You are on page 1of 7

Review TRENDS in Biochemical Sciences Vol.32 No.

How ribosomes make peptide bonds


Marina V. Rodnina1, Malte Beringer1 and Wolfgang Wintermeyer2
1
Institute of Physical Biochemistry, University of Witten/Herdecke, Witten, D-58448, Germany
2
Institute of Molecular Biology, University of Witten/Herdecke, Witten, D-58448, Germany

Ribosomes are molecular machines that synthesize ribosome and proceeds without auxiliary factors, peptide
proteins in the cell. Recent biochemical analyses and release requires specialized release factors that recognize
high-resolution crystal structures of the bacterial ribo- termination codons and promote the hydrolysis of the
some have shown that the active site for the formation P-site peptidyl-tRNA.
of peptide bonds – the peptidyl-transferase center – is Although studied for decades, it is only in the past few
composed solely of rRNA. Thus, the ribosome is the years that enormous progress has been made in understand-
largest known RNA catalyst and the only natural ribo- ing ribosome function. Crystal structures have provided
zyme that has a synthetic activity. The ribosome much information about the active site for peptide-bond
employs entropic catalysis to accelerate peptide-bond formation. Rigorous kinetic analysis and the development
formation by positioning substrates, reorganizing water of methods to produce and purify ribosomes with mutations
in the active site and providing an electrostatic network in crucial rRNA residues have revealed the nature of
that stabilizes reaction intermediates. Proton transfer catalysis and the role of ribosome residues. Computational
during the reaction seems to be promoted by a con- analysis has enabled the reaction trajectories to be modeled,
certed shuttle mechanism that involves ribose hydroxyl thereby providing direct insight into the mechanism. Struc-
groups on the tRNA substrate. tural and mutagenesis studies, enzymology, and computer
simulations converge at a consistent picture of the mechan-
The ribosome is a ribozyme ism of the peptidyl-transfer reaction of the prokaryotic
Most natural catalytic RNAs, or ribozymes, are involved in ribosome, which is the main focus of this review.
RNA maturation. They catalyze phosphoryl-transfer reac-
tions that require the activation of either a ribose hydroxyl
group (e.g. hammerhead ribozyme, hepatitis delta ribo- Structure of the peptidyl-transferase centre
zyme, hairpin ribozyme, self-splicing introns and, perhaps, The catalytic center for peptide-bond formation is located
the spliceosome) or a water molecule (e.g. RNase P) for on the large ribosomal subunit. The large subunit in
nucleophilic attack of a phosphodiester bond [1]. Compared bacteria, 50S, is composed of two RNA molecules, 23S
with protein enzymes, which are chemically much more rRNA and 5S rRNA, and >30 proteins. The 50S subunit
diverse, ribozymes possess a limited repertoire of groups alone can synthesize peptide bonds as rapidly as the 70S
that take part in catalysis. Nevertheless, ribozymes use ribosome [3]. One approach to studying peptide-bond
several mechanisms, including general acid-base catalysis, formation is to crystallize ribosomes with substrates, tran-
metal ion-assisted catalysis, and substrate-alignment by sition-state analogs and products [4–15]. The high-resolu-
base-pairing and other interactions. Thus, they act in ways tion crystal structures of ribosomes have revealed that the
that are similar to protein enzymes [2]. peptidyl-transferase center is composed of RNA only, with
The most abundant natural ribozyme is the ribosome, no protein within 15 Å of the active site, which supports
which is a ribonucleoprotein particle that synthesizes earlier biochemical evidence of the key role of rRNA, rather
proteins and is the only natural RNA-based polymerase. than proteins, in the catalysis of peptide-bond formation
The ribosome binds two tRNA substrates, one with the [4,7,10,11,14]. The only protein that might be involved is
growing peptide chain attached by a high-energy ester L27 because the deletion of as few as three amino acids at
linkage to its 30 hydroxyl (the peptidyl-tRNA in the P site), the N terminus of L27 leads to impaired activity [16]. The
and the other with a single amino acid esterified to its 30 flexible N terminus of L27, which protrudes towards the
hydroxyl (the aminoacyl-tRNA in the A site) (Figure 1). interface of the bacterial 50S subunit, might contact the 30
During peptide-bond formation, the a-amino group of the terminus of the P site tRNA. However, some organisms do
A-site aminoacyl-tRNA attacks the carbonyl carbon of the not have L27 or any protein groups where the N terminus
P-site peptidyl-tRNA to produce a new peptidyl-tRNA that of L27 is located, which indicates that L27 is not part of an
is one-amino-acid longer in the A site and a deacylated evolutionary conserved mechanism (which is expected to
tRNA in the P site. The second enzymatic activity asso- employ identical residues in all organisms). However, it is
ciated with the peptidyl-transferase center is the hydro- not excluded that L27 contributes to tRNA positioning at
lytic cleavage of the ester bond in peptidyl-tRNA during the catalytic site [16].
termination of protein synthesis. In contrast to peptide- The acceptor ends of A-site and P-site tRNAs are located
bond formation, which is an intrinsic activity of the in a cleft of the 50S subunit on the side facing the 30S
subunit [6,13]. Their universally conserved 30 -terminal
Corresponding author: Rodnina, M.V. (rodnina@uni-wh.de)
CCA residues are oriented and held in place by interactions
Available online 8 December 2006. with 23S rRNA (Figure 2a). The conserved bases A2451,
www.sciencedirect.com 0968-0004/$ – see front matter ß 2006 Elsevier Ltd. All rights reserved. doi:10.1016/j.tibs.2006.11.007
Review TRENDS in Biochemical Sciences Vol.32 No.1 21

Figure 1. Peptide-bond formation on the ribosome. (a) Reaction scheme. The a-amino group of aminoacyl-tRNA in the A site (yellow) attacks the carbonyl carbon of the
peptidyl-tRNA in the P site (orange) to produce a new peptidyl-tRNA that is one amino acid longer in the A site and a deacylated tRNA in the P site. The peptidyl-transferase
center is on the 50S subunit (green). On the 30S subunit (gray), aminoacyl-tRNA are recognized according to the match between their anticodons and the codon of mRNA in
the A site. (b) Structure of the ribosome with bound tRNAs. The model is based on the crystal structure of E. coli ribosomes [12,58]. The tRNA positions in the P site and the
A site have been adjusted according to [13]. Ribosomal protein L1 and the L12 stalk [59] are shown for orientation.

U2506, U2585, C2452 and A2602 are located at the core of Enzymology of peptidyl-transfer reaction
the peptidyl-transferase center [5,9,10] (Figure 2b). The The active sites of enzymes contain residues that
crystal structures of Haloarcula marismortui 50S subunits participate in the chemical transformation of substrates.
complexed with different transition-state analogs have The main functions of these residues are to modulate the
revealed that the reaction proceeds through a tetrahedral electrostatic environment and chemical catalysis, includ-
intermediate with S chirality [9]. The oxyanion of the ing facilitation of proton-transfer reactions and covalent
tetrahedral intermediate is stabilized by a water molecule chemistry at the reaction center. General acids, general
that is positioned by nucleotides A2602 and U2584 [9]. bases and catalytic nucleophiles represent essential active-
The only atom within hydrogen-bonding distance of the a- site residues because they participate directly in the
amino group mimic is the 20 -OH of A76 of the P-site moiety formation and rupture of covalent bonds. Further contri-
of the transition-state analog [9]. N3 of A2451, which is butions to catalysis include electrostatic and structural
within hydrogen-bonding distance of the a-amino group in complementarity to the transition state, reorganization
the pre-reaction state (see later), seems to lose this inter- of water, and the use of the binding energy for substrate
action during the course of the reaction. The rRNA back- positioning and lowering the entropy of activation [18,19].
bone in the peptidyl-transferase center occurs in similar The ribosome does not employ covalent catalysis [20], but
conformations in 50S subunits from H. marismortui with all other strategies might be involved. The aim of studying
various ligands [6,7,9–11], 50S from Deinococcus radio- the enzymology of peptidyl transfer is to assess the
durans [17], 70S ribosomes from Thermus thermophilus contribution of different catalytic strategies and to reveal
with a P-site tRNA [14,15] and vacant 70S ribosomes from the catalytic role of ribosomal groups at the active site.
Escherichia coli [12]. However, elements of the peptidyl- Before peptide-bond formation, aminoacyl-tRNA must
transferase center might assume different orientations, for enter the A site of ribosomes that carry a peptidyl-tRNA in
example when a substrate is bound to the A site [10]. Some the P site. The rate of binding of aminoacyl-tRNA to the A
active-site nucleotides are particularly mobile, such as site (accommodation) is in the range of 10 s1 [21], which is
A2602, which lies between the A and P sites [5,14]. significantly slower than the intrinsic rate of peptide-bond
www.sciencedirect.com
22 Review TRENDS in Biochemical Sciences Vol.32 No.1

Figure 3. A-site substrates and substrate analogs. Reactive nucleophilic groups are
circled. A76 is the 30 -terminal residue of tRNA to which the amino acid/growing
Figure 2. Active-site residues of 50S subunits from H. marismortui with either polypeptide is attached; the rest of the tRNA molecule is not shown for simplicity.
bound substrate (a) or a transition-state analog (b). (a) Base-pairs formed between Puromycin is O-methyl tyrosine that is linked to N6-dimethyl adenosine via an
cytosine residues of the tRNA analogs in the A site (yellow) and P site (orange) with amide bond; C-puromycin is puromycin with an additional cytidine residue that is
23S rRNA bases (green) are indicated (PDB code: 1VQN) [10]. The a-amino group analogous to C75 of tRNA.
of the A-site substrate (blue) is positioned for the attack on the carbonyl carbon of
the ester that links the peptide moiety of the P-site substrate (green). (b) Transition-
state analog (TSA) bound to the peptidyl-transferase center (PDB code: 1VQP) [9]. activation is the same for the reaction on the ribosome
The hydrogen bond between the nucleophilic nitrogen (blue) and the 20 -OH of A76
at the P site is indicated. Structures in (a) and (b) are shown in different and in solution [26] (Figure 4). By contrast, enzymes that
orientations. employ general acid–base or covalent catalysis act by low-
ering the activation enthalpy of the catalyzed reaction.
formation, which was estimated to be 300 s1 [22]. Thus, the ribosome seems to use mechanisms of catalysis
Therefore, the mechanism of peptide-bond formation can- that are largely entropic in origin, such as substrate posi-
not be studied with the native aa-tRNA substrate under tioning in the active site, desolvation and electrostatic
current experimental capabilities. This can be circum- shielding [25].
vented partially by using A-site analogs of aminoacyl- If the ribosome used chemical catalysis with ribosomal
tRNA, which are either short fragments that mimic the residues acting as either general acids or bases, the reac-
30 end, or have a weaker attacking nucleophile, or both tion rate should depend on pH. In addition, the pKa of the
(Figure 3). Short 30 -end analogs of tRNA, such as puromy- a-amino group of aminoacyl-tRNA is 8 [27] and, because
cin and C-puromycin, which contains an additional only the deprotonated form of the nucleophile is active,
cytidine residue that is analogous to C75 of tRNA, bind the reaction rate should increase as pH increases, even if
to the peptidyl-transferase center rapidly and are incorpo- other ionizing groups are not involved. In fact, measure-
rated at rates of up to 50 s1 [23,24], which enables the ments in the 1960s and 1970s indicated that the reaction
chemistry step to be studied without being limited by rate increases with pH, and the increase per pH unit
accommodation. The reactions with short A-site-substrate indicated that a single ionizing group is involved [28,29].
analogs and the natural aminoacyl-tRNA are similar At the time (before the discovery of RNA catalysis) a
because they use the same reaction chemistry and are histidine residue of a ribosomal protein was proposed to
susceptible to the same inhibitors. However, the analogs act as a catalyst, which prompted a search for catalytic
lack the tRNA body, so the details of substrate positioning histidines in ribosomal proteins [30]. However, the condi-
might differ between small analogs and full-size tRNAs. tions of these early experiments did not enable the chem-
The ribosome brings about a 107-fold enhancement in istry step to be monitored rigorously, and the observed pH
the rate of the peptidyl-transfer reaction compared with dependence was probably caused by ionization of the
the second-order reaction between model substrates in a-amino group in many of the experiments.
solution [25]. This acceleration is achieved by lowering The ambiguities caused by ionization of the a-amino
the entropy of activation, whereas the enthalpy of group have been circumvented using an aminoacyl-tRNA
www.sciencedirect.com
Review TRENDS in Biochemical Sciences Vol.32 No.1 23

Figure 4. Entropic catalysis by the ribosome. Activation parameters are shown for the second-order uncatalyzed (knon) and ribosome-catalyzed peptide-bond formation at
substrate limitation (kcat/KM) or saturation (kcat) [25,26]. Abbreviations: S1, P-site substrate; S2, A-site substrate; P, reaction products. Reproduced, with permission, from
Ref. [25].

derivative in which a hydroxyl group replaces the amino of these rRNA residues in catalysis. The effects of mutating
group as reactive nucleophile [22,24,31]. The ribosome several 23S rRNA bases that are either in the, so-called,
readily accepts hydroxy-Phe-tRNA as a substrate and inner shell of the active site (A2541, U2506, U2585 and
catalyzes the formation of an ester bond instead of a A2602) (Figure 2b) [24,32–36] or adjacent to it (G2447)
peptide bond. The accommodation of hydroxy-Phe-tRNA [34,35,37], and the non-canonical pair (A2450-C2063)
is not rate-limiting because the formation of an ester bond [32,38] have been examined. Strikingly, none of these
is slow and, therefore, hydroxy-Phe-tRNA can be used to mutant ribosomes (except those with the A2450G-
monitor the pH-dependence of the reaction with the full- C2063U mutation, see later) have defects in the rate of
length substrate. Measuring the reaction rate at pH 6–9 peptide-bond formation with either full-length, intact aa-
reveals that changes in pH do not affect the reaction rate tRNA or C-puromycin [33,36,37]. However, eight of the
[22]. This indicates that catalysis by the peptidyl-transfer- nine mutants exhibited a strong reduction in the rate of
ase center is independent of pH, which argues against the reaction with puromycin (30–9400-fold reduction com-
involvement of ionizing groups of the ribosome in chemical pared with wild-type ribosomes) [24,33,36,37]. The largest
catalysis and indicates that general acid–base catalysis is decrease in reaction rate with the native aminoacyl-tRNA
not used to a great extent. Peptide-bond formation between (200-fold) occurred with the A2450G-C2063U double muta-
full-length peptidyl-tRNA and aa-tRNAs with a native tion [32]. It is likely that replacing the ionizing A+-C pair
amino group is also independent of pH [22]. Although with a G-U pair is less isosteric than expected and that this
the accommodation step, which is rate-limiting, might disturbs the structure of the active site [32].
mask part of a potential pH effect, these results are con- The role of A2451 deserves particular comment because
sistent with protonation and deprotonation events having the notion that A2451 acts as a catalytic residue in peptidyl
either small or no influence on the reaction. transfer has entered biochemistry textbooks. Kinetic ana-
In contrast to full-length substrates, a pronounced lysis of A2451U and G2447A mutants (the latter residue
pH-dependence has been observed with the minimal A-site forms an essential part of the charge-relay system that is
substrate, puromycin. Protonation of a ribosomal group (pKa postulated to bring about the required pKa shift of A2451)
7.5) reduces the rate of reaction by 150-fold [23,24], in two organisms, E. coli and Mycobacterium smegmatis,
which is much less than expected for an essential base. This argues strongly against an essential role of A2451 in
effect reflects a conformational rearrangement of active-site peptide-bond formation [33,36,37]. Rather, the A2451U
residues that impairs catalysis but does not take place with mutation alters the structure of the peptidyl-transferase
full-length aa-tRNA. This conclusion is corroborated by the center and changes the pattern of pH-dependent rearran-
observation that the reaction between A-site C-puromycin gements, as probed by chemical modification of 23S rRNA
and P-site peptidyl-tRNA is not influenced by the ionization [36]. A2451 seems to function as a pivot point in stabilizing
of ribosomal groups, which indicates that the presence of the the ordered structure of the active site, rather than by
cytidine residue (which mimics C75 of the A-site tRNA and, taking part in chemical catalysis [36].
presumably, its interaction with G2553) (Figure 2a), is
sufficient to induce and stabilize the active conformation Which other groups might be involved?
of the peptidyl-transferase center [23]. A group that is within hydrogen-bonding distance of the
nucleophilic group of transition-state analogs is the 20 -OH
Are bases of 23S rRNA involved in catalysis? of A76 of peptidyl-tRNA in the P site [9,10]. This has a
Identification of the ribosomal residues that form the crucial role in the reaction on both isolated 50S subunits
catalytic site has raised the question of the possible roles [20] and 70S ribosomes [39] but not in the uncatalyzed
www.sciencedirect.com
24 Review TRENDS in Biochemical Sciences Vol.32 No.1

reaction [25,40]. Substitution of 20 -OH of A76 by either According to molecular-dynamics simulations, the most
hydrogen (20 -deoxy) or fluor (20 -fluoro) reduce the activity favorable mechanism does not involve general acid–base
106-fold [39]. Notably, there are no catalytic Mg2+ ions or catalysis by ribosomal groups [42]. Rather, the catalytic
monovalent metal ions in the vicinity of the 20 -OH of the P- effect is of entirely entropic origin, which is in accordance
site tRNA [9,14] that might either promote catalysis with experimental results [25], and is associated with the
directly or shift the intrinsically high pKa of the 20 -OH reduction of solvent reorganization energy rather than
towards neutrality. The essential role of the 20 -OH of A76 of with either alignment or proximity of the substrate [42].
the P-site substrate indicates that the ribosome uses sub- The 20 -OH of A76 of the P-site tRNA might take part in a
strate-assisted catalysis (i.e. a mechanism in which a proton shuttle that bridges the attacking a-amino group
functional group of the substrate contributes to catalysis). and the leaving 30 oxygen, and several shuttle pathways
Several protein enzymes use substrate-assisted catalysis, can be envisaged [9,42,43,45,46]. The attack of the a-amino
including GTPases, serine proteases, type II restriction group on the ester carbon might result in a six-membered
endonucleases, lysozyme and hexose-1-phosphate uridylyl- transition state, in which the 20 -OH group donates its
transferase [41]. proton to the adjacent 30 oxygen while simultaneously
Evidence for the involvement of additional groups in receiving one of the amino protons. Such a scenario does
catalysis comes from recent crystal structures [9] and is not require a pKa shift of the 20 -OH group because of the
supported by molecular dynamics calculations [42,43]. The concerted nature of the bond-forming and bond-breaking
20 -OH of the ribose moiety of A2451 seems to be part of the events, and is in line with earlier suggestions that are
intricate hydrogen-bond network in the active site and to based on biochemical evidence with model substrates [47]
interact directly with the crucial 20 -OH group of the P-site and quantum-dynamic simulations [45].
tRNA. Consistently, substitution of the 20 -OH of A2451 by
hydrogen impairs peptidyl-transferase activity [44]. The mechanism of peptide-bond formation
The combined evidence supports strongly the idea that
Computational analysis entropic catalysis provides the major catalytic mechanism
One role of the 20 -OH of A76 of the P-site tRNA has of peptide-bond formation on the ribosome [25,42]. The
been suggested following computational analysis [42,43]. main supporting observations from structural analysis are
Molecular-dynamics simulations and free energy-pertur- the precise alignment of the A-site and P-site substrates by
bation simulations, in combination with an empirical interactions of their CCA sequences, and of the nucleophi-
valence-bond description of the reaction energy surface lic a-amino group of the A-site substrate with residues of
have been used to examine possible catalytic mechanisms. 23S rRNA in the active site [9,10,48–50]. The most favor-
Simulations of the reactant and tetrahedral intermediate able mechanism of catalysis involves intra-reactant proton
states of the peptidyl-transferase center reveal a stable, shuttling via the 20 -OH of A76 of the P-site tRNA, which
pre-organized, hydrogen-bond network that is poised for follows the attack of the A-site a-amino group on the P-site
catalysis (Figure 5). The peptidyl-transferase center might, ester bond (Figure 5) [9,42]. The reaction does not involve
thus, be viewed as a rigid environment of pre-organized chemical catalysis by ribosomal groups but might be modu-
dipoles that do not need to rearrange during the reaction. lated by conformational changes at the active site
[22,23,33–37]. In addition to bringing the reactive groups
into close proximity and precise orientation relative to each
other, the ribosome might work by providing a pre-orga-
nized electrostatic environment that reduces the free
energy needed to form the highly polar transition state,
shielding the reaction against bulk water, helping the
proton shuttle forming the leaving group, or a combination
of these effects.
The ribosome is an ancient RNA catalyst that accelerates
the peptidyl-transfer reaction by a factor of 107 [25]. It is
much less efficient than many protein enzymes, which use
chemical catalysis and accelerate reactions by up to 1023-
fold [51]. Apparently, evolutionary pressure has had a much
larger influence on increasing the speed and fidelity of the
rate-limiting steps of protein synthesis, which do not involve
chemistry, such as substrate binding [52], than on
the chemistry step of peptide-bond formation. This has
enabled the ribosome to retain its catalytic strategy during
the evolution of a pre-biotic translational ribozyme into a
Figure 5. Concerted proton-shuttle mechanism. The P-site and A-site tRNA modern ribosome. Thus, the catalytic mechanism employed
substrates are blue and red, respectively, ribosome residues are green, and
ordered water molecules that stabilize the developing charges are gray. The attack
by the ribosome seems to be a fossil from the RNA world.
of the a-NH2 group on the ester carbon results in a six-membered transition state,
in which the 20 -OH group of the A-site A76 ribose moiety donates its proton to the Future perspectives
adjacent 30 oxygen while simultaneously receiving one of the amino protons [9,42].
Alternatively, the water molecule (*) might be used for a proton shuttle. Modified, It is presumed that the catalytic mechanism of peptide-bond
with permission, from Ref. [60]. formation on the ribosome is highly conserved in all
www.sciencedirect.com
Review TRENDS in Biochemical Sciences Vol.32 No.1 25

organisms. Given the high degree of sequence conservation 4 Ban, N. et al. (2000) The complete atomic structure of the large
ribosomal subunit at 2.4 Å resolution. Science 289, 905–920
of rRNA, in particular at the peptidyl-transferase center
5 Bashan, A. et al. (2003) Structural basis of the ribosomal machinery for
[4,53,54], the active site for the reaction is likely to consist of peptide bond formation, translocation, and nascent chain progression.
rRNA in all organisms. However, the details of the position- Mol. Cell 11, 91–102
ing of groups in the peptidyl-transferase active site might 6 Hansen, J.L. et al. (2002) Structural insights into peptide bond
differ between species [4,17]. Most of the biochemical data formation. Proc. Natl. Acad. Sci. U. S. A. 99, 11670–11675
7 Nissen, P. et al. (2000) The structural basis of ribosome activity in
available have been obtained with E. coli ribosomes and,
peptide bond synthesis. Science 289, 920–930
recently, with ribosomes from the Gram-positive bacterium 8 Schlunzen, F. et al. (2001) Structural basis for the interaction of
M. smegmatis [36], and information from other organisms is antibiotics with the peptidyl transferase centre in eubacteria.
scarce. One of the future challenges is to obtain structural Nature 413, 814–821
and mechanistic information for eukaryotic ribosomes. 9 Schmeing, T.M. et al. (2005) Structural insights into the roles of water
and the 20 hydroxyl of the P site tRNA in the peptidyl transferase
The fundamental aspects of the mechanism of reaction. Mol. Cell 20, 437–448
peptide-bond formation have been revealed, so a major 10 Schmeing, T.M. et al. (2005) An induced-fit mechanism to promote
challenge is to probe the mechanism of the second impor- peptide bond formation and exclude hydrolysis of peptidyl-tRNA.
tant function of the peptidyl-transferase center, the hydro- Nature 438, 520–524
lytic cleavage of the ester bond in peptidyl-tRNA during 11 Schmeing, T.M. et al. (2002) A pre-translocational intermediate in
protein synthesis observed in crystals of enzymatically active 50S
the termination of protein synthesis. Crystal structures subunits. Nat. Struct. Biol. 9, 225–230
indicate that binding of the tRNA CCA end to the A site, 12 Schuwirth, B.S. et al. (2005) Structures of the bacterial ribosome at
which mimics the action of termination factors by inducing 3.5 Å resolution. Science 310, 827–834
peptide release, promotes a conformational rearrangement 13 Yusupov, M.M. et al. (2001) Crystal structure of the ribosome at 5.5 Å
resolution. Science 292, 883–896
at the active site to move the ester group of peptidyl-tRNA
14 Selmer, M. et al. (2006) Structure of the 70S ribosome complexed with
into a position that enables the attack of the water mole- mRNA and tRNA. Science 313, 1935–1942
cule [10]. In the context of this model, release factors 15 Korostelev, A. et al. (2006) Crystal structure of a 70S ribosome-tRNA
presumably promote the conformational rearrangement complex reveals functional interactions and rearrangements. Cell 126,
of at least a subset of the rRNA nucleotides that are 1065–1077
16 Maguire, B.A. et al. (2005) A protein component at the heart of an RNA
responsible for activation. The characterization of these
machine: the importance of protein L27 for the function of the bacterial
structural changes is another challenge. ribosome. Mol. Cell 20, 427–435
Finally, the hydrolytic activity of the peptidyl-transferase 17 Harms, J. et al. (2001) High resolution structure of the large ribosomal
center can be influenced from inside the peptide exit tunnel, subunit from a mesophilic eubacterium. Cell 107, 679–688
presumably by inducing an inactive conformation of the 18 Jenks, W.P. (1975) Binding energy, specificity and enzymatic catalysis:
the Circe effect. Adv. Enzymol. 43, 219–410
catalytic center via allosteric effects. There are several 19 Fersht, A. (1998) Structure and Mechanism in Protein Science, W. H.
examples of regulatory peptides that traverse the exit tunnel Freeman and Co
and inhibit release factor-dependent peptidyl-tRNA hydro- 20 Krayevsky, A.A. and Kukhanova, M.K. (1979) The peptidyltransferase
lysis in the peptidyl-transferase center, including the 22- center of ribosomes. Prog. Nucleic. Acid Res. Mol. Biol. 23, 1–51
residue peptide product of an ORF upstream of the gp48 gene 21 Pape, T. et al. (1998) Complete kinetic mechanism of elongation factor
Tu-dependent binding of aminoacyl-tRNA to the A site of the E. coli
of human cytomegalovirus [55] and the TnaC leader peptide ribosome. EMBO J. 17, 7490–7497
that, together with tryptophan, regulates the transcription 22 Bieling, P. et al. (2006) Peptide bond formation does not involve acid-
of the tryptophanase operon in E. coli [56,57]. In both cases, base catalysis by ribosomal residues. Nat. Struct. Mol. Biol. 13, 423–
termination at a stop codon is blocked. This yields ribosomes 428
23 Brunelle, J.L. et al. (2006) The interaction between C75 of tRNA and
that carry unhydrolyzed peptidyl-tRNA in the P site and are
the A loop of the ribosome stimulates peptidyl transferase activity.
stalled at the end of the coding sequence of the leader RNA 12, 33–39
peptide, which inhibits transcription of the downstream 24 Katunin, V.I. et al. (2002) Important contribution to catalysis of peptide
gene by an attenuation mechanism. The mechanism of bond formation by a single ionizing group within the ribosome. Mol.
signaling to the peptidyl-transferase center is not known, Cell 10, 339–346
25 Sievers, A. et al. (2004) The ribosome as an entropy trap. Proc. Natl.
but given the progress in ribosome mutagenesis, the avail-
Acad. Sci. U. S. A. 101, 7897–7901
ability of efficient translation systems and advances in 26 Rodnina, M.V. et al. (2005) Ten remarks on peptide bond formation on
structural studies, these questions are now within reach. the ribosome. Biochem. Soc. Trans. 33, 493–498
27 Wolfenden, R. (1963) The mechanism of hydrolysis of amino acyl RNA.
Biochemistry 338, 1090–1092
Acknowledgements
28 Maden, B.E. and Monro, R.E. (1968) Ribosome-catalyzed peptidyl
We thank Niels Fischer for preparing Figure 1b, and Venki Ramakrishnan
transfer. Effects of cations and pH value. Eur. J. Biochem. 6, 309–316
and Harry Noller for providing results before publication. Work in our
29 Pestka, S. (1972) Peptidyl-puromycin synthesis on polyribosomes from
laboratories is supported by the Deutsche Forschungsgemeinschaft, the
Escherichia coli. Proc. Natl. Acad. Sci. U. S. A. 69, 624–628
Alfried Krupp von Bohlen und Halbach-Stiftung, and the Fonds der
30 Diedrich, G. et al. (2000) Ribosomal protein L2 is involved in the
Chemischen Industrie.
association of the ribosomal subunits, tRNA binding to A and P
sites and peptidyl transfer. EMBO J. 19, 5241–5250
References 31 Fahnestock, S. et al. (1970) Ribosome-catalyzed ester formation.
1 Doherty, E.A. and Doudna, J.A. (2000) Ribozyme structures and Biochemistry 9, 2477–2483
mechanisms. Annu. Rev. Biochem. 69, 597–615 32 Hesslein, A.E. et al. (2004) Exploration of the conserved A+C wobble
2 Doudna, J.A. and Lorsch, J.R. (2005) Ribozyme catalysis: not different, pair within the ribosomal peptidyl transferase center using affinity
just worse. Nat. Struct. Mol. Biol. 12, 395–402 purified mutant ribosomes. Nucleic Acids Res. 32, 3760–3770
3 Wohlgemuth, I. et al. (2006) Rapid peptide bond formation on isolated 33 Youngman, E.M. et al. (2004) The active site of the ribosome is
50S ribosomal subunits. EMBO Rep. 7, 669–703 composed of two layers of conserved nucleotides with distinct roles

www.sciencedirect.com
26 Review TRENDS in Biochemical Sciences Vol.32 No.1

in peptide bond formation and peptide release. Cell 117, for the ribosome catalytic mechanism. ChemBioChem 6,
589–599 992–996
34 Polacek, N. et al. (2001) Ribosomal peptidyl transferase can withstand 47 Dorner, S. et al. (2003) Mononucleotide derivatives as ribosomal P-site
mutations at the putative catalytic nucleotide. Nature 411, 498–501 substrates reveal an important contribution of the 20 -OH to activity.
35 Thompson, J. et al. (2001) Analysis of mutations at residues A2451 Nucleic Acids Res. 31, 6536–6542
and G2447 of 23S rRNA in the peptidyltransferase active site of the 48 Moazed, D. and Noller, H.F. (1989) Intermediate states in the
50S ribosomal subunit. Proc. Natl. Acad. Sci. U. S. A. 98, 9002– movement of transfer RNA in the ribosome. Nature 342, 142–148
9007 49 Samaha, R.R. et al. (1995) A base pair between tRNA and 23S rRNA in
36 Beringer, M. et al. (2005) Essential mechanisms in the catalysis of the peptidyl transferase centre of the ribosome. Nature 377, 309–314
peptide bond formation on the ribosome. J. Biol. Chem. 280, 36065– 50 Kim, D.F. and Green, R. (1999) Base-pairing between 23S rRNA and
36072 tRNA in the ribosomal A site. Mol. Cell 4, 859–864
37 Beringer, M. et al. (2003) The G2447A mutation does not affect 51 Radzicka, A. and Wolfenden, R. (1995) A proficient enzyme. Science
ionization of a ribosomal group taking part in peptide bond 267, 90–93
formation. RNA 9, 919–922 52 Rodnina, M.V. and Wintermeyer, W. (2001) Fidelity of aminoacyl-
38 Bayfield, M.A. et al. (2004) The A2453-C2499 wobble base pair in tRNA selection on the ribosome: kinetic and structural mechanisms.
Escherichia coli 23S ribosomal RNA is responsible for pH sensitivity Annu. Rev. Biochem. 70, 415–435
of the peptidyltransferase active site conformation. Nucleic Acids Res. 53 Gutell, R.R. et al. (1985) Comparative anatomy of 16-S-like ribosomal
32, 5512–5518 RNA. Prog. Nucleic. Acid Res. Mol. Biol. 32, 155–216
39 Weinger, J.S. et al. (2004) Substrate-assisted catalysis of peptide bond 54 Noller, H.F. and Woese, C.R. (1981) Secondary structure of 16S
formation by the ribosome. Nat. Struct. Mol. Biol. 11, 1101–1106 ribosomal RNA. Science 212, 403–411
40 Sharma, P.K. et al. (2005) What are the roles of substrate-assisted 55 Cao, J. and Geballe, A.P. (1996) Coding sequence-dependent ribosomal
catalysis and proximity effects in peptide bond formation by the arrest at termination of translation. Mol. Cell. Biol. 16, 603–608
ribosome? Biochemistry 44, 11307–11314 56 Gong, F. and Yanofsky, C. (2002) Analysis of tryptophanase operon
41 Dall’Acqua, W. and Carter, P. (2000) Substrate-assisted catalysis: expression in vitro: accumulation of TnaC-peptidyl-tRNA in a release
molecular basis and biological significance. Protein Sci. 9, 1–9 factor 2-depleted S-30 extract prevents Rho factor action, simulating
42 Trobro, S. and Åqvist, J. (2005) Mechanism of peptide bond synthesis induction. J. Biol. Chem. 277, 17095–17100
on the ribosome. Proc. Natl. Acad. Sci. U. S. A. 102, 12395–12400 57 Cruz-Vera, L.R. et al. (2006) Changes produced by bound tryptophan in
43 Trobro, S. and Åqvist, J. (2006) Analysis of predictions for the catalytic the ribosome peptidyl transferase center in response to TnaC, a
mechanism of ribosomal peptidyl transfer. Biochemistry 45, 7049–7056 nascent leader peptide. Proc. Natl. Acad. Sci. U. S. A. 103, 3598–3603
44 Erlacher, M.D. et al. (2006) Efficient ribosomal peptidyl transfer 58 Vila-Sanjurjo, A. et al. (2003) X-ray crystal structures of the WT and a
critically relies on the presence of the ribose 20 -OH at A2451 of 23S hyper-accurate ribosome from Escherichia coli. Proc. Natl. Acad. Sci.
rRNA. J. Am. Chem. Soc. 128, 4453–4459 U. S. A. 100, 8682–8687
45 Das, G.K. et al. (1999) A possible mechanism of peptide bond formation 59 Diaconu, M. et al. (2005) Structural basis for the function of the
on ribosome without mediation of peptidyl transferase. J. Theor. Biol. ribosomal L7/12 stalk in factor binding and GTPase activation. Cell
200, 193–205 121, 991–1004
46 Changalov, M.M. et al. (2005) 20 /30 -O-peptidyl adenosine as a general 60 Rodnina, M.V. et al. (2006) Mechanism of peptide bond formation on
base catalyst of its own external peptidyl transfer: implications the ribosome. Q. Rev. Biophys. 8, 1–23

How to re-use Elsevier journal figures in multimedia presentations


It’s easy to incorporate figures published in Trends, Current Opinion or Drug Discovery Today journals
into your multimedia presentations or other image-display programs.

1. Locate the article with the required figure on ScienceDirect and click on the ’Full text + links’
hyperlink
2. Click on the thumbnail of the required figure to enlarge the image
3. Copy the image and paste it into an image-display program

Permission of the publisher is required to re-use any materials from Trends, Current Opinion or Drug
Discovery Today journals or from any other works published by Elsevier. Elsevier authors can obtain
permission by completing the online form available through the Copyright Information section of
Elsevier’s Author Gateway at http://authors.elsevier.com. Alternatively, readers can access the request
form through Elsevier’s main website at:

www.elsevier.com/locate/permissions

www.sciencedirect.com

You might also like