resolution to dynamic and ordering parameters.
Despite this, modeling of the heterogeneity and of the dynamic effects to fit the ESR
spectrum can yield important insights (18, 19).

Technical Advances

New instrumentation greatly enhances these capabilities and allows ESR to be applied
more effectively to the complex structural and dynamic aspects of membranes and
proteins. There are two principal avenues of contemporary ESR developments (20, 21).
The first is the extension of cw ESR to high fields/high frequencies (22-24).
The second is time domain, or pulsed ESR (20, 21).

High-frequency (HF) ESR. The primary advantages of extending ESR to


HF (22-24) include an increased signal-to-noise ratio and improved spectral resolution.
As one moves to HF, two important features emerge. First, for a given diffusion rate, as
the ESR frequency increases, spin-label motion appears to become slower (25); that is,
HF ESR spectra act as a faster "snapshot" of the dynamics (cf. Fig. 1, right). At the low-
frequency end, one observes motionally narrowed spectra, whereas, at the high-frequency
end, the spectra display very slow motion, almost at the rigid limit (due to the increased
importance of the g tensor).

Fig. 1. Nitroxide cw ESR spectra change dramatically with frequency. (Left) Simulations
illustrate the increased orientational sensitivity of ESR on going from low (9 GHz) to
high (250 GHz) frequencies. The derivative spectra are characteristic of a spin label in a
frozen amorphous matrix (i.e., the rigid limit). The triplet hyperfine splittings (marked by
a set of three vertical lines) are due to the interaction of the unpaired electron with the 14N
nucleus. (Right) The derivative ESR simulations for characteristic ESR frequencies
demonstrate the snapshot property of ESR when studying tumbling spin-labeled
molecules. The large variations that the spectrum undergoes as the frequency is changed
are very sensitive to the character of the motion. [View Larger Version of this Image
(16K GIF file)]

The second feature is the great improvement in orientational resolution of the nitroxide
spectrum (22-24, 26), illustrated in Fig. 1, left. At 250 GHz, the regions corresponding to
nitroxides with their x axis B0 (i.e., x axis is parallel to the main magnetic field, B0), y
B B

axis B0, and z axis B0 are well separated, because of the dominant role of the g tensor;
B B

this separation is not observed at 9 GHz. Then, because one can discern the axis (or axes)
about which the motion occurs, the 250-GHz slow-motional spectra are much more
sensitive to the details of the motions than are those at microwave frequencies (26, 27).

Pulsed ESR. Pulsed ESR techniques (20, 21), especially in two dimensions, that have
many analogies to 2D NMR are a second highlight. To achieve greater reliability in
analyzing spectra, one must improve spectral resolution in studies of the dynamic
structure of biosystems. Foremost is the need to distinguish between the mechanisms of
hb and ib, which is achieved with the aid of spin echoes. A technique that accomplishes
this is 2D electron-electron double resonance (ELDOR) (28-31), which also supplies off-
diagonal peaks (cross peaks) that directly report on translational and rotational motions of
labeled biomolecules.

Double quantum coherence (DQC) is a new and powerful application of pulsed ESR for
distance measurements between interacting spins in frozen media (32, 33). The virtue of
DQC ESR shown in Fig. 2 is that it permits detection of only the weak dipolar interaction
between the two nitroxides in the ESR signal. Oscillations in echo
amplitude versus echo time are due to this dipolar interaction (Fig. 2). A
detailed analysis of DQC indicates that distances can be measured to ~80
Å, comparable to the capabilities of fluorescence energy transfer (33).

Fig. 2. Principles of the pulsed DQC ESR experiment. (Left) The doubly labeled
molecule is dissolved in a frozen amorphous matrix. (Bottom) The six-pulse DQC
sequence preserves contributions to the double quantum echo signal arising solely from
the dipolar interaction between the pair of nitroxide spin labels. One varies the time
intervals tp and t2 such that the total time of the pulse sequence remains constant. (Right)
The maximum of the echo signal is recorded as a function of t2 tp. (Top) A Fourier
transform of this echo signal yields the Pake-type dipolar spectrum in the frequency
domain. The separation of the two sharp peaks is directly related to the distance r
1/3
between the labels, r . (The echo signal and the dipolar spectrum shown here were
measured for the molecule on the left, for which r 16.2 ± 0.5 Å.) [View Larger Version
of this Image (18K GIF file)]

Illustrative Applications

HF cw ESR and complex dynamics in proteins and membranes. The snapshot feature of
cw ESR encourages a multifrequency approach to the study of the complex modes of
motion of proteins, DNA, and other polymers, enabling the decomposition of these modes
according to their different time scales (25). For example, in the case of proteins, the
higher frequency ESR spectra should "freeze out" the slow overall tumbling motions,
leaving only the faster internal modes of motion. Alternatively, ESR performed at lower
frequencies would be sensitive to the motions on a slower time scale. This is in contrast to
aqueous solution NMR, where all modes are usually in the fast motional limit and
contribute only to T1 and T2.

The virtues of such an approach were demonstrated in a study at 9 and 250 GHz on spin-
labeled mutants of T4 lysozyme in aqueous solution (34). On the short time scale of the
250-GHz ESR experiment, the overall tumbling was too slow to affect the spectrum; thus,
a MOMD analysis provided a satisfactory modeling of the overall tumbling, and HF-
enhanced spectral resolution reported on the internal dynamics (Fig. 3). The analysis of
the 9-GHz spectra was more complex, because the longer time scale did not freeze out the
overall tumbling motion. The slowly relaxing local structure (SRLS) model (25)
simultaneously incorporated both the internal and overall motions (Fig. 3, bottom). By
fixing the internal motional parameters at the values obtained from the 250-GHz data, fits
to the 9-GHz line shapes generated by SRLS successfully yielded the rate for the global
dynamics. Thus, exploiting the time scale separation allowed more information to be
extracted [see (35-38)].
Fig. 3. A multifrequency ESR study of nitroxide spin-labeled T4 lysozyme in
aqueous solution. The derivative spectra are for the spin label on residue 44 and
are taken at 10°C (34). The relevant molecular motions are shown schematically
in the lower part of the figure. The protein tumbles slowly about its principal
axes of overall diffusion; this is the SRLS. The motion of the spin label moiety is
restricted by its tether and its surroundings to be within a cone, whose main axis
makes an angle with the protein main axis. The spatial extent of the internal rotational
diffusion modes and their rates are distinguished from the protein overall tumbling rates
as described in the text. [View Larger Version of this Image (21K GIF file)]

In a demonstration of the excellent orientational resolution at 250 GHz, macroscopically


aligned membranes containing a mixture of head groups [zwitterionic
phosphatidylcholine (PC) and negatively charged phosphatidylserine (PS)] were studied
by using the cholesterol-like spin label, cholestane (CSL) (26). The direction of alignment
of the CSL molecules could be clearly discerned as the orientation of the membrane
normal was changed with respect to the magnetic field. The CSL in PC-rich membranes
exhibited typical behavior; that is, its long axis was parallel to the bilayer normal, and its
rotational diffusion rates R were slow (R ~ 106 to 107 s 1). However, the behavior in PS-
rich membranes is unusual and was interpreted in terms of a strong local biaxial
environment. Although predicted from molecular dynamics simulations, this gives
experimental evidence for local biaxiality. By contrast, given the poorer orientational
resolution at 9.1 GHz, it was not possible to distinguish such detail from the 9.1-GHz
experiment.

Two-dimensional Fourier transform (FT) ESR and dynamic structure of membranes.


Two-dimensional time domain ESR methods allow one to study both dynamics and
ordering in membrane vesicles (30, 31). In general, 2D ELDOR spectra from membrane
vesicles exhibit more dramatic variations due to changes in membrane properties than do
spectra obtained by cw ESR. A simple interpretation of these spectra is thus possible
merely by pattern recognition. For example, 2D ELDOR contour plots are shown for the
spin-labeled lipid, 16 PC, in pure lipid vesicles (Fig. 4A) and for a lipid-cholesterol (1:1)
mixture (Fig. 4B). The former is in the standard liquid crystalline (LC) phase, whereas the
latter is in a "liquid ordered" (LO) phase. This type of LO phase has been found to play
an important role in cellular signaling and in membrane trafficking (19, 39). Visually, the
spectra are qualitatively different, emphasizing that the LO phase exhibits substantially
greater ordering, but comparable rotational mobility, in comparison with the LC phase.
Quantitative spectral fits confirm that the increased microscopic ordering leads to
increased ib affecting the spectra from the LO phase. Furthermore, the restriction of the
range of orientational motion, due to the microscopic ordering in the LO phase, shows up
as a much slower development of cross peaks.
Fig. 4. The sensitivity of pulsed 2D ELDOR to the dynamic structure of
lipid membranes. Contour plots for the end-chain-labeled lipid, 16 PC,
dissolved in lipid vesicles containing (A) only di-
palmitoylphosphatidylcholine (DPPC) in the LC (55°C) phase and (B) an
equimolar mixture of DPPC and cholesterol in the LO (55°C) phase. The
2D spectra from 16 PC dissolved (C) in pure DPPC vesicles and (D) in an equimolar
mixture of DPPC and GA, both at 75°C. The spectrum in (C) is characteristic of bulk
lipid. In the spectrum in (D), it is somewhat broadened, but a second, much broader
component is apparent and is attributed to boundary lipid. All spectra are at 17.3 GHz.
[View Larger Version of this Image (40K GIF file)]

Two-dimensional FT ESR and lipid-protein interactions. The merits of 2D FT ESR


spectroscopy are well demonstrated in studies of the effect of the peptide gramicidin A
(GA) on the dynamic structure of model membranes. Besides bulk lipids, boundary lipids
that coat the peptide are present but difficult to resolve in cw ESR. Two-dimensional
ELDOR at 9.3 GHz (31) yielded dramatic changes in spectra as GA was added to
vesicles, because of changes in the dynamic structure of bulk lipids. However, newer 2D
ELDOR results at 17.3 GHz, with improved time resolution, show the presence of two
components (40). The boundary lipid appears as a broader spectral component,
characteristic of a more slowly reorienting lipid with greater motional restriction (Fig. 4,
C and D). Aligning the membranes enables resolution of boundary lipids, even in the
presence of small amounts of GA (1 to 4 mole percent) (30, 40).

Multiple-quantum ESR and distance measurements. The determination of intra- and


intermolecular distances has become an important application of contemporary ESR
spectroscopy. Areas of interest include the structure of protein complexes and the
functional dynamics of proteins that are neither soluble nor crystallized (4, 5, 10, 15, 16).

Strong DQC signals for a variety of bilabeled nitroxide molecules in both disordered and
oriented solids have been achieved recently (32, 33) (cf. Fig. 2). For a benchmark rigid
linear biradical, piperidinyl-CO2-(phenyl)4-O2C-piperidinyl, a distance of
28.8 ± 0.5 Å between the two nitroxide groups was obtained (versus 28.06 Å from
molecular modeling). For a semiflexible bilabeled synthetic peptide [MTSL-CPPPPC-
MTSL in frozen solution at 82°C (MTSL, the [1-oxyl-2,2,5,5-tetramethyl-pyrrole-3-
methyl] methanethiosulfonate spin label; C, Cys; and P, Pro) relevant to biophysical
studies, a mean separation of 22.6 Å between the two nitroxide groups and a distribution
half-width of 2 Å (which smears out the Pake doublet) were found. Current sensitivities
require ~300 pmol of bilabeled molecules, but it is estimated that, in the future, studies
with samples that are approximately an order of magnitude less will be feasible (33).

Future Prospects

Given the virtues of time domain ESR (such as direct determination of relaxation rates
and differentiation between hb and ib) compared to the virtues of HF cw ESR (enhanced
orientational selectivity and greater sensitivity to nuances of the molecular dynamics), it
would be desirable to extend pulse techniques to higher frequencies (41). However, as the
frequency is increased, spin-labeled biomolecules in fluid media would have shorter
transverse decay times, because of enhanced hb and ib. Thus, even shorter time resolution
and greater spectral bandwidths would be required, hence shorter and more intense
radiation pulses (30). Progress in developing a coherent pulsed high-power spectrometer
at 95 GHz has recently been reported (42).

The fast time scales of 2D FT ESR (30) are poised to address the functional dynamics (5)
of proteins. Nanosecond resolution is achievable in following the time evolution of a
(small) region of the spectrum, and 0.1- to 1-µs resolution is possible for the entire
nitroxide spectrum. cw ESR studies of the transient behavior of bilabeled
bacteriorhodopsin following light activation were able to resolve key distance changes
occurring on the 1-ms to 1-s time scale (43, 44). FT ESR combined with DQC should
enable studies in the submicrosecond range. Similar applications for short time dynamics
would be for rapid mixing and stopped-flow experiments (45), limited only by the time
resolution of these methods.

REFERENCES

1. L. J. Berliner, Ed., Spin Labeling: Theory and Applications (Academic Press,


New York, 1976).
2. ___, J. Reuben, Eds., Spin Labeling: Theory and Applications, vol. 8 of
Biological Magnetic Resonance (Plenum, New York, 1989).
3. L. J. Berliner, Ed., Spin Labeling: The Next Millennium, vol. 14 of Biological
Magnetic Resonance (Plenum, New York, 1998).
4. G. R. Eaton, S. S. Eaton, L. J. Berliner, Eds., Distance Measurements in
Biological Systems by EPR, vol. 19 of Biological Magnetic Resonance (Kluwer,
New York, 2000).
5. W. L. Hubbell, D. S. Cafiso, C. Altenbach, Nature Struct. Biol. 7, 735 (2000)
[CrossRef][ISI][Medline] .
6. J. H. Freed, in (1), pp. 53-132.
7. D. J. Schneider, J. H. Freed, in (2), pp. 1-76.
8. A. H. Beth, B. H. Robinson, in (2), pp. 179-253.
9. D. E. Budil, S. Lee, S. Saxena, J. H. Freed, J. Magn. Reson. A120, 155 (1996) .
10. E. Hustedt, et al., Biophys. J. 74, 1861 (1997) .
11. O. H. Griffith, P. J. Dehlinger, S. P. Van, J. Membr. Biol. 15, 159 (1974)
[ISI][Medline] .
12. K. A. Earle, et al., Biophys. J. 66, 1213 (1994) [Abstract]
13. J. H. Freed, Annu. Rev. Biophys. Biomol. Struct. 23, 1 (1994) [ISI][Medline] .
14. D. Marsh, in (2), pp. 255-303.
15. M. D. Rabenstein and Y.-K. Shin, Proc. Natl. Acad. Sci. U.S.A. 92, 823 (1995) .
16. H. J. Steinhoff, et al., Biophys. J. 73, 3287 (1997) [Abstract] .
17. A. D. Milov, A. G. Maryasov, Yu. D. Tsvetkov, Appl. Magn. Reson. 15, 107
(1998) [ISI].
18. M. Ge and J. H. Freed, Biophys. J. 76, 264 (1999) [Abstract/Full Text] .
19. M. Ge, et al., Biophys. J. 77, 925 (1999) [Abstract/Full Text] .
20. A. J. Hoff, Ed., Advanced EPR, Applications in Biology and Biochemistry
(Elsevier, New York, 1989).
21. L. Kevan, M. K. Bowman, Eds., Modern Pulsed and Continuous-Wave Electron
Spin Resonance (Wiley, New York, 1990).
22. D. E. Budil et al., in (20), pp. 307-340.
23. Y. S. Lebedev, in (21), pp. 365-404.
24. International Experts' Workshop: Present and Future of HFEPR Instrumentation,
S. S. Eaton, G. R. Eaton, Eds., Appl. Magn. Reson. 16, 159 (1999).
25. Z. Liang and J. H. Freed, J. Phys. Chem. B 103, 6384 (1999) [CrossRef][ISI] .
26. J. B. Barnes and J. H. Freed, Biophys. J. 75, 2532 (1998) [Abstract/Full Text] .
27. K. A. Earle, et al., J. Chem. Phys. 106, 9996 (1997) [CrossRef][ISI] .
28. J. Gorcester, G. L. Millhauser, J. H. Freed, in (20), pp. 177-242; in (21), pp. 119-
194.
29. M. K. Bowman, in (21), pp. 1-42.
30. P. P. Borbat, R. H. Crepeau, J. H. Freed, J. Magn. Reson. 127, 155 (1997)
[CrossRef][ISI][Medline].
31. B. Patyal, R. H. Crepeau, J. H. Freed, Biophys. J. 73, 2201 (1997) [Abstract] .
32. P. P. Borbat and J. H. Freed, Chem. Phys. Lett. 313, 145 (1999) [CrossRef][ISI].
33. P. P. Borbat, J. H. Freed, in (4), pp. 383-459.
34. J. Barnes, et al., Biophys. J. 76, 3298 (1999) [Abstract/Full Text] .
35. J. Pilar, et al., Macromolecules 33, 4438 (2000) [CrossRef][ISI] .
36. Z. Liang, A. M. Bobst, R. S. Keyes, J. H. Freed, J. Phys. Chem. B 104, 5372
(2000) [CrossRef][ISI] .
37. D. E. Budil, et al., Biophys. J. 78, 430 (2000) [Abstract/Full Text] .
38. M. Beunati, et al., J. Magn. Reson. 139, 281 (1999) [CrossRef][ISI][Medline].
39. D. A. Brown and E. London, Annu. Rev. Cell Dev. Biol. 14, 111 (1998)
[Abstract/Full Text].
40. P. P. Borbat, R. H. Crepeau, M. Ge, J. H. Freed, paper presented at the 22nd
International EPR Symposium, Denver, CO, 1 to 5 August 1999.
41. T. F. Prisner, Adv. Magn. Opt. Reson. 20, 245 (1997) .
42. K. A. Earle, C. Dunnam. P. P. Borbat and J. H. Freed, Am. Phys. Soc. Abstr. 45,
542 (2000) .
43. T. E. Thorgeirsson, et al., J. Mol. Biol. 273, 951 (1997) [CrossRef][ISI][Medline]
.
44. R. Mollaaghababa, et al., Biochemistry 39, 1120 (2000) [CrossRef][ISI][Medline]
.
45. V. M. Grigoryants, A. V. Veselov, C. P. Scholes, Biophys. J. 78, 2702 (2000)
[Abstract/Full Text] .
Electron Spin Resonance in Studies of Membranes and Proteins -- Borbat et al. 291 (5502): 266 -- Science

Jump to: Page Content, Section Navigation, Site Navigation, Site Search, Account Information, or Site Tools.

You are seeing this message because your web browser does not support basic web standards. Find out more about why this message is appearing and what you can do to
make your experience on this site better.

Site Tools



Site Search
Site Area Science Magazine Terms Advanced
Account Information
CORNELL UNIV LIBRARY Alerts | Access Rights | My Account | Sign In

Site Navigation

Readers

● Current Issue
● Previous Issues
● Science Express
● Science Products
● My Science
● About the Journal

Home > Science Magazine > 12 January 2001 > Borbat et al., pp. 266 - 269

http://www.sciencemag.org/cgi/content/full/291/5502/266 (1 of 10)2/27/2006 3:29:45 PM


Electron Spin Resonance in Studies of Membranes and Proteins -- Borbat et al. 291 (5502): 266 -- Science

Science 12 January 2001:


Vol. 291. no. 5502, pp. 266 - 269
DOI: 10.1126/science.291.5502.266 Prev | Table of Contents | Next

Review

Electron Spin Resonance in Studies of Membranes and


Proteins
P. P. Borbat,1 A. J. Costa-Filho,1 K. A. Earle,1 J. K. Moscicki,1, 2 J. H. Freed1*

We provide a review of current electron spin resonance (ESR) techniques for studying basic
molecular mechanisms in membranes and proteins by using nitroxide spin labels. In particular,
nitroxide spin label studies with high-field/high-frequency ESR and two-dimensional Fourier
transform ESR enable one to accurately determine distances in biomolecules, unravel the details of
the complex dynamics in proteins, characterize the dynamic structure of membrane domains, and
discriminate between bulk lipids and boundary lipids that coat transmembrane peptides or proteins;
these studies can also provide time resolution to studies of functional dynamics of proteins. We
illustrate these capabilities with recent examples.

1 Baker Laboratory of Chemistry and Chemical Biology, Cornell University, Ithaca, NY 14853-
1301, USA.
2 Smoluchowski Institute of Physics, Jagellonian University, Ulica Reymonta 4, 30-059 Krakow,
Poland.
* To whom correspondence should be addressed. E-mail: jhf{at}ccmr.cornell.edu

ESR, a sister technique to nuclear magnetic resonance (NMR), has some


key advantages and disadvantages in comparison to NMR. Because the
electron has a much greater magnetic moment than a nucleus, ESR is much
more sensitive per spin. In time domain experiments [e.g., free induction
decays, spin echoes, or two-dimensional (2D) spectroscopy], ESR's time
scale is nanoseconds versus NMR's time scale of milliseconds. Thus, ESR
is naturally suited for studying faster dynamics. The absence of unpaired
electrons in most chemical and biological materials would appear to be a
major impediment for the applicability of ESR, but it can be used as an
important virtue. Spin labeling of selected molecules with a small spin-
bearing moiety that has a simple ESR signal, as well as the limited number
of spins contributing to the ESR experiment, enables a broad range of
problem-targeted studies. Most biological studies have been performed
with conventional ESR (1-5), but we consider newer developments here.
ESR spectroscopy, based on nitroxide spin labeling, is largely driven by the sensitivity of the
nitroxide label to its surroundings. This comes about in a number of ways. The interaction of the
unpaired electron spin of the nitroxide with the 14N magnetic nucleus (I = 1) leads to an electron-
nuclear dipolar (or hyperfine, hf) tensor, which, for a rapidly tumbling molecule, averages to a
nonzero value (the hf splitting). This is also the case for the g tensor of the electron spin (analogous
to the chemical shift tensor in NMR), which, when averaged, yields the g shift. If the investigated
system (e.g., a membrane) is macroscopically aligned, then one can observe the different "single-
crystal-like" spectra obtained for each orientation of the system with respect to the constant magnetic
field. The homogeneous line broadening (hb) of the spectra would then reflect the motional
dynamics.

ESR spectra are known to change dramatically as the tumbling motion of the probe slows, thus
providing great sensitivity to fluidity in the neighborhood of the spin probe (1). This is different from
NMR, where the molecular motions in liquids lead to nearly complete averaging of the motion-
dependent terms in the spin Hamiltonian. Their residual effects are then reflected only in the
relaxation times, T1 and T2, and may be accounted for by Redfield perturbation theory. However, the
equivalent terms for ESR are much larger, frequently leading to effects that are too dramatic to be

http://www.sciencemag.org/cgi/content/full/291/5502/266 (2 of 10)2/27/2006 3:29:45 PM


Electron Spin Resonance in Studies of Membranes and Proteins -- Borbat et al. 291 (5502): 266 -- Science

addressed by perturbation theory. Instead, an approach based on the stochastic Liouville equation,
known as slow-motional theory, has been developed (6-10), which shows that the dramatic line
shape changes are particularly sensitive to the microscopic details of the dynamics.

These methods of analysis enable a quantitative assessment of ESR spectra in terms of motional
rates of spinning of the nitroxide moiety, the end-over-end tumbling of the tagged molecule, and
local environmental constraints to the motion. For a spin-labeled protein, for example, the motions
could be the spinning about the tether attaching the nitroxide and the overall tumbling of the protein.
In a bilayer, typical of a cell membrane, reorientation of the lipids is constrained by the surrounding
lipids and other molecules. The motional rates lead to a rotational diffusion tensor, whereas the
motional constraints lead to an orientational order parameter.

The changes in magnitude of the hf splitting and g shift can be used to monitor features of the local
surroundings, such as its polarity (11, 12). In addition, unpaired electron spins from different
nitroxides (on the same or different molecules) interact weakly through long-range magnetic dipolar
interactions or strongly through short-range Heisenberg spin exchange. In fluid media, these latter
interactions can be used to monitor microscopic translational dynamics (13, 14). In frozen or very
viscous media, the dipolar interactions can be used to measure distances (4), either by continuous
wave (cw) ESR (4, 5, 10, 15, 16) or pulsed ESR (4, 17).

Often, one must deal with the complications of samples that are MOMD [i.e., microscopically
ordered but macroscopically disordered (7, 9)]. Membrane vesicles offer a physical example of this
property. Spin-labeled moieties in the different vesicle regions may then be oriented at all angles
with respect to the magnetic field, thus providing a "powder-like" spectrum with inhomogeneous line
broadening (ib) superimposed on the hb. The degree of the ib is determined by the extent of local
(microscopic) ordering. This ib masks the hb, resulting in reduced resolution to dynamic and
ordering parameters. Despite this, modeling of the heterogeneity and of the dynamic effects to fit the
ESR spectrum can yield important insights (18, 19).

Technical Advances
New instrumentation greatly enhances these capabilities and allows ESR to
be applied more effectively to the complex structural and dynamic aspects
of membranes and proteins. There are two principal avenues of
contemporary ESR developments (20, 21). The first is the extension of cw
ESR to high fields/high frequencies (22-24). The second is time domain, or
pulsed ESR (20, 21).
High-frequency (HF) ESR. The primary advantages of extending ESR to HF (22-24) include an
increased signal-to-noise ratio and improved spectral resolution. As one moves to HF, two important
features emerge. First, for a given diffusion rate, as the ESR frequency increases, spin-label motion
appears to become slower (25); that is, HF ESR spectra act as a faster "snapshot" of the dynamics
(cf. Fig. 1, right). At the low-frequency end, one observes motionally narrowed spectra, whereas, at
the high-frequency end, the spectra display very slow motion, almost at the rigid limit (due to the
increased importance of the g tensor).

Fig. 1. Nitroxide cw ESR spectra


change dramatically with
frequency. (Left) Simulations
illustrate the increased
orientational sensitivity of ESR on
going from low (9 GHz) to high
(250 GHz) frequencies. The
derivative spectra are characteristic
of a spin label in a frozen
amorphous matrix (i.e., the rigid
limit). The triplet hyperfine splittings (marked by a set of three vertical
lines) are due to the interaction of the unpaired electron with the 14N

http://www.sciencemag.org/cgi/content/full/291/5502/266 (3 of 10)2/27/2006 3:29:45 PM


Electron Spin Resonance in Studies of Membranes and Proteins -- Borbat et al. 291 (5502): 266 -- Science

nucleus. (Right) The derivative ESR simulations for characteristic ESR


frequencies demonstrate the snapshot property of ESR when studying
tumbling spin-labeled molecules. The large variations that the spectrum
undergoes as the frequency is changed are very sensitive to the character of
the motion. [View Larger Version of this Image (16K GIF file)]

The second feature is the great improvement in orientational resolution of the nitroxide spectrum (22-
24, 26), illustrated in Fig. 1, left. At 250 GHz, the regions corresponding to nitroxides with their x

axis B0 (i.e., x axis is parallel to the main magnetic field, B0), y axis B0, and z axis B0 are well
separated, because of the dominant role of the g tensor; this separation is not observed at 9 GHz.
Then, because one can discern the axis (or axes) about which the motion occurs, the 250-GHz slow-
motional spectra are much more sensitive to the details of the motions than are those at microwave
frequencies (26, 27).

Pulsed ESR. Pulsed ESR techniques (20, 21), especially in two dimensions, that have many
analogies to 2D NMR are a second highlight. To achieve greater reliability in analyzing spectra, one
must improve spectral resolution in studies of the dynamic structure of biosystems. Foremost is the
need to distinguish between the mechanisms of hb and ib, which is achieved with the aid of spin
echoes. A technique that accomplishes this is 2D electron-electron double resonance (ELDOR) (28-
31), which also supplies off-diagonal peaks (cross peaks) that directly report on translational and
rotational motions of labeled biomolecules.

Double quantum coherence (DQC) is a new and powerful application of pulsed ESR for distance
measurements between interacting spins in frozen media (32, 33). The virtue of DQC ESR shown in
Fig. 2 is that it permits detection of only the weak dipolar interaction between the two nitroxides in
the ESR signal. Oscillations in echo amplitude versus echo time are due to this dipolar interaction
(Fig. 2). A detailed analysis of DQC indicates that distances can be measured to ~80 Å, comparable
to the capabilities of fluorescence energy transfer (33).

Fig. 2. Principles of the pulsed


DQC ESR experiment. (Left) The
doubly labeled molecule is
dissolved in a frozen amorphous
matrix. (Bottom) The six-pulse
DQC sequence preserves
contributions to the double
quantum echo signal arising solely
from the dipolar interaction
between the pair of nitroxide spin labels. One varies the time intervals tp
and t2 such that the total time of the pulse sequence remains constant.
(Right) The maximum of the echo signal is recorded as a function of
t2 tp. (Top) A Fourier transform of this echo signal yields the Pake-type
dipolar spectrum in the frequency domain. The separation of the two
sharp peaks is directly related to the distance r between the labels,
r 1/3. (The echo signal and the dipolar spectrum shown here were
measured for the molecule on the left, for which r
16.2 ± 0.5 Å.) [View Larger Version of this Image (18K GIF file)]

Illustrative Applications
HF cw ESR and complex dynamics in proteins and membranes. The

http://www.sciencemag.org/cgi/content/full/291/5502/266 (4 of 10)2/27/2006 3:29:45 PM


Electron Spin Resonance in Studies of Membranes and Proteins -- Borbat et al. 291 (5502): 266 -- Science

snapshot feature of cw ESR encourages a multifrequency approach to the


study of the complex modes of motion of proteins, DNA, and other
polymers, enabling the decomposition of these modes according to their
different time scales (25). For example, in the case of proteins, the higher
frequency ESR spectra should "freeze out" the slow overall tumbling
motions, leaving only the faster internal modes of motion. Alternatively,
ESR performed at lower frequencies would be sensitive to the motions on a
slower time scale. This is in contrast to aqueous solution NMR, where all
modes are usually in the fast motional limit and contribute only to T1 and
T2.
The virtues of such an approach were demonstrated in a study at 9 and 250 GHz on spin-labeled
mutants of T4 lysozyme in aqueous solution (34). On the short time scale of the 250-GHz ESR
experiment, the overall tumbling was too slow to affect the spectrum; thus, a MOMD analysis
provided a satisfactory modeling of the overall tumbling, and HF-enhanced spectral resolution
reported on the internal dynamics (Fig. 3). The analysis of the 9-GHz spectra was more complex,
because the longer time scale did not freeze out the overall tumbling motion. The slowly relaxing
local structure (SRLS) model (25) simultaneously incorporated both the internal and overall motions
(Fig. 3, bottom). By fixing the internal motional parameters at the values obtained from the 250-GHz
Article Views data, fits to the 9-GHz line shapes generated by SRLS successfully yielded the rate for the global
dynamics. Thus, exploiting the time scale separation allowed more information to be extracted [see
(35-38)].
● Abstract
● Full Text
(HTML) Fig. 3. A multifrequency ESR study of
nitroxide spin-labeled T4 lysozyme in
Article Tools aqueous solution. The derivative spectra
Manage This Article are for the spin label on residue 44 and
are taken at 10°C (34). The relevant
● Submit an E-
molecular motions are shown
Letter schematically in the lower part of the
● Add to Personal figure. The protein tumbles slowly about
Folders its principal axes of overall diffusion; this
● Download to
is the SRLS. The motion of the spin label ADVERTISEMENT
Citation Manager moiety is restricted by its tether and its
surroundings to be within a cone, whose
● View PubMed
main axis makes an angle with the
Citation protein main axis. The spatial extent of the internal rotational diffusion
● Alert Me When
modes and their rates are distinguished from the protein overall tumbling
Article is Cited rates as described in the
● E-mail This Page text. [View Larger Version of this Image (21K GIF file)]

● Request

Permission to
Use This Article In a demonstration of the excellent orientational resolution at 250 GHz, macroscopically aligned
membranes containing a mixture of head groups [zwitterionic phosphatidylcholine (PC) and
negatively charged phosphatidylserine (PS)] were studied by using the cholesterol-like spin label,
ADVERTISEMENT
Similar Articles In: cholestane (CSL) (26). The direction of alignment of the CSL molecules could be clearly discerned
as the orientation of the membrane normal was changed with respect to the magnetic field. The CSL
in PC-rich membranes exhibited typical behavior; that is, its long axis was parallel to the bilayer
● Science normal, and its rotational diffusion rates R were slow (R ~ 106 to 107 s 1). However, the behavior in
Magazine PS-rich membranes is unusual and was interpreted in terms of a strong local biaxial environment.
● ISI Web of Although predicted from molecular dynamics simulations, this gives experimental evidence for local
biaxiality. By contrast, given the poorer orientational resolution at 9.1 GHz, it was not possible to
Science distinguish such detail from the 9.1-GHz experiment.
● PubMed
Two-dimensional Fourier transform (FT) ESR and dynamic structure of membranes. Two-

http://www.sciencemag.org/cgi/content/full/291/5502/266 (5 of 10)2/27/2006 3:29:45 PM


Electron Spin Resonance in Studies of Membranes and Proteins -- Borbat et al. 291 (5502): 266 -- Science

dimensional time domain ESR methods allow one to study both dynamics and ordering in membrane
Search Google
vesicles (30, 31). In general, 2D ELDOR spectra from membrane vesicles exhibit more dramatic
Scholar for:
variations due to changes in membrane properties than do spectra obtained by cw ESR. A simple
interpretation of these spectra is thus possible merely by pattern recognition. For example, 2D
● Articles by ELDOR contour plots are shown for the spin-labeled lipid, 16 PC, in pure lipid vesicles (Fig. 4A)
Borbat, P. P. and for a lipid-cholesterol (1:1) mixture (Fig. 4B). The former is in the standard liquid crystalline
● Articles by (LC) phase, whereas the latter is in a "liquid ordered" (LO) phase. This type of LO phase has been
found to play an important role in cellular signaling and in membrane trafficking (19, 39). Visually,
Freed, J. H. the spectra are qualitatively different, emphasizing that the LO phase exhibits substantially greater
● Articles citing ordering, but comparable rotational mobility, in comparison with the LC phase. Quantitative spectral
this article fits confirm that the increased microscopic ordering leads to increased ib affecting the spectra from
the LO phase. Furthermore, the restriction of the range of orientational motion, due to the
microscopic ordering in the LO phase, shows up as a much slower development of cross peaks.
Search PubMed for
Articles by:
Fig. 4. The sensitivity of pulsed 2D
● Borbat, P. P. ELDOR to the dynamic structure
● Freed, J. H. of lipid membranes. Contour plots
for the end-chain-labeled lipid,
Find Citing Articles in: 16 PC, dissolved in lipid vesicles
containing (A) only di-
● ISI Web of
palmitoylphosphatidylcholine
Science (63) (DPPC) in the LC (55°C) phase
and (B) an equimolar mixture of
● HighWire Press
DPPC and cholesterol in the LO
(55°C) phase. The 2D spectra from
My Science 16 PC dissolved (C) in pure DPPC
vesicles and (D) in an equimolar mixture of DPPC and GA, both at 75°C.
● My Folders The spectrum in (C) is characteristic of bulk lipid. In the spectrum in (D), it
● My Alerts is somewhat broadened, but a second, much broader component is apparent
● My Saved and is attributed to boundary lipid. All spectra are at
Searches 17.3 GHz. [View Larger Version of this Image (40K GIF file)]
● Sign In

Two-dimensional FT ESR and lipid-protein interactions. The merits of 2D FT ESR spectroscopy are
More Information well demonstrated in studies of the effect of the peptide gramicidin A (GA) on the dynamic structure
More in Collections of model membranes. Besides bulk lipids, boundary lipids that coat the peptide are present but
difficult to resolve in cw ESR. Two-dimensional ELDOR at 9.3 GHz (31) yielded dramatic changes
● Biochemistry in spectra as GA was added to vesicles, because of changes in the dynamic structure of bulk lipids.
However, newer 2D ELDOR results at 17.3 GHz, with improved time resolution, show the presence
of two components (40). The boundary lipid appears as a broader spectral component, characteristic
Related Jobs from To Advertise | Find Products
of a more slowly reorienting lipid with greater motional restriction (Fig. 4, C and D). Aligning the
ScienceCareers membranes enables resolution of boundary lipids, even in the presence of small amounts of GA (1 to
4 mole percent) (30, 40). ADVERTISEMENT
● Biochemistry
Multiple-quantum ESR and distance measurements. The determination of intra- and intermolecular Featured Jobs
● Chemistry
distances has become an important application of contemporary ESR spectroscopy. Areas of interest
● Molecular include the structure of protein complexes and the functional dynamics of proteins that are neither
Biology soluble nor crystallized (4, 5, 10, 15, 16).

Strong DQC signals for a variety of bilabeled nitroxide molecules in both disordered and oriented
solids have been achieved recently (32, 33) (cf. Fig. 2). For a benchmark rigid linear biradical,
piperidinyl-CO2-(phenyl)4-O2C-piperidinyl, a distance of 28.8 ± 0.5 Å between the two nitroxide
groups was obtained (versus 28.06 Å from molecular modeling). For a semiflexible bilabeled
synthetic peptide [MTSL-CPPPPC-MTSL in frozen solution at 82°C (MTSL, the [1-oxyl-2,2,5,5-
tetramethyl-pyrrole-3-methyl] methanethiosulfonate spin label; C, Cys; and P, Pro) relevant to
biophysical studies, a mean separation of 22.6 Å between the two nitroxide groups and a distribution
half-width of 2 Å (which smears out the Pake doublet) were found. Current sensitivities require ~300

http://www.sciencemag.org/cgi/content/full/291/5502/266 (6 of 10)2/27/2006 3:29:45 PM


Electron Spin Resonance in Studies of Membranes and Proteins -- Borbat et al. 291 (5502): 266 -- Science

pmol of bilabeled molecules, but it is estimated that, in the future, studies with samples that are
approximately an order of magnitude less will be feasible (33).

Future Prospects
Given the virtues of time domain ESR (such as direct determination of
relaxation rates and differentiation between hb and ib) compared to the
virtues of HF cw ESR (enhanced orientational selectivity and greater
sensitivity to nuances of the molecular dynamics), it would be desirable to
extend pulse techniques to higher frequencies (41). However, as the
frequency is increased, spin-labeled biomolecules in fluid media would
have shorter transverse decay times, because of enhanced hb and ib. Thus,
even shorter time resolution and greater spectral bandwidths would be
required, hence shorter and more intense radiation pulses (30). Progress in
developing a coherent pulsed high-power spectrometer at 95 GHz has
recently been reported (42).
The fast time scales of 2D FT ESR (30) are poised to address the functional dynamics (5) of
proteins. Nanosecond resolution is achievable in following the time evolution of a (small) region of
the spectrum, and 0.1- to 1-µs resolution is possible for the entire nitroxide spectrum. cw ESR
studies of the transient behavior of bilabeled bacteriorhodopsin following light activation were able
to resolve key distance changes occurring on the 1-ms to 1-s time scale (43, 44). FT ESR combined
with DQC should enable studies in the submicrosecond range. Similar applications for short time
dynamics would be for rapid mixing and stopped-flow experiments (45), limited only by the time
resolution of these methods.

REFERENCES

1. L. J. Berliner, Ed., Spin Labeling: Theory and Applications


(Academic Press, New York, 1976).
2. ___, J. Reuben, Eds., Spin Labeling: Theory and Applications, vol.
8 of Biological Magnetic Resonance (Plenum, New York, 1989).
3. L. J. Berliner, Ed., Spin Labeling: The Next Millennium, vol. 14 of
Biological Magnetic Resonance (Plenum, New York, 1998).
4. G. R. Eaton, S. S. Eaton, L. J. Berliner, Eds., Distance
Measurements in Biological Systems by EPR, vol. 19 of Biological
Magnetic Resonance (Kluwer, New York, 2000).
5. W. L. Hubbell, D. S. Cafiso, C. Altenbach, Nature Struct. Biol. 7,
735 (2000) [CrossRef] [ISI] [Medline] .
6. J. H. Freed, in (1), pp. 53-132.
7. D. J. Schneider, J. H. Freed, in (2), pp. 1-76.
8. A. H. Beth, B. H. Robinson, in (2), pp. 179-253.
9. D. E. Budil, S. Lee, S. Saxena, J. H. Freed, J. Magn. Reson. A120,
155 (1996) [CrossRef].
10. E. Hustedt, et al., Biophys. J. 74, 1861 (1997) .
11. O. H. Griffith, P. J. Dehlinger, S. P. Van, J. Membr. Biol. 15, 159
(1974) [ISI] [Medline] .
12. K. A. Earle, et al., Biophys. J. 66, 1213 (1994) [Abstract]
13. J. H. Freed, Annu. Rev. Biophys. Biomol. Struct. 23, 1 (1994)
[CrossRef] [ISI] [Medline] .
14. D. Marsh, in (2), pp. 255-303.
15. M. D. Rabenstein and Y.-K. Shin, Proc. Natl. Acad. Sci. U.S.A. 92,
823 (1995) .
16. H. J. Steinhoff, et al., Biophys. J. 73, 3287 (1997) [Abstract] .

http://www.sciencemag.org/cgi/content/full/291/5502/266 (7 of 10)2/27/2006 3:29:45 PM


Electron Spin Resonance in Studies of Membranes and Proteins -- Borbat et al. 291 (5502): 266 -- Science

17. A. D. Milov, A. G. Maryasov, Yu. D. Tsvetkov, Appl. Magn. Reson.


15, 107 (1998) [ISI].
18. M. Ge and J. H. Freed, Biophys. J. 76, 264
(1999) [Abstract/Free Full Text] .
19. M. Ge, et al., Biophys. J. 77, 925 (1999) [Abstract/Free Full Text] .
20. A. J. Hoff, Ed., Advanced EPR, Applications in Biology and
Biochemistry (Elsevier, New York, 1989).
21. L. Kevan, M. K. Bowman, Eds., Modern Pulsed and Continuous-
Wave Electron Spin Resonance (Wiley, New York, 1990).
22. D. E. Budil et al., in (20), pp. 307-340.
23. Y. S. Lebedev, in (21), pp. 365-404.
24. International Experts' Workshop: Present and Future of HFEPR
Instrumentation, S. S. Eaton, G. R. Eaton, Eds., Appl. Magn. Reson.
16, 159 (1999).
25. Z. Liang and J. H. Freed, J. Phys. Chem. B 103, 6384 (1999)
[CrossRef] [ISI] .
26. J. B. Barnes and J. H. Freed, Biophys. J. 75, 2532
(1998) [Abstract/Free Full Text] .
27. K. A. Earle, et al., J. Chem. Phys. 106, 9996 (1997) [CrossRef]
[ISI] .
28. J. Gorcester, G. L. Millhauser, J. H. Freed, in (20), pp. 177-242; in
(21), pp. 119-194.
29. M. K. Bowman, in (21), pp. 1-42.
30. P. P. Borbat, R. H. Crepeau, J. H. Freed, J. Magn. Reson. 127, 155
(1997) [CrossRef] [ISI] [Medline].
31. B. Patyal, R. H. Crepeau, J. H. Freed, Biophys. J. 73, 2201 (1997)
[Abstract] .
32. P. P. Borbat and J. H. Freed, Chem. Phys. Lett. 313, 145 (1999)
[CrossRef] [ISI].
33. P. P. Borbat, J. H. Freed, in (4), pp. 383-459.
34. J. Barnes, et al., Biophys. J. 76, 3298
(1999) [Abstract/Free Full Text] .
35. J. Pilar, et al., Macromolecules 33, 4438 (2000) [CrossRef] [ISI] .
36. Z. Liang, A. M. Bobst, R. S. Keyes, J. H. Freed, J. Phys. Chem. B
104, 5372 (2000) [CrossRef] [ISI] .
37. D. E. Budil, et al., Biophys. J. 78, 430
(2000) [Abstract/Free Full Text] .
38. M. Beunati, et al., J. Magn. Reson. 139, 281 (1999) [CrossRef] [ISI]
[Medline].
39. D. A. Brown and E. London, Annu. Rev. Cell Dev. Biol. 14, 111
(1998) [CrossRef] [ISI] [Medline].
40. P. P. Borbat, R. H. Crepeau, M. Ge, J. H. Freed, paper presented at
the 22nd International EPR Symposium, Denver, CO, 1 to 5 August
1999.
41. T. F. Prisner, Adv. Magn. Opt. Reson. 20, 245 (1997) .
42. K. A. Earle, C. Dunnam. P. P. Borbat and J. H. Freed, Am. Phys.
Soc. Abstr. 45, 542 (2000) .
43. T. E. Thorgeirsson, et al., J. Mol. Biol. 273, 951 (1997) [CrossRef]
[ISI] [Medline] .
44. R. Mollaaghababa, et al., Biochemistry 39, 1120 (2000) [CrossRef]

http://www.sciencemag.org/cgi/content/full/291/5502/266 (8 of 10)2/27/2006 3:29:45 PM


Electron Spin Resonance in Studies of Membranes and Proteins -- Borbat et al. 291 (5502): 266 -- Science

[ISI] [Medline] .
45. V. M. Grigoryants, A. V. Veselov, C. P. Scholes, Biophys. J. 78,
2702 (2000) [Abstract/Free Full Text] .

This article has been cited by other articles:


(Search Google Scholar for Other Citing Articles)

Spin-Labeled Gramicidin A: Channel Formation and Dissociation.


B. G. Dzikovski, P. P. Borbat, and J. H. Freed (2004).
Biophys. J. 87: 3504-3517 | Abstract » | Full Text » | PDF »

Dynamic Molecular Structure of DPPC-DLPC-Cholesterol Ternary Lipid


System by Spin-Label Electron Spin Resonance.
Y.-W. Chiang, Y. Shimoyama, G. W. Feigenson, and J. H. Freed
(2004).
Biophys. J. 87: 2483-2496 | Abstract » | Full Text » | PDF »

Optimal bundling of transmembrane helices using sparse distance


constraints.
K. Sale, J.-L. Faulon, G. A. Gray, J. S. Schoeniger, and M. M.
Young (2004).
Protein Sci. 13: 2613-2627 | Abstract » | Full Text » | PDF »

High Field/High Frequency Saturation Transfer Electron Paramagnetic


Resonance Spectroscopy: Increased Sensitivity to Very Slow Rotational
Motions.
E. J. Hustedt and A. H. Beth (2004).
Biophys. J. 86: 3940-3950 | Abstract » | Full Text » | PDF »

Methods for detection of reactive metabolites of oxygen and nitrogen: in


vitro and in vivo considerations.
M. M. Tarpey, D. A. Wink, and M. B. Grisham (2004).
Am J Physiol Regulatory Integrative Comp Physiol 286: R431-444
| Abstract » | Full Text » | PDF »

Caveosomes and endocytosis of lipid rafts.


B. Nichols (2003).
J. Cell Sci. 116: 4707-4714 | Abstract » | Full Text » | PDF »

In vivo electron spin resonance spectroscopy: what use is it to


gastroenterologists?.
N S Dhanjal, I J Cox, and S D Taylor-Robinson (2003).
Gut 52: 1236-1237 | Full Text »

Lipid-Gramicidin Interactions: Dynamic Structure of the Boundary Lipid


by 2D-ELDOR.
A. J. Costa-Filho, R. H. Crepeau, P. P. Borbat, M. Ge, and J. H.
Freed (2003).
Biophys. J. 84: 3364-3378 | Abstract » | Full Text » | PDF »

http://www.sciencemag.org/cgi/content/full/291/5502/266 (9 of 10)2/27/2006 3:29:45 PM


Electron Spin Resonance in Studies of Membranes and Proteins -- Borbat et al. 291 (5502): 266 -- Science

A 2D-ELDOR Study of the Liquid Ordered Phase in Multilamellar Vesicle


Membranes.
A. J. Costa-Filho, Y. Shimoyama, and J. H. Freed (2003).
Biophys. J. 84: 2619-2633 | Abstract » | Full Text » | PDF »

Methods of Detection of Vascular Reactive Species: Nitric Oxide,


Superoxide, Hydrogen Peroxide, and Peroxynitrite.
M. M. Tarpey and I. Fridovich (2001).
Circ. Res. 89: 224-236 | Abstract » | Full Text » | PDF »

Magazine | News | STKE | SAGE KE | Careers | Collections | Help | Site Map

Subscribe | Feedback | Privacy / Legal | About Us | Advertise With Us | Contact Us

© 2001 American Association for the Advancement of Science. All Rights Reserved.

You have reached the bottom of the page. Back to top

http://www.sciencemag.org/cgi/content/full/291/5502/266
View publication stats
(10 of 10)2/27/2006 3:29:45 PM

You might also like