http://www.abbs.info e-mail:[email protected]

ISSN 0582-9879                                        ACTA BIOCHIMICA et BIOPHYSICA SINICA 2002, 34(5): 533-543                                     CN 31-1300/Q

 

Mini Review

Methods for Structural and Functional Analysis of an RNA Hexamer of Bacterial Virus phi29 DNA Packaging Motor

GUO Peixuan

( Department of Pathobiology and Purdue Cancer Center, Purdue University, West Lafayette, IN 47907, USA )

Abstract    During multiplication and maturation, the lengthy genomic DNA of dsDNA viruses is translocated with remarkable velocity into a limited space within the procapsid and packaged to crystalline density. A viral DNA-packaging motor accomplishes this energy consuming motion task. An RNA molecule of bacterial virus phi29 has been found to be a vital component of the DNA-packaging motor. Six pRNAs form a hexagonal complex to gear the DNA translocating machine using a mechanism similar to the driving of a bolt with a hex nut. Sequential action of six RNA molecules to drive the motor is similar to the consecutive firing of six cylinders of a car engine. This article reviews the structure of pRNA to demonstrate that its structure plays a vital role in its function, and focuses on methods and unique approaches that lead to the elucidation of pRNA structure.

Key words    phi29viral DNA packagingbio-motormolecule motorRNA 3D structure

The amazing diversity in RNA function is attributed to the variety of RNA species and the flexibility in RNA structure.  To elucidate how RNA molecules perform their versatile and novel functions,  it is crucial to understand the principles and rules that regulate RNA folding.  Due to its complexity and versatility,  clarifying the criterion for RNA folding and determining RNA structure is an arduous task.

One prominent feature in the assembly of all linear ds-DNA viruses is that their lengthy genome is packed with a swift velocity into the pre-formed protein coating and packaged to a near crystalline density (for review,  see references[1-5]). This DNA motion,  which is an energetically unfavorable process,  is accomplished by an ATP-hydrolyzing motor involving a connector and two nonstructural components with certain characteristics typical of ATPases. In phi29,  one of the nonstructural proteins for DNA packaging is an RNA (pRNA) molecule[6-9]. The connector is a 12-subunit hollow truncated cone cylinder having a central channel with a diameter of about 4-nm through which DNA enters the procapsid during packaging[10, 11]. The 120-base pRNA (Fig.1) encoded by the virus,  binds to the connector[12, 13] and is not present in the mature phi29 virion. The requirement for pRNA in phi29 assembly appears to be very specific in that pRNAs from other phages cannot replace the phi29 pRNA in in vitro packaging assays[14] and that a single base mutation can render the pRNA inactive[15].  

Fig.1  pRNA secondary structure

adapted from[56]  with permission from J Biol Chem.

Six copies of pRNA form a hexameric complex which serves as the essential component of the DNA translocating motor[16-21]. DNA packaging is completely blocked when one of the six slots is occupied by an inactive pRNA with a mutation at the 5 or 3 end[17, 18] since pRNA is associated with procapsids during the DNA translocation process.

Phi29 is the most efficient and the best-characterized in vitro DNA packaging model system. Most essential components required to reconstitute the phi29 in vitro DNA packaging activity have been well defined. Direct force measurements have shown that the phi29 packaging motor is the most powerful of all reported molecular motors,  producing a force of 57 pico-Newtons[22, 23]. In addition,  the crystal structure of the 12-subunit (gp10) connector has been solved[11, 24]. One nonstructural component is a protein,  gp16,  that has been shown to contain consensus ATP-binding domains[25] and is able to hydrolyze ATP. It has been found that one ATP is needed to package two base pairs of DNA[25].

This review will focus on the work produced in the author's laboratory and will concentrate on methods and approaches that were used to investigate pRNA structure and function.

1  The role of pRNA in phi29 DNA packaging

In 1997, a comprehensive paper was published to describe the role of phi29 pRNA in DNA packaging[16] which provided an elegant model for explaining the mechanism of connector (portal vertex) rotation and the quantification of energy (ATP) usage[16]. The model presented advocated that phi29 DNA packaging is accomplished by a mechanism similar to driving a bolt with a hex nut,  which consists of six DNA-packaging pRNAs (Fig.2).  Recent research[11,1920,22,23] in this area supports the conclusion about consecutive action of six pRNAs in driving the DNA rotation machine.

Fig.2  A model to depict the sequential action of pRNAs in phi29 DNA packaging motor

The hexagon represents the phi29 connector and the surrounding pentagon represents the capsid. Six protrusions represent six pRNAs. The variable pRNA patterns portray the pRNA in serial energetic states. For example,  pRNA 4 in panel A is in a contracted conformation,  and pRNA 1 is in a relaxed conformation.  Arrows marks the different transition states of pRNA 1. Steps A to G show the six steps of rotation.  Each step rotates 12°,  since a five to six-fold symmetry mismatch generates 30 equivalent positions,  and 360°/30 = 12°. The portal vertex turns 72° after six steps.  For example,  pRNA 1 moves from vertex a in A to vertex b in G,  and rotates 72°. Each step consumes one ATP to induce one conformation change of pRNA,  and six ATPs are used for the transition from one vertex to another. 30 ATPs are used for each 360° rotation (Reprinted from[16] with permission from J Virology).

All icosahedral viruses contain a five-fold symmetrical capsid vertex. In bacteriophage phi29,  the connector is embedded within this five-fold symmetrical area of the capsid,  while six copies of pRNA bind to the connector[16, 26] Rotation of the hexameric pRNA (and connector) within the 5-fold symmetrical environment could constitute a mechanical motor in which rotation may be facilitated by the symmetry mismatch. The relative motion of the two rings could thus produce a driving force to translocate viral DNA into the procapsid. Similar to a car engine in which the six cylinders fire sequentially,  the sequential action of six pRNAs is one way to achieve the turning of this DNA packaging motor. An engine could not run continuously if the cylinders fired at the same time (Fig.2). Likewise,  in the case of phi29,  sequential action of the six pRNAs sitting on the connector may drive rotation of the connector.  The pRNA may collaborate with the viral DNA-packaging enzyme gp16 to perform this rotation job.  By sequential action,  it is meant that the multiple pRNAs involved in DNA packaging appear to act in a step-by-step process,  with each pRNA exerting its individual effect alternatively (Fig.2).

The pRNA contains two functional domains one for connector binding and the other for DNA translocation.  Mutagenesis as well as chemical and nuclease probing have revealed that the pRNA binds to the connector leaving the essential  5/3domain free for interaction with other components,  such as gp16,  DNA-gp3 or other components on the procapsid. It has been reported that the C18C19A20 bulge of the pRNA is solvent-exposed when pRNA is bound to procapsid[27] and is critical for DNA translocation[16, 28-31]. The C18C19A20 bulge might be directly involved in interacting with ATP, gp16, DNA-gp3 or capsid components. It is predicted that pRNA is part of an ATPase and possesses at least two conformations a relaxed and a contracted one. Alternating between contraction and relaxation driven by ATP hydrolysis,  each member of the hexameric RNA complex helps rotate the translocating machine.

The requirement of an intermolecular loop/loop interaction between individual pRNA molecules during DNA packaging has led to the belief that pRNA forms a hexamer[19, 20] and supports a pRNA sequential action model.  The pRNAs may need to communicate with each other during DNA packaging to ensure that the motion is consecutive.  Inter-pRNA interactions via loops might serve as a link to pass a signal to adjacent pRNAs,  regulating sequential conformational changes and/or interactions. Thus base pairing between the right and left-hand loops might be necessary to transfer a conformational change from one pRNA to an adjacent one.

2  Establishment of systems to facilitate functional analysis of pRNA

Four major approaches have been undertaken in this lab to study the structure and function of pRNA.

The first approach was the development of a highly sensitive in vitro phi29 virion assembly system for the assay of pRNA activity[32, 33]. With this system,  up to 5×109 infectious virions per ml can be obtained in the presence of pRNA,  yet not a single infectious virion is detected in the absence of pRNA.  Therefore,  a system with a dynamic range of more than 9 orders of magnitude and with a sensitivity level of as few as 0 infectious virions can be used for the analysis of pRNA structure and function.

The second approach in analyzing pRNA function was the construction of circularly permuted pRNAs (cp-pRNA) in which any internal base of the pRNA could be reassigned to serve as new 5- or 3-termini[28, 34] (Fig.3). The circular permutation system greatly facilitated the construction of mutant pRNA via PCR and enabled the labeling of any specific internal base by radioisotopes,  fluorescence[35] or photoaffinity agents.

Fig.3  Tandem DNA for the synthesis of circularly permuted pRNA (cp-pRNA)

Adapted from[34] with permission from Virology.

The third approach undertaken was the construction of mutant pRNAs that were inactive in DNA packaging,  but competent to compete with wild type pRNA for procapsid binding[16-18, 29, 36, 37].

The fourth approach to analyze pRNA structure and function was the design of new methods to determine the stoichiometry of pRNA in DNA packaging[18, 19, 38]. To determine the role of pRNA in DNA packaging,  it is crucial to know how many copies of the pRNA are involved in each DNA packaging event.  Three novel methods have been developed to determine the stoichiometry of the pRNA,  and have led to the conclusion that six pRNAs are present in each DNA translocating motor. These methods include  (a),  binomial distribution (Yang Hui Triangle)[18, 38] (Fig.4)  (b),  comparing slopes of concentration dependence[18, 33]  and (c),  finding the common multiple of 2,  3,  and 6 by using a set of two interlocking pRNAs,  three interlocking pRNAs and six inter-locking pRNAs[19].  

Fig.4  Stoichiometry determination by binomial distribution

Z represents the total pRNA number,  initially assigned a theoretical value from 1 to 12,   per procapsid to be determined. The empirical curve from mutant  pRNA P8/P4 falls between the theoretical curves for  Z = 5 and Z = 6 (Reprinted from[18] with permission from J Virology).

 

3  Studies on pRNA structure

Although nucleotide derivatives have been found in RNA,  the primary sequence of the RNA molecule is nevertheless as simple as DNA,  since both are composed of four nucleotides. All DNA molecules appear as double helices,  while RNA has a diverse structure. Intriguingly,  small RNA molecules,  containing only the four nucleotides A,  G,  C,  and U,  exhibit versatile biological functions.  Such versatility is ascribed to the flexibility and complexity in RNA structural folding. NMR and X-ray crystallography have been used to obtain a physical tertiary structure of RNAs.  Currently,  NMR can only be applied to an RNA molecule with a size of less than 40 nucleotides.  X-ray crystallography of structural RNAs has proven difficult.  The difficulty,  uncertainty and time-span in obtaining a diffractable RNA crystallographic structure,  as well as the impossibility of using NMR for large RNAs,  compel the use of alternative approaches to obtain information on RNA structures.

3.1  Genetic analysis by truncation,  insertion,  deletion and mutation

The establishment of the highly sensitive in vitro phi29 assembly system (Section 2) greatly facilitated the genetic analysis of pRNA structure[32, 33]. Taking advantage of the circularly permuted pRNA system (Fig.2) (Section 2)[28, 34],  the technique of two-step PCR,  and the relatively small size of the pRNA (120 bases),  mutant pRNAs can be easily constructed with truncation,  deletion,  insertion and mutation targeting any desired position[29]. A plasmid DNA with two tandem RNA coding sequences linked with three bases AAA were used as templates to generate PCR DNA fragments with primer pairs containing either the T7 or SP6 promoter and mutations to pRNA. The DNA fragments from PCR were used to transcribe mutant or circularly permuted pRNAs in vitro with either T7 or SP6 RNA polymerase. In combination with the aforementioned in vitro assembly assay system,  dozens of mutant pRNAs can be obtained and tested in one or two weeks.  

By the use of the truncation and deletion techniques, it was revealed that three nucleotides,  U72U73U74, predicted to form a bulge located at a three-helix junction (Fig.2),  function to  provide flexibility in pRNA folding[28]. Three other nucleotides,  C18C19A20, were shown to be present on the surface of the pRNA as a bulge that is  not involved in procapsid binding but is essential for DNA packaging[30].

3.2  Phylogenetic analysis

Phylogenetic analysis of RNA is used to compare the sequences of RNA molecules with identical or similar functions from different species. A common secondary structure for RNA molecules with a similar function is deduced from such analyses. The theory behind such logic is that RNA structure plays a critical factor in RNA function.  Nature would select the most stable molecule with the best-fit structure or with acceptable base co-variations. Later on,  such phylogenetic analysis of species from nature would be expanded into molecules made artificially,  such as complementary modification or SELEX that will be described below.

Phylogenetic analysis revealed that pRNAs from bacteriophages SF5,  phi29,  PZA,  M2,  NF,  GA1 and B103,  which have a very low sequence identity and few conserved bases, very impressively show similar predicted secondary structures[14, 29]. The requirement for pRNA in phi29 assembly is very specific in that pRNAs from other phages cannot replace the phi29 pRNA in in vitro packaging[14] and that a single base mutation can render the pRNA completely inactive[15]. Thus,  similar structures do not translate into identical function. Interestingly,  phylogenetic analysis revealed that the right (upper) loop of each pRNA was complementary to the left (lower) loop within the same molecule[29]. Complementary modification studies reveal that the pairing is inter-molecular[19, 20] rather than intra-molecular (pseudoknot[39]) and that two G/C pairs are sufficient to mediate the interaction[19, 20].

3.3  Complementary modification

Ano ther approach to confirm base-pairing in predicted RNA structure is complementary modification. Before the conclusion that G pairs to C in an RNA structure is drawn,  at least three mutants should be constructed and analyzed.  First,  mutants with either the G changed to A (or U) or the C changed to U (or A) should be inactive.  In addition,  a mutant with both the Gs changed to As (or Us) and the Cs changed to Us (or As) should restore the activity.

Computer predictions of the phi29 pRNA secondary structure[40] showed that the 5 and 3 ends are paired. An extensive series of helix disruptions by base substitutions almost always resulted in the loss of DNA packaging activity.  Additional compensatory mutations that restored the predicted base pairings rescued the activity of pRNA[15, 28, 39]. Such complementary modification has led to the conclusion that bases 1-3 are paired with bases 117-115bases 7-9 are paired with bases 112-110bases 14-16 are paired with bases 103-101and bases 76-78 are paired with bases 90-88 (Fig.1). This second site suppression confirmed the existence of a helical structure that is essential for pRNA function.

Complementary modification has also been used to study inter-pRNA loop/loop interactions in dimers[19, 20, 29]. A series of mutant pRNAs carrying mutated right and/or left-hand loop sequences were constructed such that loop sequences were non-complementary. Each inactive mutant was mixed with another inactive mutant such that the loop sequences were complementary in trans, allowing the formation of intermolecular base pairing. All mutant pRNAs that had unpaired right and left loops, such as pRNA A-b, were inactive in phi29 assembly when used alone. However,  when two inactive pRNAs that were trans-complementary in their right and left loops, for example pRNA A-band B-a, were mixed in an equimolar ratio, full activity was restored. The observed activity of a mixture of two inactive mutant pRNAs confirmed that the right loop interacted with the left loop intermolecularly to form an RNA dimer.

3.4  Chemical modification

Chemical modification was employed to probe pRNA structure. The modifying agents used include dimethyl sulfate (DMS), which methylates A at N1,  G at N7 and C at N3[41, 42]kethoxal, which modifies G at N1 and N2[43]and 1-cyclohexyl-3-(2-morpholinoehtyl)-carbodiimide metho-p-toluene sulfonate (CMCT), which attacks U at N3 and G at N1[41-43]. In principle, only unpaired bases are susceptible to chemical attack. The chemicals alter unpaired specific functional groups of RNA bases and thus provide information regarding base pairing, base stacking, and the tertiary interactions of specific bases within an RNA. Locations of modified bases can be identified by primer extension with reverse transcriptase[41, 44]. Chemical modification of a base is a good indication that the base is unpaired and that the specific functional group is solvent-exposed, and thus is a possible candidate for intermolecular interactions. Lack of modification will most likely be due to base pairing, especially in helical regions,  but may also be the result of tertiary interactions or non-canonical base-base,  base-sugar, or base-phosphate interactions[43] in loop or bulge regions. Chemical modification data can provide information on base accessibility,  which is helpful in assessing predicted secondary structures,  evaluating 3-D molecular models,  and analyzing RNA/protein interactions.

Phi29 pRNAs including various mutants have been modified with DMS,  CMCT,  and kethoxal[27, 41, 43]. Chemical modification showed that the sequence C18C19A20,  which is essential for DNA packaging but dispensable for procapsid binding,  is accessible to chemicals in monomers and dimers as well as procapsid-bound pRNA[27, 30]. These results indicate that CCA,  though not involved in procapsid binding[28, 31],  is present on the surface of the pRNA as a bulge which may interact with other DNA packaging components[30] (Fig.5). This conclusion is supported by mutation studies on the CCA bulge.

Fig.5  Direct observation of pRNA three-dimensional structure with cryo-AFM (atomic force microscopy) (A and B) and shape comparision with computer models (C and D). E and F are drawings to depict the structure of monomers and dimers,  respectively

Reprinted from[27] with permission from RNA.

As noted earlier, the right (upper) loop sequence A45A46C47C48 and left (lower) loop sequence U85U84G83G82 of pRNA monomers are accessible to chemicals, as predicted by computer folding algorithms. However, when specific mutant pRNAs designed to form dimers were analyzed by chemical modification, the same sequences were protected from modification. Normally, pRNA B-awould be able to interact with A-bto form dimers. This lack of reactivity of the bases indicates that the loop sequences are involved in base pairing, and confirms that these bases are involved in inter-pRNA interaction[19, 20].  

3.5  Chemical modification interference

Chemical modification interference has been performed to determine which pRNA bases are involved in dimer formation. The monomer pRNA B-awas treated with either DMS or CMCT and then mixed with unmodified monomer A-bin order to test its competency in dimer formation. If the base is involved in dimer formation, chemical modification of this base could interfere with the ability of pRNA B-ato form a dimer with pRNA A-b, and thus this pRNA will be present in a fast migrating band representing monomers in native gels. Chemical modification was performed, and RNAs (both fast and slow migrating corresponding to pRNA monomers and dimers, respectively) were isolated from gels. After isolation,  both monomers and dimers were subjected to primer extension to identify the modified bases. The concentration of the chemicals was titrated to ensure that on the average only one base per pRNA was modified[45]. The general theory behind the experiment was that a pRNA B-acontaining an interfering modified base would appear in the fast migrating monomer band,  while pRNA B-acontaining a non-interfering modified base would appear in the slower migrating dimer band. Chemical modification interference analysis reveals that bases U54, G55, U59, C65, A66, A68, U69, A70, C71, C84, C85, C88, A89, A90 and C92 interfered with dimer formation, and thus are involved in dimerization, while bases 72-81 were not involved[45] as shown in the computer model of dimer (Fig.6).

Fig.6  Computer models of pRNA monomer (A), dimer (B), hexamer (C), and pRNA/connector complexes (E and F). D is the crytall structure of connector 11, 72

Reprinted from[56] with permission from J Biol Chem.

3.6  Photoaffinity crosslinking by psoralen

The chemical psoralen can intercalate into RNA or DNA helices and, upon irradiation with 320-400 nm light,  freeze (in helix or pseudoknot) uridines of RNA or the thymidines of DNA by covalent attachment46 if they are in close proximity (in helix or pseudoknot)[47, 48]. The sites of crosslinks can be determined by primer extension[49] and/or mung bean nuclease treatment[50]. The psoralen derivative, AMT (4-aminomethyl-4, 5, 8-trimethyl psoralen), was used to crosslink pRNA due to its solubility[49]. Psoralen crosslinks only RNA or DNA but not protein,  which is different from the azido group (see below) which crosslinks non-specifically to both protein and nucleic acids. Psoralen,  however,  can induce intra-molecular crosslinks within the pRNA even in the presence of other proteins, such as procapsid or gp16. Thus, pRNA conformations in different environments can be detected.  Psoralen crosslinks can also be reversed by 254 nm irradiation. With the use of a 2-dimensional gel electrophoresis[46, 48, 51] and 5-end radiolabeled cp-pRNAs,  pRNA conformational change in the presence of different packaging components can be investigated.  Psoralen crosslinking experiments revealed that pRNA had at least two conformations--one that was able to bind procapsid and the other that was not able to bind.  In the absence of Mg2+,  the region comprising bases C67 to A70 was in close proximity to bases U31 to U36,  since these two areas were crosslinked together by psoralen[49].

3.7  Photoaffinity crosslinking with GMPS/Aryl azide

Aryl azides contain functional groups that are chemically inert in the absence of light, but can be converted to a reactive nitrene after long wavelength UV irradiation[52, 53]. Thus, aryl azides can be incorporated into RNA to obtain structural data[54]. Aryl azide has been specifically attached to the 5-end of pRNAs or cp-pRNAs. For this 5-end labeling,  5-thiophosphate pRNA or cp-pRNA is synthesized by in vitro transcription in the presence of excess GMPS (guanine-monophosphorothioate) over GTP[53]. GMPS is an efficient primer in RNA synthesis with T7 RNA polymerase but cannot be used by this enzyme for chain elongation. The 5-thio-pRNA and the 5-thio-cpRNAs are then treated with azidophenacyl bromide to produce the 5-azido-pRNA and 5-azido-cp-pRNAs,  respectively,  by the nucleophilic displacement of bromine[53]. The azido group is converted to a reactive nitrene by long wavelength UV irradiation,  which is then inserted into nearby bonds resulting in covalent crosslinks[52]. Since it is possible to generate active cp-pRNAs by assigning certain internal sites of the pRNA as new 5- and 3-termini (Section 2)[28, 34], specific internal bases of the pRNA have been uniquely labeled with photoaffinity crosslinking agents to analyze inter- and intra-molecular interactions.  When necessary,  the 5-end of the RNA can also be labeled with [32P]. Crosslinked RNAs were separated from uncrosslinked RNAs by denaturing gel electrophoresis,  and crosslink sites were determined by primer extension[45, 55]. Bases identified as crosslink sites by primer extension indicate that these bases are in close proximity to the photoagent labeled base. The use of cp-pRNAs allows the identification of intra-molecular contacts throughout the pRNA molecule,  and such data have been used as distance constraints in molecular modeling studies[28, 34, 45, 55] (Section 3.13).

Intra-molecular crosslinking of monomers[45] revealed that G108 neighbors C10 and G11G75 is near bases 26-30,  while G78 is near U31. The azidophenacyl group is only 0.9 nm in length,  but experimental data has demonstrated that the cross-linking group can reach distances of 1.2  nm (Norman Pace,  personal communications). These distances have been used as constraints in the computer modeling of the pRNA monomer structure (Fig.7).

Fig.7  Comparison of chemical modification patterns of monomer (A) and dimer (B)

The black arrow, gray square, and double-lined arrow indicate a strong, moderate, and weak modification of bases,  respectively. C is a model to portray the formation of dimer. The four base-pairs (45-48/85-82 in gray boxes) were modified in monomers, but were protected from chemical modification in dimers (Adapted from[27, 45, 56] with permission from RNA and J Biol Chem).

Intermolecular crosslinking of dimers[55] was also achieved. Several studies revealed that G82 is in close proximity to G39,  G40, A41, C49, G62, C63, and C64[55]. Data from these crosslinking experiments have been used as distance constraints in molecular modeling of pRNA dimers[56] (Fig.7).

3.8  Photo-crosslinking by phenphi

Unlike psoralen, phenphi[(cis-Rh(phen)(phi)Cl2+ (phen = 1, 10-phenanthrolineand phi = 9, 10-phenanthrenequinone diimine)] induces covalent bonds between guanosine bases upon UV activation. Phenphi has also been shown to crosslink pRNA and has revealed the close proximity of bases G75, G28 and G30 of pRNA[57].

3.9  Ribonuclease probing

Some ribonucleases are sensitive to RNA secondary structure. For example, RNases T1 (specific for GpN linkages),  U2 (specific for ApN linkages), and S1 prefer to cleave single-stranded RNA. Nuclease V1 is specific for double stranded RNA. End-labeled pRNA and cp-pRNA in various solutions containing Mg2+ or procapsid individually or in combination,  have been probed by T1,  U2 or V1 nucleases[14, 49].  

T1 and V1 were used to distinguish the loops and helices of four RNAs with similar function[14]. RNase footprinting has also been performed to detect the sequences that bind procapsid[49, 58]. In addition, T1 nuclease has been used to study changes in pRNA conformation[49]. Since the activity of RNase T1 is Mg2+ independent, this enzyme was used to investigate the conformational change of pRNA in the presence or absence of Mg2+[49].

A Mg2+-induced pRNA conformational change was verified by T1 ribonuclease probing[49]. The pattern of partial digestion of pRNA by T1 provided strong evidence for the presence of two conformations, dependent on either the presence or absence of Mg2+. Without Mg2+, strong cleavages by T1 were seen at bases G28, G30, and G34. While in the presence of Mg2+, these three bases became more resistant to T1 attack, indicating a conformational change or refolding of pRNA stimulated by Mg2+[49].

3.10  Footprinting

Foot printing is a technique derived from nuclease probing or chemical modification and is particularly useful in probing the interaction of RNA with proteins. The procapsid/pRNA complex was probed with nucleases A, T1 and V1[58]. The optimal concentration of enzymes was determined empirically to ensure, on the average, one cleavage site per RNA molecule. Results of footprinting studies revealed that bases 22-84 were protected from enzyme digestion[58] indicating that the region from bases 22-84 contacts the procapsid.

3.11  SELEX (systematic evolution of ligands by exponential enrichment)

In vitro evolution is a powerful tool to study consensus elements of RNA structure and function. Starting with a library containing pRNA sequences with random mutations within a defined region, in vitro evolution techniques allow the selection of pRNA variants that can bind a specific ligand. Such selection for interacting species is based on different primary structures that can adopt the same structural feature as wild type RNA. SELEX allows screening for co-variation of several nucleotides and can be used to reveal noncanonical interactions that are difficult to prove by classic genetic and biochemical approaches[59, 60]. SELEX has been used for the selection of pRNA sequences that bind procapsids and are involved in intermolecular loop/loop interactions[61]. It was concluded that the wild type pRNA sequence is the most suitable sequence for procapsid binding.

3.12  Images revealed by cryo-AFM (atomic force microscopy)

    Atomic force microscopy has been used by several investigators to detect images of RNA (Fig.8) in a denatured conformation. As the first attempt to test whether this technique can be used to detect the 3D structure of RNA in native conformation, cryo-AFM has been performed on the phi29 pRNA monomer, dimer and trimer[27, 37, 45].

Fig.8  Computer model of pRNA dimer is in accord with the results of chemical modification interference

Bases that are demonstrated to interfere with dimer formation are shown as gray spacefill bases in the pRNA subunits. The dimer model is in agreement with the empirical data by showing that these bases are located at the interface of the two pRNAs of the dimer (adapted from[56] with permission from J Biol Chem).

Cryo-AFM imaging revealed that the pRNA monomer folds into a check mark-shaped structure. The native dimers appeared as elongated shapes (Fig.9)[27]. From this image, it can be suggested that head to head contact is involved in dimer formation,  resulting in a complex almost twice as long as the monomer. The trimer displays a trianglular shape under cryo-AFM[37]. The color or illumination indicates the thickness or height of the image but does not reflect the atom density observed end on. The brighter or whiter color indicates the thicker or taller the imagethe darker the color, the thinner the image. The color and contrast of the image clearly indicate that the area around the head loop (the elbow of the check mark) is the thickest or tallest, which agrees with the computer model of pRNA dimmers (see below). Cryo-AFM images of the fused dimer,  which is a pRNA construction consisting of two tandemly linked pRNAs,  exhibit a similar shape to the non-covalently linked pRNA dimer[45]. The dimensions of the covalently linked fused dimer are comparable to that of native dimer[27, 37].

Fig.9  Computer model of pRNA dimer are in accord with the results of intermolecular azidophenacyl photoaffinity crosslinking

G82 (in black spacefill) in one pRNA unit is in close proximity to G39, G40, A41, C49, G62, C63, and C64 (in gray wireframe) of the other pRNA unit, as determined by base-specific photo affinity crosslinking (adapted from[56] with permission from J Biol Chem).

3.13  Computer modeling of pRNA three-dimensional structure

The goal of modeling pRNA structure is to organize collections of structural data from crosslinking,  chemical or ribonuclease probing,  chemical modification interference,  cryo-AFM and other genetic data into a three-dimensional form.  Since a large number of structural constraints are available,  computer programs can successfully construct three-dimensional structures[56, 62, 63].  

pRNA monomer, dimer and hexamer (Fig.8) were produced on Silicon Graphics Octane and Indigo computers running IRIX 6.2 or 6.5,  using the programs NAHELIX,  MANIP,  PRENUC, NUCLIN, and NUCMULT[64, 65]. The modeling was performed based on the following assumptions(1) All helices were modeled as regular A-form double helices.  (2) Internal loops and mismatched bases were constructed by maintaining the integrity of the double helix while optimizing base pairing and stacking inside the loop, as suggested by most structural data from X-ray and NMR analysis. (3) A general rule for the modeling of the RNA hairpin loop has been proposed[66], which involves maximal stacking on the 3side of the stem and enough nucleotides stacked on the 5side to allow loop closure, as found in the anticodon loop of tRNA. (4) Bulges less than four bases in size were modeled either radiating out from stems to avoid helical distortion, while larger loops were constructed protruding from the stems or within the helical domain, causing the helical axis to bend. Parameters for stacking energy are considered in order to decide whether bulges should be protruding from or within the helical stems[67]. (5) Helix untwisting or twisting, helix-helix interactions, triple base interactions[68],  pseudoknots, or other higher order structures have been built into the model at constant geometrical distances while allowing certain torsion angle variation. The program regarding RNA flexibility has been applied to the construction of the pRNA UUU bulge at the three-helix junction. This three-base bulge has been found to provide flexibility for the appropriate folding of pRNA. Conventional computer algorithms involving the minimization of empirical energy functions have been considered. Twelve angstroms has been considered as a maximum distance constraint when bases are crosslinked by GMPS/aryl azide. Modified distance geometry and molecular mechanics algorithms using simplified pseudo atom representations have been considered to generate structures consistent with data from crosslinking, chemical modification and chemical modification interference. A constraint satisfaction algorithm is combined with discrete representations of nucleotide conformations to refine the disturbed area in order to ensure the normal representation of all atoms.

X-ray crystallography studies have revealed that the phi29 connector contains three sections, a narrow end, a central section, and a wider end, with diameters of 6.6 nm, 9.4 nm, 13.8 nm, res-pectively (Fig.8)[11, 69]. The hexameric pRNA model by Hoeprich and Guo[56] contains a central channel with a diameter of 7.6 nm, that perhaps can sheath onto the narrow end of the connector to perhaps be anchored by the central section of the connector, which is wider than the central channel of the pRNA hexamer (Fig.8).  

As noted earlier, pRNA contains two functional domains (Fig.1)one for connector binding and one for DNA translocation. The connector binding domain is located in the middle of the pRNA primary sequence,  i.e. bases 23-97,  and the DNA translocation domain is located at the 5/3 paired ends. It has been predicted that the connector protein (gp10) contains a conserved RNA recognition motif (RRM), located between residues 148-214 of each gp10 monomer.  This region of gp10 is located at the narrow end of the dodecameric connector that protrudes from the procapsid[70, 71]. The hexamer model by Hoeprich and Guo[56] complies with the aforementioned data by showing that pRNA bases 23-97 (colored green in Fig.6C,  E & F),  within the connector binding domain,  interact with the predicted RRM motifs of the connector (Fig.2E and F in blue),  while the 5/3 paired region (Fig.2E in red and cyan),  comprising the DNA translocation domain,  extends away from the connector.


Acknowledgements    I would like to thank Jane Kovach, Dan Shu and Stephen Hoeprich for the manuscript preparation,  Drs. Mark Trottier and Chaoping Chen for critical review, Dr. Zhifeng Shao for providing his AFM images, and Dr. Michael Rossmann for his permission to use his published connector structure.

References

1  Earnshaw WC,  Casjens SR. DNA packaging by the double-stranded DNA bacteriophages. Cell,  1980,  21  319-331

2  Black LW. DNA packaging in dsDNA bacteriophages. Ann Rev Microbio,  1989,  43  267-292

3  Guo P. Introduction  Principles,  perspectives,  and potential applications in viral assembly. Seminars in Virology (Editor's Introduction),  1994,  5(1)  1-3

4  Guo P. The mechanism of DNA packaging in double stranded DNA bacteriophages. Virologica Sinica,  1988,  2  229-246

5  Gong ZX. Molecular assembly of virus. Acta Biochim Biophys Sin,  1998,  30(5)  415-418

6  Guo P,  Erickson S,  Anderson D. A small viral RNA is required for in vitro packaging of bacteriophage f29 DNA. Science 1987  236  690-694

7  Guo P,  Trottier M. Biological and biochemical properties of the small viral RNA (pRNA) essential for the packaging of the double-stranded DNA of phage f29. Seminars in Virology,  1994,  5  27-37

8  Guo P. Structure and function of hexameric RNA to drive viral DNA packaging motor. Prog in Nucl Acid Res & Mole Biol,  2002  72  415-418

9  Guo P. Inhibition of novobiocin,  coumermycin A,  oxolinic acid,  nalidixic,  ethidium bromide and ATP analogue on the packaging of  f29 DNA in vitro. Virologica Sinica,  1988,  2  198-205

10  Jimenez J,  Santisteban A,  Carazo JM,  Carrascosa JL. Computer graphic display method for visualizing three-dimensional biological structures. Science,  1986,  232  1113-1115

11  Simpson AA,  Tao Y,  Leiman PG,  Badasso MO,  He Y,  Jardine PJ,  Olson NH et al. Structure of the bacteriophage phi29 DNA packaging motor. Nature,  2000,  408(6813)  745-750

12  Guo P,  Bailey S,  Bodley JW,  Anderson D. Characterization of the small RNA of the bacteriophage f29 DNA packaging machine. Nucleic Acids Res,  1987,  15  7081-7090

13  Garver K,  Guo P. Boundary of pRNA functional domains and minimum pRNA sequence requirement for specific connector binding and DNA packaging of phage phi29. RNA,  1997,  3  1068-1079

14  Bailey S,  Wichitwechkarn J,  Johnson D,  Reilly B,  Anderson D,  Bodley JW. Phylogenetic analysis and secondary structure of the Bacillus subtilis bacteriophage RNA required for DNA packaging. J Biol Chem,  1990,  265  22365-22370.

15  Zhang CL,  Tellinghuisen T,  Guo P. Conformation of the helical structure of the 5/3 termini of the essential DNA packaging pRNA of phage f29. RNA,  1995,  1  1041-1050

16  Chen C,  Guo P. Sequential action of six virus-encoded DNA-packaging RNAs during phage phi29 genomic DNA translocation. J Virol,  1997,  71(5)  3864-3871

17  Trottier M,  Zhang CL,  Guo P. Complete inhibition of virion assembly in vivo with mutant pRNA essential for phage f29 DNA packaging. J Virol,  1996,  70  55-61

18  Trottier M,  Guo P. Approaches to determine stoichiometry of viral assembly components. J Virol,  1997,  71  487-494

19  Guo P,  Zhang C,  Chen C,  Trottier M,  Garver K. Inter-RNA interaction of phage phi29 pRNA to form a hexameric complex for viral DNA transportation. Mol Cell,  1998,  2  149-155

20  Zhang F,  Lemieux S,  Wu X,  StArnaud D,  McMurray CT,  Major F,  Anderson D. Function of hexameric RNA in packaging of bacteriophage phi29 DNA in vitro. Mol Cell,  1998,  2  141-147

21  Hendrix RW. Bacteriophage DNA packaging  RNA gears in a DNA transport machine (Minireview). Cell,  1998,  94  147-150

22  Smith DE,  Tans SJ,  Smith SB,  Grimes S,  Anderson DL,  Bustamante C. The bacteriophage phi29 portal motor can package DNA against a large internal force. Nature,  2001,  413  748-752

23  Davenport R.J. Crossover research yield scents and sensitivity-watching a virus get stuffed. Science,  2001,  291  2071-2072

24  Valpuesta JM,  Fernandez JJ,  Carazo JM,  Carrascosa JL. The three-dimensional structure of a DNA translocating machine at 10 A resolution. Structure Fold Des,  1999,  7