29.04.2013 Views

Rose. Oughtred, Nathalie. Bedard, Alice. Vrielink ... - McGill University

Rose. Oughtred, Nathalie. Bedard, Alice. Vrielink ... - McGill University

Rose. Oughtred, Nathalie. Bedard, Alice. Vrielink ... - McGill University

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 273, No. 29, Issue of July 17, pp. 18435–18442, 1998<br />

© 1998 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A.<br />

Identification of Amino Acid Residues in a Class I<br />

Ubiquitin-conjugating Enzyme Involved in Determining<br />

Specificity of Conjugation of Ubiquitin to Proteins*<br />

(Received for publication, March 2, 1998, and in revised form, May 1, 1998)<br />

<strong>Rose</strong> <strong>Oughtred</strong>‡§, <strong>Nathalie</strong> Bédard‡, <strong>Alice</strong> <strong>Vrielink</strong>, and Simon S. Wing‡**<br />

From the ‡Department of Medicine, Polypeptide Laboratory, <strong>McGill</strong> <strong>University</strong>, Montreal, Quebec H3A 2B2, Canada and<br />

the Department of Biochemistry and the Montreal Joint Centre for Structural Biology, <strong>McGill</strong> <strong>University</strong>, Montreal,<br />

Quebec H3G 1Y6, Canada<br />

The ubiquitin pathway is a major system for selective<br />

proteolysis in eukaryotes. However, the mechanisms underlying<br />

substrate selectivity by the ubiquitin system<br />

remain unclear. We previously identified isoforms of a<br />

rat ubiquitin-conjugating enzyme (E2) homologous to<br />

the Saccharomyces cerevisiae class I E2 genes, UBC4/<br />

UBC5. Two isoforms, although 93% identical, show distinct<br />

features. UBC4-1 is expressed ubiquitously,<br />

whereas UBC4-testis is expressed in spermatids. Interestingly,<br />

although these isoforms interacted similarly<br />

with some ubiquitin-protein ligases (E3s) such as E6-AP<br />

and rat p100 and an E3 that conjugates ubiquitin to<br />

histone H2A, they also supported conjugation of ubiquitin<br />

to distinct subsets of testis proteins. UBC4-1<br />

showed an 11-fold greater ability to support conjugation<br />

of ubiquitin to endogenous substrates present in a testis<br />

nuclear fraction. Site-directed mutagenesis of the UBC4testis<br />

isoform was undertaken to identify regions of the<br />

molecule responsible for the observed difference in substrate<br />

specificity. Four residues (Gln-15, Ala-49, Ser-107,<br />

and Gln-125) scattered on surfaces away from the active<br />

site appeared necessary and sufficient for UBC4-1-like<br />

conjugation. These four residues identify a large surface<br />

of the E2 core domain that may represent an area of<br />

binding to E3s or substrates. These findings demonstrate<br />

that a limited number of amino acid substitutions<br />

in E2s can dictate conjugation of ubiquitin to different<br />

proteins and indicate a mechanism by which small E2<br />

molecules can encode a wide range of substrate<br />

specificities.<br />

The ubiquitin system is implicated in an ever expanding<br />

array of cellular processes that now range from DNA repair to<br />

cell cycle progression and muscle protein degradation (reviewed<br />

in Refs. 1–3). Ubiquitin is a highly conserved 76-residue<br />

protein whose many cellular functions are mediated by its<br />

covalent ligation to other proteins. Most of these functions arise<br />

from the ability of ubiquitination to lead to degradation of the<br />

* This work was supported in part by Medical Research Council<br />

Grant MT13341 (to A. V.) and Grant MT12121 (to S. S. W). The costs of<br />

publication of this article were defrayed in part by the payment of page<br />

charges. This article must therefore be hereby marked “advertisement”<br />

in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.<br />

§ Recipient of the Eileen Peters <strong>McGill</strong> Major Fellowship.<br />

Recipient of a Chercheur Boursier award from the Fonds de la<br />

Recherche en Santé du Québec.<br />

** Recipient of a Clinician Scientist award from the Medical Research<br />

Council of Canada. To whom correspondence should be addressed:<br />

Dept. of Medicine, Polypeptide Lab., <strong>McGill</strong> <strong>University</strong>, Strathcona<br />

Bldg., 3640 <strong>University</strong> St., Suite W315, Montreal, Quebec H3A 2B2,<br />

Canada. Tel.: 514-398-4101; Fax: 514-398-3923; E-mail: cxwg@musica.<br />

mcgill.ca.<br />

This paper is available on line at http://www.jbc.org 18435<br />

selected protein. Indeed, ubiquitin-mediated proteolysis is responsible<br />

for the turnover of key regulatory proteins, including<br />

mitotic cyclins (cyclin B) (4, 5), cyclin-dependent kinases (Sic1<br />

and p27) (6, 7), and transcription factors (Mat2, c-Jun, and<br />

p53) (8–10).<br />

Recognition of specific substrates occurs at the level of conjugation,<br />

which is a multistep process involving three types of<br />

enzymes (11): a ubiquitin-activating enzyme (E1), 1 ubiquitinconjugating<br />

enzymes (UBCs or E2s), and, in many cases, ubiquitin-protein<br />

ligases (E3s). Initially, ubiquitin is activated by<br />

E1 through the ATP-dependent formation of a thiol ester bond<br />

between ubiquitin and E1 (12). The activated ubiquitin is then<br />

transferred via a thiol ester linkage to a cysteine residue of an<br />

E2 (reviewed in Ref. 13). Finally, the E2 itself, or more commonly<br />

in concert with an E3, ligates the ubiquitin via its<br />

carboxyl terminus to lysine residues of a protein substrate.<br />

Successive ubiquitin molecules may be added to lysine residues<br />

of the previous ubiquitin to produce a multi-ubiquitin chain.<br />

Although the biochemical mechanisms of the pathway are<br />

becoming well defined, the molecular mechanisms by which<br />

substrates are selected by the ubiquitin-conjugating apparatus<br />

remain unclear. E3s are important for recognition and binding<br />

of the substrate (14). E3s may serve as docking proteins that<br />

bind both specific substrates and E2s (14), thereby permitting<br />

the transfer of ubiquitin from an E2 to a substrate. For example,<br />

the E3 SCF Cdc4 binds the E2 molecule Cdc34 and a specific<br />

substrate, Sic1, simultaneously, thereby facilitating the transfer<br />

of ubiquitin from Cdc34 to Sic1, an inhibitor of the yeast<br />

S-phase cyclin-dependent kinase Cln1-Cdc28 (15, 16). Alternatively,<br />

E3s may function as the final intermediate in the ubiquitin<br />

thiol ester cascade (17). The E3 E6-AP (E6-associated<br />

protein) forms a thiol ester linkage with ubiquitin prior to<br />

catalyzing the ubiquitination of p53 in the presence of the viral<br />

E6 protein (17). The catalytically active cysteine in E6-AP is<br />

found within its carboxyl terminus domain, and a number of<br />

putative E3s have been identified based on the presence of such<br />

HECT (homology to E6-AP carboxyl terminus) domains (18).<br />

Although E3s bind substrates, E2s may also be involved in<br />

substrate recognition either by conjugating substrates directly<br />

or probably more commonly by interacting only with specific<br />

E3s. Indeed, yeast genetic studies have revealed a variety of<br />

functions for different E2s indicating that they can direct conjugation<br />

of ubiquitin to specific substrates. For example, UBC2<br />

(RAD6) is required for DNA repair (19), whereas UBC4/UBC5<br />

1 The abbreviations used are: E1, ubiquitin-activating enzyme; E2<br />

and UBC, ubiquitin-conjugating enzyme; E3, ubiquitin-protein ligase;<br />

PCR, polymerase chain reaction; GST, glutathione S-transferase; DTT,<br />

dithiothreitol; AMP-PNP, 5-adenylyl imidodiphosphate; RM-Ub, reductively<br />

methylated ubiquitin.


18436<br />

FIG. 1. Comparison of rat UBC4-1,<br />

UBC4-testis, and other prototype<br />

class I E2s. Rat isoforms UBC4-1 and<br />

UBC4-testis are homologous to S. cerevisiae<br />

UBC4. The 11 amino acids that differ<br />

between UBC4-1 and UBC4-testis are underlined.<br />

The critical residues that confer<br />

a rat UBC4-1 phenotype on UBC4-testis<br />

are indicated with asterisks. The protein<br />

sequences of the mammalian homologues<br />

of yeast Rad6, HHR6A and HHR6B, and<br />

S. cerevisiae Ubc7 are also depicted. The<br />

active-site cysteine is indicated by an<br />

arrowhead.<br />

are required for the degradation of short-lived and abnormal<br />

proteins (20).<br />

Differences in E2 function evidently reflect differences in E2<br />

structure. E2 enzymes have been divided into four structural<br />

classes based on amino acid sequence comparison (21). Class I<br />

enzymes (e.g. Ubc4 and Ubc5) (20) consist of a conserved catalytic<br />

core domain of 150 amino acids that contains the activesite<br />

cysteine involved in ubiquitin transfer. Class II enzymes<br />

(e.g. Ubc2/Rad6 and Ubc3/Cdc34) (19, 22) have extra C-terminal<br />

extensions or tails attached to the core domain, whereas<br />

class III enzymes (e.g. UbcH6 and UbcD2) (23, 24) have attached<br />

N-terminal tails. Finally, class IV enzymes (e.g. E2-C)<br />

(25) possess both C- and N-terminal extensions.<br />

Jentsch et al. (21) speculated that E2 extensions play either<br />

a direct role in substrate recognition or else an indirect role<br />

through their interaction with E3s. While C- and N-terminal<br />

extensions may participate in specifying the interaction of E2s<br />

with E3s and/or substrates, a number of studies have indicated<br />

that specificity elements also reside within the E2 core. For<br />

example, the class II enzyme RAD6 is required for DNA repair,<br />

induced mutagenesis, and sporulation in yeast (19) and is<br />

capable of polyubiquitinating histones in vitro (26). However,<br />

removal of the polyacidic tail of Rad6 results only in loss of its<br />

sporulation function (27) and histone-polyubiquitinating activity<br />

(26). The core domain is sufficient for performing its DNA<br />

repair function. Since this truncated Rad6 containing only the<br />

core domain exhibits a distinct phenotype from the class I (core<br />

domain only) E2s, Ubc4 and Ubc5, which function in the turnover<br />

of short-lived and abnormal proteins (20), the core domain<br />

must possess determinants of function and by inference substrate<br />

specificity. More recently, the C-terminal tail of E2–25K<br />

was found to be necessary, but not sufficient, for some of its E2<br />

functions, indicating that the tail depends on structural features<br />

in the core for its function (28). These and other results<br />

suggest that although E2 core domains are highly conserved,<br />

they possess unique structural features that are critical for<br />

individual E2 function and specificity.<br />

Recently, we cloned and characterized a family of mammalian<br />

class I E2s homologous to S. cerevisiae Ubc4/Ubc5 (29, 30).<br />

E2 Residues Determining Substrate Specificity<br />

Two isoforms of rat UBC4, 2 although 93% identical, show distinct<br />

features. Rat UBC4-testis possesses an acidic pI and<br />

shows testis-specific RNA expression that is specifically induced<br />

in the developing spermatids (30), whereas rat UBC4-1<br />

has a basic pI and is expressed ubiquitously (29). Therefore,<br />

although the high degree of sequence similarity might suggest<br />

that these isoforms are redundant, the highly regulated and<br />

cell-specific expression suggested a unique role for the UBC4testis<br />

isoform.<br />

Therefore, we characterized carefully the abilities of rat<br />

UBC4-1 and UBC4-testis to support conjugation of ubiquitin in<br />

vitro to different subsets of testis proteins. Rat UBC4-1 shows<br />

an 11-fold greater ability to support conjugation of ubiquitin to<br />

endogenous substrates present in a testis nuclear fraction (30).<br />

We also determined whether these two isoforms interact differentially<br />

with other E3s. In addition, since rat UBC4-1 and<br />

UBC4-testis differ by only 11 amino acids (Fig. 1) and they are<br />

highly similar to yeast Ubc4, whose crystal structure has been<br />

solved (31), this provided a unique opportunity to identify, by<br />

site-directed mutagenesis of UBC4-testis, regions of the E2<br />

core domain responsible for the observed difference in substrate<br />

specificity.<br />

EXPERIMENTAL PROCEDURES<br />

Site-directed Mutagenesis—pET-11d (Novagen)-based Escherichia<br />

coli expression plasmids encoding rat UBC4-1 or UBC4-testis have<br />

been described (29, 30). Mutagenesis of selected residues in UBC4testis<br />

to those in UBC4-1 was performed using the Chameleon doublestranded<br />

site-directed mutagenesis kit (Stratagene) according to the<br />

manufacturer’s instructions. Briefly, separate mutagenic primers3 encoding<br />

site-specific mutations in UBC4-testis and a selection primer<br />

were annealed to denatured UBC4-testis-containing pET-11d plasmids,<br />

2 Previously, the rat homologue of yeast Ubc4 was referred to as<br />

E217KB and included isoforms 2E and 8A (29, 30). For purposes of clarity<br />

and to conform to a trend by workers in the field to name E2s after their<br />

apparent yeast homologues, isoform 2E will henceforth be referred to as<br />

rat UBC4-1, and isoform 8A as rat UBC4-testis. The nucleotide sequences<br />

of UBC4-1 and UBC4-testis have been submitted to Gen-<br />

BankTM with accession numbers U13177 and U56407.<br />

3 Oligonucleotide sequences and detailed PCR conditions are available<br />

on request.


and the mutant DNA strand was extended with T7 DNA polymerase<br />

and ligated with T4 DNA ligase. The selection primer, located 2<br />

kilobases from the mutagenic primers, changed the unique AccI restriction<br />

site on pET-11d to a unique KpnI restriction site on the mutant<br />

plasmid strand, thereby permitting selection of intact mutant plasmids<br />

by digestion with AccI. Initially, separate mutant UBC4-testis constructs<br />

with a D55H, E68A, or double D55H/E68A substitution were<br />

generated. Similarly, subsequent mutations were added separately or<br />

in combination onto the initial UBC4-testis D55H/E68A double mutant<br />

in the following order: Gln-15, Gln-125, Ala-49, or Ser-107. The intact<br />

mutant plasmids were transformed into E. coli XL1-Blue cells (Stratagene).<br />

All plasmids were sequenced to confirm the presence of the<br />

desired mutation using the fmol TM DNA sequencing System (Promega).<br />

Subtractive mutagenesis was then performed in a similar manner to<br />

define the minimal UBC4-1 residues on UBC4-testis that were necessary<br />

and sufficient for the UBC4-1-like conjugating activity.<br />

The converse mutagenesis of the four critical residues (Arg-15, Val-<br />

49, Cys-107, and Arg-125) of UBC4-1 to those of UBC4-testis was<br />

performed via PCR amplification (32). Mutant UBC4-1 fragments were<br />

generated using UBC4-1-containing pET-11d as a template, primers<br />

bearing the relevant base substitutions, and sense or antisense oligonucleotides<br />

encoding the amino and carboxyl termini of the protein as<br />

well as restriction sites to permit cloning into the NcoI and BamHI sites<br />

of pET-11d. The mutant UBC4-1 fragments were then purified by<br />

agarose gel electrophoresis and incorporated into full-length mutant<br />

UBC4-1 inserts via a second round of PCR amplification using the<br />

separate mutant UBC4-1 fragments as a template and both the 5-NcoI<br />

and the 3-BamHI primers. The PCR products were purified on a<br />

QIAquick PCR purification column (QIAGEN Inc.), digested with NcoI<br />

and BamHI, and then ligated into a pET-11d vector that had been<br />

digested with the same enzymes. Purified plasmids were transformed<br />

into XL1-Blue cells, and individual positive clones were sequenced<br />

using the fmol TM DNA sequencing system to confirm the presence of the<br />

desired mutation.<br />

Preparation of Proteins—The pGEX-ubiquitin plasmid encoding the<br />

GST-ubiquitin fusion protein (33) was expressed in E. coli strain DH5<br />

and induced with 0.1 mM isopropyl--D-thiogalactopyranoside for 2hat<br />

37 °C. Bacterial cell pellets resuspended in phosphate-buffered saline<br />

and 1% Triton X-100 were lysed by sonication and clarified by centrifugation<br />

at 12,000 g. Glutathione-Sepharose (Amersham Pharmacia<br />

Biotech) was added to the supernatant, and the mixture was rotated<br />

overnight at 4 °C. Beads were washed in phosphate-buffered saline and<br />

1% Triton X-100 and eluted in 20 mM glutathione in phosphate-buffered<br />

saline for 20 min at 25 °C. The GST-ubiquitin fusion protein content<br />

was estimated to be 1 mg/ml by Coomassie Blue staining.<br />

E1 was prepared from rabbit liver. Bacterially expressed recombinant<br />

UBC4-1 and UBC4-testis proteins were also purified as described<br />

previously (29, 30). The E1, UBC4-1, and UBC4-testis enzymes were<br />

quantified by measuring the initial release of radioactive pyrophosphate<br />

following incubation in the presence of [- 32 P]ATP and ubiquitin<br />

(34).<br />

The purified pET-11d-based plasmids containing the mutant<br />

UBC4-1 and UBC4-testis genes were transformed into E. coli BL21<br />

(DE3) (Novagen), and induction of the recombinant proteins with 1 mM<br />

isopropyl--D-thiogalactopyranoside was carried out for 2hat30°C.<br />

The cells were pelleted and resuspended in 0.1 volume and then lysed<br />

by sonication in 50 mM Tris, pH 7.5, and 1 mM DTT. Cellular debris was<br />

removed by centrifugation at 12,000 g. The enzymatic activities of the<br />

mutant E2-containing bacterial lysates relative to the purified recombinant<br />

UBC4-1 and UBC4-testis enzymes were determined by thiol<br />

ester assays, as described below.<br />

The [ 35 S]methionine-labeled E6-AP (35) or rat p100 (36) proteins<br />

were synthesized separately in vitro using a coupled transcription/<br />

translation kit (wheat germ extract TNT, Promega) with T7 polymerase.<br />

The translation reactions (150 l) were partially purified with<br />

DEAE-cellulose resin (Whatman DE52; 200 mg of wet resin/TNT reaction)<br />

using a batch elution procedure to remove ubiquitin, E1, and E2s<br />

homologous to rat UBC4 that were present in the break-though fraction.<br />

Briefly, the translation reactions and resin were incubated in 4.5<br />

ml of loading buffer (50 mM Tris-HCl, pH 7.5, and 1 mM DTT) for 1hat<br />

4 °C, washed two times, and eluted in loading buffer containing 0.5 M<br />

NaCl. The E3-containing eluates (2 ml) were then concentrated 10-fold<br />

by ultrafiltration with Centricon-30 concentrators (Amicon, Inc.), and 2<br />

l of each E3 preparation were utilized in each thiol ester assay.<br />

The testis cytosolic E3 and nuclear fractions eluting from a MonoQ<br />

anion-exchange column (Amersham Pharmacia Biotech) at 0.4 and 0.05<br />

M NaCl, respectively, were isolated as described previously (30). In<br />

some preparations, the nuclear fraction activity was found in the flow-<br />

E2 Residues Determining Substrate Specificity 18437<br />

through fraction instead of eluting at 0.05 M NaCl; however, it behaved<br />

identically to the original preparation eluting at 0.05 M NaCl. This<br />

nuclear fraction was concentrated 4-fold using a Centricon-10 concentrator<br />

(Amicon, Inc.). The testis cytosolic E3 activity was further purified<br />

by chromatography on a Superdex 200 gel filtration column (Amersham<br />

Pharmacia Inc.).<br />

Iodination of Proteins—The chloramine-T method was used to label<br />

bovine ubiquitin with Na 125 I to a specific radioactivity of 3000 cpm/<br />

pmol (37) and histone H2A (Boehringer Mannheim) to a specific radioactivity<br />

of 375,000 cpm/g. Unincorporated 125 I was removed by passing<br />

the reaction products over a Sephadex G-25 column.<br />

Thiol Ester Assays—The relative enzymatic activities of the purified<br />

recombinant rat UBC4-1 and UBC4-testis proteins and of the mutant<br />

E2-containing bacterial lysates were determined by incubating the<br />

following in a total volume of 10 l: 50 mM Tris-HCl, pH 7.5, 1 mM DTT,<br />

2mM MgCl 2,2mM ATP, 100 nM E1, 20 units/ml inorganic pyrophosphatase,<br />

5 M 125 I-ubiquitin (3000 cpm/pmol), and 2.5 pmol of the<br />

purified E2s (5 pmol/l) or various amounts of the mutant E2 lysates.<br />

After incubation at 37 °C for 1 min, the reaction was stopped with<br />

Laemmli sample buffer without -mercaptoethanol and resolved by<br />

12.5% SDS-polyacrylamide gel electrophoresis at 4 °C followed by autoradiography.<br />

The thiol ester bands were excised from the gels to<br />

measure the incorporated radioactivity and thereby estimate the mutant<br />

E2 enzymatic activities relative to those of the purified native E2s<br />

(5 pmol/l). Assaying dilutions of the E2-containing extracts confirmed<br />

linearity of the assays.<br />

The [ 35 S]methionine-labeled E3s E6-AP (35) and rat p100 (36), covalently<br />

bound to GST-ubiquitin fusion protein in thiol ester linkages<br />

(17, 18), were detected by incubating the enzymes in the presence of 50<br />

mM Tris-HCl, pH 7.5, 1 mM DTT, 2 mM MgCl 2,2mM ATP, 50 nM E1, 20<br />

units/ml inorganic pyrophosphatase, 500 nM E2, and 2 l of each partially<br />

purified and concentrated E3 preparation. The reaction mixtures<br />

were preincubated at 25 °C for 3 min, and then the reaction was<br />

initiated with 1 g of GST-ubiquitin fusion protein (33), incubated at<br />

25 °C for 5 min, and stopped with Laemmli sample buffer with or<br />

without -mercaptoethanol. The reactions were resolved at 4 °C on an<br />

SDS-12.5% polyacrylamide gel, which was then soaked in ENHANCE<br />

(NEN Life Science Products) and autoradiographed.<br />

Conjugation Assays—For the conjugation of ubiquitin to endogenous<br />

substrates present in the testis nuclear fraction, the reaction mixture<br />

contained the following in a final volume of 20 l: 10 l of the 0.05 M<br />

NaCl nuclear fraction, 50 mM Tris-HCl, pH 7.5, 1 mM DTT, 2 mM MgCl 2,<br />

2mM AMP-PNP, 5 M 125 I-ubiquitin (3000 cpm/pmol), 50 nM E1, and<br />

250 nM E2s. The ubiquitination rate for the nuclear fraction was linear<br />

for 1 h at 37 °C, and this assay was performed in the presence or<br />

absence of 0.5 g of ubiquitin aldehyde, an isopeptidase inhibitor, or 40<br />

M MG132 (Proscript), a proteasome inhibitor.<br />

For the conjugation of ubiquitin to the exogenous substrate histone<br />

H2A, mediated by the testis cytosolic E3, the reaction mixture contained<br />

the following in a final volume of 20 l: 3 l of the cytosolic E3<br />

Superdex 200 fraction, 50 mM Tris-HCl, pH 7.5, 1 mM DTT, 2 mM MgCl 2,<br />

2mM ATP, 0.5 units pyrophosphatase, 12.5 mM phosphocreatine, 2.5<br />

units of creatine kinase, 250 nM E1, 125 I-histone H2A (specific activity<br />

of 375,000 cpm/g), and varied concentrations of E2 as indicated. Reactions<br />

were initiated with 25 M reductively methylated ubiquitin<br />

(RM-Ub), prepared, and quantified as described (38). Since histone H2A<br />

contains a number of lysine residues, mono- to penta-RM-Ub conjugates<br />

were formed, and the rate of formation of these conjugates was found to<br />

be linear for 10 min at 30 °C, permitting their quantification.<br />

RESULTS<br />

Differential Abilities of Rat UBC4-1 and UBC4-testis to Conjugate<br />

Ubiquitin to a Fraction of Testis Nuclear Proteins—To<br />

test whether the structural differences between rat UBC4-1<br />

and UBC4-testis conferred different abilities to conjugate ubiquitin<br />

to proteins, ubiquitination assays were performed using<br />

testis extracts fractionated on a MonoQ anion-exchange column.<br />

As shown previously (30), a nuclear fraction eluting at<br />

0.05 M NaCl supported conjugation of ubiquitin to proteins<br />

essentially only with the UBC4-1 isoform (Fig. 2A). This indicated<br />

that these two isoforms showed differential substrate<br />

specificity and suggested an enhanced ability of the ubiquitous<br />

isoform to conjugate ubiquitin to endogenous proteins in this<br />

fraction. To evaluate the possibility that UBC4-testis was conjugating<br />

ubiquitin to proteins that are preferentially de-ubiq-


18438<br />

FIG. 2.Differential abilities of UBC4-1 and UBC4-testis to conjugate<br />

ubiquitin to a fraction of testis nuclear proteins. A, a<br />

nuclear testis extract was chromatographed on a MonoQ anion-exchange<br />

column. The fraction eluting at 0.05 M NaCl was incubated with<br />

E1, AMP-PNP, and 125 I-ubiquitin in the presence or absence (E2) of<br />

rat UBC4-1 (4-1) or UBC4-testis (4-T)for1hat37°C.Similar reactions<br />

were also performed in the presence of ubiquitin aldehyde (Ub Ald), an<br />

isopeptidase inhibitor, or MG132, a proteasome inhibitor. Ubiquitinated<br />

endogenous substrates were then analyzed by SDS-polyacrylamide<br />

gel electrophoresis and autoradiography. The high molecular<br />

mass ubiquitin-substrate conjugates remain within the stacking gel<br />

during electrophoresis. B, the preparations of UBC4-1 and UBC4-testis<br />

(5 pmol/l each) used in A were tested for their abilities to form thiol<br />

esters with 125 I -ubiquitin in the presence of E1 and ATP. Ub, ubiquitin.<br />

uitinated by a co-purifying isopeptidase activity, the conjugation<br />

assay was performed in the presence of the isopeptidase<br />

inhibitor ubiquitin aldehyde. Notably, the level of UBC4-testisdependent<br />

conjugation was not increased by the addition of this<br />

reagent, rendering unlikely the possibility of a co-purifying<br />

interfering isopeptidase. Likewise, to rule out the possibility of<br />

enhanced proteasomal degradation of the UBC4-testis-dependent<br />

conjugates, the conjugation assay was performed in the<br />

presence of the proteasomal inhibitor MG132. Similarly,<br />

MG132 did not increase the levels of UBC4-testis-dependent<br />

conjugates.<br />

Significantly, the observed difference in conjugating ability<br />

between UBC4-1 and UBC4-testis was not due to a difference<br />

in the ability of these E2s to accept ubiquitin from E1 because<br />

UBC4-1 and UBC4-testis formed similar amounts of ubiquitin<br />

thiol esters (Fig. 2B). These thiol ester assays are end-point<br />

assays, and the results cannot exclude the possibility of different<br />

affinities of these two isoforms for E1. However, more<br />

detailed thiol ester-based enzyme kinetic studies suggest that<br />

UBC4-1 and UBC4-testis show less than a 2-fold difference in<br />

their affinities for E1 (data not shown).<br />

Rat Isoforms UBC4-1 and UBC4-testis Interact with the E3s<br />

E6-AP and Rat p100—Since the observed difference in conjugation<br />

by rat UBC4-1 and UBC4-testis was found not to be due<br />

to significant differences in their interaction with E1, this<br />

suggested that the specificity might arise at the E2-E3 level.<br />

We therefore tested the abilities of rat UBC4-1 and UBC4testis<br />

to interact with some well defined E3s, E6-AP (10, 17)<br />

and rat p100 (18), which contain HECT domains and thereby<br />

form thiol ester linkages with ubiquitin by accepting ubiquitin<br />

from E2 thiol esters. We tested the abilities of rat UBC4-1 and<br />

UBC4-testis to transfer ubiquitin to E6-AP and rat p100 produced<br />

by in vitro translation (Fig. 3). Since these E3s are<br />

relatively large, a GST-ubiquitin fusion protein (molecular<br />

mass of 34 kDa) (18) was used to permit resolution of the<br />

E3-ubiquitin thiol ester linkage by gel electrophoresis. In the<br />

presence of E1, GST-ubiquitin, and rat UBC4-1 or UBC4-testis,<br />

bands 34 kDa larger than the expected translation products<br />

were evident. These bands were not present when the thiol<br />

ester reactions were treated with -mercaptoethanol or when<br />

the reaction was performed in the absence of rat UBC4-1 and<br />

UBC4-testis. Again, these results obtained with thiol ester<br />

E2 Residues Determining Substrate Specificity<br />

FIG. 3.UBC4-1 and UBC4-testis transfer ubiquitin similarly to<br />

the E3s E6-AP and rat p100. The rat p100 cDNA and human E6-AP<br />

cDNA were transcribed and translated in vitro using wheat germ extract.<br />

The 35 S-labeled translation products were partially purified over<br />

a DE52 column to remove endogenous E2s homologous to UBC4/UBC5.<br />

Thiol ester assays were then performed with the E3s using rat UBC4-1<br />

(4-1) or UBC4-testis (4-T) and a GST-ubiquitin fusion protein (GST-<br />

Ub). Reactions were quenched with SDS-polyacrylamide gel electrophoresis<br />

sample buffer without or with -mercaptoethanol (2-Me). After<br />

electrophoresis, the E3-ubiquitin thiol ester adducts (indicated by arrows)<br />

were detected by autoradiography.<br />

assays do not eliminate the possibility that these E2 isoforms<br />

could differ somewhat in their affinities for these E3s. More<br />

detailed kinetic assays are not feasible in these crude preparations<br />

from in vitro translation, where the concentrations of the<br />

E3s cannot be readily determined. Nonetheless, the results<br />

demonstrated that these two isoforms are both capable of interacting<br />

with ubiquitin-protein ligases as shown by their ability<br />

to transfer ubiquitin to at least two different E3s, E6-AP<br />

and rat p100.<br />

Rat Isoforms UBC4-1 and UBC4-testis Support Ubiquitination<br />

of Histone H2A Mediated by a Testis Cytosolic E3—Since<br />

the assays described above do not permit ready quantitative<br />

determinations of reaction rates, we tested both isoforms for<br />

their ability to support conjugation of ubiquitin to an exogenous<br />

substrate, histone H2A, mediated by a testis E3. To this<br />

end, a testis cytosolic E3 activity (29, 30) eluting at 0.4 M NaCl<br />

from a MonoQ anion-exchange column was further fractionated<br />

using a gel filtration column. This E3 activity was found to<br />

support conjugation of ubiquitin to the exogenous substrate<br />

125 I-labeled histone H2A in the presence of rat UBC4-1 and<br />

UBC4-testis. The effectiveness of these two isoforms in supporting<br />

this E3-dependent ubiquitination of 125 I-histone H2A<br />

was compared using E1, an ATP-regenerating system, RM-Ub<br />

(38), and different concentrations of rat UBC4-1 and UBC4testis<br />

(Fig. 4A). RM-Ub was utilized in the assays to restrict the<br />

ubiquitination of histone H2A to the mono-ubiquitinated form,<br />

which would facilitate the quantification of conjugates. Multiubiquitinated<br />

forms of histone H2A were observed, indicating<br />

that ubiquitin is attached to several different lysine residues of<br />

the molecule. In contrast to the differential ubiquitin-conjugating<br />

abilities of rat UBC4-1 and UBC4-testis observed in the


FIG. 4.UBC4-1 and UBC4-testis support conjugation of ubiquitin<br />

to histone H2A mediated by a testis cytosolic E3. A, the<br />

effectiveness of UBC4-1 or UBC4-testis in supporting E3-dependent<br />

ubiquitination of 125 I-histone H2A was compared using E1, an ATPregenerating<br />

system, RM-Ub, and different concentrations of rat<br />

UBC4-1 or UBC4-testis, incubated for 10 min at 30 °C. After electrophoresis,<br />

the RM-Ub- 125 I-histone H2A conjugates were detected by<br />

autoradiography. B, the differential conjugating abilities of rat UBC4-1<br />

and UBC4-testis observed in the nuclear fraction are not due to a<br />

UBC4-testis-specific inhibitor in the 0.05 M NaCl nuclear fraction. The<br />

above conjugation assay with UBC4-testis was performed in the presence<br />

or absence of the indicated amounts of the testis nuclear fraction<br />

eluting at 0.05 M NaCl from a MonoQ column.<br />

testis nuclear fraction, both isoforms support conjugation of<br />

ubiquitin to histone H2A mediated by the testis cytosolic E3 to<br />

similar extents. Thus, the preferential conjugating activity of<br />

UBC4-1 as compared with UBC4-testis observed in the nuclear<br />

fraction demonstrates that such specificity appears to be selective<br />

for specific E3s or substrates.<br />

To ascertain that the differential conjugating ability of rat<br />

UBC4-1 and UBC4-testis observed in the testis nuclear fraction<br />

was not due to an inhibitor of the UBC4-testis activity<br />

present in this fraction, the histone conjugation assay was<br />

performed in the presence or absence of the nuclear fraction<br />

(Fig. 4B). UBC4-testis supported E3-mediated conjugation of<br />

ubiquitin to 125 I-histone H2A to similar levels with or without<br />

E2 Residues Determining Substrate Specificity 18439<br />

the nuclear fraction. Thus, these data, in conjunction with the<br />

absence of effects of ubiquitin aldehyde and MG132, are consistent<br />

with the differential conjugating abilities of rat UBC4-1<br />

and UBC4-testis in the nuclear fraction being attributable to<br />

differences in their interactions with specific, and as yet unidentified,<br />

E3s or substrates.<br />

Four Amino Acid Differences Are Responsible for the Abilities<br />

of Rat Isoforms UBC4-1 and UBC4-testis to Conjugate Ubiquitin<br />

to Different Subsets of Testis Nuclear Proteins—Since the<br />

rat UBC4-1 and UBC4-testis isoforms differ by only 11 amino<br />

acids (Fig. 1) (30), this provided an unprecedented opportunity<br />

to identify, by site-directed mutagenesis, critical residues of the<br />

E2 molecule responsible for the observed difference in substrate<br />

specificity. Site-directed mutagenesis of the UBC4-testis<br />

isoform was undertaken to identify regions of the molecule that<br />

can confer the UBC4-1-like ability to conjugate ubiquitin to the<br />

endogenous substrates present in the nuclear fraction.<br />

Since rat UBC4-1 and UBC4-testis differ with respect to<br />

their pI values (30), mutagenesis of UBC4-testis began with the<br />

four residues responsible for the dramatic change in pI. Of<br />

particular interest were the two residues, aspartic acid 55 and<br />

glutamic acid 68, located near the active-site cysteine, that<br />

were mutated to histidine and alanine, respectively. However,<br />

mutation of UBC4-testis to the UBC4-1 residues at these two<br />

sites did not confer UBC4-1-like ability to conjugate 125 I-ubiquitin<br />

to the substrates present in the testis nuclear fraction<br />

(Fig. 5A). Additional mutagenesis of glutamines 15 and 125,<br />

the two other residues responsible for the change in pI, to<br />

arginines resulted in a small increase in conjugating activity of<br />

the mutated UBC4-testis. Since mutagenesis of the two glutamine<br />

residues appeared to increase the conjugating activity of<br />

the mutated UBC4-testis, further residues nearby were mutated.<br />

Mutation of alanine 49 to valine conferred some increase<br />

in activity, and an additional mutation of serine 107 to cysteine<br />

conferred a UBC4-1 phenotype in conjugating ability to the<br />

mutated UBC4-testis.<br />

To determine the minimal number of substitutions required<br />

for the UBC4-1-like conjugating activity of the mutated UBC4testis,<br />

subtractive mutagenesis was then performed. Since mutagenesis<br />

of Gln-15, Ala-49, Ser-107, and Gln-125 improved<br />

conjugating activity, these four mutations were tested together<br />

and found to be sufficient (Fig. 5A). Further removal of any one<br />

of these substitutions decreased conjugating ability significantly,<br />

and therefore, these four substitutions are necessary<br />

(Fig. 5B).<br />

Our observation that these residues are important would<br />

predict that mutating the rat UBC4-1 molecule at these positions<br />

to the UBC4-testis residues would decrease the conjugating<br />

activity of UBC4-1. To test this prediction, mutagenesis of<br />

UBC4-1 to UBC4-testis was performed at these four residues.<br />

As expected, mutagenesis of each of the critical residues (Arg-<br />

15, Val-49, Cys-107, or Arg-125) resulted in decreased conjugating<br />

activity of rat UBC4-1 in the testis nuclear fraction (Fig.<br />

6, A and B). Interestingly, the four residues in UBC4-testis,<br />

which were found to be necessary and sufficient for the UBC4-<br />

1-like conjugating activity, are present on surfaces away from<br />

the active site (Fig. 7) (31).<br />

DISCUSSION<br />

We have, for the first time, identified, in the core domain<br />

conserved among all ubiquitin-conjugating enzymes, specific<br />

sites that are involved in determining substrate specificity of<br />

conjugation. These studies revealed a number of intriguing<br />

insights. First, although we anticipated, based on the 93%<br />

amino acid identity between the two isoforms, that only a<br />

limited number of residues in the core domain of E2s may<br />

dictate conjugation of ubiquitin to different proteins, surpris-


18440<br />

FIG. 5. Only four residues are necessary<br />

and sufficient to confer a rat<br />

UBC4-1 phenotype on UBC4-testis. A,<br />

bacterially expressed mutant UBC4-testis<br />

proteins were tested for their ability to<br />

support conjugation with the 0.05 M NaCl<br />

testis nuclear fraction in assays as described<br />

in the legend to Fig. 2. B, the<br />

minimal number of substitutions required<br />

for UBC4-1-like conjugating activity<br />

in the testis nuclear fraction was determined.<br />

Removal of some of these<br />

substitutions appeared to decrease conjugating<br />

ability, and therefore, these four<br />

substitutions appear necessary. C, the<br />

conjugating activities of the UBC4-testis<br />

mutants relative to the UBC4-1 isoform<br />

were compared by counting the radioactivity<br />

in the gel lanes (S.E.) and were<br />

normalized to the value for UBC4-1. D,<br />

Asp-55; E, Glu-68; Q1, Gln-15; Q2, Gln-<br />

125; A, Ala-49; S, Ser-107.<br />

ingly only four substitutions (Figs. 5 and 6) were sufficient and<br />

necessary to produce the change in substrate specificity.<br />

Second, some of these substitutions were physiochemically<br />

relatively conserved, indicating that subtle changes in primary<br />

structure can be important in determining selectivity of substrates.<br />

One involved an Ala to Val conversion, whereas the<br />

other involved a Cys for Ser substitution. The other two substitutions,<br />

on the other hand, were nonconserved. The polar<br />

uncharged Gln residues in UBC4-testis have been mutated to<br />

the polar basic Arg residue. Thus, part of the interaction of<br />

UBC4-1 and a testis nuclear substrate or E3 could be due to<br />

electrostatic interactions between the molecules. However, differences<br />

in electrostatic attractions alone are unlikely to account<br />

for the specificity observed here because all four residues,<br />

including the two conserved substitutions, are necessary for<br />

conferring a UBC4-1-like phenotype on UBC4-testis. Since all<br />

four residues are on the surface, their substitution is unlikely<br />

to induce a significant conformational change on the E2 molecule.<br />

Instead, it may be that all four of these critical residues<br />

result in a surface configuration that functions to lower the free<br />

energy of binding of UBC4-1 to an E3 or substrate, thereby<br />

facilitating the UBC4-1-dependent conjugation of ubiquitin to<br />

specific proteins in the testis nuclear fraction.<br />

Third, the four critical residues in UBC4-1 are not localized,<br />

but spread over a broad surface of the E2 molecule away from<br />

the active site (Fig. 7), and so a large surface of the E2 core may<br />

be involved in binding an E3 or substrate. The critical residues<br />

are localized as follows: the N-terminal Gln-15 is in the stretch<br />

between the first -helix and the first -strand; the N-terminal<br />

Ala-49 is in the third -strand; the C-terminal Ser-107 is in the<br />

second -helix; and finally, the N-terminal Gln-125 is in the<br />

third -helix.<br />

E2 Residues Determining Substrate Specificity<br />

Although all E2s exhibit limited sequence identity and are<br />

functionally different, the overall three-dimensional folding of<br />

E2 core domains that have been crystallized to date is remarkably<br />

similar (31, 40–42). The 150 amino acids of the E2 core<br />

domains show 25% sequence identity, and notably, most of<br />

the identical residues are either buried or clustered on one<br />

surface adjacent to the active-site cysteine (31). It has been<br />

suggested that the highly conserved surface region around the<br />

active site may be specific for ubiquitin and/or E1 binding,<br />

whereas the divergent surface regions may enable individual<br />

E2 enzymes to bind their respective substrates or E3s (31). Our<br />

data now provide experimental evidence to support this<br />

hypothesis.<br />

Although we have not to date been able to identify an exogenous<br />

substrate that requires the nuclear fraction for conjugation,<br />

other studies (30, 43) showing that this family of E2s<br />

interacts extensively with E3s would suggest that the nuclear<br />

fraction probably does contain an E3 activity. Recent findings<br />

suggest that a putative E2-binding site exists in the C terminus<br />

of HECT domain-containing E3s and that a variable E3 N<br />

terminus may be involved in binding substrates (44). Since the<br />

residues found to be critical for the UBC4-1-like conjugating<br />

activity lie on a surface away from the active site, this surface<br />

may be responsible for the selective interaction of UBC4-1 with<br />

a nuclear E3. This could permit the small E2 molecule, while<br />

bound to the larger E3 protein (known E3s are 95 kDa in<br />

size), to expose its active-site cysteine, thereby facilitating the<br />

transfer of ubiquitin to the E3 if it contains a HECT domain or<br />

directly to a substrate if the E3 is functioning primarily as a<br />

docking protein (Fig. 8).<br />

Fourth, it is possible that some E2s can functionally overlap<br />

with some E3s or substrates, yet be selective for other E3s or


FIG. 6. Mutation of the critical residues in UBC4-1 confirms<br />

that they are necessary to conjugate ubiquitin to a subset of<br />

testis nuclear proteins. A, bacterially expressed mutant UBC4-1<br />

proteins were assayed for their ability to support conjugation with the<br />

0.05 M NaCl testis nuclear fraction as described in the legend to Fig. 2.<br />

B, the conjugating activities of the UBC4-1 mutants relative to the<br />

UBC4-1 isoform were compared by counting the radioactivity in the gel<br />

lanes (S.E.) and were normalized to the value for UBC4-1. R1, Arg-15;<br />

R2, Arg-125; V, Val-49; C, Cys-107.<br />

substrates. For example, the critical residues that confer a<br />

UBC4-1-like phenotype on UBC4-testis likely result in a distinctive<br />

E2 surface configuration that is selectively recognized<br />

by a specific nuclear E3. However, other E3s, such as rat p100,<br />

E6-AP, and the testis cytosolic E3 we have identified, may<br />

interact principally with residues that are conserved among<br />

these E2s (Figs. 3 and 4). Indeed, the functional specificity of<br />

distinct E2s may be contingent upon specific residues in these<br />

molecules that facilitate or impair their interaction with different<br />

E3s or substrates. In support of this, it has been shown that<br />

the mouse homologues of yeast Rad6, mHR6A and mHR6B,<br />

which show 95% amino acid sequence identity (Fig. 1), may be<br />

functionally distinct since inactivation of the mHR6B gene, but<br />

not the mHR6A gene, in mice causes male sterility (45). Thus,<br />

mHR6A appears to complement some functions of mHR6B, but<br />

not all of them. The minor sequence differences between the<br />

two isoforms may therefore be responsible for the selectivity of<br />

binding to some E3s or substrates.<br />

Recently, three-dimensional structure determination of<br />

yeast Ubc7 (41) revealed that its tertiary folding is similar to<br />

other class I enzymes that have been crystallized, with the<br />

exception of two regions where extra residues are present in<br />

Ubc7. Based on amino acid sequence alignment between 13<br />

yeast E2 enzymes, Cook et al. (41) suggested that there are four<br />

potential regions where extra residues could be inserted into<br />

the common core domain of various E2s and that these may<br />

represent hypervariable regions that confer specificity for bind-<br />

E2 Residues Determining Substrate Specificity 18441<br />

FIG. 7.The critical residues essential for UBC4-1-like conjugation<br />

are dispersed on surfaces away from the active-site cysteine.<br />

Shown is the crystal structure of yeast Ubc4 (31). Rat UBC4-1<br />

shows 80% amino acid identity to yeast Ubc4, and rat UBC4-testis<br />

shows 76% identity. Shown in red on Ubc4 are the locations of the<br />

corresponding residues of UBC4-testis required for UBC4-1-like conjugating<br />

activity in the testis nuclear fraction. These substitutions are<br />

located on surfaces away from the active-site cysteine (yellow).<br />

FIG. 8.A model for the interaction of E2s with their respective<br />

E3s. The critical residues on E2 molecules that facilitate their interaction<br />

with E3s or substrates (Sub) may be dispersed on surfaces of the E2<br />

core away from the active site, and these surfaces may be responsible<br />

for the selective interaction of E2s with different E3s. This would<br />

permit the E2 molecule, while bound to a larger E3 protein, to expose its<br />

active-site cysteine, thereby facilitating the transfer of ubiquitin (Ub)to<br />

an E3 if it contains a HECT domain (A) or directly to a substrate if the<br />

E3 is functioning primarily as a docking protein (B).<br />

ing substrates or E3s. Notably, these four hypervariable regions<br />

are located on one broad surface surrounding the activesite<br />

cysteine. It has been hypothesized that the two insertions<br />

in the Ubc7 core domain (Fig. 1) may be critical for its role in<br />

targeting specific substrates (e.g. Mat2, Sec61p, and YscY) (8,<br />

46, 47) for ubiquitin-dependent degradation. Although this hypothesis<br />

of a hypervariable surface contributing to substrate<br />

and/or E3 specificity may prove correct, functional evidence is<br />

still lacking. In contrast, functional evidence presented here<br />

demonstrates that surfaces away from the ubiquitin-accepting<br />

cysteine, and thus away from the surface containing the hypervariable<br />

regions, are critical for determining the substrate<br />

specificity of the rat UBC4 isoforms. This involvement of a<br />

large surface of the E2 (Figs. 7 and 8) in determining substrate<br />

selectivity is an attractive concept as it can explain how small<br />

E2 molecules can encode such a diverse range of specificities.


18442<br />

Significantly, these results represent the first detailed mutagenesis<br />

of an E2 molecule related to substrate specificity.<br />

Unlike most studies of structure-function relationships that<br />

use mutagenesis to create artificial mutants with distinct properties,<br />

our studies have been based on naturally occurring<br />

isoforms. Thus, the different biochemical phenotypes based on<br />

these four critical residues are likely to be biologically important.<br />

The existence of such highly similar isoforms in the same<br />

tissue may not be redundant, but rather may permit fine regulation<br />

of conjugation of ubiquitin to specific substrates. The<br />

precise induction of UBC4-testis at the round spermatid stage<br />

of spermatogenesis would also argue for a specific function of<br />

this isoform (30). Inactivation of this gene in the mouse is<br />

currently underway and will likely yield further insights into<br />

the determinants of function present in the core domains of<br />

class I E2s.<br />

Acknowledgments—We thank J. Huibregtse for the pGEX-ubiquitin<br />

plasmid, D. Muller for the cDNA encoding rat p100, P. Howley for the<br />

E6-AP construct, A. Ciechanover for ubiquitin aldehyde, Proscript for<br />

MG132, and A. Haas for the protocol for purifying E1 from rabbit liver.<br />

REFERENCES<br />

1. Ciechanover, A. (1994) Cell 79, 13–21<br />

2. Hochstrasser, M. (1996) Annu. Rev. Genet. 30, 405–439<br />

3. Coux, O., Tanaka, K., and Goldberg, A. L. (1996) Annu. Rev. Biochem. 65,<br />

801–847<br />

4. Hershko, A., Ganoth, D., Pehrson, J., Palazzo, R. E., and Cohen, L. H. (1991)<br />

J. Biol. Chem. 266, 16376–16379<br />

5. King, R. W., Peters, J.-M., Tugendreich, S., Rolfe, M., Hieter, P., and Kirschner,<br />

M. W. (1995) Cell 81, 279–288<br />

6. Schwob, E., Boehm, T., Mendenhall, M. D., and Nasmyth, K. (1994) Cell 79,<br />

233–244<br />

7. Sheaff, R. J., Groudine, M., Gordon, M., Roberts, J. M., and Clurman, B. E.<br />

(1997) Genes Dev. 11, 1464–1478<br />

8. Chen, P., Johnson, P., Sommer, T., Jentsch, S., and Hochstrasser, M. (1993)<br />

Cell 74, 357–369<br />

9. Treier, M., Staszewski, L. M., and Bohmann, D. (1994) Cell 78, 787–798<br />

10. Scheffner, M., Werness, B. A., Huibregtse, J. M., Levine, A. J., and Howley,<br />

P. M. (1990) Cell 63, 1129–1136<br />

11. Hershko, A., Heller, H., Elias, S., and Ciechanover, A. (1983) J. Biol. Chem.<br />

258, 8206–8214<br />

12. Haas, A. L., Warms, J. V. B., Hershko, A., and <strong>Rose</strong>, I. A. (1982) J. Biol. Chem.<br />

257, 2543–2548<br />

13. Haas, A. L., and Siepmann, T. J. (1997) FASEB J. 11, 1257–1268<br />

14. Reiss, Y., Heller, H., and Hershko, A. (1989) J. Biol. Chem. 264, 10378–10383<br />

15. Skowra, D., Craig, K. L., Tyers, M., Elledge, J., and Harper, J. W. (1997) Cell<br />

91, 209–219<br />

16. Feldman, R. M. R., Correll, C. C., Kaplan, K. B., and Deshaies, R. J. (1997) Cell<br />

E2 Residues Determining Substrate Specificity<br />

91, 221–230<br />

17. Scheffner, M., Nuber, U., and Huibregtse, J. M. (1995) Nature 373, 81–83<br />

18. Huibregtse, J. M., Scheffner, M., Beaudenon, S., and Howley, P. M. (1995)<br />

Proc. Natl. Acad. Sci. U. S. A. 92, 2563–2567<br />

19. Jentsch, S., McGrath, J. P., and Varshavsky, A. (1987) Nature 329, 131–134<br />

20. Seufert, W., and Jentsch, S. (1990) EMBO J. 9, 543–550<br />

21. Jentsch, S., Seufert, W., Sommer, T., and Reins, H. A. (1990) Trends Biochem.<br />

Sci. 15, 195–198<br />

22. Goebl, M. G., Yochem, J., Jentsch, S., McGrath, J. P., Varshavsky, A., and<br />

Byers, B. (1988) Science 241, 1331–1335<br />

23. Nuber, U., Schwarz, S., Kaiser, P., Schneider, R., and Scheffner, M. (1996)<br />

J. Biol. Chem. 271, 2795–2800<br />

24. Matuschewski, K., Hauser, H.-P., Treier, M., and Jentsch, S. (1996) J. Biol.<br />

Chem. 271, 2789–2794<br />

25. Aristarkhov, A., Eytan, E., Moghe, A., Admon, A., Hershko, A., and Ruderman,<br />

J. V. (1996) Proc. Natl. Acad. Sci. U. S. A. 93, 4294–4299<br />

26. Sung, P., Prakash, S., and Prakash, L. (1988) Genes Dev. 2, 1476–1485<br />

27. Morrison, A., Miller, E. J., and Prakash, L. (1988) Mol. Cell. Biol. 8, 1179–1185<br />

28. Haldeman, M. T., Xia, G., Kasperek, E. M., and Pickart, C. M. (1997) Biochemistry<br />

36, 10526–10537<br />

29. Wing, S., and Jain, P. (1995) Biochem. J. 305, 125–132<br />

30. Wing, S. S., <strong>Bedard</strong>, N., Morales, C., Hingamp, P., and Trasler, J. (1996) Mol.<br />

Cell. Biol. 16, 4064–4072<br />

31. Cook, W. J., Jeffrey, L. C., Xu, Y., and Chau, V. (1993) Biochemistry 32,<br />

13809–13817<br />

32. Higuchi, R., Krummel, B., and Saiki, R. K. (1988) Nucleic Acids Res. 16,<br />

7351–7367<br />

33. Scheffner, M., Huibregtse, J. M., Vierstra, R. D., and Howley, P. M. (1993) Cell<br />

75, 495–505<br />

34. Haas, A. L., and <strong>Rose</strong>, I. A. (1982) J. Biol. Chem. 257, 10329–10337<br />

35. Huibregtse, J. M., Scheffner, M., and Howley, P. M. (1993) Mol. Cell. Biol. 13,<br />

775–784<br />

36. Muller, D., Rehbein, M., Baumeister, H., and Richter, D. (1992) Nucleic Acids<br />

Res. 20, 1471–1475<br />

37. Ciechanover, A., Heller, H., Elias, S., Haas, A. L., and Hershko, A. (1980) Proc.<br />

Natl. Acad. Sci. U. S. A. 87, 1365–1368<br />

38. Hershko, A., and Heller, H. (1985) Biochem. Biophys. Res. Commun. 128,<br />

1079–1086<br />

39. Deleted in proof<br />

40. Cook, W. J., Jeffrey, L. C., Sullivan, M. L., and Vierstra, R. D. (1992) J. Biol.<br />

Chem. 267, 15116–15121<br />

41. Cook, W., Martin, P. D., Edwards, B. F. P., Yamazaki, R. K., and Chau, V.<br />

(1997) Biochemistry 36, 1621–1627<br />

42. Tong, H., Hateboer, G., Perrakis, A., Bernards, R., and Sixma, T. K. (1997)<br />

J. Biol. Chem. 272, 21381–21387<br />

43. Girod, P.-A., and Vierstra, R. D. (1993) J. Biol. Chem. 268, 955–960<br />

44. Hatakeyama, S., Jensen, J. P., and Weissman, A. M. (1997) J. Biol. Chem. 272,<br />

15085–15092<br />

45. Roest, H. P., van Klaveren, J., de Wit, J., van Gurp, C. G., Koken, M. H. M.,<br />

Vermey, M., van Roijen, J. H., Hoogerbrugge, J. W., Vreeburg, J. T. M.,<br />

Baarends, W. M., Bootsma, D., Grootegoed, J. A., and Hoeijmakers, J. H. J.<br />

(1996) Cell 86, 799–810<br />

46. Biederer, T., Volkwein, C., and Sommer, T. (1996) EMBO J. 15, 2069–2076<br />

47. Hiller, M. M., Finger, A., Schweiger, M., and Wolf, D. H. (1996) Science 273,<br />

1725–1728

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!