Minireview

10 downloads 110 Views 99KB Size Report
is no clear homology to α-helix 6 in the oncovirus INs. Instead, there is a proline-rich region that would not favor helix formation. (44). Sequences from this hinge ...
THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 271, No. 33, Issue of August 16, pp. 19633–19636, 1996 © 1996 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A.

Minireview Retroviral Integrase, Putting the Pieces Together* Mark D. Andrake‡ and Anna Marie Skalka§ From the Fox Chase Cancer Center, Institute for Cancer Research, Philadelphia, Pennsylvania 19111

* This minireview will be reprinted in the 1996 Minireview Compendium, which will be available in December, 1996. Work from this laboratory was supported by National Institutes of Health Grants CA-47486 and CA-06927, a grant for infectious disease research from Bristol-Myers Squibb Foundation, and also by an appropriation from the Commonwealth of Pennsylvania. ‡ Supported, in part, by United States Public Health Service Fellowship 5F32 AI08642 and Grant T32 CA09035–17. § To whom correspondence should be addressed: Fox Chase Cancer Center, Inst. for Cancer Research, 7701 Burholme Ave., Philadelphia, PA 19111. Tel.: 215-728-2490; Fax: 215-728-2778. 1 The abbreviations used are: IN, integrase; HIV-1, human immunodeficiency virus type 1; ASV, avian sarcoma virus; RAG, recombination activation gene; SH3, Src homology region 3; MuA, bacteriophage Mu transposase A.

N-terminal Domain The N-terminal domain, comprising approximately the first 50 amino acids, is characterized by a “zinc finger”-like motif (HHCC domain). However, the length of the loop between the histidines and cysteines is longer than that seen in the canonical, DNAbinding zinc fingers of several transcription factors. This motif is highly conserved in the integrases of all retroviruses and eukaryotic retrotransposons. When expressed independently, this N-terminal domain does not bind DNA (15, 16), but it may interact with DNA in the context of the intact protein (15, 17–19). The isolated domain has been shown to bind Zn21 and to assume an ordered structure upon metal binding, as expected for a zinc finger (20). Spectroscopic studies reveal that the isolated peptide coordinates metal with a tetrahedral geometry and that mutation of the conserved histidines or cysteines abolishes formation of the metalinduced ordered structure. Recent studies applying atomic absorption spectroscopy to various HIV-1 IN fragments confirm the 1:1 binding of Zn21 to this HHCC region.2 Related Zn21 binding experiments also suggest a role for this region, together with the C terminus, in promoting tetramerization.2 This agrees with previously reported evidence that the N and C termini of HIV-1 IN are likely to be in close proximity in the native structure (21). Other studies have suggested a role for this domain in determining the aggregation properties of HIV-1 IN (22). Deletion of the zinc finger motif has little or no effect on the processing and joining activities of ASV IN as measured in vitro using assays that detect single-end cleavage and joining events (23, 24). However, the high conservation of this motif and results from genetic experiments suggest that this domain is functionally important. Further biochemical studies that explore the role of this domain in coordinated cleavage and joining events may reveal its participation in interdomain or intersubunit interactions that occur in vivo. Catalytic Core Domain The catalytic domain, included in a central region of approximately 150 amino acids, is characterized by the D,D(35)E motif, a constellation of three invariant acidic amino acids, the last two separated by 35 amino acids in integrase proteins. These acidic residues are required for all catalytic functions of IN and have been proposed to bind the essential metal cofactor(s), Mn21 or Mg21 (15, 25). The recently solved crystal structures of the catalytic domains of HIV-1 IN (26) and ASV IN (27) support this view. These structures reveal that IN proteins are members of a large superfamily of nucleases and polynucleotidyltransferases (reviewed in Refs. 28 – 30) that includes RNases H from Escherichia coli and HIV-1 (31, 32), the E. coli Ruv C resolvase (33), and bacteriophage Mu A transposase (34). Catalytic and structural similarities to the type II restriction endonucleases have also been noted (35). Fig. 3 (top) shows ribbon models of monomers of the catalytic domains of HIV-1 and ASV. The ASV model includes a bound divalent cation (Mg21 or Mn21) observed after soaking crystals in the appropriate solutions (36). The overall topology of the two domains is quite similar (root mean square deviation of 1.4 Å for 107 Ca pairs) and resembles that of other members of the superfamily. The characteristic features are: (a) a prominent mixed 2

19633

R. Zheng and R. Craigie, unpublished data.

Downloaded from http://www.jbc.org/ by guest on February 13, 2017

Retroviral integrase (IN)1 mediates retroviral DNA integration, a critical step in viral replication that ensures stable expression of proviral genes in the infected cell and perpetuation of the viral genome in all the host cell progeny. The IN protein is both necessary and sufficient for the integration of a linear DNA with viral end sequences into a target DNA in vitro (1, 2). The integration reaction is known to take place in two distinct steps (see Fig. 1). IN specifically recognizes sequences at both ends of newly synthesized viral DNA, most likely as a component of a large subviral preintegration complex. The first step in integration, a processing reaction, can take place in the cytoplasm of infected cells (3). This reaction produces site-specific cuts near the viral DNA 39-ends, adjacent to a conserved CA dinucleotide, removing (generally) two nucleotides and exposing new 39-hydroxyl ends. The second step, a joining reaction, is a concerted cleavage-ligation reaction (4), which produces a staggered cut in cellular DNA when the newly exposed 39-hydroxyls of the viral DNA ends attack the phosphate bonds at the cellular DNA cleavage site. The product is an intermediate in which the 39-ends of viral DNA are covalently linked to cellular DNA and the 59-ends of viral DNA are flanked by short gaps (5, 6). Repair of the gaps and completion of integration, which can be accomplished by cellular enzymes, produce a short direct repeat of host target DNA. The length of this repeat is characteristic for each virus. For example, HIV-1 proviruses are flanked by 5-base pair repeats and ASV by 6-base pair repeats. These features are likely to reflect subtle differences in the structure or multimeric organization of the two integrases. Details of the biochemistry of IN and its catalytic mechanism have been uncovered mainly through the use of in vitro, reconstructed systems, which employ purified enzymes and model DNA substrates that consist of short oligodeoxynucleotide duplexes (1, 2, 7, 8). Unlike several site-specific recombinases that utilize a protein-DNA covalent intermediate, the cleavage-ligation reaction is a direct transesterification. The in-line nucleophilic attack by the processed viral 39-hydroxyls on the phosphate bond of the target DNA occurs with chiral inversion when appropriate substrates are used for detection (4). It has recently been discovered that a similar, but intramolecular, transesterification is catalyzed by another eukaryotic recombinase complex (RAG1 and RAG2) in the early steps of V(D)J recombination of immunoglobulin genes (9). Retroviral integrases range from 270 to 350 amino acids in length. Various lines of evidence indicate that both ASV and HIV-1 IN function as multimers (minimally, dimers) in vitro (10 –12). However, the number and arrangement of IN protomers in the active in vivo complex are still unknown. Deduced amino acid sequence alignments, limited proteolysis, site-directed mutagene-

sis studies, and complementation experiments (reviewed in Refs. 13 and 14) have revealed the presence of three distinct domains that form independent folding units within each monomer. The first two of these are highly conserved among retroviral and retrotransposon integrases (Fig. 2). Here we will review what is known about the structure of each domain and discuss current ideas concerning domain interactions and multimerization, as they relate to function.

19634

Minireview: Retroviral Integrases

b-sheet (five-stranded in the integrases) with three long anti-parallel b-strands (b1, b2, and b3) that are flanked on either side by a-helices and (b) a juxtaposition of the essential acidic residues that are presumed to form the catalytic center. Because of the differences in the metal complexes of E. coli and HIV-1 RNase H (31, 32, 37), it is still uncertain whether one or two metal ions are required for catalysis by the members of this superfamily. In the original proposal of the two-metal mechanism (38), one metal plays a role in promoting the formation of the attacking hydroxide ion, and the second functions to stabilize the pentacoordinate transition state at the phosphate to be cleaved. Observation of a single metal in the ASV IN catalytic core structure does not preclude a two-metal mechanism, because Asp-64 in the first b-strand could contribute to the coordination of two metals, in conjunction with the other acidic residues in the full-length protein, as observed in other members of this superfamily (31, 38). The binding of the second metal might also require DNA substrate or other domains of IN. The conservation of active site residue positions within a common folding topology presumably indicates a common catalytic mechanism for the members of this superfamily. Structural and biochemical data for full-length or isolated IN domains complexed to various substrates may be necessary to determine the number of metals and their specific roles in the reaction. Despite the overall similarity between the two available retroviral IN catalytic core structures, certain differences are apparent. The HIV-1 IN structure contains an a-helix (a-6) not observed in the ASV structure (see also alignment in Fig. 2). The ASV structure has an ordered active site in which all three of the invariant residues of the D,D(35)E motif are resolved (Fig. 3, top), whereas the published HIV-1 model has a disordered segment, which includes the third critical residue of this motif, Glu-152. Recent data from a different crystal form of the HIV-1 IN catalytic core domain3 reveal more of a-helix 4, including the third catalytic residue Glu152 (pictured in Fig. 3, top left panel). However, 10 amino acids between b-strand 5 and a-helix 4 remain disordered (dotted line in 3

F. Dyda, R. Craigie, and D. R. Davies, unpublished data.

FIG. 2. Structural and sequence alignment of ASV and HIV-1 IN. The HIV-1 and ASV IN sequences (denoted H and A, respectively) were aligned based on known structural information and sequence similarity. Sequence identity is noted by the vertical dash and conservative homology by a colon. The conserved His and Cys residues of the HHCC domain and the conserved acidic residues of the catalytic core domain are boxed in red. The termini of fragments used in crystallography or NMR studies are denoted by blue dashed brackets with arrowheads. The HIV-1 core fragment used for crystallization comprised amino acids 50 –212 and contained a F185K substitution. The ASV core fragment used for crystallization comprised amino acids 52–207. The question marks above the 10-amino acid segment denote the flexible loop that is disordered in the HIV-1 IN crystal. Sequences were obtained from Refs. 27 and 48 and data obtained in our laboratory for the ASV strain Schmidt-Ruppin B. The locations of various secondary structure elements (b-strands and a-helices) are noted above the HIV-1 sequence and below the ASV sequence. The colors correspond to those in the ribbon models of Fig. 3.

Fig. 3, top left panel). Flexibility is likely to be a hallmark of this loop in the isolated core domain; flexibility is also reported for the analogous portion of the ASV IN domain (27). Differences in the direction and orientation of the two catalytic Asp residues in the two structures are also observed. Several explanations for the latter difference are possible (36); however, a detailed description of the active site configuration will likely require analysis of complexes with DNA substrates. Both the HIV-1 and the ASV IN catalytic cores crystallize as dimers (see Fig. 3, middle). The primary contacts between subunits in both structures involve a-helices 1 and 5. Because the same extensive interface is seen in both IN structures, it is likely that core domain dimerization plays an important role in the multimerization required for the processing and joining activities of integrase. As might be expected from the contribution of an additional a-helix mentioned above (a-helix 6), the solvent-accessible surface area buried upon dimerization is more extensive in the HIV-1 IN dimer interface (1300 Å versus 766 Å for ASV). Correspondingly, the binding interactions between subunits are more numerous in the HIV-1 interface. These differences provide an explanation for the higher apparent association constant of the HIV-1 core dimer relative to ASV (39, 40). In each dimer, comparison of the distance between the active site residues Asp-64 (Fig. 3, middle) shows a difference of '3 Å, a value approximately equal to the length of one nucleotide in B-form DNA.

Downloaded from http://www.jbc.org/ by guest on February 13, 2017

FIG. 1. Overview of the two-step integration reaction. The conserved CA dinucleotide is shown at the ends of viral DNA, which is represented as a thin line circle. The circled P symbolizes the phosphate bond that is cleaved (position noted by red triangles) by IN releasing two nucleotides (NN). These new 39-hydroxyl ends (denoted CAOH) are the attacking nucleophiles that cleave both strands of the target DNA in the joining reaction. The number of base pairs separating the staggered cuts is characteristic for each virus. Each strand of the target DNA is shown with a different color to help visualize the reaction. Small black spheres denote the 59-end of all DNA strands. The gapped intermediate is subsequently repaired and yields the integrated provirus with the host target site duplication. Yellow rounded squares represent IN protomers arranged as dimers (top) or as a tetramer (middle) in analogy with certain transposases.

Minireview: Retroviral Integrases

19635

FIG. 3. Ribbon drawings of IN domain structures. Ribbon models of the structure of two IN domains are shown. The top two panels compare the HIV-1 and ASV IN catalytic core monomers. Corresponding b-strands and a-helices are labeled and color coded to match the alignment in Fig. 2. Side chains of conserved acidic residues of the active site are shown in an enlarged ball and stick representation (yellow). The metal bound in the ASV structure is shown in green near the two active site Asp residues. The dotted line in the HIV-1 IN monomer structure denotes the disordered region of the flexible loop. The middle two panels are ribbon diagrams of the catalytic core dimers. The orientation displayed is rotated 90° for both the x and y axis relative to the monomer panels on top. Labels use a prime suffix to denote a second subunit. The separation between Asp-64 residues is noted at the bottom of each panel. Coordinates from the new HIV-1 IN core domain crystal form3 were used to generate the top left panel; middle panels were created using published coordinates (accession codes 1ITG and 1VSF in the Brookhaven Protein Data Bank). The bottom panel shows two ribbon representations of the HIV-1 IN C-terminal domain dimer derived from the NMR data (accession code 1IHW in the Brookhaven Protein Data Bank). The left orientation (Interface) shows hydrophobic intersubunit contacts in this interface with relevant amino acid side chains in the ball and stick representation. The right orientation (Groove) shows the same dimer rotated 90° about the y axis and shows the position of several basic residues that extend into the saddle-like grove proposed to bind DNA. This figure was prepared using MOLSCRIPT (59).

This is consistent with the one-nucleotide difference in the target site duplication generated by these two enzymes during integration. However, as noted by Dyda et al. (26), the overall distance of '35 Å is too great to accommodate the 5 base pairs of B-DNA that separate the two HIV-1 IN cleavage sites in the target DNA. This difference could be accommodated if one allows for partial DNA unwinding during the reaction and coordinated but sequential cleavage at the target site. Independent lines of evidence support such a model (41, 42). Alternatively, two active sites could be positioned closer in a higher order multimer (26). A similar disparity between active site separation and the distance between two DNA scissile bonds also exists in the co-crystal of gd resolvase and a model DNA substrate (43). Partial distortion of both protein and DNA, and sequential single-strand cleavage in the target DNA, have been proposed to reconcile the gd structural data with the known biochemical properties of the reaction. C-terminal Domain The amino acid sequence of the C-terminal domain of approximately 80 –100 residues is not highly conserved among IN proteins from different families of retroviruses (see Fig. 2, bottom). Moloney murine leukemia virus IN contains a sequence insertion of unknown function in this region, and other differences in the connec-

Connecting IN Domains in the Full-length Monomer and in the Integration Complex with DNA Juxtaposition of Domains—The study of retroviral INs has benefited greatly from structural information derived from these isolated domains. However, regions that connect the three domains may also perform important functions, some of which could be specific to different retroviral IN proteins. For example, significant sequence differences are observed between oncoretrovirus and lentivirus proteins in the “hinge region” that connects the catalytic core and the C-terminal, b-barrel domain. The glycine-rich loop between a-5 and a-6 is conserved only in the lentiviruses, and there is no clear homology to a-helix 6 in the oncovirus INs. Instead, there is a proline-rich region that would not favor helix formation (44). Sequences from this hinge region in ASV IN (but not HIV-1 IN) confer nuclear localization on a heterologous reporter protein.4 This may represent a mechanism for nuclear import that is not common to all retroviral integrases. The MuA transposase catalytic core structure (34) possesses a downstream domain that may be functionally analogous to the C-terminal domain of IN. It too forms a b-barrel known to include determinants for nonspecific DNA binding. However, analysis of the MuA structure is unlikely to provide insight into specific contacts between the IN core and C-terminal domains. The MuA b-barrel is a 6-stranded structure with a topology different from that of the 5-stranded b-barrel of HIV-1 IN. Moreover, the contacts between these two MuA domains seem to be unique to this protein (34), as they involve long loops (between a-helix 2 and 3, and between catalytic core b-strands 1 and 2) not found in the retroviral integrases. Correspondingly, the MuA b-barrel does not form a dimeric interface as seen in the HIV-1 IN C-terminal b-barrel structure. Still undetermined is how the three IN domains interact with each other in the multimer-DNA complex that performs the coordinated integration reaction. While it is clear that the isolated catalytic core and C-terminal domains each can dimerize, these two interfaces are not likely to be the only determinants of multimerization. As noted above, there is biochemical evidence for interaction between the N- and C-terminal domains of HIV-1 IN. Furthermore, different IN domains exhibit trans-complementation (11, 12). As is the case with the MuA transposase (Ref. 51 and the references therein), it seems likely that a given domain of one IN monomer will interact with a different domain of a second monomer in the synaptic complex with DNA. Elucidation of such interactions is an important challenge for the future. 4

G. Kukolj and A. Skalka, unpublished data.

Downloaded from http://www.jbc.org/ by guest on February 13, 2017

tion of this domain to the catalytic core may exist between lentiviruses and oncoretroviruses (44). Mutation and deletion analyses with both ASV and HIV-1 IN indicate that the C-terminal domain contains nonspecific DNA binding activity (15, 16, 45– 47). The structure of an isolated fragment from this HIV-1 domain (amino acids 220 –270) was recently determined by NMR (48, 49). It reveals an SH3-like fold (see Fig. 3, bottom) similar to that found in signal transduction proteins (50). The significance of this feature is not yet apparent; the SH3 domain may be a ubiquitous folding unit involved in a variety of binding functions including peptide, protein, and DNA ligands. The NMR data show that the isolated SH3-like domain of IN exists as a dimer in solution (48, 49). Biochemical studies have also revealed that multimerization determinants reside in the C-terminal regions of both ASV (39) and HIV-1 IN (40). The binding interactions at the interface in the NMR structure are predominantly hydrophobic and localize to b-strands 2, 3, and 4 (purple strands), although there is a possible hydrogen bond between Gln-252 and Glu-246 of the other subunit (Fig. 3, bottom left panel). Lodi et al. (49) suggest that a DNAbinding site may be formed by the dimeric structure of this domain where the loops between b-1 and b-2 strands form a saddle-like groove that could accommodate a duplex DNA molecule. This groove contains basic residues favorably positioned for contact with DNA (see Fig. 3, bottom right panel). In support of this model, substitution of Lys-264 (to Glu) significantly reduces the DNA binding capacity of an isolated fragment from this domain of HIV-1 IN (47).

19636

Minireview: Retroviral Integrases

Perspectives The information on retroviral integrase structure acquired during the last few years has advanced the study of these enzymes significantly. However, it has also highlighted new, important questions concerning the relative orientation of IN domains, IN multimerization, and the positioning of substrate DNAs within the synaptic complex. The current structural data provide models that will aid in the rational design of active site inhibitors of these enzymes that may be used as antiviral drugs. Domain or subunit interactions represent another target that could provide new therapeutic agents. Still largely unexplored is the specific nature and relative importance of protein-protein interactions that involve IN and other viral (56, 57) and cellular proteins (58). Their study is sure to advance our understanding of the molecular biology of this important class of viruses and of the cells in which they propagate. Acknowledgments—We thank F. Dyda, R. Craigie, D. Davies, R. Zheng, M. Clore, and G. Kukolj for sharing data prior to publication. REFERENCES 1. Craigie, R., Fujiwara, T., and Bushman, F. (1990) Cell 62, 829 – 837 2. Katz, R. A., Merkel, G., Kulkosky, J., Leis, J., and Skalka, A. M. (1990) Cell 63, 87–95 3. Roth, M. J., Schwartzberg, P. L., and Goff, S. P. (1989) Cell 58, 47–54 4. Engelman, A., Mizuuchi, K., and Craigie, R. (1991) Cell 67, 1211–1221 5. Fujiwara, T., and Mizuuchi, K. (1988) Cell 54, 497–504 6. Brown, P. O., Bowerman, B., Varmus, H. E., and Bishop, J. M. (1989) Proc. Natl. Acad. Sci. U. S. A. 86, 2525–2529 7. Katzman, M., Katz, R. A., Skalka, A. M., and Leis, J. (1989) J. Virol. 63, 5319 –5327 8. Chow, S. A., Vincent, K. A., Ellison, V., and Brown, P. O. (1992) Science 255, 723–726

9. van Gent, D. C., Mizuuchi, K., and Gellert, M. (1996) Science 271, 1592–1594 10. Jones, K. S., Coleman, J., Merkel, G. W., Laue, T. M., and Skalka, A. M. (1992) J. Biol. Chem. 267, 16037–16040 11. Engelman, A., Bushman, F. D., and Craigie, R. (1993) EMBO J. 12, 3269 –3275 12. van Gent, D. C., Vink, C., Groeneger, A. A., and Plasterk, R. H. (1993) EMBO J. 12, 3261–3267 13. Vink, C., and Plasterk, R. H. (1993) Trends Genet. 9, 433– 438 14. Katz, R. A., and Skalka, A. M. (1994) Annu. Rev. Biochem. 63, 133–173 15. Khan, E., Mack, J. P., Katz, R. A., Kulkosky, J., and Skalka, A. M. (1991) Nucleic Acids Res. 19, 851– 860 16. Mumm, S. R., and Grandgenett, D. P. (1991) J. Virol. 65, 1160 –1167 17. Jonsson, C. B., and Roth, M. J. (1993) J. Virol. 67, 5562–5571 18. Vincent, K. A., Ellison, V., Chow, S. A., and Brown, P. O. (1993) J. Virol. 67, 425– 437 19. Hazuda, D. J., Wolfe, A. L., Hastings, J. C., Robbins, H. L., Graham, P. L., LaFemina, R. L., and Emini, E. A. (1994) J. Biol. Chem. 269, 3999 – 4004 20. Burke, C. J., Sanyal, G., Bruner, M. W., Ryan, J. A., LaFemina, R. L., Robbins, H. L., Zeft, A. S., Middaugh, C. R., and Cordingley, M. G. (1992) J. Biol. Chem. 267, 9639 –9644 21. Bizub, B. D., Kulkosky, J., and Skalka, A. M. (1994) AIDS Res. Hum. Retroviruses 10, 1105–1115 22. Ellison, V., Gerton, J., Vincent, K. A., and Brown, P. O. (1995) J. Biol. Chem. 270, 3320 –3326 23. Bushman, F. D., and Wang, B. (1994) J. Virol. 68, 2215–2223 24. Katz, R. A., Merkel, G., and Skalka, A. M. (1996) Virology 217, 178 –190 25. Kulkosky, J., Jones, K. S., Katz, R. A., Mack, J. P., and Skalka, A. M. (1992) Mol. Cell. Biol. 12, 2331–2338 26. Dyda, F., Hickman, A. B., Jenkins, T. M., Engelman, A., Craigie, R., and Davies, D. R. (1994) Science 266, 1981–1986 27. Bujacz, G., Jasko´lski, M., Alexandratos, J., Wlodawer, A., Merkel, G., Katz, R. A., and Skalka, A. M. (1995) J. Mol. Biol. 253, 333–346 28. Grindley, N. D., and Leschziner, A. E. (1995) Cell 83, 1063–1066 29. Yang, W., and Steitz, T. A. (1995) Structure 3, 131–134 30. Rice, P., Craigie, R., and Davies, D. R. (1996) Curr. Opin. Struct. Biol. 6, 76 – 83 31. Davies, J. F., Hostomska, Z., Hostomsky, Z., Jordan, S. R., and Matthews, D. A. (1991) Science 252, 88 –95 32. Katayanagi, K., Okumura, M., and Morikawa, K. (1993) Proteins 17, 337–346 33. Ariyoshi, M., Vassylyev, D. G., Iwasaki, H., Nakamura, H., Shinagawa, H., and Morikawa, K. (1994) Cell 78, 1063–1072 34. Rice, P., and Mizuuchi, K. (1995) Cell 82, 209 –220 35. Venclovas, C., and Siksnys, V. (1995) Nat. Struct. Biol. 2, 838 – 841 36. Bujacz, G., Jasko´lski, M., Alexandratos, J., Wlodawer, A., Merkel, G., Katz, R. A., and Skalka, A. M. (1996) Structure 4, 89 –96 37. Black, C. B., and Cowan, J. A. (1994) Inorg. Chem. 33, 5805–5808 38. Beese, L. S., and Steitz, T. A. (1991) EMBO J. 10, 25–33 39. Andrake, M. D., and Skalka, A. M. (1995) J. Biol. Chem. 270, 29299 –29306 40. Jenkins, T. M., Engelman, A., Ghirlando, R., and Craigie, R. (1996) J. Biol. Chem. 271, 7712–7718 41. van den Ent, F. M. I., Vink, C., and Plasterk, R. H. (1994) J. Virol. 68, 7825–7832 42. Kukolj, G., and Skalka, A. M. (1995) Genes & Dev. 9, 2556 –2567 43. Yang, W., and Steitz, T. A. (1995) Cell 82, 193–207 44. Cannon, P. M., Byles, E. D., Kingsman, S. M., and Kingsman, A. J. (1996) J. Virol. 70, 651– 657 45. Woerner, A. M., and Marcus, S. C. (1993) Nucleic Acids Res. 21, 3507–3511 46. Engelman, A., Hickman, A. B., and Craigie, R. (1994) J. Virol. 68, 5911–5917 47. Puras-Lutzke, R. A., Vink, C., and Plasterk, R. H. (1994) Nucleic Acids Res. 22, 4125– 4131 48. Eijkelenboom, A., Lutzke, R., Boelens, R., Plasterk, R., Kaptein, R., and Hard, K. (1995) Nat. Struct. Biol. 2, 807– 810 49. Lodi, P. J., Ernst, J. A., Kuszewski, J., Hickman, A. B., Engelman, A., Craigie, R., Clore, G. M., and Gronenborn, A. M. (1995) Biochemistry 34, 9826 –9833 50. Kuriyan, J., and Cowburn, D. (1993) Curr. Opin. Struct. Biol. 3, 828 – 837 51. Craigie, R. (1996) Cell 85, 137–140 52. Murphy, J. E., and Goff, S. P. (1992) J. Virol. 66, 5092–5095 53. Aiyar, A. P., Skalka, A. M., and Leis, J. (1996) J. Virol. 70, 3571–3580 54. Smith, J. S., and Roth, M. J. (1993) J. Virol. 67, 4037– 4049 55. West, S. C. (1994) Cell 76, 9 –15 56. Bukrinsky, M. I., Sharova, N., McDonald, T. L., Pushkarskaya, T., Tarpley, W. G., and Stevenson, M. (1993) Proc. Natl. Acad. Sci. U. S. A. 90, 6125– 6129 57. Gallay, P., Swingler, S., Song, J. P., Bushman, F., and Trono, D. (1995) Cell 83, 569 –576 58. Kalpana, G. V., Marmon, S., Wang, W. D., Crabtree, G. R., and Goff, S. P. (1994) Science 266, 2002–2006 59. Kraulis, P. J. (1991) J. Appl. Crystallogr. 24, 946 –950

Downloaded from http://www.jbc.org/ by guest on February 13, 2017

Multimerization—Although a tetrameric IN structure (see Fig. 1, middle) seems reasonable (26, 42), currently there is no direct evidence that a tetramer is required for coordinated DNA cleavage or joining reactions. Conceivably, sequential action of a dimer could also catalyze both processing and joining steps. However, IN must coordinate four DNA cleavage events in the overall recombination reaction; these include both viral DNA ends and two nearby sites in the target host DNA. It is clear that there must be communication between the active sites that perform the coordinated reactions. In vivo (52) and in coordinated reactions in vitro (42, 53), mutation at one viral DNA end will result in inefficient catalysis at the other viral end. The organization of IN protomers within the synaptic complex must determine the position of DNA substrates and the proper coordination of these four cleavages. Despite sharing a similar topology and reaction chemistry, the enzymes of this superfamily differ in substrate specificity, multimeric structure, and the number of cleavages required. For example, RNase H is known to act as a monomer (54), and correspondingly, its function does not require coordination of multiple cleavages. The position of several of the flanking a-helices would preclude a dimerization interface for RNase H similar to that seen in the IN structures (31, 32). RuvC is known to function as a dimer and performs two DNA cleavage events when resolving the Holliday junction during homologous recombination (see Ref. 55 for review). In contrast to both IN core domain interfaces, the RuvC dimeric interface involves subunit contacts between an elongated a-helix analogous to IN a-helix 3. These examples indicate that proteins which share a conserved topology and catalytic mechanism may differ in the multimeric organization that is required for their specific functions.

Retroviral Integrase, Putting the Pieces Together Mark D. Andrake and Anna Marie Skalka J. Biol. Chem. 1996, 271:19633-19636. doi: 10.1074/jbc.271.33.19633

Access the most updated version of this article at http://www.jbc.org/content/271/33/19633 Alerts: • When this article is cited • When a correction for this article is posted Click here to choose from all of JBC's e-mail alerts Downloaded from http://www.jbc.org/ by guest on February 13, 2017

This article cites 59 references, 29 of which can be accessed free at http://www.jbc.org/content/271/33/19633.full.html#ref-list-1