Structure Article · 2017. 2. 14. · Structure Article Characterization of a Trifunctional...

12
Structure Article Characterization of a Trifunctional Mimivirus mRNA Capping Enzyme and Crystal Structure of the RNA Triphosphatase Domain Delphine Benarroch, 1 Paul Smith, 1 and Stewart Shuman 1, * 1 Molecular Biology Program, Sloan-Kettering Institute, New York, NY 10021, USA *Correspondence: [email protected] DOI 10.1016/j.str.2008.01.009 SUMMARY The RNA triphosphatase (RTPase) components of the mRNA capping apparatus are a bellwether of eukaryal taxonomy. Fungal and protozoal RTPases belong to the triphosphate tunnel metalloenzyme (TTM) family, exemplified by yeast Cet1. Several large DNA viruses encode metal-dependent RTPases unrelated to the cysteinyl-phosphatase RTPases of their metazoan host organisms. The origins of DNA virus RTPases are unclear because they are structurally uncharac- terized. Mimivirus, a giant virus of amoeba, resembles poxviruses in having a trifunctional capping enzyme composed of a metal-dependent RTPase module fused to guanylyltransferase (GTase) and guanine- N7 methyltransferase domains. The crystal structure of mimivirus RTPase reveals a minimized tunnel fold and an active site strikingly similar to that of Cet1. Un- like homodimeric fungal RTPases, mimivirus RTPase is a monomer. The mimivirus TTM-type RTPase-GTase fusion resembles the capping enzymes of amoebae, providing evidence that the ancestral large DNA virus acquired its capping enzyme from a unicellular host. INTRODUCTION The m 7 G cap structure of eukaryal mRNA promotes translation initiation and protects mRNA from degradation by 5 0 exoribonu- cleases. All eukaryal species and many eukaryal viruses share a three-step capping pathway in which (1) an RNA triphosphatase (RTPase) removes the g-phosphate of the primary transcript; (2) an RNA guanylyltransferase (GTase) transfers GMP from GTP to the 5 0 -diphosphate RNA to form a GpppRNA cap; and (3) a cap-specific RNA (guanine-N7) methyltransferase (MTase) adds a methyl group from AdoMet to the cap guanine to form the m 7 GpppRNA structure (Shuman, 2002). Lower and higher eukaryal taxa differ with respect to the struc- ture and mechanism of the RTPase component of the capping apparatus. Fungi and protozoa have a metal-dependent RTPase that belongs to the triphosphate tunnel metalloenzyme (TTM) superfamily (Gong et al., 2006). The RTPase branch of the TTM superfamily is exemplified by Saccharomyces cerevisiae Cet1, which displays the signature function of hydrolyzing NTPs to NDPs in the presence of manganese (Ho et al., 1998). The crystal structure of Cet1 (Lima et al., 1999) revealed a then-novel fold in which the active site is located in the center of a topologically closed eight-stranded antiparallel b barrel (the triphosphate tun- nel). The Cet1 active site is composed of 15 essential functional groups that either coordinate the metal ion or the g-phosphate or stabilize the tunnel architecture (Bisaillon and Shuman, 2001). Biochemical characterization and comparative mutational analy- ses strongly suggest that the RTPases of fungi (Candida albicans and Schizosaccharomyces pombe) and protozoan parasites (Trypanosoma brucei, Plasmodium falciparum, Encephalitozoon cuniculi, Giardia lamblia) belong to the same tunnel enzyme fam- ily as Cet1 (Pei et al., 2000, 2001; Hausmann et al., 2002, 2005a; Gong et al., 2003, 2006). In contrast, metazoa and plants have a metal-independent RTPase that belongs to the cysteine-phos- phatase enzyme superfamily (Takagi et al., 1997; Changela et al., 2001). Mammalian RTPase catalyzes phosphoanhydride hy- drolysis via a covalent protein-cysteinyl-S-phosphoester in- termediate. The cysteine nucleophile is located within a signature HCxxxxxR(S/T) motif. The tertiary structure of mammalian RTPase, comprising a central five-stranded parallel b sheet flanked by a helices (Changela et al., 2001), is completely unre- lated to that of the TTM-type RTPases of lower eukarya. Many of the largest eukaryal DNA viruses—baculoviruses, Af- rican swine fever virus (ASFV), poxviruses, Chlorella virus, Coc- colithovirus, and certain iridoviruses—encode some or all of the enzymes responsible for synthesis and capping of viral mRNAs. Most of the DNA virus-encoded RTPases that have been characterized are metal-dependent phosphohydrolases that share with yeast Cet1 the ability to cleave NTPs to NDPs in the presence of manganese or cobalt; these include RTPases of vaccinia virus, baculovirus, and Chlorella virus PBCV-1 (Fig- ure 1; Shuman et al., 1980; Jin et al., 1998; Gross and Shuman, 1998; Ho et al., 2001). Chlorella virus RTPase closely resembles yeast Cet1 with respect to its active site, mechanism, and its pre- dicted TTM fold (Gong and Shuman, 2002). Mutational studies of poxvirus and baculovirus RTPases suggest that their metal- binding sites resemble that of Cet1 (Martins and Shuman, 2001; Gong and Shuman, 2003). However, the structural and evolutionary relationships between the baculovirus and poxvirus RTPases and the TTM clade remain unclear, because there is no structure available for any DNA virus RTPase. Mimivirus is a recently isolated parasite of Acanthamoeba polyphaga (Raoult et al., 2007). Mimivirus has an extraordinarily large genome (1.2 Mb) encoding 911 predicted proteins (Raoult et al., 2004; Claverie et al., 2006). Mimivirus appears to specify its own mRNA synthetic machinery, which includes a multisubunit Structure 16, 501–512, April 2008 ª2008 Elsevier Ltd All rights reserved 501

Transcript of Structure Article · 2017. 2. 14. · Structure Article Characterization of a Trifunctional...

Page 1: Structure Article · 2017. 2. 14. · Structure Article Characterization of a Trifunctional Mimivirus mRNA Capping Enzyme and Crystal Structure of the RNA Triphosphatase Domain Delphine

Structure

Article

Characterization of a Trifunctional MimivirusmRNA Capping Enzyme and Crystal Structureof the RNA Triphosphatase DomainDelphine Benarroch,1 Paul Smith,1 and Stewart Shuman1,*1Molecular Biology Program, Sloan-Kettering Institute, New York, NY 10021, USA

*Correspondence: [email protected]

DOI 10.1016/j.str.2008.01.009

SUMMARY

The RNA triphosphatase (RTPase) components of themRNA capping apparatus are a bellwether of eukaryaltaxonomy. Fungal and protozoal RTPases belong tothe triphosphate tunnel metalloenzyme (TTM) family,exemplified by yeast Cet1. Several large DNA virusesencode metal-dependent RTPases unrelated to thecysteinyl-phosphatase RTPases of their metazoanhost organisms. The origins of DNA virus RTPasesare unclear because they are structurally uncharac-terized. Mimivirus, a giant virus of amoeba, resemblespoxviruses in having a trifunctional capping enzymecomposed of a metal-dependent RTPase modulefused to guanylyltransferase (GTase) and guanine-N7 methyltransferase domains. The crystal structureof mimivirus RTPase reveals a minimized tunnel foldand an active site strikingly similar to that of Cet1. Un-like homodimeric fungal RTPases, mimivirus RTPaseisa monomer.ThemimivirusTTM-type RTPase-GTasefusion resembles the capping enzymes of amoebae,providing evidence that the ancestral large DNA virusacquired its capping enzyme from a unicellular host.

INTRODUCTION

The m7G cap structure of eukaryal mRNA promotes translation

initiation and protects mRNA from degradation by 50 exoribonu-

cleases. All eukaryal species and many eukaryal viruses share a

three-step capping pathway in which (1) an RNA triphosphatase

(RTPase) removes the g-phosphate of the primary transcript; (2)

an RNA guanylyltransferase (GTase) transfers GMP from GTP

to the 50-diphosphate RNA to form a GpppRNA cap; and (3)

a cap-specific RNA (guanine-N7) methyltransferase (MTase)

adds a methyl group from AdoMet to the cap guanine to form

the m7GpppRNA structure (Shuman, 2002).

Lower and higher eukaryal taxa differ with respect to the struc-

ture and mechanism of the RTPase component of the capping

apparatus. Fungi and protozoa have a metal-dependent RTPase

that belongs to the triphosphate tunnel metalloenzyme (TTM)

superfamily (Gong et al., 2006). The RTPase branch of the TTM

superfamily is exemplified by Saccharomyces cerevisiae Cet1,

which displays the signature function of hydrolyzing NTPs to

NDPs in the presence of manganese (Ho et al., 1998). The crystal

Structur

structure of Cet1 (Lima et al., 1999) revealed a then-novel fold in

which the active site is located in the center of a topologically

closed eight-stranded antiparallel b barrel (the triphosphate tun-

nel). The Cet1 active site is composed of 15 essential functional

groups that either coordinate the metal ion or the g-phosphate or

stabilize the tunnel architecture (Bisaillon and Shuman, 2001).

Biochemical characterization and comparative mutational analy-

ses strongly suggest that the RTPases of fungi (Candida albicans

and Schizosaccharomyces pombe) and protozoan parasites

(Trypanosoma brucei, Plasmodium falciparum, Encephalitozoon

cuniculi, Giardia lamblia) belong to the same tunnel enzyme fam-

ily as Cet1 (Pei et al., 2000, 2001; Hausmann et al., 2002, 2005a;

Gong et al., 2003, 2006). In contrast, metazoa and plants have

a metal-independent RTPase that belongs to the cysteine-phos-

phatase enzyme superfamily (Takagi et al., 1997; Changela et al.,

2001). Mammalian RTPase catalyzes phosphoanhydride hy-

drolysis via a covalent protein-cysteinyl-S-phosphoester in-

termediate. The cysteine nucleophile is located within a signature

HCxxxxxR(S/T) motif. The tertiary structure of mammalian

RTPase, comprising a central five-stranded parallel b sheet

flanked by a helices (Changela et al., 2001), is completely unre-

lated to that of the TTM-type RTPases of lower eukarya.

Many of the largest eukaryal DNA viruses—baculoviruses, Af-

rican swine fever virus (ASFV), poxviruses, Chlorella virus, Coc-

colithovirus, and certain iridoviruses—encode some or all of

the enzymes responsible for synthesis and capping of viral

mRNAs. Most of the DNA virus-encoded RTPases that have

been characterized are metal-dependent phosphohydrolases

that share with yeast Cet1 the ability to cleave NTPs to NDPs

in the presence of manganese or cobalt; these include RTPases

of vaccinia virus, baculovirus, and Chlorella virus PBCV-1 (Fig-

ure 1; Shuman et al., 1980; Jin et al., 1998; Gross and Shuman,

1998; Ho et al., 2001). Chlorella virus RTPase closely resembles

yeast Cet1 with respect to its active site, mechanism, and its pre-

dicted TTM fold (Gong and Shuman, 2002). Mutational studies of

poxvirus and baculovirus RTPases suggest that their metal-

binding sites resemble that of Cet1 (Martins and Shuman,

2001; Gong and Shuman, 2003). However, the structural and

evolutionary relationships between the baculovirus and poxvirus

RTPases and the TTM clade remain unclear, because there is no

structure available for any DNA virus RTPase.

Mimivirus is a recently isolated parasite of Acanthamoeba

polyphaga (Raoult et al., 2007). Mimivirus has an extraordinarily

large genome (1.2 Mb) encoding 911 predicted proteins (Raoult

et al., 2004; Claverie et al., 2006). Mimivirus appears to specify its

own mRNA synthetic machinery, which includes a multisubunit

e 16, 501–512, April 2008 ª2008 Elsevier Ltd All rights reserved 501

Page 2: Structure Article · 2017. 2. 14. · Structure Article Characterization of a Trifunctional Mimivirus mRNA Capping Enzyme and Crystal Structure of the RNA Triphosphatase Domain Delphine

Structure

Mimivirus RNA Capping Enzyme

RNA polymerase and a putative 1170 amino acid mRNA capping

enzyme (GenBank code: Q5UQX1, hereafter named MimiCE)

that resembles the ASFV capping enzyme and the D1 subunit

of vaccinia virus capping enzyme. Vaccinia D1 is a multifunc-

tional 844 amino acid polypeptide composed of N-terminal

RTPase, central GTase, and C-terminal MTase domains (Higman

et al., 1994; Cong and Shuman, 1995; Mao and Shuman, 1996;

Myette and Niles, 1996a, 1996b; Gong and Shuman, 2003).

A similar modular domain order is observed in the 868 amino

acid ASFV capping enzyme (Pena et al., 1993). The aim of the

present study was to examine the biochemical properties of

MimiCE and the extent to which it resembles or differs from other

viral and cellular capping machineries.

We find that MimiCE has the three canonical enzymatic activ-

ities of other cap-forming enzymes: (1) a metal-dependent tri-

Figure 1. Domain Organization of DNA

Virus Capping Enzymes

(Top) The indicated families of large eukaryotic

DNA viruses encode either separate capping com-

ponents or multifunctional capping enzymes com-

posed of RTPase, GTase, and MTase catalytic

modules linked in cis. The linear order of the cata-

lytic modules in the primary structure is depicted

from the amino (left module) to carboxyl (right

module) ends. (Bottom) The amino acid sequence

of MimiCE from amino acids 250–634 is aligned to

that of the GTase domain of the mouse capping

enzyme Mce1. Gaps are indicated by dashes (-).

The positions of amino acid side chain identity/

similarity are denoted by a solid circle. Motifs I,

III, IIIa, IV, V, and VI are highlighted in shaded

boxes. Positions in MimiCE that were targeted

for mutational analysis in the present study are

denoted by a vertical line.

phosphatase; (2) a metal-dependent

guanylyltransferase; and (3) an AdoMet-

dependent cap guanine-N7 MTase.

Deletion analyses defined MimiCE-(1–

668) as an autonomous RTPase-GTase

domain and MimiCE-(638–1170) as an

MTase domain. Further deletions demar-

cated MimiCE-(1–237) as an autonomous

RTPase module. Alanine substitutions at

the four predicted metal-binding gluta-

mates (Glu37, Glu39, Glu212, Glu214)

abolished the phosphohydrolase activity

of MimiCE, thereby suggesting its mem-

bership in the metal-dependent fungal/

protozoal/viral triphosphatase family.

A major goal of this project was to gain

the first atomic structure of a DNA virus

member of the metal-dependent RTPase

family. We crystallized MimiCE-(1–237)

and determined its structure at 2.0 A res-

olution. The mimivirus RTPase adopts

a tunnel fold composed of eight antiparal-

lel b strands, and its active site resembles

that of yeast Cet1. However, unlike the

homodimeric fungal RTPases, mimivirus RTPase is a monomeric

enzyme. We discuss the evolutionary connections between DNA

virus capping systems and those of unicellular eukarya.

RESULTS

Characterization of the GTase Component of MimiCEGTP:RNA GTase (EC 2.7.7.50) catalyzes a reversible two-step

ping-pong reaction (Shuman and Hurwitz, 1981). The first

step entails nucleophilic attack of the enzyme at the a phospho-

rus of GTP to form a covalent enzyme-(lysyl-N)-GMP intermedi-

ate plus pyrophosphate. In the second step, the b-phosphate

of 50 diphosphate-terminated RNA attacks the enzyme-GMP

intermediate to form the GpppRNA cap and expel the lysine

nucleophile. Both partial reactions require a divalent cation

502 Structure 16, 501–512, April 2008 ª2008 Elsevier Ltd All rights reserved

Page 3: Structure Article · 2017. 2. 14. · Structure Article Characterization of a Trifunctional Mimivirus mRNA Capping Enzyme and Crystal Structure of the RNA Triphosphatase Domain Delphine

Structure

Mimivirus RNA Capping Enzyme

Figure 2. Guanylyltransferase Activity of

MimiCE-(1–668)

(A) Aliquots (3 mg) of tag-free wild-type MimiCE-

(1–668) (WT) and the K292A, D468A, K496A, and

K498A mutants were analyzed by SDS-PAGE.

Polypeptides were visualized by staining with

Coomassie blue dye. The positions and sizes

(in kDa) of marker proteins are indicated at the left.

(B) GTase reaction mixtures (20 ml) containing

50 mM Tris-HCl (pH 8.0), 5 mM DTT, 5 mM MgCl2,

82 nM [a-32P]GTP (5 mCi; 3000 Ci/mmol), and

100 ng of the different proteins were incubated

for 10 min at 37�C. The products were resolved

by SDS-PAGE. An autoradiograph of the gel is

shown. The radiolabeled enzyme-guanylate adduct

(EpG) is indicated at the left.

(C) GTP dependence. Reaction mixtures (20 ml)

containing 50 mM Tris-HCl (pH 8.0), 5 mM DTT,

5 mM MgCl2, 100 ng MimiCE-(1–668), and serial

2-fold dilutions of [a-32P]GTP (from a stock solu-

tion of 50 mM [a-32P]GTP; 200 Ci/mmol) to attain

the final concentrations as specified were incu-

bated for 10 min at 37�C. Enzyme-GMP formation

is plotted as a function of GTP concentration.

(D) Sedimentation of tag-free wild-type MimiCE-

(1–668) in a glycerol gradient with internal stan-

dards was performed as described in the Experi-

mental Procedures. Aliquots of even-numbered

fractions were analyzed by SDS-PAGE; the bot-

tom (heaviest) gradient fraction is at the left, and

the top (lightest) fraction is at the right. The identi-

ties of the polypeptides are indicated.

cofactor. Cellular and DNA virus GTases (i.e., poxvirus, Chlor-

ella virus, ASFV, and baculovirus GTases) comprise a branch

of the covalent nucleotidyl transferase superfamily, which in-

cludes DNA ligases (Shuman and Lima, 2004). Chlorella virus

and fungal GTases are monofunctional enzymes that collabo-

rate with separately encoded TTM-type RTPase enzymes to

perform the initial steps of the mRNA capping pathway (Fig-

ure 1). Baculovirus GTase comprises the C-terminal domain

of a bifunctional RTPase-GTase fusion protein in which the

metal-dependent baculovirus RTPase is located at the N termi-

nus (Figure 1).

Crystal structures of Chlorella virus and C. albicans GTases

showed that they are composed of an N-terminal nucleotidyl-

transferase domain and a C-terminal OB-fold domain (Hakans-

son et al., 1997; Fabrega et al., 2003). Within the nucleotidyl-

transferase domain is an NMP-binding pocket composed of

five peptide motifs (I, III, IIIa, IV, and V) that define the ligase/cap-

ping enzyme superfamily (Shuman and Lima, 2004). Another

conserved element, motif VI, is located within the OB domain.

Motif I (KxDG) contains the lysine nucleophile to which GMP be-

comes covalently linked in the first step of the capping reaction.

The other five motifs also contribute essential constituents of the

GTase active site (Sawaya and Shuman, 2003).

The segment of MimiCE from amino acid 250 to amino acid

634 resembles a segment found in members of the conserved

family of cellular and viral GTases. A pairwise alignment of this

portion of MimiCE to the GTase domain of mammalian capping

enzyme Mce1 highlights 69/397 positions of identity (20%) and

the presence in MimiCE of all 6 nucleotidyltransferase motifs

(Figure 1). Indeed, 14 of the amino acids that were found by

Structu

mutagenesis to be essential for the GTase activity of Mce1 (Sa-

waya and Shuman, 2003) are conserved in MimiCE.

To study the function and structure of MimiCE, we initially

attempted to produce the full-length protein in bacteria as

a His10-Smt3 fusion (Mossessova and Lima, 2000), but the yield

of soluble recombinant protein was unsatisfactory. Efforts

to produce a soluble His10-Smt3 fusion to the central portion of

MimiCE (amino acids 233–668) in bacteria were also unsuccess-

ful. Therefore, we expressed the N-terminal fragment MimiCE-

(1–668) that includes the entire segment homologous to GTases,

plus an N-terminal extension that has putative counterparts

of the signature metal-binding motifs of the TTM-like RTPase

superfamily (discussed below). We purified His10-Smt3-Mimi-

CE-(1–668) from a soluble bacterial lysate by Ni2+-agarose chro-

matography. The His10-Smt3 tag was removed by treatment

with the Smt3-specific protease Ulp1 (Mossessova and Lima,

2000), and the tag-free MimiCE-(1–668) protein was separated

from the tag by a second Ni2+-agarose chromatography step.

SDS-PAGE showed that the preparation contained a predomi-

nant polypeptide migrating at 74 kDa that corresponds to

MimiCE-(1–668), which has a calculated mass of 77.8 kDa

(Figure 2A).

GTase activity can be detected by label transfer from

[a-32P]GTP to the enzyme. Incubation of MimiCE-(1–668) with

[a-32P]GTP and magnesium resulted in the formation of an

SDS-stable �75 kDa nucleotidyl-protein adduct (Figure 2B).

Formation of the covalent intermediate required a divalent cation

cofactor (not shown). Magnesium supported the reaction with

5 mM [a-32P]GTP at pH 8.0 over a broad concentration range,

from 0.3 to 10 mM (not shown). Manganese and cobalt (5 mM)

re 16, 501–512, April 2008 ª2008 Elsevier Ltd All rights reserved 503

Page 4: Structure Article · 2017. 2. 14. · Structure Article Characterization of a Trifunctional Mimivirus mRNA Capping Enzyme and Crystal Structure of the RNA Triphosphatase Domain Delphine

Structure

Mimivirus RNA Capping Enzyme

also supported activity, although the yield of enzyme-GMP

adduct was �30% of the level supported by magnesium (not

shown). Cadmium, calcium, and zinc were even less effective

at a 5 mM concentration, and no activity was detected in the

presence of 5 mM copper (not shown). Enzyme-GMP formation

with 5 mM [a-32P]GTP and 5 mM magnesium was optimal at pH

7.0–8.0 in 50 mM Tris buffer; the yields at pH 5.5 (Tris acetate)

and pH 9.5 (Tris-HCl) were 15% and 45%, respectively, of the

peak value at pH 7.5 Tris-HCl (data not shown). The extent of

the reaction increased with [a-32P]GTP concentration up to 2.5 mM

and attained saturation at R5 mM (Figure 2C). We estimated that

�34% of the input MimiCE-(1–668) was labeled with GMP during

the in vitro reaction with 5 mM [a-32P]GTP.

The native size of MimiCE-(1–668) was gauged by zonal veloc-

ity sedimentation through a 15%–30% glycerol gradient. Marker

proteins catalase (248 kDa), BSA (66 kDa), and cytochrome c

(12 kDa) were included as internal standards. MimiCE-(1–668)

sedimented as a single discrete component that paralleled the

BSA standard (Figure 2D). We surmise that MimiCE-(1–668)

is a monomer. The GTase activity cosedimented with the

MimiCE-(1–668) polypeptide (not shown).

To probe whether the structural requirements of the mimivirus

GTase adhere to those established previously for homologous

cellular and DNA virus GTases (Wang et al., 1997; Sawaya and

Shuman, 2003), we introduced alanine mutations at four amino

acids of MimiCE-(1–668) that are putative constituents of the

GTase active site: Lys292 in motif I, Asp468 in motif IV, and

Lys496 and Lys498 in motif V (Figure 1). The Ala mutants were

produced in bacteria and were purified in parallel with wild-

type MimiCE-(1–668) (Figure 2A). The K292A, D468A, K496A,

and K498A changes abolished GTase activity (Figure 2B). The

requirement for Lys292 is expected in light of its imputed func-

tion as the site of GMP attachment to the GTase. By analogy

to structurally characterized members of the covalent nucleoti-

dyltransferase superfamily, Asp468 is implicated in binding the

divalent cation and, Lys496 and Lys498 are implicated in coordi-

nating the GTP phosphates (Shuman and Lima, 2004).

Characterization of the C-Terminal CapMethyltransferase DomainThe segment of MimiCE from amino acids 638–1006 strongly

resembles a segment in members of the conserved family of

RNA (guanine-N7) MTase enzymes found in all eukarya and pox-

viruses. Pairwise alignment of this portion of MimiCE to vaccinia

D1 reveals 82/377 positions of side chain identity (22%) (not

shown). Alignment to human cap MTase Hcm1 highlights 81/

402 positions of identity (20%; Figure S1, see the Supplemental

Data available with this article online). Six amino acids identified

by alanine scanning as essential for Hcm1 activity (Asp203,

Gly207, Asp211, Asp227, Arg239, and Tyr289 in Hcm1; Saha

et al., 1999) are conserved in the putative MTase domain of

MimiCE (Figure S1).

We produced the C-terminal MimiCE-(638–1170) polypeptide

that includes the MTase-like segment as a His10-Smt3 fusion and

purified it from a soluble bacterial lysate by Ni2+-agarose chro-

matography. The His10-Smt3 tag was removed, and the tag-

free MimiCE-(638–1170) was separated from the tag by a second

step of Ni2+-agarose chromatography. Initial experiments entail-

ing reaction of the protein with 50 mM [3H-CH3]AdoMet and 1 mM

504 Structure 16, 501–512, April 2008 ª2008 Elsevier Ltd All rights r

GDP, GTP, or GpppA and analysis of the reaction mixtures by

PEI-cellulose TLC and fluorography (Hausmann et al., 2005b;

Hausmann and Shuman, 2005) revealed that only GpppA served

as a methyl acceptor, and that the radiolabeled reaction product

comigrated with m7GpppA (not shown). The premethylated

cap dinucleotide m7GpppA did not serve as a methyl acceptor

(not shown).

The extent of methyl transfer from 50 mM [3H-CH3]AdoMet to

1 mM GpppA during a 60 min reaction at 37�C was proportional

to the amount of MimiCE-(638–1170) added. At saturating

enzyme, 80% of the input AdoMet was consumed. From the

slope of the protein titration curve, we estimated that 120 pmol

m7GpppA was formed per pmol protein (Figure 3A). This value

reflects a yield not a rate, insofar as we found that m7GpppA

increased during the initial 3 min and accumulated more slowly

thereafter (not shown), perhaps indicating inhibition by the

AdoHcy product. The extent of methyltransfer was proportional

to the concentration of the GpppA substrate up to 250 mM and

plateaued at R500 mM GpppA (Figure 3B). Half-maximal activity

was attained at �110 mM GpppA.

The native size of MimiCE-(638–1170) was gauged by sedi-

mentation in a 15%–30% glycerol gradient with internal stan-

dards. The MTase activity sedimented as a single component

with a peak at fraction 18 (Figure 3C), coincident with the sedi-

mentation profile of the 63 kDa MimiCE-(638–1170) polypeptide

(not shown) and close to the BSA peak in fraction 17 (Figure 3C).

A plot of the S values of the three standards versus fraction num-

ber yielded a straight line (not shown). An S value of 4.3 was de-

termined for MimiCE-(638–1170) by interpolation to the internal

standard curve. These results suggest that MimiCE-(638–1170)

is a monomer in solution.

An Autonomous N-Terminal TriphosphataseDomain of MimiCEThe N-terminal segment of MimiCE is notable for the presence

of two peptide motifs —36LEFEV40 and 209YELELEL215 —com-

posed of alternating hydrophobic and acidic residues that are

putative counterparts of the signature metal-binding motifs of

the fungal/protozoal/viral RTPase enzyme family. These motifs

were initially defined by mutational analysis of the vaccinia virus

and budding yeast RTPases (Ho et al., 1998), and they have

since been detected as essential components of the active sites

of all TTM superfamily phosphohydrolases. Thus prompted, we

tested a series of serially truncated N-terminal fragments of

MimiCE for the functional signature of the RTPase clade: the

hydrolysis of NTPs in the presence of manganese or cobalt.

We initially analyzed His10-Smt3-tagged versions of MimiCE-

(1–668), MimiCE-(1–275), MimiCE-(12–275), MimiCE-(1–265),

MimiCE-(1–255), and MimiCE-(1–237) that had been expressed

in bacteria and isolated from a soluble extract by Ni2+-agarose

chromatography. Each of the recombinant proteins catalyzed

the release of 32Pi from 1 mM [g-32P]ATP in the presence of

2.5 mM cobalt. The extent of ATP hydrolysis was proportional

to input enzyme, and the specific activities (32Pi released/ng pro-

tein) were similar for the six proteins tested (Figure S2A). When

the activities were calculated per pmol of input enzyme and

normalized to that of MimiCE-(1–668), defined as 100%, we

obtained the following relative activities for the truncated

derivatives: MimiCE-(1–275), 73%; MimiCE-(12–275), 64%;

eserved

Page 5: Structure Article · 2017. 2. 14. · Structure Article Characterization of a Trifunctional Mimivirus mRNA Capping Enzyme and Crystal Structure of the RNA Triphosphatase Domain Delphine

Structure

Mimivirus RNA Capping Enzyme

Figure 3. Characterization of the Mimivirus Cap Methyltransferase

(A) Methyltransferase reaction mixtures (10 ml) containing 50 mM Tris-HCl (pH 8.0), 5 mM DTT, 50 mM [3H-CH3]AdoMet, 1 mM GpppA, and protein as specified

were incubated for 60 min at 37�C. Aliquots (4 ml) were spotted on polyethyleneimine cellulose TLC plates, which were developed with 0.05 M (NH4)2SO4. The

AdoMet- and m7GpppA-containing portions of the lanes were cut out, and the radioactivity in each was quantified by liquid scintillation counting.

(B) Reaction mixtures (10 ml) containing 50 mM Tris-HCl (pH 8.0), 5 mM DTT, 50 mM [3H-CH3]AdoMet, 250 ng enzyme, and GpppA as specified were incubated for

60 min at 37�C. The extent of 3H-m7GpppA formation is plotted as a function of the GpppA concentration.

(C) Glycerol gradient sedimentation was performed as described in the Experimental Procedures. Aliquots (1 ml) of the even-numbered fractions were assayed for

cap MTase activity with 1 mM GpppA as the methyl acceptor. The sedimentation peaks of the internal markers catalase, BSA, and cytochrome c are denoted by

arrows.

MimiCE-(1–265), 66%, MimiCE-(1–255), 55%; and MimiCE-(1–

237), 52%. We surmise that the N-terminal 237 amino acid seg-

ment of MimiCE comprises an autonomous metal-dependent tri-

phosphatase domain, and that the N-terminal 11 amino acid

peptide is not essential for triphosphatase activity.

Further characterization of the phosphohydrolase activity was

performed by using tag-free MimiCE-(1–668). The rate of release

of 32Pi from 2 mM [g-32P]ATP in 2.5 mM cobalt at pH 6.5 was

identical to the rate of conversion of [a-32P]ATP to [a-32P]ADP

in a parallel reaction mixture containing the same concentration

of MimiCE-(1–668) (Figure S2B). We detected no formation of

[a-32P]AMP during the reaction. We conclude that MimiCE cata-

lyzes the hydrolysis of ATP to ADP and Pi. From the initial rates,

we calculated a turnover number of 165 s�1. In light of previous

studies of baculovirus RTPase showing that the divalent cation

cofactor requirement was strongly pH dependent (Martins and

Shuman, 2001), we surveyed MimiCE-(1–668) for its optimal

pH and optimal levels of cobalt, magnesium, and manganese

cofactors. Cobalt-dependent ATP hydrolysis was optimal at

pH 6.5 in 50 mM Tris buffer (not shown) at a concentration of

1.25–2.5 mM cobalt (Figure S2C). Magnesium-dependent

ATPase was optimal at pH 8.5 (not shown) at 10 mM magnesium

(Figure S2C). Activity with manganese was optimal at pH 7.5 (not

shown) at 1.25 mM manganese (Figure S2C). No activity was

detected in the absence of a divalent cation (Figure S2C).

The nucleotide specificity of MimiCE-(1–668) was tested

by malachite green colorimetric assay of the release of Pi from

1 mM ATP, GTP, CTP, UTP, dATP, dGTP, dCTP, and dTTP by

10 ng MimiCE-(1–668) at pH 6.5 in the presence of 2.5 mM

cobalt. All eight common nucleotides were hydrolyzed to the

following extents: ATP, 7 nmol; dATP, 6.9 nmol; UTP, 6.8 nmol;

dTTP, 8.0 nmol; GTP, 10.5 nmol; dGTP, 9.9 nmol; CTP, 7.5 nmol;

dCTP, 4.7 nmol. Thus, the MimiCE triphosphatase is relatively in-

different to the nucleoside base and deoxyribose versus ribose

Structur

sugars. The ATPase activity of MimiCE-(1–668) cosedimented

with the GTase activity in a glycerol gradient (data not shown).

We surmise that MimiCE has a metal-dependent NTP phospho-

hydrolase function similar to that of other DNA virus mRNA cap-

ping enzymes.

Mutational Effects on the TriphosphataseActivity of MimiCE-(1–237)Glu37, Glu39, Glu212, and Glu214 in the putative metal-binding

motifs, 36LEFEV40 and 209YELELEL215, were replaced individu-

ally by alanine in the context of the autonomous N-terminal do-

main MimiCE-(1–237). The E37A, E39A, E212A, and E214A pro-

teins were produced as His10-Smt3-tagged fusions and were

purified from soluble bacterial lysates by Ni2+-agarose chroma-

tography. The tags were removed, and the native proteins

were separated from the tags. The purity of the recombinant mu-

tant proteins was comparable to that of wild-type MimiCE-

(1–237), which has a calculated mass of 27.5 kDa (Figure 4A).

The E37A, E39A, E212A, and E214A mutants were unable to hy-

drolyze [g-32P]ATP, even at a level of input enzyme (50 ng) that

sufficed for quantitative release of 32Pi by wild-type MimiCE-

(1–237) (Figure 4B). The four essential glutamates are broadly

conserved in the RTPases encoded by fungi, protozoa, and other

DNA viruses. In the crystal structure of Cet1, three of the essen-

tial glutamates directly coordinate the manganese at the active

site, and the fourth essential glutamate coordinates a water mol-

ecule bound to the metal (Lima et al., 1999). The two b strands of

Cet1 containing the metal-binding glutamates comprise the

‘‘floor’’ of the tunnel and consist of alternating charged side

chains interdigitated with alternating aliphatic/aromatic side

chains. The hydrophilic face points into the tunnel and comprises

the metal-binding site, whereas the hydrophobic face points

down into the globular core of the protein upon which the tunnel

rests. This hydrophilic/hydrophobic sequence pattern is reprised

e 16, 501–512, April 2008 ª2008 Elsevier Ltd All rights reserved 505

Page 6: Structure Article · 2017. 2. 14. · Structure Article Characterization of a Trifunctional Mimivirus mRNA Capping Enzyme and Crystal Structure of the RNA Triphosphatase Domain Delphine

Structure

Mimivirus RNA Capping Enzyme

in MimiCE, suggesting that the structural context for its metal-

binding site is probably similar to that of Cet1.

Crystal Structure of the RTPase Domain of MimiCECrystals of the catalytically active native MimiCE-(1–237) protein

were grown by vapor diffusion against a precipitant solution con-

taining PEG4000 plus citrate and acetate buffers. Crystals be-

longing to hexagonal and monoclinic space groups grew under

identical conditions. SAD phases from a single hexagonal crystal

of SeMet-substituted MimiCE-(1–237) were exploited to derive

an initial model, which was then refined by using the diffraction

data from native hexagonal and monoclinic crystals. Both crystal

forms contained two protomers in the asymmetric unit, albeit

with different packing arrangements. The refined model of the

hexagonal form at 2.0 A resolution (R = 21.2; Rfree = 25.3) con-

tained 439 of the 474 amino acid residues of protomers A and

B (Table 1). There was no interpretable electron density in either

protomer for the N-terminal decapeptide, which was not essen-

tial for triphosphatase activity (Figure S1A). The refined structure

of the RTPase A protomer in the hexagonal space group was

used as a search model to derive phases for the monoclic crystal

Figure 4. Effects of Alanine Mutations on MimiCE-(1–237) Triphos-

phatase Activity

(A) Aliquots (5 mg) of tag-free wild-type MimiCE-(1–237) and the indicated mu-

tants were analyzed by SDS-PAGE. Polypeptides were visualized by staining

with Coomassie blue dye. The positions and sizes (in kDa) of marker proteins

are indicated on the left.

(B) Reaction mixtures (20 ml) containing 50 mM Tris acetate (pH 6.5), 2.5 mM

CoCl2, 1 mM [g-32P]ATP, and MimiCE-(1–237) proteins as specified were

incubated for 10 min at 37�C. 32Pi formation is plotted as a function of input

enzyme.

(C) Sedimentation of tag-free wild-type MimiCE-(1–237) in a glycerol gradient

with internal standards was performed as described in the Experimental Pro-

cedures. Aliquots of even-numbered fractions were analyzed by SDS-PAGE;

the bottom (heaviest) gradient fraction is at the left, and the top (lightest) frac-

tion is at the right. The identities of the polypeptides are indicated.

506 Structure 16, 501–512, April 2008 ª2008 Elsevier Ltd All rights

form. Refinement statistics for the monoclinic RTPase at 2.9 A

resolution are listed in Table 1. Our discussion of the RTPase

structure will focus on protomer A in the hexagonal space group

(Table 1).

The RTPase is composed of eight b strands and five a helices

(Figure 5). The strands comprise an antiparallel b barrel (Fig-

ure 5A) homologous to the triphosphate tunnel fold first de-

scribed for yeast Cet1 (Lima et al., 1999). Each of the b strands

serves as a barrel stave, as seen in the side views of the RTPase

fold shown in Figures 6C and 6D. The 121DIEIVYKN128 and133KLIGI137 elements of the b4 strand are interrupted by a short

non-b 129RGSG132 peptide (Figures 5A and 5E). (The Arg129 res-

idue is disordered in the B protomer of the hexagonal crystal; the130GSG132 tripeptide is disordered in the A protomer of the

monoclinic form.) Nonetheless, the b4 strand will be considered

as a single b element (Figure 5E) that corresponds to the fourth

b strand of the eight-stranded triphosphate tunnel of yeast

Cet1. Thus, the MimiCE RTPase tunnel has the characteristic

eight-strand architecture of the TTM superfamily (Figure 5F).

A space-filling surface model of the RTPase highlights the

through-and-through tunnel passage (Figure 5B). A trigonal-

shaped electronic density observed in the center of the tunnel

was modeled as acetate (present in the crystallization buffer)

and is depicted in the image in Figure 5B. Two a helices (a3

and a4) that connect the third and fourth tunnel strands are situ-

ated on the ‘‘roof’’ of the tunnel (Figures 5A and 5F). The three

helices (a1, a2, and a5) located below the tunnel ‘‘floor’’ provide

a hydrophobic core on which the hydrophobic back sides of

several of the tunnel strands rest. The a-helical base of the mim-

ivirus RTPase is analogous to the much larger globular pedestal

domain that supports the b barrel in yeast Cet1 (Lima et al.,

1999). The more complex structure of the Cet1 pedestal (Fig-

ure 5F) reflects the fact that it includes all of the components

of the extensive Cet1 homodimer interface as well as a short

module that mediates heterodimerization of Cet1 with the yeast

guanylyltransferase.

Mimivirus RTPase crystallized with two protomers in the

asymmetric unit. The A and B promoters of the hexagonal crys-

tal form were related by a 150� rotation and could be superim-

posed with an rsmd of 1.62 A over 214 Ca positions; this value

is higher than the 0.5 A root-mean-square deviation (rmsd)

value typical of protein structures with noncrystallographic

symmetry refined at 2.0 A resolution (Kleywegt, 1996). We

suspect that the mimivirus RTPase is flexible, owing to the lim-

ited number of buttressing structural elements for the tunnel

walls. The conformational freedom of the enzyme is also likely

to be higher when there is no triphosphate substrate filling the

tunnel aperture, as is the case presently. The interface of the

A and B protomers in the asymmetric unit of the hexagonal

crystal lattice comprised a buried surface area of 306 A2 per

protomer. A more extensive interaction surface was apparent

between two B protomers in adjacent asymmetric units that

were related by a crystallographic 2-fold axis (buried surface

area of 1745 A2 per protomer). However, no extensive contact

surface between protomers within or across the asymmetric

unit was evident in the monoclinic crystal lattice, suggesting

that the observed protomer-protomer interactions are the result

of crystal packing rather than a higher-order quaternary struc-

ture. The consensus for a monomeric structure in the crystals

reserved

Page 7: Structure Article · 2017. 2. 14. · Structure Article Characterization of a Trifunctional Mimivirus mRNA Capping Enzyme and Crystal Structure of the RNA Triphosphatase Domain Delphine

Structure

Mimivirus RNA Capping Enzyme

Table 1. Crystallographic Data and Refinement Statistics

Crystal SeMet–P64 Native–P64 Native–P21 Native–P1

Lattice (A) a = b = 154.18,

c = 39.59

a = b = 154.46,

c = 39.62

a = 39.18 b = 92.40,

c = 72.81, b = 102.6�a = 39.64, b = 48.46,

c = 69.70, a = 74.62�,

b = 86.48�, g = 76.36�

Crystallographic Data Quality (<I> � 1.0sI)

Resolution A–Overall

(highest shell)

40.0�2.6 (2.64�2.6) 40.0�2.0 (2 .03�2.0) 100.0�2.9 (3.02�2.9) 40.0�1.65 (1.68�1.65)

Beamline NSLS-X29 NSLS-X29 NSLS-X12C NSLS-X29

Wavelength 0.9793 A 1.1 A 1.6 A 1.21 A

Rsym 11.8% (62.5%) 6.0% (56.2%) 8.4% (29.7%) 4.8% (43.5%)

Unique reflectionsa 32,287 (1,643) 36,854 (1,763) 10,635 (1,077) 50,008 (1,929)

Mean redundancy 9.81 (5.5) 9.0 (4.2) 4.43 (2.5) 5.35 (1.2)

Completeness 99.6% (98.6%) 99.7% (97.4%) 94.4% (84.7%) 89.1% (78.6%)

Mean I/sI 17.73 (2.0) 41.1 (1.9) 14.91 (2.5) 58.0 (3.6)

Phasing Statistics

Phasing method SeMet SAD – MR-AMORE MR-AMORE

Resolution (A) 20.0�2.6 – 15.0�3.5 15.0�4.0

Anomalous S/Nb 1.64 – – –

Figures-of-meritc 0.26/0.69 – 0.48 0.38

Correlations from MRd – – 0.56(F) 0.65(I) 0.66(F) 0.68(I)

Refinement Statistics (<I> 1.0sI)

Resolution A (highest shell) – 40.0�2.0 (2.07�2.0) 50.0�2.9 (2.90�3.00) 40.0�1.65 (1.68�1.65)

Completeness – 93.4% (65.87%) 87.7% (61.7%) 88.6% (76.7%)

Rfreee – 25.3% (29.9%) 27.7% (29.9%) 21.2% (31.5%)

Rwork – 21.2% (27.8%) 22.3% (31.0%) 18.4% (29.4%)

Model Statistics

Rms bond length deviation – 0.0126 A 0.015 A 0.014 A

Rms bond angle deviation – 1.55� 1.80� 1.70�

NCS deviationf (Ca atoms) – 1.62 A (214) 1.02 A (217) 1.06 A (223)

Ramachandran

Core – 93.1% 90.4% 94.2%

Allowed – 6.9% 9.6% 5.8%

B factors (A2)

Chain A – 33.9 51.4 23.9

Chain B – 58.3 52.8 28.9

Heteroatoms – 51.0 41.0 41.6

Model Contents

Residues, chain A/B – 222/217 223/225 226/224

Alternate conformations – 31 residues – 30 residues

Heteroatoms – 1 citrate; 2 acetate – 5 acetate

Water molecules – 345 91 580

PDB ID – 2QY2 2QZE 3BGY

Standard definitions are used for all of the parameters. Figures in parentheses refer to data in the highest-resolution bin. The data collection statistics

come from SCALEPACK. The refinement and geometric statistics come from CNS, and the Ramachandran analyses were performed with

PROCHECK.a F+ and F� were treated as equivalent observations in all native data sets, but as distinct in the SeMet P64 data set.b Anomalous signal to noise ratio as output by SOLVE.c SAD phasing FOMs are from SOLVE/RESOLVE. The FOMs for molecular replacement structures are calculated by using the sA method as imple-

mented in CNS.d Correlation coefficients were derived by using AMORE after rigid body fitting.e Rfree sets for crossvalidation consisted of 10% of the data for the P64 crystal forms and 5% for both the P1 and P21 crystal forms chosen at random.f No NCS restraints were applied during refinement.

Structure 16, 501–512, April 2008 ª2008 Elsevier Ltd All rights reserved 507

Page 8: Structure Article · 2017. 2. 14. · Structure Article Characterization of a Trifunctional Mimivirus mRNA Capping Enzyme and Crystal Structure of the RNA Triphosphatase Domain Delphine

Structure

Mimivirus RNA Capping Enzyme

Figure 5. Tertiary Structure of Mimivirus RTPase

(A–F) The fold of mimivirus RTPase (amino acids 11–237) is depicted as a ribbon diagram in (A), (C), and (D); a helices are colored cyan, and b strands are colored

magenta. A view into the triphosphate tunnel is highlighted in (A). The N and C termini are indicated. The images in (C) and (D) are rotated clockwise and coun-

terclockwise, respectively, with respect to (A) in order to highlight side views of the staves of the b barrel. (B) shows a space-filling surface model in the same

orientation as (A) that highlights the tunnel aperture and an acetate molecule (depicted as a stick model) in the center of the tunnel. The primary structure is dis-

played in (E); secondary structure elements are highlighted in cyan for a helices and magenta for b strands. The putative metal-binding motifs are located

in strands b1 and b8; the essential glutamates are denoted by dots (�). The 121DIEIVYKN128 and 133KLIGI137 b segments that are interrupted by a short

non-b 129RGSG132 peptide (indicated by an asterisk in [A]) together comprise one of the barrel staves, which will be considered as a single b element (indicated

by brackets in [E]) that corresponds to the fourth b strand of the triphosphate tunnel of yeast Cet1. (F) shows a comparison of the topologies of mimivirus RTPase

and yeast Cet1. Tunnel b strands are shown as magenta pentagons oriented in the flat plane according to the view in (A), such that pentagons with the apices

pointing into the tunnel in (F) correspond to strands that project out from the page toward the viewer in (A), while pentagons with apices pointing away from the

tunnel in (F) are ones that project into the plane of the page in (A). The mimivirus RTPase a helices are shown as cyan circles, as are the corresponding a helices in

Cet1. Additional secondary structure elements unique to Cet1 are colored gray. A disordered chain break on the tunnel roof of mimivirus RTPase (from amino acid

155 to amino acid 157 in the loop connecting strands 5 and 6) is indicated by a dashed line. This segment is ordered in the A protomer of the monoclinic crystal, as

a result of crystal packing contacts unique to the monoclinic lattice. Chains breaks occurring at different sites in Cet1 are denoted by dashed lines.

was in agreement with glycerol gradient sedimentation analy-

sis, indicating that MimiCE-(1–237) is a monomer in solution

(Figure 4C). The 28 kDa MimiCE-(1–237) polypeptide sedi-

mented in a 15%–30% glycerol gradient as a single component

508 Structure 16, 501–512, April 2008 ª2008 Elsevier Ltd All rights re

with a peak between the cytochrome c (12 kDa) and BSA

(66 kDa) internal standards. The ATPase activity profile in the

glycerol gradient paralleled the abundance of the MimiCE-

(1–237) protein (data not shown).

served

Page 9: Structure Article · 2017. 2. 14. · Structure Article Characterization of a Trifunctional Mimivirus mRNA Capping Enzyme and Crystal Structure of the RNA Triphosphatase Domain Delphine

Structure

Mimivirus RNA Capping Enzyme

A DALI search (Holm and Sander, 1993) of the protein struc-

tural database recovered Cet1 as the closest homolog of mimi-

virus RTPase (z score, 13.4; rmsd, 3.9 A at 175 Ca positions), fol-

lowed by other tunnel-family proteins such as the Med18 subunit

of yeast mediator (z score 9.0; rmsd 3.5 A at 152 Ca positions),

Pyrococcus furiosus 1YEM (z score, 8.7; rmsd, 3.1 A at 131 Ca

positions), and the yeast Med20 mediator subunit (z score,

6.5; rmsd, 3.7 A at 133 Ca positions; Lariviere et al., 2006;

Gong et al., 2006).

The RTPase Active SiteFigure 6A shows a stereo view into the triphosphate tunnel of

MimiCE. The tunnel interior is dominated by a constellation of

hydrophilic amino acid side chains that project toward the tunnel

center and form a network of direct and water-mediated hydro-

gen bonds. A rim of positively charged functional groups ema-

nating from the tunnel roof and side walls surrounds an acetate

anion in the tunnel center (Figure 6A). We speculate that the

acetate mimics one of the phosphate groups of the triphosphate

substrate. The acetate in the MimiCE tunnel is analogous to the

sulfate anion coordinated by a similar network of positively

charged amino acids in the center of the Cet1 tunnel

(Figure 6B). The Cet1 sulfate is proposed to mimic the hydro-

lyzed g-phosphate of a Cet1-Pi product complex (Lima et al.,

Figure 6. MimiCE RTPase Active Site and

Comparison to Cet1

(A and B) Stereo views of the tunnel interiors of (A)

MimiCE-(1–237) and (B) yeast Cet1. Waters are

depicted as red spheres. The Cet1-bound manga-

nese ion is a cyan sphere. Acetate and sulfate ions

in the tunnels are rendered as stick models, as are

side chains emanating from the b strands that

comprise the tunnel walls.

1999). The acetate in MimiCE is situated

�5 A closer to the tunnel entrance than

the Cet1 sulfate when the two structures

are superimposed, suggesting that the

acetate might therefore be a more plausi-

ble mimetic of the b-phosphate. The

positively charged residues of MimiCE

that surround the acetate include Arg85

(in b3), Lys127 and Arg129 (in b4), and

Arg173, Lys175, and Arg177 (in b6; Fig-

ure 6A). These are the counterparts of

Cet1 tunnel residues Arg393, Lys409,

His411, Arg454, Lys456, and Arg458,

respectively (Figure 6B). Mutational analy-

ses of Cet1 have established that Arg454,

Arg393, Lys409, Lys456, and Arg458 are

critical for Cet1 activity (Pei et al., 1999; Bi-

saillon and Shuman, 2001).

Acidic residues predominate on the

floor of the MimiCE tunnel. These include

Glu37 and Glu39 (in b1), Glu149 (in b5),

Asp189 (in b7), and Glu210, Glu212, and

Glu214 (in b8; Figure 6A). Glu37, Glu39,

Glu212, and Gu214, which are required

for MimiCE TPase activity (Figure 4), correspond to Cet1 resi-

dues Glu305, Glu307, Glu494, and Glu496, respectively, which

coordinate the essential divalent cation (Figure 6B). Whereas

the Glu39, Glu212, and Gu214 side chains in MimiCE are dis-

posed in a manner consistent with metal coordination when

superimposed on Cet1, the Glu37 side chain is pointed away

from the putative metal-binding site (Figure 6A). Because Mim-

iCE was crystallized in the absence of divalent cation, we pre-

sume that Glu37 reorients when an effective metal cofactor is

available. Asp189 in MimiCE forms a salt bridge to its neighbor

Arg187 (Figure 6A). Asp189 corresponds to Cet1 Asp471, an es-

sential side chain that forms an ion pair with neighbor Arg469 and

makes a water-mediated contact to the manganese ion (Fig-

ure 6B). Glu149 in MimiCE is the counterpart of Cet1 Glu433,

an essential residue that is suggested to coordinate and activate

the nucleophilic water for its attack on the g-phosphate (Fig-

ure 6).

We made strenuous efforts to gain a structural snapshot of

a substrate complex, a product complex, a transition-state mi-

metic, or an inhibitor complex by growing crystals of mimivirus

RTPase in the presence of various combinations of cobalt, man-

ganese, ATP, AMPPNP, ADP, tungstate, vanadate, aluminum

fluoride, or tripolyphosphate. We also performed a full spectrum

of crystal soaks involving these ligands. Diffraction data were

Structure 16, 501–512, April 2008 ª2008 Elsevier Ltd All rights reserved 509

Page 10: Structure Article · 2017. 2. 14. · Structure Article Characterization of a Trifunctional Mimivirus mRNA Capping Enzyme and Crystal Structure of the RNA Triphosphatase Domain Delphine

Structure

Mimivirus RNA Capping Enzyme

collected on scores of candidate cocrystals and soaked crystals,

but we were frustrated in every case to find no convincing

evidence of occupancy of the active site by the relevant ligands

or metals. However, these trials did yield RTPase crystals in a

new triclinic space group (P1), grown from protein that had

been preincubated with 10 mM each of manganese, ADP, and

tungstate prior to mixing with precipitant solution containing

PEG4000 and citrate/acetate buffers. The triclinic crystal dif-

fracted to higher resolution (1.65 A) than any other crystals grown

previously (Table 1). The unit cell contained two RTPase proto-

mers. Other than the enhanced resolution limit, their refined

structure (R = 18.4; Rfree = 21.2; Table 1) did not differ signifi-

cantly from that of the hexagonal form. One of the RTPase pro-

tomers in the triclinic crystal had two planar-shaped densities in

the tunnel that we modeled as two acetate anions (not shown).

No density corresponding to nucleotide or tungstate was appar-

ent. Likewise, anomalous difference Fourier analyses with com-

plete data sets collected at appropriate wavelengths showed no

peaks for either manganese or tungsten.

DISCUSSION

The present study provides functional and structural insights to

mimivirus mRNA capping enzyme. Available knowledge of mimi-

virus biology and gene expression strategies is sparse com-

pared to other large DNA viruses such as vaccinia, the AcNPV

baculovirus, and Chlorella virus PBCV-1. Most publications con-

cerning mimivirus are driven by the genome sequence and pre-

dicted proteome. Several mimivirus enzymes involved in nucleic

acid metabolism have been studied as recombinant proteins

(Benarroch et al., 2005; Benarroch and Shuman, 2006; Jeudy

et al., 2006; Bandaru et al., 2007), and a few have been crystal-

lized (Abergel et al., 2007; Jeudy et al., 2005). The demonstration

here that MimiCE has the three catalytic functions found in cellu-

lar and viral capping systems underscores the likelihood that

MimiCE is indeed responsible for capping mimivirus mRNAs dur-

ing a natural infection. Because it is not yet feasible to genetically

engineer mimivirus, the role of MimiCE in the virus replication

cycle cannot be evaluated. Previous studies have exploited bud-

ding yeast as a surrogate system to demonstrate the bioactivity

of poxvirus capping enzymes (Ho et al., 2000; Saha et al., 2003).

However, our attempts to complement null mutations of the

yeast capping enzymes by expression of MimiCE (or component

domains) have not been successful.

The tripartite domain organization of MimiCE is most similar

to that of the well-characterized D1 subunit of vaccinia virus

capping enzyme. Like MimiCE, vaccinia D1 can be split into

two functional fragments: an N-terminal RTPase-GTase and

a C-terminal MTase (Cong and Shuman, 1992, 1995; Higman

et al., 1994; Mao and Shuman, 1994; Myette and Niles, 1996a,

1996b). The mimivirus and poxvirus RTPase-GTase fragments

are biochemically and structurally akin to the baculovirus

RTPase-GTase that caps late mRNAs synthesized by the bacu-

lovirus RNA polymerase, of which the capping enzyme is a sub-

unit (Guarino et al., 1998). A common feature of the mimivirus

and baculovirus RTPase-GTase enzymes is that their N-terminal

modules comprise autonomous triphosphatase catalytic do-

mains (Martins and Shuman, 2001; present study). This is not

the case for vaccinia capping enzyme (Myette and Niles,

510 Structure 16, 501–512, April 2008 ª2008 Elsevier Ltd All rights

1996a, 1996b). The MTase domain of vaccinia D1 has a feeble

intrinsic MTase activity on its own and requires a stimulatory

subunit (D12) for full biological and biochemical activity in cap

methylation (Higman et al., 1994; Mao and Shuman, 1994;

Schwer et al., 2006). Although we do not exclude the existence

of an accessory subunit for MimiCE, it is noteworthy that (1)

the mimivirus proteome includes no apparent homolog of poxvi-

rus D12 and (2) the MTase activity of the recombinant MimiCE-

(638–1170) protein in vitro is higher than that of the vaccinia

D1 MTase domain alone (Schwer et al., 2006). The mimivirus cap

MTase differs in another interesting respect from the poxvirus

MTase, in that it requires a cap dinucleotide as a methyl accep-

tor. By contrast, vaccinia cap MTase is adept at methylating

GTP, which is not an effective substrate for MimiCE. In this

regard, mimivirus MTase resembles T. brucei RNA (guanine-N7)

MTase, which uses GpppA or GpppG, but not GTP, as a sub-

strate (Hall and Ho, 2006).

The most instructive aspect of our study of MimiCE concerns

the triphosphatase module, which we demonstrate functionally

and structurally to be a bona fide member of the TTM family. A

salient difference between mimivirus and yeast RTPases is that

MimiCE-(1–237) is a monomer, whereas Cet1 is a homodimer

(Lehman et al., 1999; Lima et al., 1999). The RTPase of fission

yeast Schizosaccharomyces pombe is also a homodimer (Pei

et al., 2000). Homodimerizerization of fungal RTPases is not

required for their catalytic activity (Lehman et al., 1999). Nonethe-

less, dimer formation stabilizes the fungal RTPase against

thermal inactivation and is critical for RTPase function in vivo

(Lehman et al., 2001; Hausmann et al., 2003). The T. brucei

RTPase TbCet1 is the only case to date in which a cellular

TTM-type RTPase was reported to have a monomeric quaternary

structure (Gong et al., 2003). TbCet1 is extraordinarily thermosta-

ble in vitro (Gong et al., 2003).

The mimivirus RTPase module (amino acids 11–237) repre-

sents a minimized TTM-family triphosphatase. It is comparable

in size to the smallest catalytically active fragment (amino acids

34–252) of TbCet1 (Gong et al., 2003). The slightly greater size of

the minimal active fragment of yeast Cet1 (amino acids 275–539;

Lehman et al., 1999) is attributable to the larger pedestal domain

underlying the Cet1 triphosphate tunnel, which includes an extra

a helix distal to the last tunnel strand that is not present in Mim-

iCE and a 30 amino acid segment between b1 and a2 that has no

counterpart in MimiCE (Figure 5F). The topology of the tunnel

roof also differs in MimiCE versus Cet1. Whereas the mimivirus

roof consists of two a helices that connect the third and fourth

tunnel strands, the Cet1 roof comprises a helix-containing seg-

ment that connects the fifth and sixth strands of the tunnel

(Lima et al., 1999; Figure 5F). Notwithstanding these differences

in how the tunnel is embellished, a detailed comparison of the

MimiCE and Cet1 structures shows that the active sites are con-

served with respect to their tunnel fold and constituent functional

groups (Figure 6). This is striking given that a psi-blast search

with MimiCE failed to retrieve Cet1 (or any other fungal RTPase).

The present evidence that MimiCE belongs to the TTM

RTPase clade consolidates the case for a common evolutionary

origin for fungal/protozoal and DNA virus capping systems. It is

striking that the RTPase(TTM)-GTase enzymes of poxviruses,

baculoviruses, and Chlorella virus have a different evolutionary

history from the RTPase(cysteinyl-phosphatase)-GTase

reserved

Page 11: Structure Article · 2017. 2. 14. · Structure Article Characterization of a Trifunctional Mimivirus mRNA Capping Enzyme and Crystal Structure of the RNA Triphosphatase Domain Delphine

Structure

Mimivirus RNA Capping Enzyme

enzymes of their respective mammalian, arthropod, and green

algal hosts. The implication is that these metazoan and algal

DNA viruses evolved from an ancestor that replicated in (and

acquired its capping apparatus from) a unicellular eukaryon.

We have discussed previously the hypothesis that omnivorous

amoebae serve as a ‘‘melting pot’’ for horizontal gene transfer

among intracellular pathogens (Benarroch et al., 2005). Now,

with firmer knowledge of the domain organization and structure

of the mimivirus capping enzyme, and with the newly available

genome sequences of two amoebal species, Entamoeba histoly-

tica and Dictyostelium discoideum (Loftus et al., 2005; Eichinger

et al., 2005), we can query the relationship between candidate

capping enzymes of amoeba (readily identified in a database

search with any guanylyltransferase) and that of the amoebal

virus. Figure S3 shows an alignment of E. histolytica and

D. discoideum capping enzymes. Both are bifunctional proteins

composed of a C-terminal GTase domain fused to a putative

N-terminal TTM-type RTPase domain. The amoebal RTPase do-

main includes the signature metal-binding motifs located in the

first and eighth strands of the triphosphate tunnel, as well as pu-

tative equivalents of the fifth, sixth, and seventh tunnel strands

(Figure S3). To our knowledge, this is the first example of concor-

dance of viral and host cell mRNA capping enzymes with respect

to fusion of the triphosphatase to the guanylyltransferase and

membership of the RTPase in the TTM family.

EXPERIMENTAL PROCEDURES

See the Supplemental Data for a description of Experimental Procedures.

ACCESSION NUMBERS

Coordinates have been deposited in the Protein Data Bank with accession

codes 2QY2, 2QZE, and 3BGY.

SUPPLEMENTAL DATA

Supplemental Data include the methods employed for the production and

purification of recombinant MimiCE; biochemical assays of triphosphatase,

guanylyltransferase, and MTase activities; crystallization of the MimiCE

RTPase domain; and determination of the RTPase structure and are available

at http://www.structure.org/cgi/content/full/16/4/501/DC1/.

ACKNOWLEDGMENTS

This work was supported by National Institutes of Health grant GM42498. S.S.

is an American Cancer Society Research Professor.

Received: August 21, 2007

Revised: January 2, 2008

Accepted: January 3, 2008

Published: April 8, 2008

REFERENCES

Abergel, C., Rundinger-Thirion, J., Giege, R., and Claverie, J.M. (2007). Virus-

encoded amino-acyl tRNA synthetases: structural and functional characteriza-

tion of mimivirus TyrRS and MetRS. J. Virol. 81, 12406–12417.

Bandaru, V., Zhao, X., Newton, M.R., Burrows, C.J., and Wallace, S.S. (2007).

Human endonuclease VIII-like (NEIL) proteins in the giant DNA mimivirus. DNA

Repair (Amst.) 6, 1629–1641.

Benarroch, D., and Shuman, S. (2006). Characterization of mimivirus

NAD+-dependent DNA ligase. Virology 353, 133–143.

Structu

Benarroch, D., Claverie, J.M., Raoult, D., and Shuman, S. (2005). Characteriza-

tion of mimivirus DNA topoisomerase IB suggests horizontal gene transfer

between eukaryal viruses and bacteria. J. Virol. 80, 314–321.

Bisaillon, M., and Shuman, S. (2001). Structure-function analysis of the active

site tunnel of yeast RNA triphosphatase. J. Biol. Chem. 276, 17261–17266.

Changela, A., Ho, C.K., Martins, A., Shuman, S., and Mondragon, A. (2001).

Structure and mechanism of the RNA triphosphatase component of mamma-

lian mRNA capping enzyme. EMBO J. 20, 2575–2586.

Claverie, J.M., Ogata, H., Audic, S., Abergel, C., Suhre, K., and Fournier, P.E.

(2006). Mimivirus and the emerging concept of ‘‘giant’’ virus. Virus Res. 117,

133–144.

Cong, P., and Shuman, S. (1992). Methyltransferase and subunit association

domains of vaccinia virus mRNA capping enzyme. J. Biol. Chem. 267,

16424–16429.

Cong, P., and Shuman, S. (1995). Mutational analysis of mRNA capping

enzyme identifies amino acids involved in GTP binding, enzyme-guanylate

complex formation, and GMP transfer to RNA. Mol. Cell. Biol. 15, 6222–6231.

Eichinger, L., Pachebat, J.A., Glockner, G., Rajandream, M.A., Sucgang, R.,

Berriman, M., Song, J., Olsen, R., Szafranski, K., Xu, Q., et al. (2005). The

genome of the social amoeba Dictyostelium discoideum. Nature 435, 43–57.

Fabrega, C., Shen, V., Shuman, S., and Lima, C.D. (2003). Structure of an

mRNA capping enzyme bound to the phosphorylated carboxyl-terminal

domain of RNA polymerase II. Mol. Cell 11, 1549–1561.

Gong, C., and Shuman, S. (2002). Chlorella virus RNA triphosphatase:

mutational analysis and mechanism of inhibition by tripolyphosphate. J. Biol.

Chem. 277, 15317–15324.

Gong, C., and Shuman, S. (2003). Mapping the active site of vaccinia virus RNA

triphosphatase. Virology 309, 125–134.

Gong, C., Martins, A., and Shuman, S. (2003). Structure-function analysis of

Trypanosoma brucei RNA triphosphatase and evidence for a two-metal mech-

anism. J. Biol. Chem. 278, 50843–50852.

Gong, C., Smith, P., and Shuman, S. (2006). Structure-function analysis of

Plasmodium RNA triphosphatase and description of a triphosphate tunnel

metalloenzyme superfamily that includes Cet1-like RNA triphosphatases and

CYTH proteins. RNA 12, 1468–1474.

Gross, C.H., and Shuman, S. (1998). RNA 50-triphosphatase, nucleoside

triphosphatase, and guanylyltransferase activities of baculovirus LEF-4

protein. J. Virol. 72, 10020–10028.

Guarino, L.A., Jin, J., and Dong, W. (1998). Guanylyltransferase activity of the

LEF-4 subunit of baculovirus RNA polymerase. J. Virol. 72, 10003–10010.

Hakansson, K., Doherty, A.J., Shuman, S., and Wigley, D.B. (1997). X-ray crys-

tallography reveals a large conformational change during guanyl transfer by

mRNA capping enzymes. Cell 89, 545–553.

Hall, M.P., and Ho, C.K. (2006). Characterization of a Trypanosoma brucei RNA

cap (guanine N-7) methyltransferase. RNA 12, 488–497.

Hausmann, S., and Shuman, S. (2005). Giardia lamblia RNA cap guanine-N2

methyltransferase (Tgs2). J. Biol. Chem. 280, 32101–32106.

Hausmann, S., Vivares, C.P., and Shuman, S. (2002). Characterization of the

mRNA capping apparatus of the microsporidian parasite Encephalitozoon

cuniculi. J. Biol. Chem. 277, 96–103.

Hausmann, S., Pei, Y., and Shuman, S. (2003). Homodimeric quaternary struc-

ture is required for the in vivo function and thermal stability of Saccharomyces

cerevisiae and Schizosaccharomyces pombe RNA triphosphatases. J. Biol.

Chem. 278, 30487–30496.

Hausmann, S., Altura, M.A., Witmer, M., Singer, S.M., Elmendorf, H.G., and

Shuman, S. (2005a). Yeast-like mRNA capping apparatus in Giardia lamblia.

J. Biol. Chem. 280, 12077–12086.

Hausmann, S., Zheng, S., Fabrega, C., Schneller, S.W., Lima, C.D., and

Shuman, S. (2005b). Encephalitozoon cuniculi mRNA cap (guanine-N7) meth-

yltransferase: methyl acceptor specificity, inhibition by AdoMet analogs, and

structure-guided mutational analysis. J. Biol. Chem. 280, 20404–20412.

Higman, M.A., Christen, L.A., and Niles, E.G. (1994). The mRNA (guanine-7-)

methyltransferase domain of the vaccinia virus mRNA capping enzyme:

re 16, 501–512, April 2008 ª2008 Elsevier Ltd All rights reserved 511

Page 12: Structure Article · 2017. 2. 14. · Structure Article Characterization of a Trifunctional Mimivirus mRNA Capping Enzyme and Crystal Structure of the RNA Triphosphatase Domain Delphine

Structure

Mimivirus RNA Capping Enzyme

expression in Escherichia coli and structural and kinetic comparison to the in-

tact capping enzyme. J. Biol. Chem. 269, 14974–14981.

Ho, C.K., Pei, Y., and Shuman, S. (1998). Yeast and viral RNA 50 triphospha-

tases comprise a new nucleoside triphosphatase family. J. Biol. Chem. 273,

34151–34156.

Ho, C.K., Martins, A., and Shuman, S. (2000). A yeast-based genetic system

for functional analysis of viral mRNA capping enzymes. J. Virol. 74, 5486–5494.

Ho, C.K., Gong, C., and Shuman, S. (2001). RNA triphosphatase component of

the mRNA capping apparatus of Paramecium bursaria Chlorella virus 1.

J. Virol. 75, 1744–1750.

Holm, L., and Sander, C. (1993). Protein structure comparison by alignment of

distance matrices. J. Mol. Biol. 233, 123–138.

Jeudy, S., Coutard, B., Lebrun, R., and Abergel, C. (2005). Acanthamoeba

polyphaga mimivirus NDK: preliminary crystallographic analysis of the first viral

nucleoside diphosphate kinase. Acta Crystallogr. Sect. F Struct. Biol. Cryst.

Commun. 61, 569–572.

Jeudy, S., Claverie, J.M., and Abergel, C. (2006). The nucleoside diphosphate

kinase from mimivirus: a peculiar affinity for deoxypyrimidine nucleotides.

J. Bioenerg. Biomembr. 38, 247–254.

Jin, J., Dong, W., and Guarino, L.A. (1998). The LEF-4 subunit of baculovirus

RNA polymerase has RNA 50- triphosphatase and ATPase activities. J. Virol.

72, 10011–10019.

Kleywegt, G. (1996). Use of non-crystallographic symmetry in protein structure

refinement. Acta Crytallogr. D Biol. Crystallogr. 52, 842–857.

Lariviere, L., Geiger, S., Hoeppner, S., Rother, S., Strasser, K., and Cramer, P.

(2006). Structure and TBP binding of the mediator head subcomplex Med8-

Med18-Med20. Nat. Struct. Mol. Biol. 13, 895–901.

Lehman, K., Schwer, B., Ho, C.K., Rouzankina, I., and Shuman, S. (1999).

A conserved domain of yeast RNA triphosphatase flanking the catalytic core

regulates self-association and interaction with the guanylyltransferase compo-

nent of the mRNA capping apparatus. J. Biol. Chem. 274, 22668–22678.

Lehman, K., Ho, C.K., and Shuman, S. (2001). Importance of homodimeriza-

tion for the in vivo function of yeast RNA triphosphatase. J. Biol. Chem. 276,

14996–15002.

Lima, C.D., Wang, L.K., and Shuman, S. (1999). Structure and mechanism of

yeast RNA triphosphatase: an essential component of the mRNA capping

apparatus. Cell 99, 533–543.

Loftus, B., Anderson, I., Davies, R., Alsmark, U.C., Samuelson, J., Amedeo, P.,

Roncaglia, P., Berriman, M., Hirt, R.P., Mann, B.J., et al. (2005). The genome of

the protist parasite Entamoeba histolytica. Nature 433, 865–868.

Mao, X., and Shuman, S. (1994). Intrinsic RNA (guanine-7) methyltransferase

activity of the vaccinia virus capping enzyme D1 subunit is stimulated by the

D12 subunit: identification of amino acid residues in the D1 protein required

for subunit association and methyl group transfer. J. Biol. Chem. 269,

24472–24479.

Mao, X., and Shuman, S. (1996). Vaccinia virus mRNA (guanine-7-) methyl-

transferase: mutational effects on cap methylation and AdoHcy-dependent

photocrosslinking of the cap to the methyl acceptor site. Biochemistry 35,

6900–6910.

Martins, A., and Shuman, S. (2001). Mutational analysis of baculovirus capping

enzyme Lef4 delineates an autonomous triphosphatase domain and structural

determinants of divalent cation specificity. J. Biol. Chem. 276, 45522–45529.

Mossessova, E., and Lima, C.D. (2000). Ulp1-SUMO crystal structure and ge-

netic analysis reveal conserved interactions and a regulatory element essential

for cell growth in yeast. Mol. Cell 5, 865–876.

Myette, J.R., and Niles, E.G. (1996a). Domain structure of the vaccinia virus

mRNA capping enzyme: expression in Escherichia coli of a subdomain pos-

512 Structure 16, 501–512, April 2008 ª2008 Elsevier Ltd All rights r

sessing the RNA 50-triphosphatase and guanylyltransferase activities and a ki-

netic comparison to the full-size enzyme. J. Biol. Chem. 271, 11936–11944.

Myette, J.R., and Niles, E.G. (1996b). Characterization of the vaccinia virus

RNA 50-triphosphatase and nucleoside triphosphate phosphohydrolase activ-

ities: demonstration that both activities are carried out at the same active site.

J. Biol. Chem. 271, 11945–11952.

Pei, Y., Ho, C.K., Schwer, B., and Shuman, S. (1999). Mutational analyses of

yeast RNA triphosphatases highlight a common mechanism of metal-depen-

dent NTP hydrolysis and a means of targeting enzymes to pre-mRNAs

in vivo by fusion to the guanylyltransferase component of the capping appara-

tus. J. Biol. Chem. 274, 28865–28874.

Pei, Y., Lehman, K., Tian, L., and Shuman, S. (2000). Characterization of

Candida albicans RNA triphosphatase and mutational analysis of its active

site. Nucleic Acids Res. 28, 1885–1892.

Pei, Y., Schwer, B., Hausmann, S., and Shuman, S. (2001). Characterization of

Schizosaccharomyces pombe RNA triphosphatase. Nucleic Acids Res. 29,

387–396.

Pena, L., Yanez, R.J., Revilla, Y., Vinuela, E., and Salas, M.L. (1993). African

swine fever virus guanylyltransferase. Virology 193, 319–328.

Raoult, D., Audic, S., Robert, C., Abergel, C., Renesto, P., Ogata, H., La Scola,

B., Suzan, M., and Claverie, J.M. (2004). The 1.2-megabase genome sequence

of Mimivirus. Science 306, 1344–1350.

Raoult, D., La Scola, B., and Birtles, R. (2007). The discovery and characteriza-

tion of mimivirus, the largest known virus and putative pneumonia agent. Clin.

Infect. Dis. 45, 95–102.

Saha, N., Schwer, B., and Shuman, S. (1999). Characterization of human,

Schizosaccharomyces pombe and Candida albicans mRNA cap methyltrans-

ferases and complete replacement of the yeast capping apparatus by mam-

malian enzymes. J. Biol. Chem. 274, 16553–16562.

Saha, N., Shuman, S., and Schwer, B. (2003). Yeast-based genetic system for

functional analysis of poxvirus mRNA cap methyltransferase. J. Virol. 77,

7300–7307.

Sawaya, R., and Shuman, S. (2003). Mutational analysis of the guanylyltrans-

ferase component of mammalian mRNA capping enzyme. Biochemistry 42,

8240–8249.

Schwer, B., Hausmann, S., Schneider, S., and Shuman, S. (2006). Poxvirus

cap methyltransferase: bypass of the requirement for the stimulatory subunit

by mutations in the catalytic subunit and evidence for intersubunit allostery.

J. Biol. Chem. 281, 18953–18960.

Shuman, S. (2002). What messenger RNA capping tells us about eukaryotic

evolution. Nat. Rev. Mol. Cell Biol. 3, 619–625.

Shuman, S., and Hurwitz, J. (1981). Mechanism of mRNA capping by vaccinia

virus guanylyltransferase: characterization of an enzyme-guanylate intermedi-

ate. Proc. Natl. Acad. Sci. USA 78, 187–191.

Shuman, S., and Lima, C.D. (2004). The polynucleotide ligase and RNA cap-

ping enzyme superfamily of covalent nucleotidyltransferases. Curr. Opin.

Struct. Biol. 14, 757–764.

Shuman, S., Surks, M., Furneaux, H., and Hurwitz, J. (1980). Purification and

characterization of a GTP:pyrophosphate exchange activity from vaccinia

virions. J. Biol. Chem. 255, 11588–11598.

Takagi, T., Moore, C.R., Diehn, F., and Buratowski, S. (1997). An RNA

50-triphosphatase related to the protein tyrosine phosphatases. Cell 89,

867–873.

Wang, S.P., Deng, L., Ho, C.K., and Shuman, S. (1997). Phylogeny of mRNA

capping enzymes. Proc. Natl. Acad. Sci. USA 94, 9573–9578.

eserved