Anda di halaman 1dari 12

NIH Public Access

Author Manuscript
Curr Opin Struct Biol. Author manuscript; available in PMC 2011 February 1.
Published in final edited form as:
NIH-PA Author Manuscript

Curr Opin Struct Biol. 2010 February ; 20(1): 114–120. doi:10.1016/j.sbi.2009.12.006.

Genome packaging in viruses

Siyang Sun1, Venigalla B. Rao2, and Michael G. Rossmann1


1Department of Biological Sciences, 915 W. State Street, Purdue University, West Lafayette, IN

47907-2054, USA
2Department of Biology, The Catholic University of America, 620 Michigan Avenue NE, Washington,

DC 20064, USA

Summary of recent advances


Genome packaging is a fundamental process in a viral life cycle. Many viruses assemble preformed
capsids into which the genomic material is subsequently packaged. These viruses use a packaging
motor protein that is driven by the hydrolysis of ATP to condense the nucleic acids into a confined
space. How these motor proteins package viral genomes had been poorly understood until recently,
NIH-PA Author Manuscript

when a few X-ray crystal structures and cryo-electron microscopy structures became available. Here
we discuss various aspects of genome packaging and compare the mechanisms proposed for
packaging motors based on structural information.

Introduction
There are two main strategies for viral genome packaging. Many viruses assemble their capsids
around the viral genomes, such as the ssRNA helical tobacco mosaic virus [1], the ssRNA
icosahedral bacteriophage R17 [2], the ssRNA conical HIV [3], and the dsRNA unsegmented
icosahedral Totiviridae [4] viruses (e.g. L-A virus of yeast). Similarly small dsDNA viruses,
with a less than 20 kb genome, assemble the shell around condensed DNA [5,6].

On the other hand, some dsDNA viruses (e.g. tailed bacteriophages and herpesviruses) and
dsRNA viruses (e.g. φ6 and φ12 bacteriophages) form protein capsid shells first and then
package their genomes into the procapsids. For the latter viruses, a motor is required to perform
the packing with energy supplied by hydrolysis of ATP [7].
NIH-PA Author Manuscript

The final genome organization varies independently of the mode of packaging. In most
icosahedral viruses there is no apparent order in the packaged genome, whereas in some viruses
the genome takes on partial icosahedral symmetry in its secondary structure. Examples are the
plant ssRNA bean-pod mottle virus [8], ssRNA satellite tobacco mosaic virus [9], to a lesser
extent the ssDNA bacteriophage φX174 [10] and the ssDNA canine parvovirus [11]. Many
viruses, especially DNA bacteriophages, show systematic layers of packaged DNA [12–14].
Yet other, mostly filamentous, viruses package their genome as helices surrounded by proteins
[15,16]. Some viruses neutralize the negative charge with polyamines (e.g. some RNA plant
viruses) [17] and/or cations (e.g. bacteriophages) [18], whereas other viruses neutralize the
charge with a positively charged domain of the capsid protein [19–21].

Here we review recent structural studies on certain well-characterized viral packaging motor
proteins. A variety of methods have been used, including x-ray crystallography, cryo-electron
microscopy (cryo-EM) and single-molecule measurements using optical tweezers. We will

Corresponding author: Rossmann, Michael G. (mr@purdue.edu).


Sun et al. Page 2

compare models of packaging mechanisms derived from these structural studies for both
dsRNA and dsDNA viruses. We will not discuss in detail the genome organization inside the
viruses and the way genomes are unpackaged after infection has been initiated.
NIH-PA Author Manuscript

Packaging initiation
Viral genome packaging must differentiate between host and viral nucleic acid. A strategy
employed by many viruses is that the viral capsid protein contains a binding site which
recognizes a specific sequence of the genome. For example, in unsegmented Totiviridae
dsRNA viruses, such as the L-A virus of yeast, there is a secondary structure (stem-loop) and
a specific sequence at the 5' end of the genome [22] that are used for recognition by the
polymerase-group antigen (pol-gag) fusion protein [23]. However, segmented dsRNA viruses
such as Reoviridae, Orthomyxoviridae (influenza), Bunyaviruses and Arenaviruses need to
package one of each of the dsRNA segments. It has been proposed that there is complementarity
between regions of the segments which helps the virus to include one of each of the genomic
RNA molecules. Alternatively, for the segmented dsRNA bacteriophage φ6, the procapsid may
have specific binding sites for each of the different segments [22,24].

Double-stranded DNA viruses have evolved yet other strategies for packaging their own
genome. Most bacteriophages and herpesviruses replicate their genomes as head-to-tail
concatemers. A “terminase” complex of two proteins, which is also a part of the packaging
NIH-PA Author Manuscript

motor, recognizes a specific sequence or structure on the concatemeric DNA and makes the
initial cut to generate the free end at which packaging is initiated. After one or slightly more
than one genome length of DNA has been packaged into the head, the same nuclease makes
another cut to terminate packaging. Hence, historically, these proteins that are required for the
generation of genome termini were named “terminases”. Certain dsDNA viruses such as
bacteriophage φ29 and adenoviruses employ protein-primed DNA replication and do not
produce concatemers. These viruses recognize their own DNAs through the terminal primer
proteins that are covalently linked to the genome ends [25].

Mechanism for dsRNA packaging


In dsRNA viruses such as φ6 and φ12, the positive-sense strand is packaged and then replicated
to form dsRNA inside the capsid. Therefore, the substrate for the packaging motor in these
dsRNA viruses is ssRNA. Structurally, the best studied packaging motor protein for dsRNA
viruses is the P4 ATPase in bacteriophage φ12. The hexameric P4 is a multi-functional
molecular unit, which is involved in procapsid assembly [26], actively packages ssRNA
molecules [27], and acts as a passive channel when newly synthesized mRNA molecules are
extruded from the virus during infection [28]. The φ12 genome contains three segmented
dsRNAs. The positive strand of each RNA molecule is recognized sequentially by the procapsid
NIH-PA Author Manuscript

[24,29]. During the packaging process, the procapsid undergoes conformational changes and
expands successively to accommodate all RNA molecules [30].

The P4 ATPase is structurally similar to RecA/F1-ATPase-like motors [31]. For multi-subunit


motors there are two fundamental questions. How is the ATP hydrolysis coupled to
translocation and how do the motor subunits communicate with each other to achieve efficient
packaging? In hexameric P4, the central channel is lined with a portion of a helix α6 and loops
L1 and L2. Mutagenesis and hydrogen-deuterium exchange experiments have shown that L1
and L2 are essential for RNA binding and translocation [32]. The phosphate-binding P-loop,
where the conserved Walker A motif resides, is in a “down” conformation before the bound
ATP is hydrolyzed, and adopts an “up” conformation after ATP has been hydrolyzed. This is
accompanied by the transition of α6 and loop L2 from an “up” position to a “down” position
by about 6 Å, which was proposed to be the driving force for RNA translocation [31]. The
efficient coordination between the substrate and the motor requires a symmetry match. The

Curr Opin Struct Biol. Author manuscript; available in PMC 2011 February 1.
Sun et al. Page 3

ssRNA was proposed to have a structure similar to the A form dsRNA, which has about 11
bases per turn with a rise of 2.6–2.8 Å per base. It was proposed that the slight symmetry
mismatch of 11 bases per turn ssRNA to six motor subunits is alleviated by the flexible loop
NIH-PA Author Manuscript

L1 which acts as a “grommet” to correct the position of the substrate ssRNA [31]. The
sequential hydrolysis of ATP by neighboring P4 subunits is achieved through an “arginine”
finger that is commonly seen in hexameric ATPase motors [33–35]. ATP hydrolysis in one
ATPase subunit induces a conformational change that places the “trans” arginine finger into
the ATPase active center of the neighboring subunit, triggering the subsequent ATP hydrolysis
(Figure 1).

Mechanism for dsDNA packaging


Genome packaging motors in dsDNA viruses are powerful molecular machines. The density
of DNA genomes in bacteriophages is near crystalline and the resultant pressure inside the
viral capsid is about 60 atmospheres [36]. Thus, dsDNA packaging motors are required to
generate enough force to counter this pressure towards the end of the packaging process [36–
39]. The most powerful molecular motor known to date is the bacteriophage T4 genome
packaging motor, which can generate a force of up to 60 piconewtons and package DNA at an
average rate of about 700 bp/sec [37].

The packaging machine in dsDNA viruses usually consists of a dodecameric portal protein at
NIH-PA Author Manuscript

one special 5-fold vertex [40–46] and a pentameric motor protein (historically named the large
terminase) which hydrolyzes ATP to package DNA [40,47–49]. For viruses that replicate their
genomes as concatemers, a regulator protein, historically named the small terminase, works
together with the large terminase in the packaging initiation process and fine tunes the activity
of the large terminase during packaging [50,51]. Unlike the dsRNA motor, the dsDNA
packaging motor is a transient component of the viral particle and does not remain with the
final virion. Rather, the motor complex binds to the empty procapsid at the portal and
dissociates once packaging is completed. For bacteriophage φ29, a viral genome encoded RNA
component (pRNA) forms a bridge between the capsid and the motor protein [40,47].

Rotary motor mechanism


Much of the early structural work on DNA packaging motors was obtained from bacteriophage
φ29. The first atomic structure of a packaging machine component is that of the dodecameric
portal protein gp10 of φ29 [40]. The portal has a wider end that is inside the capsid and a
narrower end that is protruding from the capsid. It has a central channel formed by α-helices
and lined with negative charges, which is suitable for the smooth passage of DNA. Cryo-EM
reconstruction of the procapsid showed that five pRNA molecules bind to the procapsid [40].
The symmetry mismatch of DNA (101 screw), portal (12-fold) and procapsid/pRNA/ATPase
NIH-PA Author Manuscript

(5-fold) would be consistent with an earlier proposed rotary mechanism for DNA packaging
[52]. In this mechanism the portal rotates using the energy from ATP hydrolysis, driving the
DNA into the procapsid. However, single molecule fluorescence spectroscopy experiments
show no portal rotation, making this mechanism unlikely [53].

Linear motor mechanism based on electrostatic forces


The packaging motor protein in dsDNA viruses generally has two domains, an N-terminal
ATPase domain and a C-terminal domain that has nuclease activity [54]. The two domains
need to be physically linked together to retain packaging activity [55]. The linker between the
two domains consists of small amino acids, presumably providing the flexibility the motor
requires to accomplish packaging. Recent crystal structures of the bacteriophage T4 packaging
ATPase gp17 showed that the N-terminal domain has a nucleotide binding fold and is
structurally more similar to monomeric helicases than hexameric ATPases [48,49], whereas

Curr Opin Struct Biol. Author manuscript; available in PMC 2011 February 1.
Sun et al. Page 4

the C-terminal domain belongs to the RnaseH/resolvase/intergrase superfamily [49,56]. The


interactions between the N- and C-terminal domains are quite extensive that involve six charge
pairs and a hydrophobic core. An in vitro active packaging complex [57] that is composed of
NIH-PA Author Manuscript

T4 proheads and gp17 was examined by cryo-EM which showed five gp17 molecules bound
below the special vertex [49]. In contrast to the crystal structure, where the two domains are
in close contact with each other, the cryo-EM reconstruction showed that the two domains are
well-separated. It was proposed that the two structures represent two states of the motor [49].
Comparison of the crystal structure (post-translocation state) with the structure fitted into the
EM density (pre-translocation state) showed an about 7 Å translation of the C-terminal domain
of gp17 along the central axis of the phage head. This is consistent with the bulk measurement
in phages φ29 and T3, where each ATP consumed packages two base pairs of DNA [58,59].
Therefore, a mechanism based on electrostatic interactions was proposed for the bacteriophage
T4 packaging motor. Upon binding of dsDNA to the C-terminal domain of one gp17 subunit,
a cis “arginine” finger in the N-terminal domain is properly positioned into the ATPase active
center which triggers ATP hydrolysis. The subsequent conformation change aligns the
opposing charges in the N- and C-terminal domains, causing the C-terminal domain to be pulled
7 Å towards the N-terminal domain by the electrostatic forces, packaging two base pairs of
dsDNA. Because of the symmetry match between the five gp17s and the 101 B-form DNA,
the neighboring gp17 is now aligned with dsDNA substrate for the next round of translocation
(Figure 2).
NIH-PA Author Manuscript

Non-integer step size motor mechanism


Although it would be logical to assume that each motor subunit would package an integral
number of base pairs for each ATP hydrolyzed, a recent publication by Moffitt et al. reported
a step size of 2.5 bps/ATP for the φ29 packaging motor using high resolution optical tweezer
measurements [60]. The φ29 procapsid/pRNA/ATPase and DNA were tethered to microbeads
held in two optical traps. Packaging was measured by the distance change between the two
traps. External forces were applied to slow down the motor so that individual steps of packaging
could be resolved. Bursts of 10 base pairs packaging were observed separated by dwells. Each
burst consisted of four steps, which was interpreted as four ATP hydrolysis and DNA
translocation events. The dwells between bursts were interpreted as the time required to reload
the ATP molecules onto the motor subunits that had “fired”. The presence of five motor
subunits and four ATP hydrolysis events during each burst demands some asymmetry within
the pentameric motor. Two models were proposed, a piston-like model and an inchworm-like
model. The piston-like model assumes that four of the subunits hydrolyze ATP while the fifth
one retains its ATP and also holds onto the DNA while the other subunits are being reloaded.
This model would require that each motor subunit binds DNA in a non-specific manner or has
multiple binding sites on the DNA. However, it is not clear how the motor would “know”
NIH-PA Author Manuscript

which subunit needs to be special at a certain time. The inchworm-like model proposes that
only two subunits make contact with the DNA while conformational changes from all subunits
create a distortion that drives the packaging of 10bps. This model requires the subunits in the
pentameric motor to change its organization from a symmetrical ring to a staircase during each
burst of 10bps. This seems unlikely since the motor would be attached to the pRNA in the case
of φ29 or to the capsid/portal in the case of T4. Changing the organization of the subunits from
a ring to a one-turn staircase with a 34 Å pitch would dramatically reduce the contact surface
of the motor with the procapsid, which might result in dissociation of the motor (Figure 3). A
very recent paper [61] discusses the type of chemical contact between the DNA and the motor
parts that are required for successful packaging for bacteriophage φ29.

Curr Opin Struct Biol. Author manuscript; available in PMC 2011 February 1.
Sun et al. Page 5

Closing remarks
A few models with common themes for virus genome packaging have emerged based on results
NIH-PA Author Manuscript

from structural studies. In these models packaging occurs by linear movement of the viral
nucleic acid driven by conformational changes in the ATPase motor protein. In addition, the
subunits of the motor are all proposed to function in a highly coordinated manner. However,
the details of the mechanisms are quite different, partly due to the structural difference of the
ligands that are translocated, e.g. ssRNA vs dsDNA. For dsDNA packaging motors, the recent
finding that the packaging step could be non-integral has severely challenged other proposed
mechanisms. Although this is only observed in the phage φ29 packaging system, the only
system that has an RNA component, it would be evolutionarily more likely that all phages
share similar packaging mechanisms. Hopefully, future experiments will shed light on this
puzzle.

Acknowledgments
We are grateful to Sheryl Kelly for assistance in the preparation of this manuscript. The work was supported by National
Science Foundation grants to MGR (MCB-0443899) and VDB (MCB-0923873) as well as National Institutes of Health
grant R56AI081726 to MGR, VDB and SS.

References
NIH-PA Author Manuscript

1. Buck KW. Replication of tobacco mosaic virus RNA. Philos. Trans. R. Soc. Lond. B Biol. Sci
1999;354:613–627. [PubMed: 10212941]
2. Valegård K, Murray JB, Stockley PG, J SN, Liljas L. Crystal structure of an RNA bacteriophage coat
protein-operator complex. Nature 1994;371:623–626. [PubMed: 7523953]
3. Russell RS, Liang C, Wainberg MA. Is HIV-1 RNA dimerization a prerequisite for packaging? Yes,
no probably? Retrovirology 2004;1:23. [PubMed: 15345057]
4. Fujimura T, Esteban R, Esteban LM, Wickner RB. Portable encapsidation signal of the L-A double-
stranded RNA virus of S. cerevisiae. Cell 1990;62:819–828. [PubMed: 2117501]
5. Burroughs AM, Iyer LM, Aravind L. Comparative genomics and evolutionary trajectories of viral ATP
dependent DNA-packaging system. Genome Dyn 2007;3:48–65. [PubMed: 18753784]
6. Roitman-Shemer V, Stokrova J, Forstova J, Oppenheim A. Assemblages of simian virus 40 capsid
proteins and viral DNA visualized by electron microscopy. Biochem. Biophys. Res. Commun
2007;353:424–430. [PubMed: 17189615]
7. Rao VB, Feiss M. The bacteriophage DNA packaging motor. Annu. Rev. Genet 2008;42:642–681.
8. Chen Z, Stauffacher C, Li Y, Schmidt T, Bomu W, Kamer G, Shanks M, Lomonossoff G, Johnson JE.
Protein-RNA interactions in an icosahedral virus at 3.0 Å resolution. Science 1989;245:154–159.
[PubMed: 2749253]
9. Larson SB, Koszelak S, Day J, Greenwood A, Dodds JA, McPherson A. Three-dimensional structure
NIH-PA Author Manuscript

of satellite tobacco mosaic virus at 2.9 Å resolution. J. Mol. Biol 1993;231:375–391. [PubMed:
8510153]
10. McKenna R, Xia D, Willingmann P, Ilag LL, Rossmann MG. Structure determination of the
bacteriophage φX174. Acta Crystallogr. sect. B 1992;48:499–511. [PubMed: 1418820]
11. Chapman MS, Rossmann MG. Structural refinement of the DNA-containing capsid of canine
parvovirus using RSRef, a resolution-dependent stereochemically restrained real-space refinement
method. Acta Crystallogr D Biol. Crystallogr 1996;52:129–142. [PubMed: 15299734]
12. Fokine A, Kostyuchenko VA, Efimov AV, Kurochkina LP, Sykilinda NN, Robben J, Volckaert G,
Hoenger A, Chipman PR, Battisti AJ, et al. A three-dimensional cryo-electron microscopy structure
of the bacteriophage φKZ head. J. Mol. Biol 2005;352:117–124. [PubMed: 16081102]
13. Fokine A, Chipman PR, Leiman PG, Mesyanzhinov VV, Rao VB, Rossmann MG. Molecular
architecture of the prolate head of bacteriophage T4. Proc. Natl. Acad. Sci. U.S.A 2004;101:6003–
6008. [PubMed: 15071181]

Curr Opin Struct Biol. Author manuscript; available in PMC 2011 February 1.
Sun et al. Page 6

14. Choi KH, McPartland J, Kaganman I, Bowman VD, Rothman-Denes LB, Rossmann MG. Insight
into DNA and protein transport in double-stranded DNA viruses: the structure of bacteriophage N4.
J. Mol. Biol 2008;378:726–736. [PubMed: 18374942]
NIH-PA Author Manuscript

15. Sachse C, Chen JZ, Coureux P-D, Stroupe ME, Fändrich M, Grigorieff N. High-resolution electron
microscopy of helical specimens: a fresh look at tobacco mosaic virus. J. Mol. Biol 2007;371:812–
835. [PubMed: 17585939]
16. Goulet A, Blangy S, Redder P, Prangishvili D, Felisberto-Rodrigues C, Forterre P, Campanacci V,
Cambillau C. Acidianus filamentous virus 1 coat proteins: a helical fold spanning the filamentous
archaeal viruses lineage. Proc. Natl. Acad. Sci. U.S.A. 2009 in press.
17. Balint R, Cohen SS. The incorporation of radiolabeled polyamines and methionine into turnip yellow
mosaic virus in protoplasts from infected plants. Virology 1985;144:181–193. [PubMed: 4060587]
18. Yu T-Y, Schaefer J. REDOR NMR characterization of DNA packaging in bacteriophage T4. J. Mol.
Biol 2008;382:1031–1042. [PubMed: 18703073]
19. Warrier R, Linger BR, Golden BL, Kuhn RJ. Role of Sindbis virus capsid protein region II in
nucleocapsid core assembly and encapsidation of genomic RNA. J. Virol 2008;82:4461–4470.
[PubMed: 18305029]
20. Hacker DL. Identification of a coat protein binding site on southern bean mosaic virus RNA. Virology
1995;210:562–565. [PubMed: 7886960]
21. Hillman BI, Hearne P, Rochon D, Morris TJ. Organization of tomato bushy stunt virus genome:
characterization of the coat protein gene and the 3' terminus. Virology 1989;169:42–50. [PubMed:
2922927]
NIH-PA Author Manuscript

22. Wickner RB. Double-stranded RNA virus replication and packaging. J. Biol. Chem 1993;268:3797–
3800. [PubMed: 8440674]
23. Fujimura T, Ribas JC, Makhov AM, Wickner RB. Pol of gag-pol fusion protein required for
encapsidation of viral RNA of yeast L-A virus. Nature 1992;359:746–749. [PubMed: 1436038]
24. Olkkonen VM, Gottlieb P, Strassman J, Qiao XY, Bamford DH, Mindich L. In vitro assembly of
infectious nucleocapsids of bacteriophage φ6: formation of a recombinant double-stranded RNA
virus. Proc. Natl. Acad. Sci. U.S.A 1990;87:9173–9177. [PubMed: 2251260]
25. Rodriguez I, Lázaro JM, Salas M, de Vega M. φ29 DNA polymerase-terminal protein interaction.
Involvement of residues specifically conserved among protein-primed DNA polymerases. J. Mol.
Biol 2004;337:829–841. [PubMed: 15033354]
26. Poranen MM, Paatero AO, Tuma R, Bamford DH. Self-assembly of a viral molecular machine from
purified protein and RNA constituents. Mol. Cell 2001;7:845–854. [PubMed: 11336707]
27. Juuti JT, Bamford DH. RNA binding, packaging and polymerase activities of the different incomplete
polymerase complex particles of dsRNA bacteriophage φ6. J. Mol. Biol 1995;249:545–554.
[PubMed: 7783210]
28. Kainov DE, Lísal J, Bamford DH, Tuma R. Packaging motor from double-stranded RNA
bacteriophage φ12 acts as an obligatory passive conduit during transcription. Nucleic Acids Res
2004;32:3515–3521. [PubMed: 15247341]
NIH-PA Author Manuscript

29. Qiao X, Casini G, Qiao J, Mindich L. In vitro packaging of individual genomic segments of
bacteriophage φ6 RNA: serial dependence relationships. J. Virol 1995;69:2926–2931. [PubMed:
7707518]
30. Butcher SJ, Dokland T, Ojala PM, Bamford DH, Fuller SD. Intermediates in the assembly pathway
of the double-stranded RNA virus φ6. EMBO J 1997;16:4477–4487. [PubMed: 9250692]
31. Mancini EJ, Kainov DE, Grimes JM, Tuma R, Bamford DH, Stuart DI. Atomic snapshots of an RNA
packaging motor reveal conformational changes linking ATP hydrolysis to RNA translocation. Cell
2004;118:743–755. [PubMed: 15369673] **This work described the packaging mechanism for
ssRNA packaging in a dsRNA virus.
32. Lísal J, Lam TT, Kainov DE, Emmett MR, Marshall AG, Tuma R. Functional visualization of viral
molecular motor by hydrogen-deuterium exchange reveals transient states. Nat. Struct. Mol. Biol
2005;12:460–466. [PubMed: 15834422]
33. Sawaya MR, Guo S, Tabor S, Richardson CC, Ellenberger T. Crystal structure of the helicase domain
from the replicative helicase-primase of bacteriophage T7. Cell 1999;99:167–177. [PubMed:
10535735]

Curr Opin Struct Biol. Author manuscript; available in PMC 2011 February 1.
Sun et al. Page 7

34. Hishida T, Han Y-W, Fujimoto S, Iwasaki H, Shinagawa H. Direct evidence that a conserved arginine
in RuvB AAA+ ATPase acts as an allosteric effector for the ATPase activity of the adjacent subunit
in a hexamer. Proc. Natl. Acad. Sci. U.S.A 2004;101:9573–9577. [PubMed: 15210950]
NIH-PA Author Manuscript

35. Abrahams JP, Leslie AGW, Lutter R, Walker JE. Structure at 2.8 Å resolution of F1-ATPase from
bovine heart mitochondria. Nature 1994;370:621–628. [PubMed: 8065448]
36. Smith DE, Tans SJ, Smith SB, Grimes S, Anderson DL, Bustamante C. The bacteriophage φ29 portal
motor can package DNA against a large internal force. Nature 2001;413:748–752. [PubMed:
11607035] * This work provides background towards single molecule investigations of packaging
motors using optical tweezers.
37. Fuller DN, Raymer DM, Kottadiel VI, Rao VB, Smith DE. Single phage T4 DNA packaging motors
exhibit large force generation, high velocity, and dynamic variability. Proc. Natl. Acad. Sci. U.S.A
2007;104:16868–16873. [PubMed: 17942694] * Examines the properties of the T4 DNA packaging
motor using single molecule techniques.
38. Fuller DN, Raymer DM, Rickgauer JP, Robertson RM, Catalano CE, Anderson DL, Grimes S, Smith
DE. Measurements of single DNA molecule packaging dynamics in bacteriophage λ reveal high
forces, high motor processivity, and capsid transformations. J. Mol. Biol 2007;373:1113–1122.
[PubMed: 17919653]
39. Fuller DN, Rickgauer JP, Jardine PJ, Grimes S, Anderson DL, Smith DE. Ionic effect on viral DNA
packaging and portal motor function in bacteriophage φ29. Proc. Natl. Acad. Sci. U.S.A
2007;104:11245–11250. [PubMed: 17556543]
40. Simpson AA, Tao Y, Leiman PG, Badasso MO, He Y, Jardine PJ, Olson NH, Morais MC, Grimes
S, Anderson DL, et al. Structure of the bacteriophage φ29 DNA packaging motor. Nature
NIH-PA Author Manuscript

2000;408:745–750. [PubMed: 11130079] * The first structure of a phage portal protein.


41. Lebedev AA, Krause MH, Isidro AL, Vagin AA, Orlova EV, Tavares JT, Antson AA. Structural
framework for DNA translocation via the viral portal protein. EMBO J 2007;26:1984–1994.
[PubMed: 17363899]
42. Orlova EV, Gowen B, Dröge A, Stiege A, Weise F, Lurz R, van Heel M, Tavares P. Structure of a
viral DNA gatekeeper at 10 Å resolution by cryo-electron microscopy. EMBO J 2003;22:1255–1262.
[PubMed: 12628918]
43. Leiman PG, Chipman PR, Kostyuchenko VA, Mesyanzhinov VV, Rossmann MG. Three-dimensional
rearrangement of proteins in the tail of bacteriophage T4 on infection of its host. Cell 2004;118:419–
429. [PubMed: 15315755]
44. Chang J, Weigele P, King J, Chiu W, Jiang W. Cryo-EM asymmetric reconstruction of bacteriophage
P22 reveals organization of its DNA packaging and infection machinery. Structure 2006;14:1073–
1082. [PubMed: 16730179]
45. Lander GC, Tang L, Casjens SR, Gilcrease EB, Prevelige P, Poliakov A, Potter CS, Carragher B,
Johnson JE. The structure of an infectious P22 virion shows the signal for headful DNA packaging.
Science 2006;312:1791–1795. [PubMed: 16709746]
46. Jiang W, Chang J, Jakana J, Weigele P, King J, Chiu W. Structure of epsilon15 bacteriophage reveals
genome organization and DNA packaging/injection apparatus. Nature 2006;439:612–616. [PubMed:
NIH-PA Author Manuscript

16452981]
47. Morais MC, Koti JS, Bowman VD, Reyes-Aldrete E, Anderson DL, Rossmann MG. Defining
molecular and domain boundaries in the bacteriophage φ29 DNA packaging motor. Structure
2008;16:1267–1274. [PubMed: 18682228]
48. Sun S, Kondabagil K, Gentz PM, Rossmann MG, Rao VB. The structure of the ATPase that powers
DNA packaging into bacteriophage T4 procapsids. Mol. Cell 2007;25:943–949. [PubMed:
17386269]
49. Sun S, Kondabagil K, Draper B, Alam TI, Bowman VD, Zhang Z, Hegde S, Fokine A, Rossmann
MG, Rao VB. The structure of the phage T4 DNA packaging motor suggests a mechanism dependent
on electrostatic forces. Cell 2008;135:1251–1262. [PubMed: 19109896] ** Describes the mechanism
of the T4 packaging motor based on crystallographic and cryo-EM structures.
50. Němeček D, Lander GC, Johnson JE, Casjens SR, Thomas GJ Jr. Assembly architecture and DNA
binding of the bacteriophage P22 terminase small subunit. J. Mol. Biol 2008;383:494–501. [PubMed:
18775728]

Curr Opin Struct Biol. Author manuscript; available in PMC 2011 February 1.
Sun et al. Page 8

51. Al-Zahrani AS, Kondabagil K, Gao S, Nelly N, Ghosh-Kumar M, Rao VB. The small terminase,
gp16, of bacteriophage T4 is a regulator of the DNA packaging motor. J. Biol. Chem
2009;284:24490–24500. [PubMed: 19561086]
NIH-PA Author Manuscript

52. Hendrix RW. Symmetry mismatch and DNA packaging in large bacteriophages. Proc. Natl. Acad.
Sci. U.S.A 1978;75:4779–4783. [PubMed: 283391] * The original basis for considering a rotary
motion for the portal protein.
53. Hugel T, Michaelis J, Hetherington CL, Jardine PJ, Grimes S, Walter JM, Falk W, Anderson DL,
Bustamante C. Experimental test of connector rotation during DNA packaging into bacteriophage
φ29 capsids. PLoS Biology 2007;5:e59. [PubMed: 17311473]
54. Mitchell MS, Matsuzaki S, Imai S, Rao VB. Sequence analysis of bacteriophage T4 DNA packaging/
terminase genes 16 and 17 reveals a common ATPase center in the large subunit of viral terminases.
Nucleic Acids Res 2002;30:4009–4021. [PubMed: 12235385]
55. Kanamaru S, Kondabagil K, Rossmann MG, Rao VB. The functional domains of bacteriophage T4
terminase. J. Biol. Chem 2004;279:40795–40801. [PubMed: 15265872]
56. Alam TI, Draper B, Kondabagil K, Rentas FJ, Ghosh-Kumar M, Sun S, Rossmann MG, Rao VB. The
headful packaging nuclease of bacteriophage T4. Mol. Microbiol 2008;69:1180–1190. [PubMed:
18627466]
57. Kondabagil KR, Zhang Z, Rao VB. The DNA translocating ATPase of bacteriophage T4 packaging
motor. J. Mol. Biol 2006;363:786–799. [PubMed: 16987527]
58. Guo P, Peterson C, Anderson D. Prohead and DNA-gp3-dependent ATPase activity of the DNA
packaging protein gp16 of bacteriophage φ29. J. Mol. Biol 1987;197:229–236. [PubMed: 2960820]
NIH-PA Author Manuscript

59. Morita M, Tasaka M, Fujisawa H. DNA packaging ATPase of bacteriophage T3. Virology
1993;193:748–752. [PubMed: 8460483]
60. Moffitt JR, Chemla YR, Aathavan K, Grimes S, Jardine PJ, Anderson DL, Bustamante C. Intersubunit
coordination in a homomeric ring ATPase. Nature 2009;457:446–450. [PubMed: 19129763] ** A
mechanistic interpretation of the bursts and dwells that occur during DNA packaging in φ29,
concluding that there is a non-integral number of bases packaged for each ATP hydrolyzed.
61. Aathavan K, Politzer AT, Kaplan A, Moffitt JR, Chemla YR, Grimes S, Jardine PJ, Anderson DL,
Bustamante C. Substrate interactions and promiscuity in a viral DNA packaging motor. Nature. 2009
in press. ** An investigation of the types of interactions between the packaging motor and dsDNA
in φ29.
NIH-PA Author Manuscript

Curr Opin Struct Biol. Author manuscript; available in PMC 2011 February 1.
Sun et al. Page 9
NIH-PA Author Manuscript
NIH-PA Author Manuscript

Figure 1.
Model for translocation along P4 by the ssRNA in bacteriophage φ12 [31]. The relative
NIH-PA Author Manuscript

movements of the RNA binding loops in six adjacent subunits as the ATP hydrolysis-induced
conformational changes ripple around the ring. (a) Diagram showing the hexameric molecule
represented as a cylinder associated with the viral capsid shell (gray). The direction of ssRNA
(yellow) translocation is depicted by a yellow arrow, while the cyan arrow shows the direction
of sequential ATP hydrolysis. To obtain the view shown in b, the coordinates are projected
onto the cylinder shown and the cylinder unwrapped to lie across the page. The effect of this
is that the P4 hexamer is peeled open and viewed from inside the ring looking outward. (b)
Cylindrical polar projection showing a snapshot of the hexamer just prior to hydrolysis at
subunit j (marked with the flash). The RNA binding loops are colored green and the ssRNA
yellow with phosphates represented by balls. Phosphates interacting with the RNA binding
loops are identified by a green outer glow. Diphosphate and triphosphate moieties are colored
according to their conformation: AMPcPP inactive “I” (orange), AMPcPP active “A” (red),
and product “P” (blue). The power stroke associated with the hydrolysis step about to occur
will translocate RNA downward (from the solid RNA position to the semitransparent position).

Curr Opin Struct Biol. Author manuscript; available in PMC 2011 February 1.
Sun et al. Page 10

Two turns of the RNA spiral are shown. [Reprinted from Cell, Volume 118, Mancini EJ, Kainov
DE, Grimes JM, Tuma R, Bamford DH, and Stuart DI: Atomic snapshots of an RNA
packaging motor reveal conformational changes linking ATP hydrolysis to RNA
NIH-PA Author Manuscript

translocation, pages 743–755, copyright 2004, with permission from Elsevier].


NIH-PA Author Manuscript
NIH-PA Author Manuscript

Curr Opin Struct Biol. Author manuscript; available in PMC 2011 February 1.
Sun et al. Page 11
NIH-PA Author Manuscript

Figure 2.
NIH-PA Author Manuscript

DNA packaging mechanism in bacteriophage T4 [49]. Panels (a)–(d) relate to the sequence of
events that occur in a single gp17 molecule. The gp17 N-terminal subdomain I, subdomain II,
and C-terminal domain are represented as green, yellow, and cyan ovals, respectively. The
five-pointed stars show the charge interactions between the N-terminal subdomain I and the
C-terminal domain. The four-pointed stars show the charge interaction between the N-terminal
subdomain II and the C-terminal domain. The flexible linker between N- and C-terminal
domain is represented by a wiggly cyan line.(a) The gp17 C-terminal domain is ready to bind
DNA.(b) The C-terminal domain, when bound to the DNA, brings the DNA closer to the N-
terminal domain of the same subunit. Conformational change in the N-terminal domain causes
Arg162 to be placed into the ATPase active center in preparation for hydrolysis.(c) Hydrolysis
of ATP has rotated the N-terminal subdomain II by about 6°, thereby aligning the charge pairs
resulting in an electrostatic attraction that moves the C-terminal domain and the DNA 6.8 Å
(equivalent to the distance between two base pairs) closer to the N-terminal domain and into
the capsid.(d) ADP and Pi are released and the C-terminal domain returns to its original
position. DNA is released and is aligned to bind the C-terminal domain of the neighboring
subunit. (e) This panel relates to the synchronization of the five gp17 molecules located around
the special vertex of the procapsid. Successive DNA base pairs are indicated by yellow arrows
outside the procapsid, red entering the procapsid, and white inside the procapsid. The
NIH-PA Author Manuscript

surrounding five gp17 molecules are shown as stars. The red star represents gp17 ATPase
hydrolyzing ATP, the blue star represents ATPase that is ready to hydrolyze ATP, and the
white star represents ATPase that has already hydrolyzed ATP. Left: Hydrolysis of ATP at
position 1 translocating two base pairs into the procapsid. Middle: Hydrolysis of ATP at
position 3 causing the translocation of further two base pairs into the procapsid. Right: The
ATP at position 5 is ready to be hydrolyzed. [Reprinted from Cell, Volume 135, Sun S,
Kondabagil K, Draper B, Alam TI, Bowman VD, Zhang Z, Hegde S, Fokine A, Rossmann
MG, and Rao VB, The structure of the phage T4 DNA packaging motor suggests a
mechanism dependent on electrostatic forces, pages 1251–1262, copyright 2008.]

Curr Opin Struct Biol. Author manuscript; available in PMC 2011 February 1.
Sun et al. Page 12
NIH-PA Author Manuscript
NIH-PA Author Manuscript

Figure 3.
Packaging models that produce a non-integer step size in bacteriophage φ29 [60]. (a) Depiction
of a translocation model in which all subunits eventually contact the DNA (cyan spheres). The
contacting subunit is outlined in black (top view). (b) In such a model the size of internal
conformational changes set the step size (side view). (c) Depiction of a translocation model in
which only two subunits contact the DNA (black outline). (d) In such a model, one subunit
maintains contact with the DNA (the latch) while the loading of each ATP introduces relative
subunit–subunit rotations which distort the ring. This distortion extends one subunit (the lever)
along the DNA by ~10 bp. The DNA contact point is then transferred from the latch to the
lever, and the release of hydrolysis products relaxes the ring, retracting the lever and the DNA.
The DNA contact is then transferred back to the latch, the ring resets and the cycle begins
again. Because there are four subunits, the ring is retracted in four steps, dividing a 10 bp step
into four ~2.5 bp substeps. The subunit is colored based on its substrate binding state as follows:
NIH-PA Author Manuscript

ATP docking (green, T), tight ATP binding and activation (red, T*), ADP bound (blue) and
apo (purple). [Reprinted by permission from Macmillan Publishers Ltd: Moffitt JR, Chemla
YR, Aathavan K, Grimes S, Jardine PJ, Anderson DL, Bustamante C: Intersubunit
coordination in a homomeric ring ATPase. Nature 2009, 457:446–450., www.nature.com]

Curr Opin Struct Biol. Author manuscript; available in PMC 2011 February 1.

Anda mungkin juga menyukai