Você está na página 1de 13

Structure

Article

The Signaling Pathway of Rhodopsin


Yifei Kong1 and Martin Karplus1,2,*
1
Department of Chemistry and Chemical Biology, Harvard University, Cambridge, MA 02138, USA
2
Laboratoire de Chimie Biophysique, ISIS, Universite Louis Pasteur, 67000 Strasbourg, France
*Correspondence: marci@tammy.harvard.edu
DOI 10.1016/j.str.2007.04.002

SUMMARY more than 50 years (Hubbard and Wald, 1952; Wald,


1968), the mechanism by which this occurs is still obscure.
The signal-transduction mechanism of rho- Rhodopsin consists of the protein opsin, which is com-
dopsin was studied by molecular dynamics posed of seven transmembrane helices (helices IVII),
(MD) simulations of the high-resolution, inactive a short additional helix (helix VIII) approximately parallel
structure in an explicit membrane environment. to the membrane, and a set of connecting loops on the
The simulations were employed to calculate two sides of the membrane, plus the 11-cis retinylidene
chromophore bound covalently to Lys296 through a pro-
equal-time correlations of the fluctuating inter-
tonated Schiff-base linkage (Okada et al., 2004; Palczew-
action energy of residue pairs. The resulting
ski et al., 2000; Teller et al., 2001). The chromophore is
interaction-correlation matrix was used to de- buried in the middle of the transmembrane helical bundle,
termine a network that couples retinal to the far from the rhodopsin cytoplasmic surface, which is the
cytoplasmic interface, where transducin binds. binding interface between rhodopsin and the G protein.
Two highly conserved motifs, D(E)RY and The isomerization of retinal takes place on the 200 fs
NPxxY, were found to have strong interaction timescale (Schenkl et al., 2005), while metarhodopsin II
correlation with retinal. MD simulations with re- (meta II), the active species, is formed on the millisecond
straints on each transmembrane helix indicated timescale through a complex cycle comprised of a series
that the major signal-transduction pathway in- of intermediates including bathorhodopsin, lumirhodop-
volves the interdigitating side chains of helices sin, and metarhodopsin I (meta I), which are formed after
retinal isomerization on a nanosecond, microsecond,
VI and VII. The functional roles of specific resi-
and millisecond timescale, respectively (Menon et al.,
dues were elucidated by the calculated effect
2001; Okada et al., 2001). They have been identified pri-
of retinal isomerization from 11-cis to all-trans marily through spectral shifts of the primary retinal absorp-
on the residue-residue interaction pattern. It is tion band (Okada et al., 2001). The overall structures of
suggested that Glu134 may act as a signal bathorhodopsin (ns) and lumirhodopsin (ms) appear to be
amplifier and that Asp83 may introduce a very similar to the inactive protein, as shown by recent
threshold to prevent background noise from X-ray crystallography (Nakamichi and Okada, 2006a,
activating rhodopsin. 2006b). Meta I keeps essentially the same helical posi-
tions, an orientation illustrated by cryo-electron micros-
copy (Ruprecht et al., 2004). Only when the system
reaches meta II, which is required for normal activation
INTRODUCTION of the G protein, transducin, do larger structural changes
appear (Nakamichi and Okada, 2006b). The states have
Rhodopsin is a membrane protein that detects light in the also been studied by a variety of other methods, including
rod photoreceptor cell. Like other G protein-coupled re- Cys scanning mutagenesis and site-directed spin labeling
ceptors (GPCRs), rhodopsin exists in equilibrium between (for a review, see Hubbell et al. [2003]). There is indirect ev-
its activated and inactivated forms in vivo. This equilibrium idence, based primarily on spin-label mobility changes,
is controlled by the isomerization of retinal, the cofactor that there are significant displacements of certain helices
covalently bound to rhodopsin. In the dark state, retinal (particularly helix VI) at their cytoplasmic ends. The pro-
is in its 11-cis form and stabilizes rhodopsin in its inactive posed outward motion of helix VI was confirmed by dis-
conformation. In the presence of light, retinal is photoiso- tance change estimates from spin-label interactions and
merized to the all-trans form, which activates rhodopsin. disulfide crosslinking (Fritze et al., 2003). Although it has
Activated rhodopsin catalyzes the replacement of GDP been suggested that helix VI is particularly flexible due
by GTP on the a subunit of a heterotrimetric G protein, to looser packing, there is no evidence for this from the
transducin, which is bound to rhodopsin. This, in turn, trig- X-ray B factors of the inactive structure. Some of the
gers the response of the rod cells in the retina (Filipek changes observed in photoactivation are also observed
et al., 2003; Gether and Kobilka, 1998; Meng and Bourne, by corresponding techniques in constitutively active mu-
2001). Although the activation of rhodopsin by retinal tants of rhodopsin (Kim et al., 1997). How the outward
isomerization from 11-cis to all-trans has been known for motion of helix VI and the smaller motions of other helices

Structure 15, 611623, May 2007 2007 Elsevier Ltd All rights reserved 611
Structure
Signaling Pathway of Rhodopsin

lead to activation, presumably through the cytoplasmic sequences (Lockless and Ranganathan, 1999; Suel et al.,
loops of rhodopsin in contact with the a subunit of trans- 2003). However, the simulations permit us to determine
ducin, is not clear. An essential question is how the isom- the energetic origins of the coupling, which are not avail-
erization signal propagates efficiently from the middle of able from the conservation analysis.
rhodopsin, where retinal is located, to the cytoplasmic Comparisons of amino acid sequences for GPCRs have
surface. identified two highly conserved sequence motifs that have
After high-resolution structures of inactive rhodopsin been shown to play critical roles (Stenkamp et al., 2005).
became available (Okada et al., 2004; Palczewski et al., One of these is the D(E)RY motif (residues 134136) at
2000; Teller et al., 2001), a series of molecular-simulation the cytoplasmic terminus of helix III, and the other one is
and normal-mode studies were performed (Crozier et al., the NPxxY motif (residues 302306) in helix VII. The pres-
2003, 2007; Faraldo-Gomez et al., 2004; Grossfield ent simulations show how these two motifs are coupled to
et al., 2007; Huber et al., 2004; Isin et al., 2006; Lemaitre retinal and its isomerization. Also, we find that, for certain
et al., 2005; Rohrig et al., 2002; Saam et al., 2002). The charged residues, there is a strong anisotropic effect of
results are described briefly in the Discussion. retinal isomerization on the interaction network. Finally,
It is generally presumed that an allosteric-type mecha- an analysis of the role of the helices in transmitting the
nism is involved, in which there is a pre-existing equilib- signal shows that, interestingly, it is the fluctuations of
rium between the inactive and active forms (Changeux the side chains rather than main chains that play the es-
and Edelstein, 2005), in accord with the model proposed sential role.
by Monod, Wyman, and Changeux (MWC) in their land-
mark paper published 40 years ago (Brunori et al., 2005).
Although the MWC model was proposed for multisubunit RESULTS
proteins, in a monomer like the GPCR rhodopsin, only ter-
tiary structural changes, by definition, are involved in acti- Overall Simulation Characteristics
vation; thus, the flexibility and dynamics of the protein Eight independent MD simulations of rhodopsin were
must play the essential roles. This suggests that inactive performed with a total simulation length of 22 ns (see
rhodopsin (with 11-cis retinal) and meta II (also with Experimental Procedures). For the trajectories, the root-
11-cis retinal) are in equilibrium. The equilibrium constant mean-square deviations (rmsds) of Ca atoms (Figure S1)
is very far on the side of the former to avoid false signals, and the residue-based root-mean-square fluctuation
although constitutively active rhodopsins exist (Fritze (rmsf) (Figure S2) are comparable to those of previous
et al., 2003; Kim et al., 1997; Weitz and Nathans, 1993). studies (Lemaitre et al., 2005) and experimental B factors
The equilibrium constant shifts to meta II when isomeriza- (Teller et al., 2001). This indicates that analysis of the MD
tion to the all-trans form has taken place, though there are simulations can provide a meaningful description of the
no details on the relative concentrations. With the isomer- dynamics of inactive rhodopsin, at least on the nano-
ization occurring in 200 fs, while the change to the active second timescale. For details, see Supplemental Data 1
conformation requires a time course on the order of milli- available with this article online. While the current simula-
seconds, it is very likely that the transition is not all or tions were in progress, a higher-resolution structure of
none, but involves a propagation of more localized struc- rhodopsin (Okada et al., 2004) was released. Although
tural alterations. This can be termed tertiary allosteric we completed our analysis with the older structure (PDB
coupling. The investigation into how retinal isomerization ID 1HZX), in part to permit comparison with the work of
provides a signal that propagates to the cytoplasmic end others (Crozier et al., 2003, 2007; Huber et al., 2004;
of the protein is the objective of this paper. To determine Saam et al., 2002), we have performed simulations (four
the source of the tertiary allosteric coupling, we use equal-length trajectories, totaling 16 ns) on that structure
a method based on equal-time correlations of the interac- with same setup. The results indicate that the interaction
tion-energy fluctuations (i.e., the correlation between two correlation pattern is not sensitive to the choice of
sets of interaction energies between pairs of amino acids initial crystal structure (PDB ID 1HZX or 1U19) or to the
at a given time, averaged over the simulation) between all presence of internal water molecules. For details, see
sets of residue pairs in equilibrium molecular dynamics Supplemental Data 11.
simulations of rhodopsin in the inactive (11-cis retinal)
structure and the changes in the interaction energies Residue-Based Interaction-Energy Correlation
that occur on a short timescale (i.e., 10 ns) after the isom- The signal transduction provided by rhodopsin from the
erization of retinal to the all-trans form. The fluctuations of chromophore to the cytoplasmic regions requires that
the interaction energies are obtained from ensembles of there be long-range communication within the protein.
rhodopsin conformers generated by molecular dynamics One approach is to investigate the correlation of the dy-
(MD) simulation of the inactive rhodopsin structure in namic behavior between different structural units at equi-
a membrane. By use of the interaction-energy correlation librium in the inactive, high-resolution structure (2.6 A).
matrix, we are able to identify a network that extends from Alternatively, it is possible to examine the changes in the
the retinal-binding pocket to the cytoplasmic surface. The correlation induced by chromophore isomerization. We
rationale for the present study is similar to that used in the begin with the former and discuss the later in the following
conservation correlation analysis based on multiple GPCR subsections. To initiate the analysis, we calculated the

612 Structure 15, 611623, May 2007 2007 Elsevier Ltd All rights reserved
Structure
Signaling Pathway of Rhodopsin

Figure 1. Interaction Correlation


(A) Four pairs of residue-residue interactions are shown in an equilibrated rhodopsin structure (gray). The Ca atoms of each residue are rendered as
spheres. The interactions between residues 139248, 135249, 94296, and 135139 are illustrated by double-headed arrows in green, blue, red, and
orange, respectively. The interaction-energy profiles between pairs of interactions during the 22 ns simulation are shown in (B)(D), with the same
colors used in this panel.
(B) The interaction energy between residues 139248 and 135249 shows a strong positive correlation (0.673).
(C) The interaction energy between residues 135249 and 94296 shows a strong negative correlation (0.626).
(D) The interaction energy between residues 135139 and 135249 shows little correlation (0.038).

interaction energies between all residue pairs as a function To illustrate the interaction correlation analysis, we
of time in the 22 ns MD runs and then used the results to show the residue pair interaction-energy profile and inter-
determine equal-time correlations of the interaction ener- action-energy correlation between four residue pairs
gies of any two residue pairs throughout the molecule. (Arg135/Glu249, Val139/Lys245, Thr94/Lys249, and Arg135/
Because of the helical structure of rhodopsin and experi- Val13) (Figure 1). A control simulation on the same set of
mental analyses that have focused on certain helices, two residue pairs in the absence of the protein showed
we then determined which of the secondary structural that the long-range interaction-energy correlation be-
elements play an important role in signal transduction by tween 94/296 and 135/249 is essentially zero. For details,
restraining them. see Supplemental Data 2.

Structure 15, 611623, May 2007 2007 Elsevier Ltd All rights reserved 613
Structure
Signaling Pathway of Rhodopsin

Figure 2. Comparison of Interaction Correlation with Displacement Correlation from Quasi-Harmonic Analysis
(A) Residue correlation matrix of rhodopsin (348 3 348).
(B) Residue-based displacement correlation of rhodopsin. Both quantities are normalized; the former is normalized between 0 and 1, and the later is
normalized between 1 and +1. In both matrices, which are symmetric, the column/row index is the sequence index of rhodopsin from 1 to 348, and
the positions of the seven transmembrane helices are shown at the bottom. The arrow indicates the position (residue 296) to which retinal is bound.

Interaction-Energy Correlation Matrix and Residue For comparison, the residue-residue displacement
Correlation Matrix of Rhodopsin crosscorrelation matrix (Figure 2B) based on a quasi-
A total of 1026 residue-residue interactions with average harmonic analysis (Levy et al., 1984) of the 22 ns MD sim-
interaction energies larger than 1 kcal/mol in magnitude ulation trajectories (see Experimental Procedures) was
were identified among the 348 residues of rhodopsin constructed. The usual correlations (i.e., close residues
(see Experimental Procedures). The 1026 3 1026 interac- and residues within a helix have large correlations) are
tion-energy correlation matrix was constructed from these seen, but there is no evidence for long-range correlations
data. There are a total of 525,825 non-zero data points in (see Supplemental Data 4).
this matrix with an average value of 0.104 and a standard
deviation of 0.102. The distribution of the absolute values Correlation of Retinal Interactions:
of the correlations of the pairwise interaction energies is Signal-Transduction Pathway
shown in Figure S3A. To eliminate background noise, To obtain information about the signal-transduction
a condensed correlation matrix (993 3 993) was built by mechanism, we map the correlated interactions between
introducing a correlation cutoff of 0.206 (see Supplemen- the origin of the signal (Lys296, including retinal) and
tal Data 3). The resulting interaction matrix was projected the rest of rhodopsin on the structure. Figure 3A shows
to obtain a residue-based correlation matrix (see Experi- a color-coded scheme of the magnitude of the interaction
mental Procedures), which gives the correlations between correlation. Many residues (90, 94 113, 122, 181, 190, 208,
residues. In the residue correlation matrix, each column 297, and 298) in the retinal-binding pocket (within 5 A)
and row represents a specific residue in the protein; the show relatively strong coupling with retinal, due to local
resulting matrix is shown in Figure 2A. The residue corre- correlations of the type illustrated in Figures 1A and 1B;
lation matrix shows the coupling between any two resi- in particular, Glu113 forms a salt bridge to the protonated
dues in the system, independent of how it arises (i.e., Schiff base. All of the residues in the retinal-binding
whether it is direct or is transmitted through other residue pocket that are highlighted in Figure 3A are also perturbed
interactions). In the process of intramolecule signal trans- by retinal isomerization (see below); also see Supplemen-
duction, the initial perturbation, such as retinal isomeriza- tal Data 5. On the extracellular side, the loop (190201)
tion in rhodopsin, is likely to first change the residue-inter- between the b sheet and helix V shows relatively strong
action pattern around the retinal-binding pocket and correlation with retinal, although there is no direct inter-
eventually reach the rhodopsin cytoplasmic side. There- action. Residue 181 is also correlated with residue 296,
fore, the residue correlation matrix, which represents the which is consistent with experimental observations (Patel
interaction coupling between any residue pairs in the et al., 2004; Yan et al., 2003). On the cytoplasmic side, we
structure, is the essential quantity for analysis. identified two groups of residues with strong coupling to

614 Structure 15, 611623, May 2007 2007 Elsevier Ltd All rights reserved
Structure
Signaling Pathway of Rhodopsin

Lys296 (see Figure 3A). The first group, which is located its tight packing, and the torsion engendered by the isom-
around the D(E)RY motif, includes the cytoplasmic end of erization was distributed to several C-C single bonds of
helix III (134, 135, and 139), Arg147 on the second intracel- retinal. The main effect is that the retinal molecule be-
lular loop, and the cytoplasmic end of helix VI (247252). comes more extended along its longitudinal axis due to
The second group, which is located around the NPxxY the double-bond cis/trans isomerization (Nakamichi and
motif, includes Asp83 on helix II, the C-terminal half of he- Okada, 2006b). Rotation of the C19 methyl group was ob-
lix VII (301309), and the N-terminal portion of helix VIII served due to the rotation about the C8-C9 single bond.
(311314). Experimental studies have suggested that res- This appears to be one source of the signal from retinal
idues 139, 251 (Franke et al., 1990), and 310313 (Bae to the protein that arises from significant alterations of
et al., 1997, 1999; Kostenis et al., 1997; Onrust et al., the interaction partners. This result is in accord with the
1997) are potential binding sites for transducin. We found experimental observation that 9-demethyl-retinal has
a continuous coupling pathway from retinal to the NPxxY only weak activation properties and with its photo-isomer-
motif through helix VII; i.e., residues 83, 297, 301, 302, ization results in a lower fraction of the rhodopsin in the
303, 307, and 308 (Figure 3A). The pathway to the meta II state, relative to meta I (Fritze et al., 2003).
D(E)RY motif from retinal is more indirect; residues on he- The interaction-energy changes between Lys296 and
lices III and VI between this region and retinal do not show its neighboring residues, relative to those from the inactive
an obvious signal (Figure 3A); however, see below. structure obtained by averaging the eight simulations, are
Given the helix-bundle structure of rhodopsin, it is im- shown in Figure 4B and Table S1. Weaker interactions
portant to investigate directly the possible role of the heli- were found between residue 296 (retinal) and residues
ces in the transmission of the signal from retinal to the cy- 86, 90, 94, 113, 114, 116, 117, 124, 188, 207, 208, 212,
toplasmic loops. Seven independent MD simulations were 265, 268, and 294, while stronger interactions resulted
performed, in which each transmembrane helix was for residues 43, 91, 93, 95, 118, 122, 167, 178, 180, 186,
restrained in the region between retinal and cytoplasmic 187, 189, 211, 264, 289, 291, 292, 293, and 298. Of these,
interface; for details, see Supplemental Data 6. Based on only 14 interactions (86, 91, 93, 113, 114, 116, 178, 208,
the restrained dynamics trajectories, the residue coupling 212, 265, 268, 292, 293, and 294296 [retinal]) are re-
strength to residue 296 (retinal) was constructed (Figures garded as significant (the underlined ones become
3B3H) analogously to the unrestrained analysis (Fig- weaker, and the other ones stronger) (see Experimental
ure 3A). Restraints on helix VI (254259) (Figure 3G) or helix Procedures).
VII (303308) (Figure 3H) were found to significantly
weaken the coupling between retinal and the rhodopsin Effects of Retinal Isomerization on the Interaction
cytoplasmic interface, especially the D(E)RY motif, indicat- Network: Importance of Charged Residues
ing that they are essential for the signal transduction from To analyze the changes in the interaction network due to
retinal to the rhodopsin cytoplasmic interface. Results of retinal isomerization, it is useful to focus on specific inter-
restraining other helices, which have smaller or no effects, actions rather than the residues; see Equation 5 in Exper-
are given in Supplemental Data 7. Additional simulations imental Procedures. Figure 5A shows how the interaction
(Supplemental Data 8) showed that the long-range signal network is altered by retinal isomerization; a correlation
transduction is achieved by the cooperative motions of threshold of 0.204 was used. Most pair interactions close
the side chains of the two helices; such interdigitated to retinal are significantly altered, in accord with the results
side chain-mediated coupling of helices has been ob- shown in Figure 4B, as expected. More strikingly, the net-
served in hemoglobin simulations (Gelin et al., 1983). work of changes spreads from retinal to more distant re-
gions, including both the extracellular and cytoplasmic
Local Perturbation Introduced sides of rhodopsin. Densely clustered groups of changes
by Retinal Isomerization in the interactions are observed around both the D(E)RY
To complement the interaction coupling analysis based on motif and the NPxxY motif (Cai et al., 2001; Fritze et al.,
the equilibrium fluctuations of rhodopsin in its inactive 2003; Itoh et al., 2001). To focus on the most important
structure, we determined how retinal isomerization per- residue interactions coupled to retinal isomerization, we
turbed the interactions described above. Eight indepen- show in Figure 5B only the perturbations of interactions
dent 1 ns and one 9 ns MD simulation of all-trans retinal with an average interaction energy greater than 5 kcal/
were performed (see Experimental Procedures). The mol in absolute value. A total of 107 interactions appear
results correspond to a stage in the rhodopsin cycle in Figure 5B. Not surprisingly, the charge residues
between bathorhodopsin, which appears in about 200 fs Asp83, Glu122, Glu134, Arg135, Arg147, Lys248, and
(Schenkl et al., 2005), as mimicked in the simulation Arg314 (see Figure 5B) are important (see also below). In-
from 11-cis to all-trans retinal, and lumirhodopsin; for de- terestingly, the changes in the interaction coupling
tails, see Supplemental Data 9. The overall structure does network are anisotropic. Specifically, the weakened or in-
not have any obvious deviations (rmsd < 3.4 A) from the cis tensified interactions tend to be clustered in certain direc-
retinal crystal structure even after a 9 ns simulation with tions, instead of being randomly distributed. This obser-
all-trans retinal; similar results were found in previous cal- vation suggests a mechanism for conduction of a signal.
culations (Lemaitre et al., 2005; Saam et al., 2002). As for Compared with hydrophobic interactions, salt bridges
retinal, the b-ionone ring did not move significantly due to and hydrogen bonds are more specific in direction, which

Structure 15, 611623, May 2007 2007 Elsevier Ltd All rights reserved 615
Structure
Signaling Pathway of Rhodopsin

616 Structure 15, 611623, May 2007 2007 Elsevier Ltd All rights reserved
Structure
Signaling Pathway of Rhodopsin

Figure 4. Local Effect of Retinal Isomerization


For more details, see text.
(A) Diagram of retinal isomerization from 11-cis to all-trans.
(B) The pattern of interaction-energy deviation in the neighborhood of residue 296 after retinal isomerization; all residues shown have an average
interaction energy with residue 296 (including retinal) larger than 1 kcal/mol in magnitude. This view is from the extracellular side to the cytoplasmic
side. Residue 296 (including retinal) is colored gray and is rendered as a space-filling model. The residues having weaker interaction after retinal isom-
erization (86, 90, 94, 113, 114, 117, 124, 188, 207, 208, 212, 265, 268, and 294) are colored black, and residues having stronger interaction upon retinal
isomerization (43, 91, 93, 95, 118, 122, 167, 178, 180, 186, 187, 189, 211, 264, 289, 291, 292, 293, and 298) are colored white. A residue is rendered as
a space-filling model (86, 91, 93, 113, 114, 116, 178, 208, 212, 265, 268, 292, 293, and 294) if its interaction energy with residue 296 (including retinal)
has a large difference after retinal isomerization (the deviation is larger than 0.5 kcal/mol plus one-fourth of its standard deviation in the ground-state
simulation); otherwise, it is shown as a bond. The numerical values of the interactions with significant changes are listed in Table S1. (Residue 86,
which has significant changes, is not shown, because it is not visible in the chosen orientation of rhodopsin.)

could make them effective as molecular switches that structure of bathorhodopsin (Nakamichi and Okada,
respond to the distant (allosteric) retinal perturbation, in 2006b), response of charged residue interactions to retinal
accord with Figure 5. Interestingly, in a very new crystal isomerization was also observed.

Figure 3. Correlation Maps of Each Residue with Residue 296, Including the Bound Retinal, Projected on the Inactive Rhodopsin
Structure in a Ca Ribbon Representation for the Helices and Threads for the Loops
(AL) Helices, certain residues, and the extent of the membrane are labeled. Residue 296 (including retinal) is rendered as a space-filling model in
orange. All figures use the same color scale with no correlation (blue) to the strongest coupling (red) (same as Figure 2A). The magnitude of coupling
is determined by line/row 296 of the residue correlation matrix. The Ca atoms of the conservative NPxxY and D(E)RY motifs are shown by purple and
pink spheres, respectively. (A) Results for the unrestrained system averaged over 22 ns. (BE) Results with various restraints applied to the system
during the dynamics simulation (also, see text); the restrained residues are rendered in purple, and restrained Ca atoms are shown as red spheres.
Restraints are as follows: (B) restrained helix I (residues 5660), (C) helix II (residues 7082), (D) helix III (residues 124132), (E) helix IV (residues 158
162), (F) helix V (residues 214226), (G) helix VI (residues 254259), (H) helix VII (residues 303308), (I) Ca atoms of helix VI (residues 254259), and
(J) Ca atoms of helix VII (residues 303308). (K) The correlation map based on simulation with the Asp83 side chain charge scaled down to one-fourth
of its original value. Asp83 is shown in red bonds. (L) Correlation map based on simulation with protonated Glu134.

Structure 15, 611623, May 2007 2007 Elsevier Ltd All rights reserved 617
Structure
Signaling Pathway of Rhodopsin

Figure 5. The Impact of Retinal Isomerization on the Residue-Residue Interactions


(A and B) The rhodopsin backbone is rendered as a gray thread, residue 296 (including retinal) is shown as orange bonds, and Ca atoms of involved
residues are shown as spheres. Positively charged residues (Arg, Lys) are in red, negatively charged residues (Asp, Glu) are in blue, and neutral res-
idues are shown by gray spheres. Helices III, VI, and VII are pink, green, and blue, respectively. The second intracellular loop (between helices III and
IV) is shown in purple, and the third intracellular loop (between helices V and VI) is shown in cyan. (A) Interactions with a cumulative correlation to retinal
isomerization larger than 0.204 and an average interaction energy with an absolute value larger than 1.0 kcal/mol are shown as dashed lines. For
clarity, no interactions involving residue 296, which mostly have strong correlation, are shown. The red and green dashed lines indicate interactions
that become stronger or weaker, respectively, in response to retinal isomerization. Strong signals are observed between charged residues at the rho-
dopsin cytoplasmic interface, especially at the D(E)RY motif. (B) An enlarged stereo view in which interactions with an average interaction energy less
than 5.0 kcal/mol in magnitude are omitted. Important charged residues identified by both this study and previous works are labeled. The interaction
deviation of the D(E)RY motif clearly shows the effect of change of electrostatic energy involving Glu134; its interaction with Arg147 becomes weaker,
and its interaction with Glu247 becomes stronger (see text).

Of the charged residues cited above, only Asp83 in i.e., for a protein in a membrane, the surrounding solvent
helix II and Glu122 in helix III are inside the membrane (membrane and more distant water) has a smaller
between retinal and the cytoplasmic interface (Figure 3A). shielding effect than for the corresponding interactions
Since they are in the interior of the portion of the in aqueous solution. Mutagenesis studies have shown
protein buried in the membrane, they are ex- that after retinal isomerization, D83N introduce a shift
pected to play an important role in signal transduction; in the meta I 4 meta II equilibrium toward meta II

618 Structure 15, 611623, May 2007 2007 Elsevier Ltd All rights reserved
Structure
Signaling Pathway of Rhodopsin

(DeCaluwe et al., 1995), which leads to rhodopsin lar loop between helices III and IV, including residue 147,
activation. which interacts with Glu134 (see above). This suggests
Based on the rhodopsin structure (Palczewski et al., the possibility that Glu134 is a molecular switch, which
2000; Teller et al., 2001) and the simulations, Asp83 has is turned on by retinal cis/trans isomerization and
very weak interaction with the lipid molecules (Figure S7A), amplifies the signal.
and their desolvation energy is compensated by favor- To further investigate the potential role that Glu134
able interactions with their neighboring residues (Fig- might play in long-range coupling, we did a 4 ns simulation
ure S7B); i.e., residue 55, 86, 87, 297, 298, 299, 302, of rhodopsin with protonated Glu134 (see Experimental
and 303 for Asp83 (Figure 5B). The changes in its interac- Procedures). The overall structural fluctuations (maximum
tions (Figure 5B) suggest that anisotropic forces act on rmsd < 2.4 A) are very similar to those with Glu134 unpro-
Asp83. In response to retinal isomerization, the interac- tonated (see Figure S2). The retinal residue-based cou-
tion-pattern changes for Asp83 result in stronger inter- pling map with Glu134 protonated (Figure 3L) shows that
actions with residues 86, 87, 302, and 303 and weaker it has a weaker interaction correlation with retinal than in
interactions with residues 297, 298, and 299, indicative the unprotonated state (Figure 3A). Interestingly, signifi-
of a displacement toward the cytoplasmic side, relative cant decreases of coupling were also observed between
to helix VII. The mutation D83N (DeCaluwe et al., 1995) retinal and the NPxxY motif (Figure 3L). Together with
reduces the polarity of the side chain and weakens the re- the interaction correlation analysis on rhodopsin with re-
straints on it, lowering the barrier of the inactive state / straints on the NPxxY motif (Figure 3H), this result sug-
meta II transition and perhaps favoring the meta II state gests that perturbing either the D(E)RY motif (Figure 3L)
thermodynamically. A possible function, following from or the NPxxY motif (Figure 3H) would significantly affect
the findings described above, is that Asp83 acts as the coupling between the other motif and retinal. This
threshold controller to help in reducing the (thermal) sig- suggests that the NPxxY motif and the D(E)RY motif are
nal to a low level; it is known experimentally that the hu- intrinsically coupled in their dynamics, in accord with their
man eye can detect a single photon (Baylor et al., 1979). suggested role in signal transduction.
Constrained by the Asp83 interactions, thermal fluctuation
would be less likely to overcome the energetic barriers DISCUSSION
and produce a signal on the cytoplasmic side. To evaluate
this hypothesis, a 4 ns MD simulation was performed with The primary function of signal-transduction proteins, such
the side chain charge of D83 scaled by 0.25. Retinal- as rhodopsin, and more generally the class of GPCRs, is to
related correlation factors were built with the same propagate a signal from an upstream perturbation toward
protocol used in Figure 3A. As shown in Figure 3K, in the the downstream partner in the signal-transduction path-
simulation with the charge-scaled D83, residues on the way. When the perturbation and target sites are widely
cytoplasmic side of rhodopsin have a stronger correlation separated, this process is thought to generally involve
with retinal than that calculated from wild-type simulation an allosteric transition. For monomeric proteins, like rho-
(Figure 3A), especially those on the cytoplasmic end of dopsin, the coupling of conformational changes between
helix VI and the C-terminal loop. remote regions involves tertiary structural changes and
Another charged residue, Glu134 in the D(E)RY motif, is can be termed tertiary allosteric coupling. Instead of
highly conserved in the GPCR family (Fritze et al., 2003), spreading out isotropically and decaying with the dis-
and its protonation has been shown to be necessary for tance, the signal triggers specific responses and may
rhodopsin to reach the meta II state (Arnis and Hofmann, even be amplified at the target region. This feature is a con-
1995; Fahmy et al., 2000; Kuwata et al., 2001); rhodopsin sequence of the inhomogeneous nature of protein struc-
with an E134Q mutation assumes a partially active confor- tures that were designed for signal transmission by their
mation at the cytoplasmic surface without photoactivation evolutionary development. Given the sensitivity of the ver-
(Arnis and Hofmann, 1995; Scheer et al., 1996). In Fig- tebrate photocycle, one expects that rhodopsin is a mole-
ure 5B, residue-residue interactions around the D(E)RY cule optimized for signal transduction upon activation by
motif showed highly ordered responses to retinal isomer- light via retinal isomerization and, at the same time, is
ization. In particular, the D(E)RY motif moves such that the inhibited from transmitting a signal via thermal motion in
interaction between Glu134 and Arg147 becomes weaker the dark-adapted state. Thus, rhodopsin is an ideal
and that between Glu134 and Glu247 becomes stronger. system for investigating tertiary allosteric coupling at an
As both of these interactions have relatively high energies atomic level of detail.
(absolute value larger than 5 kcal/mol) in the 22 ns simula- The signal-transduction pathway of GPCRs has been
tions, this change is expected to increase the pKa of studied by an insightful statistical analysis of amino acid
Glu134, which would be a factor leading to proton uptake conservation in multiple sequence alignments (MSA)
by Glu134. Since the D(E)RY motif is surrounded by a large (Lockless and Ranganathan, 1999; Suel et al., 2003). The
number of charged residues, Glu134 protonation could, in related conservation behavior of different residues was in-
turn, have a strong effect on the local interaction network terpreted as describing a physically connected network of
and cause significant conformational changes. Spin-label coupling between retinal and the cytoplasmic interface.
experiments (Farahbakhsh et al., 1995) reported activa- As in the present analysis, the D(E)RY and NPxxY motifs
tion-induced mobility change in residues on the intracellu- were found to be important in the statistical approach.

Structure 15, 611623, May 2007 2007 Elsevier Ltd All rights reserved 619
Structure
Signaling Pathway of Rhodopsin

However, the coupling network inferred from the statisti- and their coupling via the side chains, are directly involved
cal analysis (Suel et al., 2003) is less well defined than in signal transmission. Therefore, crosslinking experi-
that from the MD results; i.e., the former appears to involve ments, e.g., restraining side chains of residues on helices
a large part of the molecule. The interaction-energy corre- VI and VII, would provide additional information about the
lation analysis is computationally more expensive than signal-transduction pathway. Moreover, with the refine-
that based on the MSA, but it deals directly with the ener- ment of single-molecule spectroscopy and its combina-
getics and its role in generating the signaling pathway. For tion with FRET measurements, fluorescence labels could
further discussion, see Supplemental Data 10. be attached to residues on helices VI and VII to monitor
A number of MD simulations investigating rhodopsin are how the rhodopsin conformation evolves after retinal
in the literature; some of them appeared while this paper isomerization. Also, the two charged residues, Asp83
was under review. They focused on the inactive state (Cro- and Glu134, are suggested to play a special role. Since
zier et al., 2003; Huber et al., 2004), on the convergence these two residues function through their charged side
of membrane protein simulations (Faraldo-Gomez et al., chains, any mutations, which perturb their local electro-
2004; Grossfield et al., 2007), on the process of retinal static environment, could regulate rhodopsin activity.
isomerization (Lemaitre et al., 2005), or on the structural More generally, an experimental alanine scanning analysis
changes a short time after isomerization (Crozier et al., of rhodopsin would be of great interest.
2007; Rohrig et al., 2002; Saam et al., 2002). Only the
last three are, in some sense, related to the results of the EXPERIMENTAL PROCEDURES
present work. The three studies characterized the local
perturbations of the retinal-binding site brought about by Molecular Dynamics Simulation
retinal isomerization and the structural changes of the For the analysis of the interaction correlation (see below), a large num-
ber of coordinate sets are necessary to obtain statistically significant
transmembrane helices within 10 ns or 150 ns. Also,
results. A series of MD simulations was used to generate an ensemble
a model of the meta II state was proposed based on an of rhodopsin structures for the analysis. Recent studies indicated that
elastic network normal mode (Isin et al., 2006). Although satisfactory conformational sampling of helical membrane proteins
retinal isomerization is a fast process (200 fs) (Schenkl can be achieved from MD simulations (Faraldo-Gomez et al., 2004);
et al., 2005; Zgrablic et al., 2005), the consequential con- the use of multiple simulations to achieve better convergence rather
formational change of rhodopsin occurs on a timescale than a single long simulation (Caves et al., 1998) is supported by a re-
cent work (Grossfield et al., 2007). The initial coordinates of rhodopsin
from microseconds to milliseconds (Menon et al., 2001;
were taken from the Protein Data Bank, PDB ID 1HZX, at 2.6 A resolu-
Okada et al., 2001). With standard analyses of simulations tion (Palczewski et al., 2000; Teller et al., 2001). Chain A of the dimer in
on the MD timescale, as used in these papers, it was not the crystal was used in this study because it has fewer missing resi-
possible to find the essential coupling involved in the rho- dues. For details of the system setup, see Supplemental Experimental
dopsin cycle, although some data concerning structural Procedures 1. In addition to the protein, the system contained 138
changes of short duration were obtained. To overcome DOPC molecules, 5,422 water molecules, and 24 ions; thus, the entire
this limitation, we introduced a novel, to our knowledge, system has 34,662 atoms.
NPT (constant pressure and constant temperature) MD simulations
approach for investigating tertiary allosteric coupling by
(Andersen, 1980; Hoover, 1984; Nose and Klein, 1983) were per-
estimating the correlations between residue-residue inter- formed; for details, see Supplemental Experimental Procedures 2.
actions in equilibrium MD simulations. In actual allosteric The potential function for retinal was obtained from a published ab ini-
transitions, both the signal initiation (e.g., ligand binding, tio quantum mechanical/molecular mechanical calculation (Tajkhor-
retinal isomerization) and the remote-site response (e.g., shid et al., 1997, 2000). Retinal isomerization was induced by imposing
conformational changes) are much stronger than what is a harmonic potential with a force constant of 50 kcal/mol/radian2,
which switches the C11 = C12 double bond from cis to trans in
expected to be observed in equilibrium thermal fluctua-
200 fs, in accord with the experimental timescale (Rohrig et al.,
tions. The signal propagation is expected to be a serial, 2002). After that, a 1 ns simulation was performed with the retinal in
multistep process that is transmitted through the protein. the all-trans configuration. This process was repeated eight times,
In most cases, as in rhodopsin, it must overcome ener- starting from eight independent cis-retinal rhodopsin simulations. To
getic barriers, which makes it too slow to be captured investigate further the global structural relaxation due to retinal isom-
by current state of the art MD simulations. However, if erization, one simulation was continued for 9 ns with all-trans retinal.
To study whether or not the long-range correlations observed in the
the allosteric transition does not involve a large conforma-
simulations are brought about by the protein scaffold, a control
tional change (e.g., rhodopsin dark state to meta I [Ru-
simulation was performed with only two residue pairs, 94/296 and
precht et al., 2004]), the pathway and the mechanism of 135/249. For details, see Supplemental Experimental Procedures 3.
long-distance coupling in the actual signal propagation The NPT simulation of rhodopsin with regional restraints, protonated
is likely to be encoded in the protein structure and can Glu134, and charge-scaled Asp83 followed the procedure of the
be determined by looking at its equilibrium behavior (Kar- rhodopsin simulation described above. For details, see Supplemental
plus and Gao, 2004). Therefore, we are able to use the Experimental Procedures 4.

coupling of interactions observed in nanosecond simula-


tions to determine the actual signaling pathway that Interaction Energy Correlation
The analysis is based on the energetic coupling between pairs of res-
produces the allosteric transition.
idues and the equal-time correlation function between the pairs. The
In this study, we identified a network of coupled resi- latter corresponds to the equilibrium correlation between the pairs,
dues extending from the retinal-binding pocket to the which provides information about how the interactions are coupled
cytoplasmic surface. We showed that helices VI and VII, with each other. We begin by determining the residue-residue

620 Structure 15, 611623, May 2007 2007 Elsevier Ltd All rights reserved
Structure
Signaling Pathway of Rhodopsin

interaction-energy time series from MD trajectories. Only amino acids The error estimates of the values in Figure 3A are shown in Figure S8;
were included in the analysis, i.e., the membrane lipids, water mole- the error is defined as the difference between residuewise correlation
cules, and ions were not considered. However, retinal was treated values to residue 296 independently calculated from two sets of coor-
as part of the side chain of residue 296. For each recorded dynamics dinates, which were generated by randomly partitioning the recorded
frame (see below), the nonbonded interaction energy, Ei,j, between coordinates into two sets.
two residues, i and j, is defined as
Effects of Retinal Isomerization
Ei; j = Ei;elec vdw
j + Ei; j ji  jj>1; (1)
The isomerization of retinal, as part of residue 296, is expected to
change the interaction pattern between residue 296 and its neighbors.
where Ei;elec
j and Ei;vdw
j are the electrostatic and van der Waals interac-
However, the effect on the interactions between more distal residues is
tion energies, respectively, between residues i and j, summed over
of more interest for the signal-transduction mechanism. The potential
the main chain and side chain atoms; the PME potential was not in-
perturbation of distal residue interaction by retinal isomerization was
cluded here because it is expected to provide only a small correlation
calculated, projecting the changes of retinal-involved interactions on
to the pairwise interaction energy between residues within a single
the interaction correlation matrix; details are given in Supplemental
rhodopsin molecule. The interaction energies for amino acid residue
Experimental Procedures 5.
pairs that are neighbors in sequence are not included because they
are covalently bonded. The average interaction energy between resi-
Supplemental Data
due i and residue j is defined as
Supplemental Data include one table and Experimental Procedures
1X f and are available at http://www.structure.org/cgi/content/full/15/5/
E i; j = E t ji  jj>1; (2) 611/DC1/.
f t = 1 i; j

where Ei;t j is the interaction energy between residues i and j in coordi- ACKNOWLEDGMENTS
nate set t, and f is the number of coordinate sets; as stated above, we
saved coordinates every 0.2 ps. The correlation between two sets of The part of the work done at Harvard was supported by the National
residue-residue interactions, i,j and k,l, Ci; jjk; l , is defined as Institutes of Health.
Pf  t  
t = 1 Ei; j  Ei; j Ek;t l  Ek; l Received: December 13, 2006
Ci; jjk;l = r
2  2 : (3) Revised: March 30, 2007
Pf
t=1 Ei;t j  Ei; j Ek;t l  Ek; l Accepted: April 6, 2007
Published: May 15, 2007
We include a given pair only if its average interaction energy is
greater in magnitude than 1 kcal/mol; this gives 1026 pairs. In the pres- REFERENCES
ent simulation of rhodopsin, the absolute value of the average interac-
tion energies varies from 5.85 3 108 (i.e., 0) to 79.23 kcal/mol. Andersen, H.C. (1980). Molecular dynamics simulations at constant
pressure and/or temperature. J. Chem. Phys. 72, 23842393.
Matrix Assembly Arnis, S., and Hofmann, K.P. (1995). Photoregeneration of bovine rho-
From Equation 3, the interaction-energy correlation matrix has col- dopsin from its signaling state. Biochemistry 34, 93339340.
umns and rows corresponding to all nonsequential neighboring resi- Bae, H., Anderson, K., Flood, L.A., Skiba, N.P., Hamm, H.E., and
due-residue interactions that are included. To reduce the size of the Graber, S.G. (1997). Molecular determinants of selectivity in 5-hydroxy-
matrix to a reasonable value, a correlation cutoff (Ccutoff) was intro- tryptamine1B receptor-G protein interactions. J. Biol. Chem. 272,
duced in addition to the interaction-energy cutoff (see Figure S3). In 3207132077.
the current study, the Ccutoff was set to 0.204, which generates a Bae, H., Cabrera-Vera, T.M., Depree, K.M., Graber, S.G., and Hamm,
993 3 993 interaction correlation matrix (Figure S4). It is useful to H.E. (1999). Two amino acids within the a4 helix of Gai1 mediate cou-
also define a residue correlation matrix, which is essentially a projec- pling with 5-hydroxytryptamine1B receptors. J. Biol. Chem. 274,
tion of the interaction energy correlation matrix on the residue space. 1496314971.
By this, we mean that for any pair of residues, I and J, the residue cor-
Baylor, D.A., Lamb, T.D., and Yau, K.W. (1979). Responses of retinal
relation matrix is obtained by summing over all interacting pairs involv-
rods to single photons. J. Physiol. 288, 613634.
ing I and J. This is, of course, different from the direct I,J interaction,
since it includes indirect interactions, which are likely to be important Brunori, M., Careri, G., Changeux, J.P., and Schachman, H.K. (2005).
for distant residues. If the dimension of the interaction correlation ma- Allosteric Proteins: 40 Years with Monod-Wyman-Changeux, Volume
trix is N (here N = 993), the correlation between residues I and J, RCI,J, 17 (Roma: Academia Nationale dei Lincei).
is defined as Cai, K., Itoh, Y., and Khorana, H.G. (2001). Mapping of contact sites in
complex formation between transducin and light-activated rhodopsin
X
N N 
X 
 I; J  by covalent crosslinking: use of a photoactivatable reagent. Proc. Natl.
RCI; J = Cmjn 3 dmjn ; (4)
m=1 n=1 Acad. Sci. USA 98, 48774882.
Caves, L.S., Evanseck, J.D., and Karplus, M. (1998). Locally accessi-
ble conformations of proteins: multiple molecular dynamics simula-
where dI;mjn
J
is equal to 1 only if residues I and J are involved in interac- tions of crambin. Protein Sci. 7, 649666.
tions m and n or n and m, respectively; otherwise, dI;mjn J
is 0. Cmjn is the
Changeux, J.P., and Edelstein, S.J. (2005). Allosteric mechanisms of
correlation between interactions m and n obtained from the interaction
signal transduction. Science 308, 14241428.
correlation matrix. For example, if the correlation between interaction
pairs (10,25j17,30) is 0.35 and that between pairs (10,14j17,35) is Crozier, P.S., Stevens, M.J., Forrest, L.R., and Woolf, T.B. (2003).
0.45, the residue correlation between 10 and 17 would be 0.70 due Molecular dynamics simulation of dark-adapted rhodopsin in an ex-
to these two interaction-energy correlations; this means that the resi- plicit membrane bilayer: coupling between local retinal and larger
due correlation can be greater than unity. The dimension of the residue scale conformational change. J. Mol. Biol. 333, 493514.
correlation matrix is equal to the number of residues in rhodopsin (348). Crozier, P.S., Stevens, M.J., and Woolf, T.B. (2007). How a small
Figure 3 shows the projection of row/column 296 (i.e., that involving change in retinal leads to G-protein activation: initial events suggested
retinal) of the residue correlation matrix on the rhodopsin structure. by molecular dynamics calculations. Proteins 66, 559574.

Structure 15, 611623, May 2007 2007 Elsevier Ltd All rights reserved 621
Structure
Signaling Pathway of Rhodopsin

DeCaluwe, G.L., Bovee-Geurts, P.H., Rath, P., Rothschild, K.J., and formation of the active form of Meta II. Biochemistry (Mosc.) 66,
de Grip, W.J. (1995). Effect of carboxyl mutations on functional prop- 12831299.
erties of bovine rhodopsin. Biophys. Chem. 56, 7987. Lemaitre, V., Yeagle, P., and Watts, A. (2005). Molecular dynamics
Fahmy, K., Sakmar, T.P., and Siebert, F. (2000). Transducin-depen- simulations of retinal in rhodopsin: from the dark-adapted state to-
dent protonation of glutamic acid 134 in rhodopsin. Biochemistry 39, wards lumirhodopsin. Biochemistry 44, 1266712680.
1060710612.
Levy, R.M., Martin, K., Kushick, J., and Perahia, D. (1984). Evaluation
Farahbakhsh, Z.T., Ridge, K.D., Khorana, H.G., and Hubbell, W.L. of the configurational entropy for proteins: application to molecular dy-
(1995). Mapping light-dependent structural changes in the cytoplas- namics simulations of an a-helix. Macromolecules 17, 13701374.
mic loop connecting helices C and D in rhodopsin: a site-directed
Lockless, S.W., and Ranganathan, R. (1999). Evolutionarily conserved
spin labeling study. Biochemistry 34, 88128819.
pathways of energetic connectivity in protein families. Science 286,
Faraldo-Gomez, J.D., Forrest, L.R., Baaden, M., Bond, P.J., Domene, 295299.
C., Patargias, G., Cuthbertson, J., and Sansom, M.S. (2004). Confor-
Meng, E.C., and Bourne, H.R. (2001). Receptor activation: what does
mational sampling and dynamics of membrane proteins from 10-nano-
the rhodopsin structure tell us? Trends Pharmacol. Sci. 22, 587593.
second computer simulations. Proteins 57, 783791.
Filipek, S., Stenkamp, R.E., Teller, D.C., and Palczewski, K. (2003). G Menon, S.T., Han, M., and Sakmar, T.P. (2001). Rhodopsin: structural
protein-coupled receptor rhodopsin: a prospectus. Annu. Rev. Phys- basis of molecular physiology. Physiol. Rev. 81, 16591688.
iol. 65, 851879. Nakamichi, H., and Okada, T. (2006a). Crystallographic analysis of pri-
Franke, R.R., Konig, B., Sakmar, T.P., Khorana, H.G., and Hofmann, mary visual photochemistry. Angew. Chem. Int. Ed. Engl. 45, 4270
K.P. (1990). Rhodopsin mutants that bind but fail to activate transdu- 4273.
cin. Science 250, 123125. Nakamichi, H., and Okada, T. (2006b). Local peptide movement in the
Fritze, O., Filipek, S., Kuksa, V., Palczewski, K., Hofmann, K.P., and photoreaction intermediate of rhodopsin. Proc. Natl. Acad. Sci. USA
Ernst, O.P. (2003). Role of the conserved NPxxY(x)5,6F motif in the 103, 1272912734.
rhodopsin ground state and during activation. Proc. Natl. Acad. Sci. Nose, S., and Klein, M.L. (1983). Constant pressure molecular-dynam-
USA 100, 22902295. ics for molecular-systems. Mol. Phys. 50, 10551076.
Gelin, B.R., Lee, A.W., and Karplus, M. (1983). Hemoglobin tertiary Okada, T., Ernst, O.P., Palczewski, K., and Hofmann, K.P. (2001).
structural change on ligand binding. Its role in the co-operative mech- Activation of rhodopsin: new insights from structural and biochemical
anism. J. Mol. Biol. 171, 489559. studies. Trends Biochem. Sci. 26, 318324.
Gether, U., and Kobilka, B.K. (1998). G protein-coupled receptors. II.
Okada, T., Sugihara, M., Bondar, A.N., Elstner, M., Entel, P., and Buss,
Mechanism of agonist activation. J. Biol. Chem. 273, 1797917982.
V. (2004). The retinal conformation and its environment in rhodopsin in
Grossfield, A., Feller, S.E., and Pitman, M.C. (2007). Convergence of light of a new 2.2 A crystal structure. J. Mol. Biol. 342, 571583.
molecular dynamics simulations of membrane proteins. Proteins 67,
Onrust, R., Herzmark, P., Chi, P., Garcia, P.D., Lichtarge, O., Kingsley,
3140.
C., and Bourne, H.R. (1997). Receptor and bg binding sites in the a sub-
Hoover, W.G. (1984). Canonical dynamics: equilibrium phase-space unit of the retinal G protein transducin. Science 275, 381384.
distributions. Phys. Rev. A. 31, 16951697.
Palczewski, K., Kumasaka, T., Hori, T., Behnke, C.A., Motoshima, H.,
Hubbard, R., and Wald, G. (1952). Cis-trans isomers of vitamin A and Fox, B.A., Le Trong, I., Teller, D.C., Okada, T., Stenkamp, R.E., et al.
retinene in the rhodopsin system. J. Gen. Physiol. 36, 269315. (2000). Crystal structure of rhodopsin: a G protein-coupled receptor.
Hubbell, W.L., Altenbach, C., Hubbell, C.M., and Khorana, H.G. (2003). Science 289, 739745.
Rhodopsin structure, dynamics, and activation: a perspective from
Patel, A.B., Crocker, E., Eilers, M., Hirshfeld, A., Sheves, M., and
crystallography, site-directed spin labeling, sulfhydryl reactivity, and
Smith, S.O. (2004). Coupling of retinal isomerization to the activation
disulfide cross-linking. Adv. Protein Chem. 63, 243290.
of rhodopsin. Proc. Natl. Acad. Sci. USA 101, 1004810053.
Huber, T., Botelho, A.V., Beyer, K., and Brown, M.F. (2004). Membrane
Rohrig, U.F., Guidoni, L., and Rothlisberger, U. (2002). Early steps of
model for the G-protein-coupled receptor rhodopsin: hydrophobic in-
the intramolecular signal transduction in rhodopsin explored by molec-
terface and dynamical structure. Biophys. J. 86, 20782100.
ular dynamics simulations. Biochemistry 41, 1079910809.
Isin, B., Rader, A.J., Dhiman, H.K., Klein-Seetharaman, J., and Bahar,
Ruprecht, J.J., Mielke, T., Vogel, R., Villa, C., and Schertler, G.F.
I. (2006). Predisposition of the dark state of rhodopsin to functional
(2004). Electron crystallography reveals the structure of metarhodop-
changes in structure. Proteins 65, 970983.
sin I. EMBO J. 23, 36093620.
Itoh, Y., Cai, K., and Khorana, H.G. (2001). Mapping of contact sites in
complex formation between light-activated rhodopsin and transducin Saam, J., Tajkhorshid, E., Hayashi, S., and Schulten, K. (2002). Molec-
by covalent crosslinking: use of a chemically preactivated reagent. ular dynamics investigation of primary photoinduced events in the
Proc. Natl. Acad. Sci. USA 98, 48834887. activation of rhodopsin. Biophys. J. 83, 30973112.

Karplus, M., and Gao, Y.Q. (2004). Biomolecular motors: the F1- Scheer, A., Fanelli, F., Costa, T., De Benedetti, P.G., and Cotecchia, S.
ATPase paradigm. Curr. Opin. Struct. Biol. 14, 250259. (1996). Constitutively active mutants of the a 1B-adrenergic receptor:
role of highly conserved polar amino acids in receptor activation.
Kim, J.M., Altenbach, C., Thurmond, R.L., Khorana, H.G., and Hubbell,
EMBO J. 15, 35663578.
W.L. (1997). Structure and function in rhodopsin: rhodopsin mutants
with a neutral amino acid at E134 have a partially activated conforma- Schenkl, S., van Mourik, F., van der Zwan, G., Haacke, S., and Chergui,
tion in the dark state. Proc. Natl. Acad. Sci. USA 94, 1427314278. M. (2005). Probing the ultrafast charge translocation of photoexcited
retinal in bacteriorhodopsin. Science 309, 917920.
Kostenis, E., Conklin, B.R., and Wess, J. (1997). Molecular basis of re-
ceptor/G protein coupling selectivity studied by coexpression of wild Stenkamp, R.E., Teller, D.C., and Palczewski, K. (2005). Rhodopsin:
type and mutant m2 muscarinic receptors with mutant G a(q) subunits. a structural primer for G-protein coupled receptors. Arch. Pharm.
Biochemistry 36, 14871495. (Weinheim) 338, 209216.
Kuwata, O., Yuan, C., Misra, S., Govindjee, R., and Ebrey, T.G. (2001). Suel, G.M., Lockless, S.W., Wall, M.A., and Ranganathan, R. (2003).
Kinetics and pH dependence of light-induced deprotonation of the Evolutionarily conserved networks of residues mediate allosteric com-
Schiff base of rhodopsin: possible coupling to proton uptake and munication in proteins. Nat. Struct. Biol. 10, 5969.

622 Structure 15, 611623, May 2007 2007 Elsevier Ltd All rights reserved
Structure
Signaling Pathway of Rhodopsin

Tajkhorshid, E., Paizs, B., and Suhai, S. (1997). Role of isomerization Weitz, C.J., and Nathans, J. (1993). Rhodopsin activation: effects on
barriers in the pKa control of the retinal Schiff base: a density functional the metarhodopsin I-metarhodopsin II equilibrium of neutralization or
study. J. Phys. Chem. B 103, 45184527. introduction of charged amino acids within putative transmembrane
Tajkhorshid, E., Baudry, J., Schulten, K., and Suhai, S. (2000). Molec- segments. Biochemistry 32, 1417614182.
ular dynamics study of the nature and origin of retinals twisted struc-
Yan, E.C., Kazmi, M.A., Ganim, Z., Hou, J.M., Pan, D., Chang, B.S.,
ture in bacteriorhodopsin. Biophys. J. 78, 683693.
Sakmar, T.P., and Mathies, R.A. (2003). Retinal counterion switch in
Teller, D.C., Okada, T., Behnke, C.A., Palczewski, K., and Stenkamp,
the photoactivation of the G protein-coupled receptor rhodopsin.
R.E. (2001). Advances in determination of a high-resolution three-
Proc. Natl. Acad. Sci. USA 100, 92629267.
dimensional structure of rhodopsin, a model of G-protein-coupled
receptors (GPCRs). Biochemistry 40, 77617772. Zgrablic, G., Voitchovsky, K., Kindermann, M., Haacke, S., and
Wald, G. (1968). Molecular basis of visual excitation. Science 162, Chergui, M. (2005). Ultrafast excited state dynamics of the protonated
230239. Schiff base of all-trans retinal in solvents. Biophys. J. 88, 27792788.

Structure 15, 611623, May 2007 2007 Elsevier Ltd All rights reserved 623

Você também pode gostar