Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2012 Feb 25.
Published in final edited form as: J Mol Biol. 2011 Jan 6;406(3):467–478. doi: 10.1016/j.jmb.2010.12.034

Crystal structure of arrestin-3 reveals the basis of the difference in receptor binding between two non-visual subtypes

Xuanzhi Zhan 1, Luis E Gimenez 1, Vsevolod V Gurevich 1,*, Benjamin W Spiller 1,2,*
PMCID: PMC3034793  NIHMSID: NIHMS263280  PMID: 21215759

Abstract

Arrestins are multi-functional proteins that regulate signaling and trafficking of the majority of G protein-coupled receptors (GPCRs), as well as sub-cellular localization and activity of many other signaling proteins. Here we report the first crystal structure of arrestin-3, solved at 3.0Å. Arrestin-3 is an elongated two-domain molecule with the overall fold and key inter-domain interactions that hold free protein in the basal conformation similar to the other subtypes. Arrestin-3 is the least selective member of the family, binding wide variety of GPCRs with high affinity and demonstrating lower preference for active phosphorylated forms of the receptors. In contrast to the other three arrestins, part of the receptor-binding surface in the arrestin-3 C-domain does not form a contiguous β-sheet, consistent with increased flexibility. By swapping the corresponding elements between arrestin-2 and -3 we show that the presence of this loose structure correlates with reduced arrestin selectivity for activated receptor, consistent with a conformational change in this β-sheet upon receptor binding.

Keywords: arrestin, stability, BRET, self-association, GPCR

Introduction

Mammals express four arrestin proteins1. In contrast to highly specialized visual arrestin-1 and -4 that are selectively expressed at very high levels in photoreceptors and bind rhodopsin and cone opsins2; 3; 4, the two non-visual subtypes are present in virtually every cell, and interact with hundreds of G protein-coupled receptors (GPCRs)5. In most cases arrestin-2 outnumbers arrestin-3 by ~10–20:16; 7. Non-visual arrestins also serve as multi-functional adaptors, linking GPCRs to the endocytic machinery8; 9 and directing signaling to mitogen-activated protein kinases and Src family kinases, as well as orchestrating protein ubiquitination (reviewed in5; 10). Direct studies of the binding of different arrestins to the four functional states of their cognate receptors (active phosphorylated, inactive phosphorylated, active unphosphorylated, and inactive unphosphorylated) revealed striking differences in their selectivity. Arrestin-1 has a remarkable preference for active phosphorylated receptor over inactive phosphoreceptor with >10-fold difference in binding11; 12, while non-visual arrestins show <2-fold difference in binding levels at best13; 14; 15. Non-visual arrestins also show a much smaller preference for phosphorylated over unphosphorylated active receptors than arrestin-113; 16. In both regards, arrestin-3 is the least selective.

Single non-visual arrestin knockout mice are grossly normal, whereas the double arrestin-2/3 knockout is embryonic lethal17, suggesting that to some extent these subtypes are redundant. However, arrestin-3 demonstrates higher affinity for many G protein-coupled receptors (GPCRs)17; 18, as well as clathrin8, and has a number of unique functions (reviewed in5; 10). Crystal structures of arrestin-119; 20, arrestin-221; 22; 23; 24, and arrestin-425 have been determined, making the structure of this intriguing subtype the only one missing. Here we report the crystal structure of bovine arrestin-3 at 3.0Å resolution (Table 1). While the overall fold of arrestin-3 resembles that of the other family members, we identified an element on the receptor-binding concave surface of the C-domain26 that is significantly different from its closest homologue, arrestin-2. By swapping this element between non-visual arrestins, we demonstrate that this structural difference significantly contributes to the differential receptor binding of the two subtypes. The structure of arrestin-3 also allowed us to identify a highly conserved self-association interface in the arrestin family.

Table 1.

Data collection and refinement statistics

Data Collection
Space Group P212121
Cell Dimensions (Å)
 a, b, c (Å) 73.18, 73.32, 201.97
 α, β, γ(°) 90.00, 90.00, 90.00
Resolution (Å) 3.0
Rsym (%) 4.4(83)
I/σI 24(1.8)
Completeness (%) 92.7(94.4)
Redundancy 5.2(5.2)
Refinement
Resolution range (Å) 35−3.0
No. of reflections 20,879
Rfactor (%) 22.1
Rfree (%) 28.6
Total Protein Atoms 5,611
RMSD from ideality
 Bond lengths 0.011
 Bond angles 1.508
Mean Coordinate Error (Å)
 Luzzati plot 0.36
 SigmaA 0.37
 Maximum Likelihood 0.40
Ramachandran (%)
 Favored 97.6
 Allowed 1.3
 Outliers 1.1

RESULTS

The conservation of the “arrestin fold”

The arrestin-3 structure is largely similar to other arrestins (Fig. 1A). It is composed of two 7-stranded beta sandwiches, termed the N-terminal and C-terminal domains, connected by a ten-residue linker (hinge region27). The carboxy terminus folds back toward the N-terminal domain, becomes unstructured for ~35 residues and forms a highly conserved tripartite interaction with the N-terminal domain consisting of two hydrophobic interactions with β-strand I and α-helix I and one buried ion pair constituting part of the main arrestin phosphate sensor, the polar core20; 28. Sequence and structural alignments of arrestin-1, -2, and -3 are shown in Fig. 1, with key functional elements highlighted. The structure of the polar core, as well as the positions of other known phosphate-binding residues are well conserved between arrestin-3 and -2, as could be expected based on the functional conservation of phosphate sensing mechanisms in the arrestin family14; 15; 28. The concave sides of the two arrestin domains also contain elements that interact with other parts of the receptor13; 26; 29; 30; 31. The two key “receptor discriminator” regions (N- and C-domain elements) responsible for receptor preference of arrestin proteins constitute parts of this surface in the N- and the C-domain26. A key difference between arrestin-3 and arrestin-2 is that the outer most “strand” in the C-domain element, residues 250–259, although well-ordered (Fig. 2A) is displaced from the position it occupies in arrestin-2. Displacements of Cα atoms for V257, E258, Q259, and D260 are 1.3Å, 2.2Å, 2.0Å, and 2.0Å, respectively. Although these shifts are relatively small, they are ~3–5 times greater than the estimated positional error of the coordinates (Table 1), and occur in a region known to be essential for receptor binding26. The shift results in the loss of two H-bonds within the C-domain element, (Fig. 2B), consistent with a lower energy barrier for reorganization of this region. The distorted H-bonds are between Q259 and I232, resulting in distances of 4.9Å between the Q259 carbonyl oxygen and the Q232 amide nitrogen and 4.2Å between the Q259 amide nitrogen and I232 carbonyl oxygen. The alignment of previously determined structures indicate that the N-domain element exists in multiple conformations when multiple views of identical proteins are afforded by either non-crystallographic symmetry or from multiple solved structures19; 20; 21; 22; 23; 24 (Fig. 2C), indicating that the N-domain element is inherently flexible, and therefore structural differences in this region are unlikely to be particularly informative. In contrast, the structure of the C-domain element is highly conserved (Fig. 2C). An overlay of the C-domain from all non-visual arrestins (there are ten crystallographically unique arrestin-2 structures: 1g4r, 1zsh, 1jsy, 3gc3, and two non-identical molecules in each of 1g4m 3gdi, and 2wtr)21; 22; 23; 24, shows that in all cases, arrestin-2 forms a complete β-sheet in the C-domain element, whereas both versions of arrestin-3 in our asymmetric unit form a distorted β-sheet (Fig. 2D).

Figure 1. Comparison of arrestin-3 to other arrestins.

Figure 1

Figure 1

The N-domain element and C-domain element that are important for receptor specificity are shown in red. Residues important for locking the C-terminus to the N-terminal domain and critical polar core residues are shown as sticks in panel A and as colored residues in panel B, with the polar core residues shown in green and the C-terminal tail residues shown in majenta. The residues in the arrestin-3 C-element that do not form a proper β-sheet are shown in grey in panel A and white on black highlight in panel B. A. Structural overlay of arrestin-2 (blue) and arrestin-3 (salmon). B. Sequence alignment of arrestins with solved structures.

Figure 2. Arrestin-3 has a uniquely distorted C-domain element.

Figure 2

Figure 2

Figure 2

Figure 2

The N-domain element and C-domain element that are important for receptor specificity are shown in red. The residues in the arrestin-3 C-domain element that do not form proper β-sheet hydrogen bonding are shown in grey. A. The distorted region of the C-domain element is well ordered and displays clear electron density. Shown is a composite omit map drawn at 1.5 sigma. B. Overlay of arrestin-2 (1g4r, blue) and arrestin-3 (salmon), showing the deviation from standard β-sheet geometry. Distances between Cα positions in Arr2 and Arr3 are shown. C. Sausage representation of a calculated average arrestin-2 structure, based on the ten crystallographically independent arrestin-2 structures. The sausage thickness corresponds to average deviation of the eight structures from the average structure, as calculated with THESEUS 80. Receptor discriminator elements are shown in red, with the C-domain element on the left and the N-domain element on the right, showing that although the N-domain element is variable, the C-domain element in arrestin-2 is structurally conserved. D. Alignment of ten distinct arrestin-2 structures (blue, from pdbs 1g4r, 1g4m, 1zsh, 1jsy, 3gdi, 3gc3, and 2wtr) with the two arrestin-3 structures present in the asymmetric unit (salmon), indicating that all other arrestins form a contiguous β-sheet in the C-terminal domain.

Structurally distinct receptor-binding surface contributes to decreased arrestin-3 specificity for active receptor

The arrestin-3 structure reveals that despite very similar overall architecture, the C-domain beta-sandwich is distorted, resulting in a distinct conformation of the C-domain region implicated in receptor binding26 (Fig. 2). This difference occurs in a region of the structure that is otherwise well ordered, as judged by the difference electron density (Fig. 2A). The distorted strand does not make canonical β-sheet hydrogen bond contacts with the rest of the C-domain element (Fig. 2B), consistent with a less stable sheet. This sets arrestin-3 apart from the other three vertebrate arrestin subtypes with known structure. Since this is the first distinguishing structural feature of the receptor-binding elements in any arrestin, we experimentally tested the functional consequences of the “relaxed” conformation of this element in arrestin-3.

Arrestins are highly homologous proteins demonstrating remarkable conservation of the positions of all known functionally important residues from C. elegans to mammals1. Due to this conservation, chimeras between different arrestin family members fold properly and retain functionality11; 13; 32. Therefore, we exchanged the region with distinct structure (residues 233–261 and 234–262 in arrestin-2 and -3, respectively) between the two non-visual arrestins, generating arrestin-2-6M and arrestin-3-6M mutants. Due to high homology, this is equivalent to the mutation of six residues (V234I, S246N, Q256M, V257E, Q259A, Q262T in arrestin-3, and I233V, N245S, M255Q, E256V, A258Q, T261Q in arrestin-2). Receptor binding of these chimeras was compared to parental wild type proteins in intact cells using arrestins tagged with Venus (enhanced YFP33) on the N-terminus and β2-adrenergic receptor (β2AR) C-terminally tagged with Renilla luciferase in COS-7 cells. We chose β2AR because both non-visual arrestins interact with this receptor in vitro, Xenopus oocytes, and cultured cells13; 14; 15; 16; 17; 18; 34. One distinct functional feature of arrestin-3 is that it demonstrates much smaller difference than arrestin-2, in binding to inactive and active GPCRs when the receptor is phosphorylated13; 14; 15. Arrestin recruitment to the receptor is reflected in bioluminescence resonance energy transfer (BRET) from luciferase to Venus, which saturates with increasing Venus-arrestin expression35; 36; 37 (Fig. 3).This assay allows measurement of arrestin interactions with receptors with or without agonist stimulation in the physiological conditions of an intact cell. We found that in this cell-based assay arrestin-3 shows a stronger agonist-independent binding to β2AR than arrestin-2, and a correspondingly much smaller agonist-induced increase (Fig. 3). The introduction of six arrestin-2 residues into arrestin-3 significantly boosted agonist-induced arrestin-β2AR interaction (Fig. 3B), whereas introducing corresponding arrestin-3 residues into arrestin-2 somewhat reduced it (Fig. 3A). The data suggest that the absence of a proper β-strand in this region of arrestin-3 likely destabilizes the receptor binding face of the beta sandwich, allowing it to more readily adopt the “active” conformation, thereby reducing the requirement for receptor activation. Interestingly, charge reversal mutation K257E in a homologous region of arrestin-1 was found to greatly decrease its selectivity, dramatically increasing the binding to both non-preferred forms of its cognate receptor, phosphorylated inactive and light-activated unphosphorylated rhodopsin31.

Figure 3. Exchange of the structurally different C-domain receptor-binding element between arrestin-2 and -3 changes their binding to ß2AR in intact cells.

Figure 3

Figure 3

A. BRET signal as a function of WT Venus-arrestin-2 (filled circles, solid line) and the Venus-arrestin-2-6M mutant (I233V, N245S, M255Q, E256V, A258Q and T261Q; open circles, dashed line) co-expression with ß2AR-RLuc in COS-7 cells. B. BRET signal as a function of WT Venus-arrestin-3 (filled triangles, solid line) and the Venus-arrestin-3-6M mutant (V234I, S246N, Q256M, V257E, Q259A and Q262A; open circles, dashed line) co-expression with ß2AR-RLuc in COS-7 cells. Shown is the difference between BRET signal in the presence and absence of 25 μM ß2AR agonist isoproterenol, which reflects agonist-induced increase in Venus-arrestin interaction with fixed amount of ß2AR-RLuc. Means ± SE of one representative experiment (out of three) performed in quadruplicate are shown in A and B. C. The average BRETMAX as estimated from the fits of a one-site binding hyperbola to the data of the binding of the indicated arrestins to the ß2AR. Means ± S.E. of three independent experiments are shown. The differences between the BRETMAX values were assessed by one-way ANOVA, followed by Bonferroni post hoc test. Significance of the differences is indicated as follows: *p<0.05, **p<0.01, ***p<0.001.

Inter-subunit interface in arrestin-3 crystals

The oligomeric interactions of arrestins are thought to be functionally important, although for most arrestins, the nature of the biologically relevant oligomers is unknown. Nevertheless, self-association appears to be a key feature of arrestins, and has been documented for arrestin-138; 39, -2, and -323; 40. Arrestin-4 is the only family member known to be a constitutive monomer40. The analysis of all arrestin structures reveals one common highly conserved interface present in crystal forms of arrestins-1, -2, and -3. This interface is formed via a parallel beta-sheet interaction between the first strand on the concave face of the N-terminal domain and the first strand on the convex face of the C-terminal domain (Fig. 4). This interface is observed in multiple crystal forms of arrestin-1 (1cf1 and 1ary)19; 20, arrestin-2 (1g4r, and the related 1g4m, 1zsh, and 1jsy)21; 23; 24, the co-crystal of arrestin-2 with clathrin NT domain (3gc3)22, as well as the arrestin-3 crystals reported here. The only crystal forms where this interface is not conserved are arrestin-4 (1suj)25 and an alternative crystal of arrestin-2-clathrin NT domain complex (3gd1)22. The surface area buried by each partner in this conserved interface varies from ~1017 Å2 for arrestin-1 (1cf1) to ~494 Å2 for arrestin-2 bound to clathrin (3gc3). Frequently (85% of the time), interfaces of greater than ~850Å2 (per partner) represent biologically relevant dimerization interfaces41, indicating that the 1cf1 interface is likely to be biologically important. We note here that while this interface is frequently small, ~666Å2 in arrestin-3, it is conserved among all arrestins known to form oligomers, and not conserved in the cone arrestin, which is a constitutive monomer25; 40, or among the arrestin-like retromer subunit Vps26 proteins42; 43. Further, inositol-6-phosphate (IP6), which is known to promote oligomerization of arrestin-2 and -3, binds between protomers in a binding pocket formed by oligomerization at this interface (Fig. 4c)22; 40.

Figure 4. Conserved packing interface.

Figure 4

Figure 4

Figure 4

Figure 4

The interface (colored yellow) is a parallel ß-sheet interaction between the first strand on the concave face of the N-terminal domain and the first strand on the convex face of the C-terminal domain. The length of the interaction and the residues involved varies. A. In arrestin-1 (1cf1, green), six residues from each monomer participate, and each monomer buries ~1017Å2. B. In arrestin-2 (1g4r, light blue) three residues from each monomer participate, with each arrestin-2 monomer burying ~726Å2. C. In the arrestin-2-IP6 structure (1zsh, dark blue), four residues from each monomer participate, with each arrestin-2 monomer burying ~585Å2. IP6 is drawn as space-filling spheres and is visible in the background, behind the C-terminal β-sandwich, packed between N-terminal and C-terminal domains, in a binding site formed by dimerization. D. In arrestin-3 (3P2D, salmon), three residues from each monomer participate and each monomer buries 666 Å2.

DISCUSSION

Several lines of evidence suggest partial functional redundancy between arrestin-2 and -3: the expression of the two non-visual arrestins in many tissues and cell types significantly overlaps6; 7; 44; 45, both bind many GPCRs with comparable affinity13; 34, and they can substitute for each other in single isoform knockout mice17. However, recent findings also revealed significant functional differences: arrestin-2 promotes RhoA activation in conjunction with Gaq/1146 and IGF-1-mediated activation of phosphatidylinositol-3-kinase47. A difference between the receptor-bound conformations of arrestin-2 and -3 was recently reported48. This study also noted a conserved conformational change that occurs upon activation of both arrestin-2 and -3 and involves increased protease accessibility of the region between amino acids 180 and 250 in both isoforms, consistent with increased flexibility of this region upon arrestin activation. Furthermore, different trafficking routes for agonist-activated gastrin-releasing peptide receptor (GRP-R)-arrestin-2 and GRP-R-arrestin-3 complexes were reported49. Arrestin-3 internalizes with GRP-R to endosomes, whereas arrestin-2 dissociates from the GRP-R near the plasma membrane. However, there was virtually no mechanistic and structural insight into the functional differences between arrestin-2 and -3.

The structure of bovine arrestin-3 reported here completes the structural inventory of this protein family. The overall structure is very similar to other arrestins, both in terms of the individual domains and the structural relationship between these domains (rmsd=0.8Å for arrestin-3 to arrestin-2 (1G4M) using 282 core Cα positions). The key phosphate-sensitive interactions holding the relative orientation of the two domains, the inter-domain polar core and the three-element interaction between the C-tail, β-strand I and α-helix I, are well conserved (Fig. 2). Although inter-domain flexibility appears to be an important functional feature of arrestins27, the relative orientation of the two domains in the basal state is conserved in the family (Fig. 1A). Differences in inter-domain orientation have been reported only in the active receptor-bound state, which is expected to be quite different from the basal conformation of free arrestin50. The relaxed confromation within the C-domain receptor-binding element (Fig. 2B) of arrestin-3 sets it apart from other isoforms. Our experiments show that swapping this element between arrestin-2 and -3 shifts the preference of the resulting chimera for the active receptor towards that of the donor arrestin (Fig. 3), indicating an important role of the flexibility of this region in receptor binding. Indeed, K257E mutation in this area in the most specific member of the family, arrestin-1, resulted in several-fold increases in binding to phosphorylated inactive and active unphosphorylated rhodopsin, thereby significantly reducing its selectivity for activated phosphorylated form31. Interestingly, within one residue of this distorted element one finds one or two amino acid insertions in the invertebrate arrestins, from C. elegans to Drosophila1, some of which were shown to bind unphosphorylated receptors. Collectively, these data support the idea that the rigidity of the C-domain element is a prerequisite for high arrestin selectivity for active phosphoreceptors. Importantly, this element is part of the more compact C-domain, which invariably is more rigid, as judged by differences between atomic positions of multiple structures (Fig. 2C), than the N-domain, which contains key phosphate-binding elements involved in arrestin activation28; 51. It is tempting to speculate that inherently more flexible receptor-binding surface contributes to a better ability of arrestin-3 to “fit” numerous GPCRs, as reflected in its higher affinity for many receptors17; 18. As this structural difference is likely to affect the stability of the β-sandwich core of the C-domain, it may also underlie known differences in the binding of non-receptor partners engaging this part of the arrestin molecule. MAP kinases c-Raf1, MEK1, ERK2, ASK1, MKK4, and JNK352, as well as ubiquitin ligase Mdm253, interact with both domains of the two non-visual arrestins. Despite similar binding54, arrestin-3, but not arrestin-2, promotes JNK3 activation32; 52; 55, indicating that it arranges the three kinases in the complex differently. The distinct structure of arrestin-3 may contribute to the biologically important functional differences. The identification of the binding sites for these kinases and other non-receptor partners of arrestin proteins is necessary to test this idea56.

All arrestins, with the exception of cone-specific arrestin-440, oligomerize. Self-association of arrestin-1 (under the name of S-antigen) was described even before its role in the regulation of rhodopsin was established57. In solution arrestin-1 cooperatively forms tetramers39; 58, where the receptor-binding elements of all monomers are shielded by sister subunits59. Not surprisingly, only monomeric arrestin-1 binds rhodopsin58, although the oligomers retain the ability to bind microtubules58, which serve as the default arrestin-1 “parking space” in the dark-adapted rod60; 61. Oligomerization of arrestin-2 has been observed in some crystal forms21; 22; 24, including one (1zsh)23 where inositol-6-phosphate (IP6), interacting with parts of the receptor-binding surface, was found to fit perfectly between monomers in a pocket formed by a conserved interface (Fig. 4C)23. Indeed, IP6 was shown to promote self-association of purified arrestin-240, and elimination of IP6-binding positive charges reduced its oligomerization in cells23. Similar observations with arrestin-3 led to the idea that its inter-subunit interface resembles that of arrestin-223. As shown in Figure 4, this oligomeration interface is conserved in the arrestin-3 structure described here. Although various ideas have been proposed23; 40; 62, the biological role of homo- and possibly hetero-oligomerization of non-visual arrestins and the ability of the oligomers to bind GPCRs remains unknown (reviewed in63). The role of this interface in arrestin-3 oligomerization in solution needs to be tested experimentally.

Methods

Expression and Purification of Arrestin-3 and Its Truncation mutant

cDNA encoding bovine arrestin-3 was cloned in frame in the pTrcHisB vector (Invitrogen) between Nco I and Hind III sites. A stop codon (TGA) has been introduced at L394 to generate arrestin3-(1-393) truncation mutant by PCR. Arrestin-3 proteins were expressed in BL21 Gold bacterial cells. Full-length arrestin-3 was purified as described64. Arrestin3 1-393, which failed to interact with the Q-Sepharose, was purified by a modified procedure. Arrestin3-(1-393) eluted from heparin-Sepharose was loaded onto a Q-Sepharose column while diluting the sample with column buffer (10 mM Tris/HCl, pH 7.5, 2 mM EDTA, 2 mM EGTA, 2 mM benzamidine, 1 mM PMSF, 4mM DTT) (CB) to a final concentration of 10 mM NaCl. The flow-through containing the majority of arrestin-3-(1-394) was loaded directly onto an SP-Sepharose column. The coluimn was washed with CB containing 100 mM NaCl, and then eluted with a 400 ml linear gradient (100 mM NaCl to 500 mM NaCl). Eluted arrestin-3-(1-394) (peak at ~320 mM NaCl) was concentrated and further purified by gel filtration on a Superdex S200 column equilibrated in 10 mM Tris (pH 7.5), 150 mM NaCl, 2mM TCEP.

Crystallization and structure determination of arrestin-3-(1-394)

Arrestin3-(1-394) was concentrated to 12 mg/mL and crystallized by hanging drop vapor diffusion from 50mM HEPES, 675mM Na/K tartrate, pH 7.5. Crystals appeared in 48 hours and were harvested after two weeks. The crystals were cryo-preserved by stepwise (5% steps) transfer from mother liquor to mother liquor containing 30% glycerol, followed by freezing in liquid nitrogen. Data were collected at the 24-ID-C beamline at the Advanced Photon Source, in Argonne, IL. Data were integrated and scaled with HKL200065. The phases were determined by molecular replacement using PHASER66 and the 1G4M21 structure. The search model included residues 5–357. Waters and the C-terminal tail were removed manually, and non-conserved side chains were truncated to Cb using the program CHAINSAW67 as implemented in CCP468. The structure was manually rebuilt in COOT69 using simulated annealing composite omit maps calculated in CNS70; 71. The structure was refined in Phenix72, using non-crystallographic restraints between the two chains, and using 10 TLS groups identified with the TLS motion determination webserver73; 74. The refinement strategy included positional refinement, simulated annealing, and group B-factor refinement. This strategy, as implemented in Phenix, makes use of maximum likelihood weighting of the errors associated with measured reflections, allowing proper weighting of weak data and hence its inclusion in refinement75. Estimates for upper limits of mean positional error were obtained by the method of Luzzatti (a plot of R-factor vs reciprocal resolution)76 with improvements allowing for the potential incompleteness of the model (SigmaA)77 and for different error distributions for different atoms (maximum likelihood)78. Refinement statistics and error estimates are given in Table 1. The structure was analyzed using PISA79, THESEUS80, and Pymol81, and figures were prepared using Pymol81. Coordinates and structure factors have been deposited in the Protein Data Bank with the accession number 3P2D, and will be released upon publication.

Plasmid construction for bioluminescence energy transfer (BRET)

The plasmid g3NVE-1 containing the sequence of arrestin-3 N-terminally tagged with Venus (a variant of enhanced yellow fluorescent protein33; generous gift from Dr. J. A. Javitch, Columbia University), was constructed using a modified pGEM2 in vitro transcription vector (Promega; Madison, WI) that contains under control of SP6 promoter “idealized” 5'-untranslated region82 with upstream EcoRI site followed by the coding sequence of bovine arrestin-3 between NcoI and HindIII sites. Venus was amplified by PCR using the 5'-AGTCAGAATTCGCGATCGCGGCCACGATGGTGAGCAAGGGCGA-3' forward primer that adds EcoRI and AsiSI sites upstream of the start codon and the 5'-TCTCCCCCATGGAGTCGAGCGCTCGCCGAGACTTAAGTCCGGAGGTGGCCT-3' reverse primer that codes for a short spacer with the “SGLKSRRALDS” sequence and an in-frame NcoI site. Venus was subcloned between EcoRI – NcoI restriction sites. Different arrestins were subcloned in frame with the Venus-spacer sequence using NcoI and HindIII sites. The Venus-arrestin fusion proteins were subcloned into a modified pcDNA3 mammalian expression vector (Invitrogen; Carlsbad, CA) using the EcoRI and HindIII restriction sites to generate the P3VEA2-5 and P3VEA3-1 plasmids encoding arrestin-2 and -3, respectively. Clones encoding arrestin-2-6M and arrestin-3-6M mutants were made by PCR in pGEM2, sequenced, and subcloned into pcDNA3. A plasmid encoding Renilla luciferase variant 8 (RLuc8)83 was a generous gift of Dr. Nevin A. Lambert (Medical College of Georgia). RLuc8 was fused in frame with the sequence of triple HA-tagged human ß2 adrenergic receptor (ß2AR, cDNA resource center, www.cdna.org). To this end, the coding sequence of ß2AR was amplified by PCR using the 5'-GCTAGAATTCTGCGATCGCACCACCATGGCGTACCCATACGATGTTCCA-3' forward primer that introduces EcoRI and AsiSI restriction sites upstream of the receptor start codon and the 5'-AGCGGAAGCTTCTAGCCTGCAGGTGCCAGCAGTGAGTCATTTG-3' reverse primer that introduces an in-frame SbfI restriction site, which was subcloned using EcoRI and SbfI sites in-frame with C-terminal RLuc8, yielding the P3HB2ALuc-2 plasmid.

BRET assays

The well-established BRET1 assay35; 36; 37; 84 with Venus as the acceptor and RLuc8 as the donor was used to characterize the binding of arrestins to ß2AR-RLuc8. COS-7 cells were transfected with indicated plasmids using Lipofectamine™ 2000 (Invitrogen; Carlsbad, CA), according to the manufacturers protocol (3 μL of Lipofectamine™ 2000 per 1 μg of DNA). Increasing amounts of indicated Venus-arrestin constructs (0–12 μg) along with 250 ng of P3HB2ALuc-2 and empty pcDNA3 to equalize DNA were used to transfect 80–90% confluent COS-7 cells on 60 mm dishes. Twenty-four hours after transfection, cells were trypsinized and re-seeded at 100,000 to 200,000 cells per well into white opaque 96 well microplates (Nunc, Rochester, NY) for luminescence measurements, or black opaque microplates (Nunc) for fluorescence determination. Forty-eight hours post-transfection, the media was replaced with PBS with Ca2+ and Mg2+ containing 0.01% glucose (w/v), 36 mg/L sodium pyruvate and 25 mmol/L HEPES, pH = 7.2. Coelenterazine-h (DiscoveRx, Fremont, CA) to a final concentration of 5 μM was added 8 min after agonist (25 μM isoproterenol) stimulation, and luminescence was measured immediately using a POLARstar Optima dual channel luminometer and fluorimeter microplate reader (BMG Labtech, Cary, NC). The light emitted by coelenterazine-h and Venus in each well was measured simultaneously five times through a 465-485 nm bandpass filter and through a 522.5–547.5 nm bandpass filter, respectively. The net BRET ratio was calculated as the long wavelength emission divided by the short wavelength emission and expressed as the relative change compared to unstimulated cells. The expression of each Venus-arrestin was evaluated using fluorescence at 535 nm upon excitation by 485 nm. The Venus-arrestin fluorescence which is directly proportional to the expression levels was normalized by the basal luminescence from the ß2AR-RLuc8 to account for variations in cell number and expression levels. The curves resulting from the titration of the various amounts of Venus-arrestin were fit by non-linear regression to a one-site hyperbola model using Prism version 5.04 (GraphPad Software, San Diego, CA). To assess the difference between the respective WT and mutant arrestins curves, a global fit algorithm was used.

Acknowledgements

The authors are grateful to Dr. Jonathan A. Javitch for expert advice on receptor-arrestin BRET and the plasmid encoding Venus, Dr. Nevin A. Lambert for plasmid encoding Renilla luciferase variant 8, and Dr. Carl Johnson for use of POLARstar Optima dual channel luminometer and fluorimeter microplate reader. This study was supported in part by NIH grants GM077561, GM081756 (VVG), and GM081778 (BWS).

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

ACCESSION NUMBERS: Coordinates and structure factors have been deposited in the Protein Data Bank with accession number 3P2D.

1

We use systematic names of arrestin proteins: arrestin-1 (historically called S-antigen, 48 kDa protein, and visual or rod arrestin), arrestin-2 (b-arrestin or b-arrestin1), arrestin-3 (b-arrestin2), and arrestin-4 (cone or X-arrestin).

References

  • 1.Gurevich EV, Gurevich VV. Arrestins are ubiquitous regulators of cellular signaling pathways. Genome Biol. 2006;7:236. doi: 10.1186/gb-2006-7-9-236. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Strissel KJ, Sokolov M, Trieu LH, Arshavsky VY. Arrestin translocation is induced at a critical threshold of visual signaling and is superstoichiometric to bleached rhodopsin. J Neurosci. 2006;26:1146–1153. doi: 10.1523/JNEUROSCI.4289-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Hanson SM, Gurevich EV, Vishnivetskiy SA, Ahmed MR, Song X, Gurevich VV. Each rhodopsin molecule binds its own arrestin. Proc Natl Acad Sci U S A. 2007;104:3125–8. doi: 10.1073/pnas.0610886104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Nikonov SS, Brown BM, Davis JA, Zuniga FI, Bragin A, Pugh EN, Jr, Craft CM. Mouse cones require an arrestin for normal inactivation of phototransduction. Neuron. 2008;59:462–74. doi: 10.1016/j.neuron.2008.06.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Gurevich VV, Gurevich EV. The structural basis of arrestin-mediated regulation of G-protein-coupled receptors. Pharmacol Ther. 2006;110:465–502. doi: 10.1016/j.pharmthera.2005.09.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Gurevich EV, Benovic JL, Gurevich VV. Arrestin2 and arrestin3 are differentially expressed in the rat brain during postnatal development. Neuroscience. 2002;109:421–436. doi: 10.1016/s0306-4522(01)00511-5. [DOI] [PubMed] [Google Scholar]
  • 7.Gurevich EV, Benovic JL, Gurevich VV. Arrestin2 expression selectively increases during neural differentiation. J Neurochem. 2004;91:1404–1416. doi: 10.1111/j.1471-4159.2004.02830.x. [DOI] [PubMed] [Google Scholar]
  • 8.Goodman OB, Jr., Krupnick JG, Santini F, Gurevich VV, Penn RB, Gagnon AW, Keen JH, Benovic JL. Beta-arrestin acts as a clathrin adaptor in endocytosis of the beta2-adrenergic receptor. Nature. 1996;383:447–50. doi: 10.1038/383447a0. [DOI] [PubMed] [Google Scholar]
  • 9.Laporte SA, Oakley RH, Zhang J, Holt JA, Ferguson s. S. G., Caron MG, Barak LS. The 2-adrenergic receptor/arrestin complex recruits the clathrin adaptor AP-2 during endocytosis. Proc Nat Acad Sci USA. 1999;96:3712–3717. doi: 10.1073/pnas.96.7.3712. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.DeWire SM, Ahn S, Lefkowitz RJ, Shenoy SK. Beta-arrestins and cell signaling. Annu Rev Physiol. 2007;69:483–510. doi: 10.1146/annurev.physiol.69.022405.154749. [DOI] [PubMed] [Google Scholar]
  • 11.Gurevich VV, Richardson RM, Kim CM, Hosey MM, Benovic JL. Binding of wild type and chimeric arrestins to the m2 muscarinic cholinergic receptor. J Biol Chem. 1993;268:16879–82. [PubMed] [Google Scholar]
  • 12.Gurevich VV. The selectivity of visual arrestin for light-activated phosphorhodopsin is controlled by multiple nonredundant mechanisms. J Biol Chem. 1998;273:15501–15506. doi: 10.1074/jbc.273.25.15501. [DOI] [PubMed] [Google Scholar]
  • 13.Gurevich VV, Dion SB, Onorato JJ, Ptasienski J, Kim CM, Sterne-Marr R, Hosey MM, Benovic JL. Arrestin interaction with G protein-coupled receptors. Direct binding studies of wild type and mutant arrestins with rhodopsin, b2-adrenergic, and m2 muscarinic cholinergic receptors. J Biol Chem. 1995;270:720–731. doi: 10.1074/jbc.270.2.720. [DOI] [PubMed] [Google Scholar]
  • 14.Celver J, Vishnivetskiy SA, Chavkin C, Gurevich VV. Conservation of the phosphate-sensitive elements in the arrestin family of proteins. J Biol Chem. 2002;277:9043–9048. doi: 10.1074/jbc.M107400200. [DOI] [PubMed] [Google Scholar]
  • 15.Kovoor A, Celver J, Abdryashitov RI, Chavkin C, Gurevich VV. Targeted construction of phosphorylation-independent b-arrestin mutants with constitutive activity in cells. J Biol Chem. 1999;274:6831–6834. doi: 10.1074/jbc.274.11.6831. [DOI] [PubMed] [Google Scholar]
  • 16.Pan L, Gurevich EV, Gurevich VV. The nature of the arrestin x receptor complex determines the ultimate fate of the internalized receptor. J Biol Chem. 2003;278:11623–11632. doi: 10.1074/jbc.M209532200. [DOI] [PubMed] [Google Scholar]
  • 17.Kohout TA, Lin FS, Perry SJ, Conner DA, Lefkowitz RJ. beta-Arrestin 1 and 2 differentially regulate heptahelical receptor signaling and trafficking. Proc Natl Acad Sci U S A. 2001;98:1601–6. doi: 10.1073/pnas.041608198. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Oakley RH, Laporte SA, Holt JA, Caron MG, Barak LS. Differential affinities of visual arrestin, barrestin1, and barrestin2 for G protein-coupled receptors delineate two major classes of receptors. J Biol Chem. 2000;275:17201–17210. doi: 10.1074/jbc.M910348199. [DOI] [PubMed] [Google Scholar]
  • 19.Granzin J, Wilden U, Choe HW, Labahn J, Krafft B, Buldt G. X-ray crystal structure of arrestin from bovine rod outer segments. Nature. 1998;391:918–921. doi: 10.1038/36147. [DOI] [PubMed] [Google Scholar]
  • 20.Hirsch JA, Schubert C, Gurevich VV, Sigler PB. The 2.8 A crystal structure of visual arrestin: a model for arrestin's regulation. Cell. 1999;97:257–269. doi: 10.1016/s0092-8674(00)80735-7. [DOI] [PubMed] [Google Scholar]
  • 21.Han M, Gurevich VV, Vishnivetskiy SA, Sigler PB, Schubert C. Crystal structure of beta-arrestin at 1.9 A: possible mechanism of receptor binding and membrane translocation. Structure. 2001;9:869–880. doi: 10.1016/s0969-2126(01)00644-x. [DOI] [PubMed] [Google Scholar]
  • 22.Kang DS, Kern RC, Puthenveedu MA, von Zastrow M, Williams JC, Benovic JL. Structure of an arrestin2-clathrin complex reveals a novel clathrin binding domain that modulates receptor trafficking. J Biol Chem. 2009;284:29860–72. doi: 10.1074/jbc.M109.023366. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Milano SK, Kim YM, Stefano FP, Benovic JL, Brenner C. Nonvisual arrestin oligomerization and cellular localization are regulated by inositol hexakisphosphate binding. J Biol Chem. 2006;281:9812–23. doi: 10.1074/jbc.M512703200. [DOI] [PubMed] [Google Scholar]
  • 24.Milano SK, Pace HC, Kim YM, Brenner C, Benovic JL. Scaffolding functions of arrestin-2 revealed by crystal structure and mutagenesis. Biochemistry. 2002;41:3321–3328. doi: 10.1021/bi015905j. [DOI] [PubMed] [Google Scholar]
  • 25.Sutton RB, Vishnivetskiy SA, Robert J, Hanson SM, Raman D, Knox BE, Kono M, Navarro J, Gurevich VV. Crystal Structure of Cone Arrestin at 2.3Å: Evolution of Receptor Specificity. J Mol Biol. 2005;354:1069–1080. doi: 10.1016/j.jmb.2005.10.023. [DOI] [PubMed] [Google Scholar]
  • 26.Vishnivetskiy SA, Hosey MM, Benovic JL, Gurevich VV. Mapping the arrestin-receptor interface: structural elements responsible for receptor specificity of arrestin proteins. J Biol Chem. 2004;279:1262–1268. doi: 10.1074/jbc.M308834200. [DOI] [PubMed] [Google Scholar]
  • 27.Vishnivetskiy SA, Hirsch JA, Velez M-G, Gurevich YV, Gurevich VV. Transition of arrestin in the active receptor-binding state requires an extended interdomain hinge. J Biol Chem. 2002;277:43961–43968. doi: 10.1074/jbc.M206951200. [DOI] [PubMed] [Google Scholar]
  • 28.Vishnivetskiy SA, Paz CL, Schubert C, Hirsch JA, Sigler PB, Gurevich VV. How does arrestin respond to the phosphorylated state of rhodopsin? J Biol Chem. 1999;274:11451–11454. doi: 10.1074/jbc.274.17.11451. [DOI] [PubMed] [Google Scholar]
  • 29.Pulvermuller A, Schroder K, Fischer T, Hofmann KP. Interactions of metarhodopsin II. Arrestin peptides compete with arrestin and transducin. J Biol Chem. 2000;275:37679–37685. doi: 10.1074/jbc.M006776200. [DOI] [PubMed] [Google Scholar]
  • 30.Hanson SM, Francis DJ, Vishnivetskiy SA, Klug CS, Gurevich VV. Visual arrestin binding to microtubules involves a distinct conformational change. J Biol Chem. 2006;281:9765–72. doi: 10.1074/jbc.M510738200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Hanson SM, Gurevich VV. The differential engagement of arrestin surface charges by the various functional forms of the receptor. J Biol Chem. 2006;281:3458–3462. doi: 10.1074/jbc.M512148200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Miller WE, McDonald PH, Cai SF, Field ME, Davis RJ, Lefkowitz RJ. Identification of a motif in the carboxyl terminus of beta -arrestin2 responsible for activation of JNK3. J Biol Chem. 2001;276:27770–27777. doi: 10.1074/jbc.M102264200. [DOI] [PubMed] [Google Scholar]
  • 33.Nagai T, Ibata K, Park ES, Kubota M, Mikoshiba K, Miyawaki A. A variant of yellow fluorescent protein with fast and efficient maturation for cell-biological applications. Nat Biotechnol. 2002;20:87–90. doi: 10.1038/nbt0102-87. [DOI] [PubMed] [Google Scholar]
  • 34.Barak LS, Ferguson SS, Zhang J, Caron MG. A beta-arrestin/green fluorescent protein biosensor for detecting G protein-coupled receptor activation. J Biol Chem. 1997;272:27497–27500. doi: 10.1074/jbc.272.44.27497. [DOI] [PubMed] [Google Scholar]
  • 35.Kuravi S, Lan TH, Barik A, Lambert NA. Third-party bioluminescence resonance energy transfer indicates constitutive association of membrane proteins: application to class a g-protein-coupled receptors and g-proteins. Biophys J. 2010;98:2391–9. doi: 10.1016/j.bpj.2010.02.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Namkung Y, Dipace C, Urizar E, Javitch JA, Sibley DR. G Protein-coupled Receptor Kinase-2 Constitutively Regulates D2 Dopamine Receptor Expression and Signaling Independently of Receptor Phosphorylation. J Biol Chem. 2009;284:34103–34115. doi: 10.1074/jbc.M109.055707. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Namkung Y, Dipace C, Javitch JA, Sibley DR. G Protein-coupled Receptor Kinase-mediated Phosphorylation Regulates Post-endocytic Trafficking of the D2 Dopamine Receptor. J Biol Chem. 2009;284:15038–15051. doi: 10.1074/jbc.M900388200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Schubert C, Hirsch JA, Gurevich VV, Engelman DM, Sigler PB, Fleming KG. Visual arrestin activity may be regulated by self-association. J Biol Chem. 1999;274:21186–21190. doi: 10.1074/jbc.274.30.21186. [DOI] [PubMed] [Google Scholar]
  • 39.Imamoto Y, Tamura C, Kamikubo H, Kataoka M. Concentration-dependent tetramerization of bovine visual arrestin. Biophys J. 2003;85:1186–1195. doi: 10.1016/S0006-3495(03)74554-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Hanson SM, Vishnivetskiy SA, Hubbell WL, Gurevich VV. Opposing effects of inositol hexakisphosphate on rod arrestin and arrestin2 self-association. Biochemistry. 2008;47:1070–5. doi: 10.1021/bi7021359. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Ponstingl H, Henrick K, Thornton JM. Discriminating between homodimeric and monomeric proteins in the crystalline state. Proteins. 2000;41:47–57. doi: 10.1002/1097-0134(20001001)41:1<47::aid-prot80>3.3.co;2-#. [DOI] [PubMed] [Google Scholar]
  • 42.Shi H, Rojas R, Bonifacino JS, Hurley JH. The retromer subunit Vps26 has an arrestin fold and binds Vps35 through its C-terminal domain. Nat Struct Mol Biol. 2006;13:540–8. doi: 10.1038/nsmb1103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Collins BM, Norwood SJ, Kerr MC, Mahony D, Seaman MN, Teasdale RD, Owen DJ. Structure of Vps26B and mapping of its interaction with the retromer protein complex. Traffic. 2008;9:366–79. doi: 10.1111/j.1600-0854.2007.00688.x. [DOI] [PubMed] [Google Scholar]
  • 44.Attramadal H, Arriza JL, Aoki C, Dawson TM, Codina J, Kwatra MM, Snyder SH, Caron MG, Lefkowitz RJ. Beta-arrestin2, a novel member of the arrestin/beta-arrestin gene family. J Biol Chem. 1992;267:17882–17890. [PubMed] [Google Scholar]
  • 45.Sterne-Marr R, Gurevich VV, Goldsmith P, Bodine RC, Sanders C, Donoso LA, Benovic JL. Polypeptide variants of beta-arrestin and arrestin3. J Biol Chem. 1993;268:15640–15648. [PubMed] [Google Scholar]
  • 46.Barnes WG, Reiter E, Violin JD, Ren XR, Milligan G, Lefkowitz RJ. beta-Arrestin 1 and Galphaq/11 coordinately activate RhoA and stress fiber formation following receptor stimulation. J Biol Chem. 2005;280:8041–8050. doi: 10.1074/jbc.M412924200. [DOI] [PubMed] [Google Scholar]
  • 47.Povsic TJ, Kohout TA, Lefkowitz RJ. Beta-arrestin1 mediates insulin-like growth factor 1 (IGF-1) activation of phosphatidylinositol 3-kinase (PI3K) and anti-apoptosis. J Biol Chem. 2003;278:51334–9. doi: 10.1074/jbc.M309968200. [DOI] [PubMed] [Google Scholar]
  • 48.Nobles KN, Guan Z, Xiao K, Oas TG, Lefkowitz RJ. The active conformation of beta-arrestin1: direct evidence for the phosphate sensor in the N-domain and conformational differences in the active states of beta-arrestins1 and -2. J Biol Chem. 2007;282:21370–81. doi: 10.1074/jbc.M611483200. [DOI] [PubMed] [Google Scholar]
  • 49.Schumann M, Nakagawa T, Mantey SA, Howell B, Jensen RT. Function of non-visual arrestins in signaling and endocytosis of the gastrin-releasing peptide receptor (GRP receptor) Biochem Pharmacol. 2008;75:1170–85. doi: 10.1016/j.bcp.2007.11.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Gurevich VV, Gurevich EV. The molecular acrobatics of arrestin activation. Trends Pharmacol Sci. 2004;25:59–112. doi: 10.1016/j.tips.2003.12.008. [DOI] [PubMed] [Google Scholar]
  • 51.Vishnivetskiy SA, Schubert C, Climaco GC, Gurevich YV, Velez M-G, Gurevich VV. An additional phosphate-binding element in arrestin molecule: implications for the mechanism of arrestin activation. J Biol Chem. 2000;275:41049–41057. doi: 10.1074/jbc.M007159200. [DOI] [PubMed] [Google Scholar]
  • 52.Song X, Coffa S, Fu H, Gurevich VV. How does arrestin assemble MAP kinases into a signaling complex? J Biol Chem. 2009;284:685–95. doi: 10.1074/jbc.M806124200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Song X, Gurevich EV, Gurevich VV. Cone arrestin binding to JNK3 and Mdm2: conformational preference and localization of interaction sites. J Neurochem. 2007;103:1053–62. doi: 10.1111/j.1471-4159.2007.04842.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Song X, Raman D, Gurevich EV, Vishnivetskiy SA, Gurevich VV. Visual and both non-visual arrestins in their "inactive" conformation bind JNK3 and Mdm2 and relocalize them from the nucleus to the cytoplasm. J Biol Chem. 2006;281:21491–9. doi: 10.1074/jbc.M603659200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.McDonald PH, Chow CW, Miller WE, Laporte SA, Field ME, Lin FT, Davis RJ, Lefkowitz RJ. Beta-arrestin 2: a receptor-regulated MAPK scaffold for the activation of JNK3. Science. 2000;290:1515–1518. doi: 10.1126/science.290.5496.1574. [DOI] [PubMed] [Google Scholar]
  • 56.Gurevich VV, Gurevich EV. Custom-designed proteins as novel therapeutic tools? The case of arrestins. Expert Rev Mol Med. 2010;12:e13. doi: 10.1017/S1462399410001444. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Donoso LA, Gregerson DS, Smith L, Robertson S, Knospe V, Vrabec T, Kalsow CM. S-antigen: preparation and characterization of site-specific monoclonal antibodies. Curr Eye Res. 1990;9:343–55. doi: 10.3109/02713689008999622. [DOI] [PubMed] [Google Scholar]
  • 58.Hanson SM, Van Eps N, Francis DJ, Altenbach C, Vishnivetskiy SA, Arshavsky VY, Klug CS, Hubbell WL, Gurevich VV. Structure and function of the visual arrestin oligomer. EMBO J. 2007;26:1726–36. doi: 10.1038/sj.emboj.7601614. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Hanson SM, Dawson ES, Francis DJ, Van Eps N, Klug CS, Hubbell WL, Meiler J, Gurevich VV. A Model for the Solution Structure of the Rod Arrestin Tetramer. Structure. 2008;16:924–34. doi: 10.1016/j.str.2008.03.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Nair KS, Hanson SM, Kennedy MJ, Hurley JB, Gurevich VV, Slepak VZ. Direct binding of visual arrestin to microtubules determines the differential subcellular localization of its splice variants in rod photoreceptors. J Biol Chem. 2004;279:41240–8. doi: 10.1074/jbc.M406768200. [DOI] [PubMed] [Google Scholar]
  • 61.Nair KS, Hanson SM, Mendez A, Gurevich EV, Kennedy MJ, Shestopalov VI, Vishnivetskiy SA, Chen J, Hurley JB, Gurevich VV, Slepak VZ. Light-dependent redistribution of arrestin in vertebrate rods is an energy-independent process governed by protein-protein interactions. Neuron. 2005;46 doi: 10.1016/j.neuron.2005.03.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Storez H, Scott MG, Issafras H, Burtey A, Benmerah A, Muntaner O, Piolot T, Tramier M, Coppey-Moisan M, Bouvier M, Labbé-Jullié C, Marullo S. Homo- and hetero-oligomerization of beta-arrestins in living cells. J Biol Chem. 2005;280:40210–5. doi: 10.1074/jbc.M508001200. [DOI] [PubMed] [Google Scholar]
  • 63.DeFea KA. Beta-arrestin multimers: does a crowd help or hinder function? Biochem J. 2008;413:e1–3. doi: 10.1042/BJ20081009. [DOI] [PubMed] [Google Scholar]
  • 64.Gurevich VV, Benovic JL. Arrestin: mutagenesis, expression, purification, and functional characterization. Methods Enzymol. 2000;315:422–37. doi: 10.1016/s0076-6879(00)15859-8. [DOI] [PubMed] [Google Scholar]
  • 65.Otwinowski Z, M. W. Processing of X-ray Diffraction Data Collected in Oscillation Mode. Methods in Enzymology. 1997;276A:307–326. doi: 10.1016/S0076-6879(97)76066-X. [DOI] [PubMed] [Google Scholar]
  • 66.McCoy AJ, Grosse-Kunstleve RW, Storoni LC, Read RJ. Likelihood-enhanced fast translation functions. Acta Crystallogr D Biol Crystallogr. 2005;61:458–64. doi: 10.1107/S0907444905001617. [DOI] [PubMed] [Google Scholar]
  • 67.Stein N. CHAINSAW: a program for mutating pdb files used as templates in molecular replacement. J. Appl. Cryst. 2008;41:641–643. [Google Scholar]
  • 68.Bailey S. The CCP4 Suite - Programs for Protein Crystallography. Acta Crystallographica Section D-Biological Crystallography. 1994;50:760–763. doi: 10.1107/S0907444994003112. [DOI] [PubMed] [Google Scholar]
  • 69.Emsley P, Cowtan K. Coot: model-building tools for molecular graphics. Acta Crystallogr D Biol Crystallogr. 2004;60:2126–32. doi: 10.1107/S0907444904019158. [DOI] [PubMed] [Google Scholar]
  • 70.Brünger AT, Adams PD, Rice LM. New applications of simulated annealing in X-ray crystallography and solution NMR. Structure. 1997;5:325–36. doi: 10.1016/s0969-2126(97)00190-1. [DOI] [PubMed] [Google Scholar]
  • 71.Brünger AT, Adams PD, Clore GM, DeLano WL, Gros P, Grosse-Kunstleve RW, Jiang JS, Kuszewski J, Nilges M, Pannu NS, Read RJ, Rice LM, Simonson T, Warren GL. Crystallography & NMR system: A new software suite for macromolecular structure determination. Acta Crystallogr. D. 1998;54:905–921. doi: 10.1107/s0907444998003254. [DOI] [PubMed] [Google Scholar]
  • 72.Adams PD, Grosse-Kunstleve RW, Hung LW, Ioerger TR, McCoy AJ, Moriarty NW, Read RJ, Sacchettini JC, Sauter NK, Terwilliger TC. PHENIX: building new software for automated crystallographic structure determination. Acta Crystallogr D Biol Crystallogr. 2002;58:1948–54. doi: 10.1107/s0907444902016657. [DOI] [PubMed] [Google Scholar]
  • 73.Painter J, Merritt EA. Optimal description of a protein structure in terms of multiple groups undergoing TLS motion. Acta Crystallogr D Biol Crystallogr. 2006;62:439–50. doi: 10.1107/S0907444906005270. [DOI] [PubMed] [Google Scholar]
  • 74.Painter J, Merritt EA. TLSMD web server for the generation of multi-group TLS models. Journal of Applied Crystallography. 2006;39:109–111. [Google Scholar]
  • 75.Pannu NS, Read RJ. Improved structure refinement through maximum likelihood. Acta Cryst A. 1996;52:659–668. [Google Scholar]
  • 76.Luzzati VP. Resolution d'une Structure Crisalline Lorsque les Positions d'une Partie des Atomes sont Connues: Tratement Statistique. Acta Cryst. 1953;6 [Google Scholar]
  • 77.Read RJ. Improved Fourier coefficients for maps using phases from partial structures with errors. Acta Cryst. 1986;A42:140–144. [Google Scholar]
  • 78.Adams PD, Afonine PV, Grosse-Kunstleve RW, Read RJ, Richardson JS, Richardson DC, Terwilliger TC. Recent developments in phasing and structure refinement for macromolecular crystallography. Curr Opin Struct Biol. 2009;19:566–72. doi: 10.1016/j.sbi.2009.07.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Krissinel E, Henrick K. Inference of macromolecular assemblies from crystalline state. J Mol Biol. 2007;372:774–97. doi: 10.1016/j.jmb.2007.05.022. [DOI] [PubMed] [Google Scholar]
  • 80.Theobald DL, Wuttke DS. THESEUS: maximum likelihood superpositioning and analysis of macromolecular structures. Bioinformatics. 2006;22:2171–2. doi: 10.1093/bioinformatics/btl332. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.DeLano WL. The PyMOL Molecular Graphics System. DeLano Scientific; San Carlos, CA: 2002. [Google Scholar]
  • 82.Gurevich VV. Use of bacteriophage RNA polymerase in RNA synthesis. Methods Enzymol. 1996;275:382–97. doi: 10.1016/s0076-6879(96)75023-1. [DOI] [PubMed] [Google Scholar]
  • 83.Loening AM, Fenn TD, Wu AM, Gambhir SS. Consensus guided mutagenesis of Renilla luciferase yields enhanced stability and light output. Protein Eng Des Sel. 2006;19:391–400. doi: 10.1093/protein/gzl023. [DOI] [PubMed] [Google Scholar]
  • 84.Kocan M, Pfleger KD. Detection of GPCR/beta-arrestin interactions in live cells using bioluminescence resonance energy transfer technology. Methods Mol Biol. 2009;552:305–17. doi: 10.1007/978-1-60327-317-6_22. [DOI] [PubMed] [Google Scholar]

RESOURCES