Cowan 1998 Metal Activation of Enzymes in Nucleic Acid Biochemistry
Cowan 1998 Metal Activation of Enzymes in Nucleic Acid Biochemistry
J. A. Cowan
The Evans Laboratory of Chemistry, The Ohio State University, 100 West 18th Avenue, Columbus, Ohio 43210
Received August 15, 1997 (Revised Manuscript Received January 15, 1998)
Contents
I. Introduction 1067
II. Physicochemical Properties of Magnesium 1068
III. Protein Binding Constants 1068
IV. Magnesium Analogues 1069
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
V. Kinetics/Inhibition 1070
VI. Mechanistic Issues 1072
A. Stoichiometry 1073
Downloaded via IISER BERHAMPUR on January 14, 2024 at 13:11:34 (UTC).
Figure 1. Hexahydrated Mg2+ binds to the wobble base pair GT in the DNA strand 5′-CGCGTG-3′ (left) and a site on
tRNAPhe (right).161,162 Hydrogen bonds are shown by the dotted lines, and the distances (in Å) are oxygen-oxygen internuclear
spacings. Key: W, water molecule; OPO, a backbone phosphate.
Enzymes in Nucleic Acid Biochemistry Chemical Reviews, 1998, Vol. 98, No. 3 1069
Figure 2. Typical protein coordination modes for magnesium- and calcium-binding proteins, illustrating the tendency for
extensive hydration for the former. The Ca2+ site in staphylococcal nuclease (left) shows four direct binding contacts from
Asp21, Asp40, the backbone carbonyl of Thr41, and a phosphate from the nucleotide substrate mimic.163,167 One of two
additional bound H2O molecules is also shown. For magnesium-dependent nucleases the hexaaquo Mg2+ site in bacteriophage
E. coli T4 ribonuclease H (middle),102 and the Mg2+ site in avian sarcoma viral integrase (right) are shown.49
Table 2. Magnesium Binding Motifs metal must also vary, while maintaining a KD at a
motifa enzyme ref(s) physiologically balanced level. The availability of
-NADFDGD- RNA polymerase 19
high-resolution structural data on several metal-
-GDD- RNA polymerase 21 dependent nuclease enzymes allows a critical analy-
-D-NSLYP- and DNA polymerase φ29 129, 130 sis of the coordination chemistry of the bound cofac-
-K-NS(L/V)YG- tor, and the implications for an understanding of the
-YGDTDS- DNA polymerase R 22 functional mechanism.
-YXDD- reverse transcriptase 147
-LXDD- telomerase 166 In contrast to the situation for Ca2+ binding
a
proteins, where metal-binding motifs can be identi-
Letters in bold are metal binding ligands. fied from primary protein sequences,17,18 few such
general motifs are known for Mg2+ (Table 2, and
Figures 3 and 4). Some of these include the homolo-
gous -NADFDGD- motif found in RNA polymerases
isolated from E. coli, M. thermoautotrophicum, S.
oleracea chloroplast, and S. cerevisiae Pol I, II, and
III. By use of Fe2+ as a probe with hydroxyl radical
footprinting, and studies of Asp mutants, Mustaev
and co-workers have provided evidence that the
conserved triad of carboxylates illustrated in Figure
4 is involved in Mg2+ coordination.19 Such residues
are also observed in DNA Pol I, and HIV-RT.20 A
-GDD- motif has been observed for poliovirus RNA-
dependent polymerase, and it has been proposed that
the Asp-Asp pair is involved in metal binding.
Overall, magnesium binding sites appear to show
a greater homology in their three-dimensional struc-
ture, rather than in primary sequence. This is
Figure 3. Comparison of magnesium coordination spheres especially so in the case of the restriction enzymes,
for selected magnesium proteins (the metal binding do- and the family of bacterial and viral RNase H
mains of E. coli ribonuclease H, the chemotaxal protein domains and viral integrases, illustrated in Figure
CheY, and the DNA repair enzyme exonuclease III). 5.1,8,10
in KD, many metal-binding proteins, and magnesium-
dependent enzymes in particular, show a rich variety IV. Magnesium Analogues
of coordination motifs (Table 2 and Figures 3 and
4).8,10 These variations tailor the KD as described, Divalent magnesium is essentially spectroscopi-
but also accommodate the diverse functional roles of cally invisible and has such a low electron density
the metal cofactor. For magnesium, the importance that it is difficult to distinguish by most spectroscopic
of metal-bound H2O in the catalytic mechanism of and crystallographic experiments. For this reason,
many enzymes has already been illustrated8,15,16 and the chemistry of magnesium-dependent enzymes has
can extend to a requirement for specific coordination been studied with the use of transition metal probes
geometries for bound waters (Figure 2). Conse- and analogues. The underlying assumption in many
quently, the coordination mode for the protein-bound studies of metal cofactor requirement in nucleic acid
1070 Chemical Reviews, 1998, Vol. 98, No. 3 Cowan
Figure 4. The conserved -NADFDGD- metal-coordination motif from a selection of RNA polymerases from E. coli, M.
thermoautotrophicum, S. oleracea chloroplast, and S. cerevisiae Pol I, II, and III (adapted from ref 19). The three aspartate
residues most likely form a metal binding core. Note the different sequence spacing of the three binding residues relative
to the three binding residues shown for E. coli RNase H and Che Y in Figure 3.
biochemistry is that magnesium analogues (espe- ion which promotes activity levels that are greater
cially manganese) show similar chemistry to that than the requirements for metabolism. For example,
displayed by Mg2+. It will be seen that this is not a Myers and co-workers have explored the metal
good general assumption. In fact, the stoichiometry dependence of the ribonuclease H activity of Thermus
and coordination mode of other metals may, and thermophilus DNA polymerase and have shown that
frequently do, differ from Mg2+. In particular, it is Mn2+ promotes higher levels of RNase H activity than
usually observed that Mn2+ (or other transition Mg2+.32 This activity is manifest by 5′-3′ exonu-
metals such as Co2+, Fe2+, or Zn2+) confers higher clease digestion of RNA/DNA hybrids. Mn2+ acti-
levels of activity for both native enzyme, and Asp or vates optimally over Mg2+; however, Mg2+ is most
Glu mutants (since these typically coordinate more likely the natural cofactor. In fact, many class I
tightly than the natural Mg2+ cofactor). For example, RNases H are activated by Mn2+ and can utilize
the plant-stimulated nuclease from Fusarium solani Mn2+, but class II RNases H are inhibited by this
is stimulated by Mn2+ > Ca2+ > Co2+ > Mg2+, but is metal ion.33
inhibited strongly by Zn2+.24 These transition metal In summary, while the use of analogues can
ions may also show changes in coordination geom- provide insight on function, there are differences in
etry, or induce conformation perturbations in active- coordination chemistry that can lead to changes in
site residues that influence substrate specificity. substrate selectivity and metal ion stoichiometry.
The basis for metal ion mutagenicity and nucle- Such problems need to be recognized and accounted
otide selectivity in human DNA polymerase β has for during data analysis.
been examined.25,26 It has been demonstrated that
distinct nucleotidyl reactivity patterns arise with a V. Kinetics/Inhibition
variety of divalent metal ions, while these also
promote a change in side-chain position of Asp192 Measurement of enzyme activity as a function of
relative to Mg2+. E. coli DNA polymerase I shows solution conditions provides an effective means of
reverse transcriptase ability; however, the fidelity is characterizing the metallobiochemistry of magnesium-
lower with Mn2+ as the catalytic cofactor,27 and so dependent enzymes. The dependence of activity on
Mn2+ and other divalent analogues such as Be2+ are solution pH and cofactor concentration can be used
mutagenic and carcinogenic.28 Mn2+ mutagenicity of to report directly on the ligand environment in the
T7 DNA polymerase and E. coli DNA polymerase I active site, and on metal cofactor stoichiometry.34-36
arises as a result of reduced nucleotidyl discrimina- This is particularly pertinent to the issue of the metal
tion, even though the activity levels are higher.29 cofactor requirement to effect catalysis, where there
In keeping with the aforementioned trends for is inconsistency between crystallographic and solu-
polymerase enzymes, loss of specificity for the TaqI tion measurements.10 This issue will be addressed
restriction endonuclease is observed with Mn2+ rela- later, but can be readily approached in solution
tive to the natural Mg2+ cofactor.30 Evidence for studies by a combination of kinetic and thermody-
three types of ternary complex with Mn2+ are pre- namic measurements.
sented, presumably each reflecting distinct Mn2+ Several methods for monitoring activity have been
coordination modes. Such loss of sequence specificity reported. Often these involve the use of radiolabeled
with the use of metal cofactors other than Mg2+ is substrates and evaluation of the rate of formation of
common for EcoRI and other restriction nucleases.31 labeled fragments.37-39 A more convenient method,
It has been earlier noted that Mn2+ will often when applicable, is direct monitoring of the change
promote higher levels of enzyme activity than Mg2+. in absorbance that results from hydrolysis of a
This should not be misconstrued as an indication that nucleic acid oligo- or polynucleotide substrate.16,34,36,40
Mn2+ is in fact the natural cofactor, but simply A variety of mechanistic models have been pro-
reflects the distinct chemistry of this transition metal posed and used to rationalize the metal dependence
Enzymes in Nucleic Acid Biochemistry Chemical Reviews, 1998, Vol. 98, No. 3 1071
Figure 5. Homology in the tertiary structural arrangement of the active-site metal binding residues in some magnesium-
binding proteins. (A) (Left) Schematic illustration of the two Mn2+ ions identified in the active site of the HIV-RT RNase
H domain after doping the cofactor-free crystal with 45 mM Mn2+. The positions of key carboxylate residues are shown.
(Right) The structurally homologous sites in E. coli RNase H and avian sarcoma viral integrase, showing the positions of
conserved carboxylate residues (the numbering for ASV-IN is shown in parentheses). Only one metal-binding site (Mg2+
or Mn2+) has been identified.49,97 The metal ions in bold are the major homolgous sites common to the enzymes shown in
A and B. (B) A comparison of the metal cofactor binding domains, and other features of the catalytic pockets of EcoRI and
EcoRV, illustrating the high degree of local homology for these two restriction endonucleases.38
of nuclease activity. This issue has been most form of competitive binding between the metal ions
thoroughly addressed by the laboratories of and enzyme, to the substrate. Such a viewpoint is
Pingoud,38,41 Halford,39,42 and Cowan.17,34 A general supported by the similarity of this response over a
observation in almost all plots of enzyme activity as diverse group of enzymes,34-36,43 and the similarity
a function of metal cofactor concentration is an initial of apparent binding constants for metal-substrate
increase in activity with increasing [Mg2+], followed complexation determined from these kinetic and
by a gradual decrease in activity (Figure 6). While thermodynamic measurements. The effect is clear
direct enzyme inhibition is a possibility, no direct in the case of E. coli RNase H, where there is no
evidence has been advanced for such, and the author evidence for secondary magnesium binding sites on
of this review is inclined to see this effect as a general the enzyme,34 while metal-hybrid and metal-
example of substrate inhibition, where the metal ions enzyme binding constants (KI ≈ 14 mM, and KMET ≈
bind to the nucleic acid substrate, perturbing the 0.2 mM, respectively), obtained by kinetic measure-
charge density of the substrate and its interaction ments for E. coli RNase H34 and by independent NMR
with the enzyme. This may also be thought of as a and calorimetric titration methods, are found to be
1072 Chemical Reviews, 1998, Vol. 98, No. 3 Cowan
Figure 6. Variation of initial velocity (vo) with magnesium concentration for E. coli RNase H (left) and the exonuclease
activity of Klenow (right). The data clearly shows the inhibitory influence of high concentrations of Mg2+, and was fit to
a model that assumes metal-mediated substrate inhibition.34,62
affinity metal ion (site B) was discussed in terms of However, this question can be addressed by kinetics
electrostatic stabilization of the negative charge that methods. One strategy that has been used success-
is transferred from hydroxide to phosphate. Such a fully involves determination of metal binding param-
model was also favored for one metal ion enzymes eters from kinetics experiments, and comparison of
(such as Staphylococcal nuclease) where the single the fitted binding parameters with results from
cation carries out both roles.58 In this, and other independent calorimetry experiments carried out in
cases where single metal ion cofactors have been the presence of substrate analogues.34,62 The results
identified, the metal is typically divalent calcium. are analyzed with a view to identifying specific
Inasmuch as all Ca2+-dependent nucleases act by one mechanistic models that yield consistent results from
metal ion catalytic pathways, it seems reasonable to each independent approach and, importantly, that
infer that magnesium, with a charge density that is are consistent with the physiological availability of
significantly higher than that of Ca2+,67 would also the divalent metal ions. Nature imposes well-defined
show a requirement for one metal cofactor. This is limits on metal binding stoichiometry. This is con-
indeed generally substantiated by available evidence. trolled by both the magnitude of the dissociation
constant and the bioavailability of metal cofactor in
A. Stoichiometry vivo. For example, a 40 mM binding affinity is of
The stoichiometry of metal cofactor that is required little significance if the available concentration of the
to mediate phosphate ester hydrolysis is an impor- cofactor is 0.5 mM. In effect, the dissociation con-
tant variable for mechanistic considerations. This stant for a biologically relevant divalent magnesium
problem has been the subject of debate as a result of cofactor should be ∼0.5 mM, since this is the con-
contradictions between solution and crystallographic centration of free magnesium in a cellular environ-
studies47-50 and has fueled recent efforts by chemists ment.14 This is a powerful constraint in assessing
to construct synthetic models to address this issue.51-53 metal ion stoichiometry and distinguishing kinetic
Several binuclear metallohydrolases that contain a models.
pair of bridged transition metal ions at the active site Evaluations of metal cofactor stoichiometry by
have been characterized.2,54 However, while there crystallographic and solution studies have been made
are a few examples of metabolic enzymes that appear for several enzymes (summarized in Table 3). There
to utilize two magnesium centers in close proximity, is now extensive data available for E. coli Klenow
such as xylose (glucose) isomerase,55 and glutamine fragment, RNase H, Exo III, and EcoRV. Some of
synthase,56 the evidence for binuclear magnesium in these results will be presented in greater detail in
nucleic acid hydrolysis is not so clear-cut and is the sections dealing with specific enzymes that follow;
supported by a rather small body of crystallographic however, some general comments can be made that
data with relatively few confirmatory results from shed considerable light on this dichotomy. X-ray
solution experiments. The two metal ion mechanism structures of proteins with labile Mg2+ cofactors are
has been proposed for numerous metal hydrolase,2 usually obtained by doping preexisting crystals of
nuclease,39,57,84 polymerase,58 and ribozyme reac- “apo” enzyme with very high concentrations of metal
tions;59,60 however, solution thermodynamic and ki-
cofactor to permit population of metal binding sites
netic measurements on substrate-free enzyme typi-
in a kinetically reasonable time frame. Often the
cally indicate a single metal cofactor (Table 3). An
metal used is not the natural magnesium cofactor,
important caveat to note is that simple binding
which has a low electron density and is difficult to
experiments do not directly indicate the number of
metal ions required for catalytic turnover; for ex- distinguish from oxygen in water, but rather a
ample, it is possible that an additional metal binding transition metal ion carrying a higher electron den-
site is found in the enzyme-substrate complex, as sity which is more easily observed. Increasingly,
suggested for the Klenow fragment of DNA pol I.48,61 crystallographic studies are using the natural Mg2+
cofactor and making use of the strong pattern of
Table 3. Summary of Metal Binding Constants and octahedral coordination by metal-bound H2O to dis-
Stoichiometries Evaluated from Solution tinguish Mg2+ from general solvent water. Although
Measurements and Crystallographic Analysis10 much useful information has come from studies with
crystallography solution magnesium analogues, a few simple caveats must be
no. of [M2+] no. of KD
kept in mind. (1) Coordination sites in a preexisting
enzyme M2+ used (mM) M2+ (mM) structure of an apoenzyme may not exist in the
S. nuclease 1 0 (Ca2+)a 1 0.005
structure of the active metalloenzyme. (2) Coordina-
Klenow exonuclease 2 12 to 50 (Mn2+) 1 0.40 tion preferences for Mg2+ and analogues are often
Exo III 1 20 (Mn+) 1 0.11 different. (3) Under conditions of high metal con-
EcoRV 1 or 2 10 to 30 (Mg2+) 1 NDb centration used in doping experiments, weak and
EcoRI 1 1 ND physiologically irrelevant sites may be occupied.
HIV-RNase H 2 45 (Mn2+) ND
E. coli RNase H 1 50 to 100 (Mg2+) 1 0.5 The paradigm of the two metal ion model is
Che Y 1 50 (Mn2+) 1 0.2 perhaps best defined as the requirement for two
ASV-IN 1 200 (Mg2+) ND metal ions, located in close proximity (<4 Å), that
a The Ca2+ cofactor binds sufficiently tightly that it can be are bridged by a common substrate. This allows a
isolated in the metal-bound state and requires no additional clear distinction from other enzymes that may bind
solution Ca2+(aq) to maintain saturation of this site. b ND, not two metal cofactors in the same catalytic domain, but
determined.
do not function as a coherent catalytic unit. For
1074 Chemical Reviews, 1998, Vol. 98, No. 3 Cowan
Figure 13. (A) Schematic illustration of the catalytic pocket for the EcoRV-substrate complex. Note that only one metal
cofactor is identified. (B) Proposed model for the transition state, showing recruitment of a second metal cofactor. However,
in other work, the critical residue Glu45 has been mutated to residues that do not bind magnesium and the mutants were
found to be active. (C) Schematic illustration of the catalytic pocket for the EcoRV-product complex. Note that now two
metal ions are observed, although metal ion concentrations in the range of 10 to 30 mM were used. Adapted from ref 50.
metal ion transition-state model. Later, Halford and ions is required for catalysis; the other appears to
co-workers also examined the metal requirement for promote binding of the substrate molecule. This
EcoRI and EcoRV and correctly pointed out that the most likely corresponds to the additional metal
Pingoud analysis incorrectly assumed total cooper- cofactor suggested by the stopped-flow kinetic studies
ativity.39 They suggested a catalytic requirement of of Halford and co-workers.39,42
one metal cofactor for the former, and a two metal Specific recognition of DNA by EcoRV can be
ion mechanism for the latter. It was observed that promoted by Ca2+ ion. By analyzing gel shifts with
Ca2+ inhibits Mg2+ activation of EcoRV, but stimu- a permuted set of DNA fragments, the degree of DNA
lates Mn2+ activation of the same enzyme, suggesting bending was shown to be similar to that seen in the
that Ca2+ can displace a metal from one of the two crystal structure of the cognate DNA-protein com-
sites. Such a model was supported by stopped-flow plex in the presence of Mg2+.57 Calcium ion, there-
fluorescence experiments that examined the influ- fore, mimics the ability of Mg2+ to generate a specific
ence of metal ions on DNA binding to EcoRV.42 Two protein-metal-DNA complex, but is incapable of
Mg2+-dependent transitions were observed and in- inducing the cleavage reaction. This may also reflect
terpreted in terms of a model where one Mg2+ binds the tendency of Ca2+ to engage in inner-sphere rather
to the active site before phosphodiester hydrolysis, than outer-sphere interactions, if the latter are
with a second binding subsequently to a preformed required for catalysis. Outer-sphere activation is also
enzyme-DNA complex. consistent with the absence of any significant influ-
These facts notwithstanding, several problems ence on the relative cleavage rates for Mg2+- versus
nevertheless remained with this interpretation. First, Mn2+-induced activity with phosphorothioate sub-
the crystal complex with the substrate lacked activ- strates, since direct coordination to the metal cofactor
ity, even in the presence of saturating magnesium, should discriminate between a hard Mg2+ ion and
and the active site bound only one divalent magne- softer Mn2+ ion.79
sium ion (Figure 13A). Second, a Mg2+ concentration Recent results for an Ile91Leu mutant show a
of g10 mM was required to populate both sites; greater than 1000-fold decrease in activity.80 A
however, the key issue is whether these sites can be change in metal ion dependency was observed, with
populated under physiological conditions (<1 mM). a preference for Mn2+ rather than Mg2+ as for the
Third, and perhaps most importantly, the critical native enzyme. The mutant shows evidence of non-
residues that define the binding site of the second specific DNA binding in gel-shift experiments, as does
metal ion (Mg22+ in Figure 13B), especially Glu45, the native. At noncognate sites that differ from
have been mutated to nonbinding residues (such as EcoRV by one base pair, Mn2+ gives higher cleavage
Ala) and the mutants were found to be active.78 It rates than Mg2+, but the effect is reversed for the
has therefore been suggested that a second, but Ile91 Leu mutant.81 Since each mutant requires the
spatially distinct and noncatalytic site might exist,77 same carboxylate residues for binding, compared to
since the DNA binding specificity is increased by native, it is likely that the switch in metal preference
Mg2+ binding to a site that is distinct from that of arises from a structural perturbation of the metal
the catalytic center (bound by Glu 45, Asp 74, Asp binding pocket, as described in section IV.
90). Also, the Ala triple mutant which cannot bind
catalytic metal, was found to bind specifically to DNA B. Other Restriction Nucleases
in the presence of Mg2+. Phosphorothioate deriva-
tives provided further evidence for a Mg2+ site There is a paucity of structural and mechanistic
distinct from the catalytic center. Only one of these information concerning other restriction enzymes,
1078 Chemical Reviews, 1998, Vol. 98, No. 3 Cowan
Zn2+ known to inhibit apoptosis also inhibit the clease or ribonuclease functions that are required for
activity of these Ca2+/Mg2+ nucleases, which are their overall operation. A large number of poly-
found extensively in lymphocyte tissue. Other than merases have now been structurally characterized
inhibition by zinc, the activity of these nucleases may and can be divided into three major categories;
be regulated at the transcriptional level. For ex- namely RNA and DNA polymerases, and reverse
ample, an 18 kD Ca2+/Mg2+-dependent nuclease transcriptases. Divalent magnesium is an essential
(putatively a cyclophilin) identified in lymphoid cells cofactor for these three classes of polymerase enzyme,
undergoing apoptosis, is regulated by the glucocor- and serves a variety of roles.
ticoid receptor.111,112 RNA polymerases are commonly used in transcrip-
For many nucleases the metal cofactors appear to tion, where the information encoded in a genomic
mediate not only direct catalytic activity, but also DNA (or in some retroviruses, a genomic RNA)
structural changes that regulate substrate recogni- molecule is transcribed into a single strand of mes-
tion and binding specificity. In the absence of Mg, senger RNA. DNA polymerases replicate double-
the periplasmic nuclease from Streptomyces antibi- strand DNA, and usually possess an exonuclease
oticus does not form tight specific complexes with activity as a proof-reading function. Finally, reverse-
cognate (dG) sequences (n g 4).113 Changes in site transcriptases are commonly associated with retro-
specificity of single strand specific endonucleases, viruses. This enzyme catalyzes three reactions: first
such as mung bean nuclease, have been observed the conversion of single-strand genomic RNA to a
with a change in Mg2+ concentration, and also by double-strand RNA-DNA hybrid (DNA polymerase
varying temperature and ionic strength.114 Confor- function); second, the digestion of the RNA strand of
mational changes of supercoiled DNA were proposed. the hybrid (ribonuclease H function); and third, a
As previously noted for nuclear nucleases, many second DNA polymerase reaction that converts the
fungal multisubunit nucleases require two types of single-strand DNA to double-strand DNA prior to
metal for optimal activity. Studies with metal- integration into the host cell’s genomic DNA.
selective chelating agents suggest a Ca2+ requirement
to stabilize quaternary structure and a Mg2+ require- A. RNA Polymerases
ment to promote activity.115 The fungal nuclease In addition to its well-known catalytic role, Mg2+
from Aspergillus may be related to Serratia nuclease, induces structural changes in the E. coli RNA poly-
and again Ca2+ has been found to increase the merase open complex with the λPR promoter that
thermostability of this enzyme.36,116 The acid-soluble have been examined by chemical and enzymatic
nuclease γ has been purified from U. maydis cell footprinting.119-121 Sigman and co-workers122 have
extracts.117 Either Mg2+ or Ca2+ promotes nicking demonstrated a Mg2+ requirement for E. coli RNA
of one strand of the substrate DNA duplex, while polymerase to form single-stranded DNA structures
Mn2+, Co2+, or Zn2+ promotes double-strand cleavage. that resemble open transcription complexes. The
Many nucleases serve as general hydrolytic agents influence of Mg2+ on formation of the transcription
to break down nucleic acids. One of the best-studied bubble upon binding of E. coli RNA polymerase to
examples is DNase, although this class of enzyme the T7A1 promoter has also been examined by use
uses Ca2+ as the principal catalytic site, with Mg2+ of chemical probes, including hydroxyl radical foot-
and other Ca2+ ions serving as structural cofactors. printing.19,120 These data suggest that formation of
Early work on pancreatic DNase A demonstrated the transcription active complex requires two steps:
distinct Mg2+ and Ca2+ sites, with the number of sites one Mg2+-dependent, and the other Mg2+-indepen-
varying with solution pH.118 Aaqvist and Warshel dent. The data also suggests that one or two Mg2+
have examined the effect of metal ion substitution ions activate polymerase activity, while other ions are
on the activity of staphylococcal nuclease by free responsible for enlarging the transcription bubble.
energy perturbation calculations.58 Their results Magnesium binding motifs have been identified in
suggest a protein architecture that is designed to primase, a ssDNA-dependent RNA polymerase (Table
optimize the activity of Ca2+. More electrophilic 2 and Figure 4).123 Also, the Mg2+ binding site of E.
cations with large hydration free energies increase coli RNA polymerase has been investigated by use
the activation barrier for hydrolysis as a result of of hydroxyl radical footprinting,19 and mutagenesis
overstabilization of the putative HO- nucleophile. studies have implicated a -NADFDGD- motif in Mg2+
Optimization of metal cofactors is related to the dual binding (Table 2 and Figure 4).
requirement of stabilizing both the entering nucleo- The structure of bacteriophage T7 RNA polymerase
phile, and also the transition state for the reaction. has been reported at 3.3 Å resolution.124 Both
One obvious problem in over-generalizing these is- Asp537 and Asp812 have been proposed as metal-
sues stems from the discrete coordination chemistries binding residues; however, the Asp537Asn and
and pKas for metal-bound H2O that must be consid- Asp812Asn mutants demonstrate only slightly lower
ered when comparing results from transition metals Kds for Mn2+ relative to native enzyme, although the
versus alkaline earths. activity was greatly reduced.125 Scatchard analysis
of EPR Mn2+ binding data has suggested two bound
ions;125 however, such an approach is not a reliable
VIII. Polymerases method for the estimation of metal ion stoichiometry.
Polymerases are enzymes that catalyze the replica- It is proposed that the Asp537 and Asp812 residues
tion and synthesis of strands of DNA or RNA from a bridge the metal cofactors and that subtle changes
single-strand or double-strand template polynucle- in the geometry of these residues might have a strong
otide. Some of these enzymes contain other exonu- influence on catalytic activity.
Enzymes in Nucleic Acid Biochemistry Chemical Reviews, 1998, Vol. 98, No. 3 1081
Table 5. Thermodynamic Binding Parameters from cofactor rather than Mn2+ or any other transition
Calorimetric Studies of Mg2+ Binding to Klenow62 metal ion; however, studies with Mn2+ allow a direct
Mg2+ Mn2+ comparison with crystallographic data. In the pres-
parameters Mg2+ (+ TMP2-) Mn2+ (+ TMP2-) ence of TMP2-, one Mg2+ binds at each of the active
Polymerase sites (one at the 3′-5′ exonuclease site, and one at the
no. of bound cations 1 1 1 1 polymerase site) and the binding constants are
K1 (mM) 0.16 0.08 0.02 0.03
slightly augmented relative to binding in the absence
3′-5′ Exonuclease of TMP2-. A similar result is obtained with Mn2+ and
no. of bound cations 1 1 1 1
K2 (mM) 0.40 0.20 0.60 0.07 is supported for both cofactors by a detailed kinetic
analysis.62
Mutation of the aspartic acid residues of the -GDD- Warshel and co-workers have carried out a theo-
acid sequence motif of poliovirus RNA-dependent retical analysis of the role of metal ions in phosphate
RNA polymerase results in enzymes with different ester hydrolysis catalyzed by the Klenow fragment
metal ion requirements for optimal activity.21 The of E. coli DNA polymerase I64 and argued against a
repeated carboxylate motif is a common feature of metal-bound hydroxide as the active nucleophile, but
polymerase domains (Table 2 and Figure 4). Both favored an external hydroxide that is stabilized
aspartate residues are required for Mg2+ binding; electrostatically by the metal cofactor. Stabilization
however, Asn mutations are found to be active with of this hydroxide could be accounted for in terms of
transition metal replacements. Mutation of either the electrostatic influence of the high-affinity metal
Asp to Asn knocks out activity with Mg2+ as cofactor, ion alone (Figure 7A).
however, the -GDN- mutation is active with Mn2+.21
The conserved amino acids in region I (-YGDTDG-)
Inasmuch as an amide linkage may serve as a ligand,
of human DNA polymerase R are most likely involved
especially to Mn2+, it is possible that the activity of
this mutant reflects retention of the metal cofactor. in formation of a metal binding domain (Table 2 and
Figure 4). Both Asp1002Asn and Thr1003Ser mu-
B. DNA Polymerases tants were found to utilize Mn2+ more effectively than
Mg2+;22,23 however, these mutants show enhanced
The Klenow fragment is a proteolysis product of misinsertion fidelity as a result of a 850- and 62-fold
DNA polymerase I, an important enzyme in DNA higher KD. Mutation of Asp1004 resulted in a loss
replication. This fragment contains both the 3′-5′ of activity for either metal cofactor, presumably
exonuclease and polymerase domains126 and has arising from loss of metal binding. The results are
become a paradigm for a class of metalloenzymes that consistent with metal binding to each Asp residue;
follow a two metal ion mechanism of phosphate ester more critically for Asp1004. In the case of Mn2+
hydrolysis.48,61 This is based extensively on the activation, a significant increase in misincorporation
crystallographic work of Steitz and co-workers.48,61
of nucleotide bases was observed, but not for Mg2+,
Early studies of Klenow demonstrated two metal ions
even for the active Asp/Asn and Thr/Ser mutants
(illustrated in Figure 7B) bound in the exonuclease
active site in the presence of the substrate mimic described earlier. The effect for Mn2+ was manifest
thymidine monophosphate (TMP2-).61 In the absence by the influence of this cation on KMs, and the results
of TMP2- only one metal ion (A in Figure 7B) was suggest that subtle changes in Mn2+ coordination
observed to bind. Multiple bound ions have also been chemistry resulting from these mutations gives rise
proposed on the basis of solution studies of Klenow to structural changes in the active-site pocket that
by kinetic, optical, and EPR methods with Co2+ and diminish its ability to recognize specific nucleotide
Mn2+ cofactors.127,128 After these studies, other en- bases. The consequences of these structural pertur-
zymes have been crystallographically characterized bations for increased mutagenicity were discussed in
with evidence of two active-site metal ions.97 A section IV.
common feature of each of these studies is the use of The DNA polymerase φ29 shares many of the
Mn2+, and occasionally Zn2+ and Co2+, as analogues characteristics of R-like DNA polymerases, including
for the presumed natural Mg2+ cofactor, since the two conserved regions of amino acid similarity (D-
heavier manganous ion facilitates discrimination of
NSLYP and K-NS(L/V)YG) that have been proposed
the metal ion from solvent water.49 Subsequent
to form part of the active site.129,130 Tyr to Phe
characterization of bound Mg2+ is often carried out
by difference methods, by subtracting electron den- mutations of the Tyr residues in these two domains
sity maps of an enzyme from those obtained with a were found to influence the polymerization activity
manganese or heavy metal complex of an enzyme;48 of Mg-NTP substrates, but not Mn-NTP substrates.
however, the difficulties associated with the analysis In the first domain, an Asp to Glu mutation greatly
of metal cofactor stoichiometry from crystallography influenced the polymerase activity, but not the exo-
have been alluded to earlier. nuclease function. Polymerase activity could be
The binding stoichiometries to Klenow for both partially recovered by use of Mn-NTPs. These results
Mg2+ and Mn2+ have been examined by titration are consistent with involvement of these two amino
calorimetry (Table 5), both in the presence and acid regions in the polymerase, rather than the
absence of the substrate analogue thymidine mono- exonuclease domain. Also, the recovery of partial
phosphate, TMP2-.62 Considering their physiological activity for the Asp to Glu mutation with Mn2+ is
availablity, divalent magnesium is the physiological consistent with direct interaction with the NTP-
1082 Chemical Reviews, 1998, Vol. 98, No. 3 Cowan
Figure 17. R-Carbon trace structures of rat DNA polymerase β, HIV-RT, E. coli DNA pol I, and T7 RNA polymerase,
showing the positioning of common aspartate residues. Adapted from ref 132.
Figure 18. Stereoview of the polymerase active site of rat DNA polymerase β, and a schematic model of the reaction
chemistry. Adapted from ref 131.
bound metal cofactor: a more favorable interaction reactions, one divalent cofactor is carried into the
arising in the case of Mn2+, relative to Mg2+. active site as a tight complex with dNTPs (shown
In addition to DNA polymerases R, the polymerases with ddCTP in Figure 18).131 There is an additional
β have also been the subject of intense structural magnesium binding site that is most likely stabilized
investigation. Ternary complexes of rat DNA poly- by the high charge density resulting from the mul-
merase β complexed with a DNA template primer tiple carboxylates and the triphosphate. The Mg2+
and dideoxycytidine (ddCTP) have been determined bound to the nucleotide triphosphate is well-known
at 2.9 and 3.6 Å and yield significant mechanistic to facilitate nucleophilic attack,133-136 although the
insight.131,132 Three conserved aspartates are located 3′-OH of the attacking sugar need not necessarily be
at the active polymerase site of all DNA polymerases directly coordinated to the magnesium center as
thus far characterized, while RNA polymerases show shown schematically in Figure 18, but may be
two of these (Figure 17). For nucleotidyl transferase stabilized electrostatically as an outer-sphere com-
Enzymes in Nucleic Acid Biochemistry Chemical Reviews, 1998, Vol. 98, No. 3 1083
plex.16,64 Also, the use of the metal-bound carboxylate activity and the ability to catalyze specific removal
as a catalytic base is not appealing and has been of the tRNA replication primer. These observations
discussed in the literature.137,138 Previously it has point to obvious structural and/or coordination changes
been shown that a DNA polymerase can be converted for Mg2+, relative to Mn2+ activation, and may even
to an RNA polymerase by changing the metal cofactor be indicative of distinct coordination sites.
from Mg2+ to Mn2+.139 Figure 18 shows that all three Conservative mutation Q151N in the polymerase
phosphates on ddCTP bind to one or other of the two domain of HIV-1 RT demonstrated similar activity to
metal cofactors, and so it is likely that the metal ions that of WT,143 although the Q151A mutant demon-
help to position the nucleotide. It is argued that a strated reduced activity. The preference for divalent
slight change in orientation of ddCTP would be cation changed from Mg2+ to Mn2+. The influence of
sufficient to reduce steric hindrance at the C2′ ribose each of the p66 and p51 subunits on substrate and
of ddCTP and allow RNA to become as good a inhibitor binding, and metal ion preference, has been
substrate as DNA.131,139 This could be brought about tested by subunit-selective mutagenesis studies on
by subtle changes in the coordination chemistry of Glu 89. Evaluation of Michaelis and inhibition
Mg2+ relative to Mn2+, as suggested earlier in section constants, and metal preference was evaluated for
IV on the origins of mutagenicity. E89G mutants of HIV-1 RT (for both the p66 and p51
In a later study of rat DNA pol β, using Mn2+ as domains).144 Only those mutants containing a p66
metal cofactor, two metal centers were observed even WT domain demonstrated a preference for Mg2+ (i.e.,
in the absence of a nucleotide triphosphate.140 These p51 mutants had no effect), while substrate and
two cations were held to the enzyme by a single inhibitor binding were also unchanged. However,
bridging aspartate: a mode of binding that seems absolute activity was influenced by mutation of the
unreasonable for Mg2+. It is likely that these sites p66 Glu89 residue. Clearly, the p66 domain contrib-
reflect specific binding preferences for Mn2+ under utes the most important side-chain functionality to
the high metal concentrations used. Other mecha- metal-mediated chemistry.
nistic aspects of polymerase activity, relating to The inhibitory influence of the antiviral drugs AZT
orientation of the primer template, have been de- (azidothymidylate) and N-ethylmaleimide (NEM) on
bated; however, this issue is not relevant to the focus the RNase H activity of HIV-1 RT is distinct when
of this review.137,138 using either Mg2+ or Mn2+ as cofactors.145 AZT
monophosphate is a competitive inhibitor in the
C. Reverse Transcriptase presence of Mg2+, but is noncompetitive with a Mn2+
Blain and Goff have examined the RNase H activ- cofactor. This observation is also consistent with
ity of native and mutant Moloney murine leukemia distinct metal-binding sites for Mg2+ and Mn2+ in the
virus reverse transcriptase.141 Two of these mutants, RNase H active site. Furthermore, the mode of
designated ∆5 and ∆C, contain deletions in the substrate cleavage is distinct, with endonucleolytic
RNase H domain. These were found to be active with hydrolysis of a poly(rA)/poly(dT) substrate in the
in vitro assays using Mn2+, but were found to be presence of Mg2+ and both endo- and exonucleolytic
inactive in vivo, where the physiological cofactor is hydrolysis in the presence of Mn2+.145,146 With poly-
Mg2+. These results are consistent with perturbation (rG)/poly(dC) both exo- and endonucleolytic cleavage
of the metal cofactor binding site in the RNase H are observed for Mg2+ and Mn2+. The authors sug-
domain, but apparently the deletion mutations do not gest that the RNase H domain of HIV-1 RT under-
completely remove the Mn2+ binding propensity. goes conformational changes upon binding hybrid
Three point mutants (R657S, Y598V, and S526A) substrates and that these changes depend on the
were found to show full activity with Mn2+, and divalent cofactor.
partial activity in the presence of Mg2+. Again, All RT enzymes possess a common stretch of seven
mutation of these residues are likely to perturb the amino acids termed the -YXDD- box. Amino acid
metal-binding site (Table 2 and Figure 4). Two other substitution of X results in changes in divalent metal
point mutants (Y586F and D524N) showed no activ- ion preference and enzyme activity, and so the
ity for Mg2+ activation, and only minimal levels (5%) -YXDD- box appears to play a role at the active center
for Mn2+ in the Y586F mutant. In these cases, the of RTs by binding the metal cofactor.147
metal-binding capability has clearly been more seri- The metal cofactor carries out functions in addition
ously damaged, suggestive of direct coordination by to catalysis. Mg2+ stabilizes HIV-1 RT binding to
Asp524 and Tyr586. Variations in DNA polymerase DNA-DNA substrates, with a 20-60-fold decrease
and processivity were also observed with the use of in koff, and an increase in productive interaction with
Mn2+ in these RNase H-deficient mutants, consistent DNA.148-150 Other charged species may enhance or
with the mutagenic effects of this cation described diminish the influence of magnesium. For example,
earlier in section IV. Ca2+ competitively inhibits Mg2+-promoted func-
The stimulatory effect of Mg2+ compared to Mn2+ tions.151 Spermine enhances, but cannot replace
has been examined for HIV-RT.142 Wild-type enzyme Mg2+ in reactions catalyzed by avian myeloblastosis
shows normal functional properties for both cations, virus reverse transcriptase,152 in a manner consistent
although RNase H* activity (hydrolysis of double- with the function of polyamines as described in
strand RNA) was observed with Mn2+. An E478Q section V.
mutant in the p66 domain, which participates in Finally, a reverse transcriptase motif has been
metal binding, was found to be inactive with Mg2+, identified in the catalytic subunit of telomerase, the
although Mn2+ restored both its endoribonuclease ribonucleoprotein required for replication of chromo-
1084 Chemical Reviews, 1998, Vol. 98, No. 3 Cowan
some ends, isolated from Euplotes aediculatus and contacts to Asp26 and Asp24 on each monomer. At
yeast (S. cerevisiae).166 Many metal binding residues higher metal concentrations the phosphatase is ob-
are conserved, and mutations in these motifs results served to bind up to 4.5 Mg2+ per monomer, however,
in loss of activity and telomere shortening. Interest- it is not clear that any site other than the “tight” site
ingly the -YXDD- motif described earlier for RT is physiologically relevant. The structure of Saccha-
enzymes, appears as -LXDD- in telomerase. romyces cerevisiae yeast inorganic pyrophosphatase
has also been obtained at 2.4 Å resolution (Figure
IX. Phosphatases 19) using Mn2+ as an analogue for Mg2+.154 This
Crystallographic studies of E. coli inorganic pyro- structure showed four bound Mn2+ ions and two
phosphatase, a hexameric soluble enzyme, provide bound phosphates that putatively are derived from
evidence for 1.5 Mg2+ cofactors per monomer (Figure pyrophosphate. The same group had previously
19).153 One ion resides in a “tight” binding site and structurally characterized the E. coli enzyme and
is coordinated by three aspartates (Asp65, Asp70, and identified one active-site Mn2+ ion,164 as found earlier
Asp102). The other Mg2+ is shared by two mono- by Kankare et al.153 It remains uncertain whether
mers, and appears to stabilize the interfacial contact these two enzymes do in fact follow distinct mecha-
region, and is coordinated by outer-sphere H-bond nistic pathways, or at least differ in their use of metal
Figure 19. (Top) The catalytic (left) and interfacial (right) Mg2+ sites of inorganic pyrophosphatase. Adapted from ref
153. (Bottom) Schematic illustration of the proposed transition state for the pyrophosphatase activity of the Saccharomyces
cerevisiae enzyme. The four Mn2+ ions and two phosphates, identified crystallographically, are indicated. Adapted from
ref 154.
Enzymes in Nucleic Acid Biochemistry Chemical Reviews, 1998, Vol. 98, No. 3 1085
cofactors, or if the additional metal cofactors for the phering the chemistry of this essential metal cofactor.
yeast enzyme arise from the use of Mn2+ ion in the This fact contributes to the design of magnesium
structural study. In passing, it can be noted that a binding pockets in proteins and enzymes. The im-
structural role is played by the Mg2+ ion cofactor for portance of reading the relevance of metal binding
the dinuclear zinc E. coli enzyme, alkaline phos- chemistry within the context of natural abundance
phatase.155 has been repeatedly emphasized.
(39) Vipond, I. B.; Baldwin, G. S.; Halford, S. E. Biochemistry 1995, (85) Cheng, X.; Balendiran, K.; Schildkraut, I.; Anderson, J. E. EMBO
34, 697-704. J. 1994, 13, 3927-3935.
(40) Hale, S. P.; Poole, L. B.; Gerlt, J. A. Biochemistry 1993, 32, (86) Aggarwal, A. K. Curr. Opin. Struct. Biol. 1995, 5, 11-19.
7479-7487. (87) Gromova, E. S.; Vinogradova, M. N.; Uporova, T. M.; Gryaznova,
(41) Jeltsch, A.; Alves, J.; Maass, G.; Pingoud, A. FEBS Lett. 1992, O. I.; Isagulyants, M. G.; Kosykh, V. G.; Nikol’skaya, I. I.;
304, 4-8. Shabarova, Z. A. Bioorg. Khim. 1987, 13, 269-72.
(42) Baldwin, G. S.; Vipond, I. B.; Halford, S. E. Biochemistry 1995, (88) Weiss, B. In The Enzymes, 3rd ed.; Academic Press: New York,
34, 705-714. 1981; Vol. 14, pp 203-231.
(43) Demple, B.; Johnson, A.; Fung, D. Proc. Natl. Acad. Sci. U.S.A. (89) Mol, C. D.; Kuo, C.-F.; Thayer, M. M.; Cunningham, R. P.;
1986, 83, 7731-7735. Tainer, J. A. Nature 1995, 374, 381-386.
(44) Huang, H.-W.; Cowan, J. A. Eur. J. Biochem. 1994, 219, 253- (90) Kow, Y. W. Biochemistry 1989, 28, 3280-3287.
260. (91) Katayanagi, K.; Miyagawa, M.; Matsushima, M.; Isikawa, M.;
(45) Shimamura, S.; Hibasami, H.; Kano, U.; Watanabe, S.; Suzuki, Kanaya, S.; Ikehara, M.; Matsuzaki, T.; Morikawa, K. Nature
S.; Nakashima, K. Int. J. Biochem. 1990, 22, 545-549. 1990, 347, 306-309.
(46) Bachrach, U.; Abu-Elheiga, L. Eur. J. Biochem. 1990, 191, 633- (92) Kashiwagi, T.; Jeanteur, D.; Haruki, M.; Katayanagi, K.;
637. Kanaya, S.; Morikawa, K. Protein Eng. 1996, 9, 857-867.
(47) Casareno, R. L. B.; Cowan, J. A. J. Chem. Soc., Chem. Commun. (93) Katayanagi, K.; Miyagawa, M.; Matsushima, M.; Ishikawa, M.;
1996, 1813-1814. Kanaya, S.; Nakamura, H.; Ikehara, M.; Matsuzaki, T.; Morika-
(48) Beese, L. S.; Steitz, T. A. EMBO J. 1991, 10, 25-33. wa, K. J. Mol. Biol. 1992, 223, 1029-1052.
(49) Bujacz, G.; Jaskólski, M.; Alexandratos, J.; Wlodawer, A.; (94) Katayanagi, K.; Ishikawa, M.; Okumura, M.; Ariyoshi, M.;
Merkel, G.; Katz, R. A.; Skalka, A. M. Structure 1996, 4, 89- Kanaya, S.; Kawano, Y.; Suzuki, M.; Tanaka, I.; Morikawa, K.
96. J. Biol. Chem. 1993, 268, 22092-22099.
(50) Kostrewa, D.; Winkler, F. K. Biochemistry 1995, 34, 683-696. (95) Needham, J. V.; Chen, T. Y.; Falke, J. J. Biochemistry 1993, 32,
(51) Bruice, T. C.; Tsubouchi, A.; Dempcy, R. O.; Olson, L. P. J. Am. 3363-3367.
Chem. Soc. 1996, 118, 9867-9875. (96) Yang, W.; Hendrickson, W. A.; Crouch, R. J.; Satow, Y. Science
(52) Deal, K. A.; Hengge, A. C.; Burstyn, J. N. J. Am. Chem. Soc. 1990, 249, 1398-1406.
1996, 118, 1713-1718. (97) Davies, J. F.; Hostomska, Z.; Hostomsky, Z.; Jordan, S. R.;
(53) Jenkins, L. A.; Bashkin, J. K.; Autrey, M. E. J. Am. Chem. Soc. Matthews, D. A. Science 1991, 252, 88-95.
1996, 118, 6882-6825. (98) Oda, Y.; Yamazaki, T.; Nagayama, K.; Kanaya, S.; Kuroda, Y.;
(54) Kanyo, Z. F.; Scolnick, L. R.; Ash, D. E.; Christianson, D. W. Nakamura, H. Biochemistry 1994, 33, 5275-5284.
Nature 1996, 383, 554. (99) Oda, Y.; Nakamura, H.; Kanaya, S.; Ikehara, M. J. Biomol. NMR
(55) Jenkins, J.; Janin, J.; Rey, F.; Chiadmi, M.; van Tilbeurgh, H.; 1991, 1, 247-255.
Lasters, I.; De Maeyer, M.; Van Belle, D.; Wodak, S. J.; (100) Beese, L. S.; Friedman, J. M.; Steitz, T. A. Biochemistry 1993,
Lauwereys, M.; Stanssens, P.; Mrabet, N. T.; Snauwaert, J.; 32, 14095-10141.
Matthyssens, G.; Lambeir, A.-M. Biochemistry 1992, 31, 5449- (101) Keck, J. L.; Marqusee, S. J. Biol. Chem. 1996, 271, 19883-19887.
5458. (102) Mueser, T. C.; Nossal, N. G.; Hyde, C. C. Cell 1996, 85, 1101-
(56) Almassey, R. J.; Janson, C. A.; Hamlin, R.; Xuong, X. H.; 1112.
Eisenberg, D. Nature 1986, 323, 304. (103) Haruki, M.; Noguchi, E.; Kanaya, S.; Crouch, R. J. J. Biol. Chem.
(57) Vipond, I. B.; Halford, S. E. Biochemistry 1995, 34, 1113-1119. 1997, 272, 22015-22022.
(58) Aaqvist, J.; Warshel, A. J. Am. Chem. Soc. 1990, 112, 2860- (104) Hibino, Y.; Iwakami, N.; Sugano, N. Biochim. Biophys. Acta
2868. 1991, 1088, 305-307.
(59) Pyle, A. M. Science 1993, 261, 7089. (105) Nikonova, L. V.; Beletsky, I. P.; Umansky, S. R. Eur. J. Biochem.
(60) Steitz, T. A.; Steitz, J. A. Proc. Natl. Acad. Sci. U.S.A. 1993, 90, 1993, 215, 893-901.
6498-6502. (106) Giannakis, C.; Forbes, I. J.; Zalewski, P. D. Biochem. Biophys.
Res. Commun. 1991, 181, 915-20.
(61) Ollis, D. L.; Brick, P.; Hamlin, R.; Xuong, N. G.; Steitz, T. A.
(107) Chang, M. P.; Baldwin, R. L.; Bruce, C.; Wisnieski, B. J. Science
Nature (London) 1985, 313, 762-766.
(Washington, D.C.) 1989, 246, 1165-1168.
(62) Black, C. B.; Cowan, J. A. Manuscript submitted.
(108) Beletskii, I. P.; Matyasova, J.; Nikonova, L. V.; Skalka, M.;
(63) Ceska, T. A.; Sayers, J. R.; Stier, G.; Suck, D. Nature 1996, 382,
Umanskii, S. R. Gen. Physiol. Biophys. 1989, 8, 381-398.
90-93.
(109) Nikonova, L. V.; Zotova, R. N.; Umanskii, S. R. Biokhimiya
(64) Fothergill, M.; Goodman, M. F.; Petruska, J.; Warshel, A. J. Am. (Moscow) 1989, 54, 17709-17718.
Chem. Soc. 1995, 117, 11619-11627. (110) Khodarev, N. N.; Sokolova, I. A.; Aleksandrova, S. S.; Volgina,
(65) Christianson, D. W.; Lipscomb, W. N. Acc. Chem. Res. 1989, 22, V. V.; Votrin, I. I. Byul. Eksp. Biol. Med. 1987, 103, 312-314.
62-69. (111) Caron-Leslie, L. A. M.; Schwartzman, R. A.; Gaido, M. L.;
(66) Black, C. B.; Cowan, J. A. J. Am. Chem. Soc. 1994, 116, 1174- Compton, M. M.; Cidlowski, J. A. J. Steroid Biochem. Mol. Biol.
1178. 1991, 40, 661-671.
(67) Shriver, D. F.; Atkins, P. W.; Langford, C. H.; Inorganic (112) Montague, J. W.; Hughes, F. M.; Cidlowski, J. A. J. Biol. Chem.
Chemistry; 2nd ed.; Freeman: New York, 1994. 1997, 272, 6677-6684.
(68) Rawlings, J.; Speckhard, D. C.; Cleland, W. W. Biochemistry (113) Cal, S.; Nicieza, R. G.; Connolly, B. A.; Sanchez, J. Biochemistry
1993, 32, 11204-11210. 1996, 35, 10828-10836.
(69) Cornelius, R. D.; Cleland, W. W. Biochemistry 1978, 17, 3279. (114) Kowalski, D. Nucleic Acids Res. 1984, 12, 7071-7086.
(70) Kim, S.; Cowan, J. A. Inorg. Chem. 1992, 31 3495-3496. (115) Ito, K.; Hara, C.; Matsuura, Y.; Minamiura, N. Arch. Biochem.
(71) Hampel, A.; Cowan, J. A. Chem. Biol. 1997, 4, 513-517. Biophys. 1995, 317, 25-32.
(72) Casareno, R.; Li, D.; Cowan, J. A. J. Am. Chem. Soc. 1995, 117, (116) Ito, K.; Matsuura, Y.; Minimiura, N. Arch. Biochem. Biophys.
11011-11012. 1994, 309, 160-167.
(73) Katayanagi, K.; Okumura, M.; Morikawa, K. Proteins: Struct., (117) Yarnall, M.; Rowe, T. C.; Holloman, W. K. J. Biol. Chem. 1984,
Funct. Genet. 1993, 17, 337-346. 259, 3026-3032.
(74) Heydenreich, A.; Koellner, G.; Choe, H.-W.; Cordes, F.; Kisker, (118) Price, P. A. J. Biol. Chem. 1972, 274, 2895-2899.
C.; Schindelin, H.; Adamiak, R.; Hahn, U.; Saenger, W. Eur. J. (119) Craig, M. L.; Suh, W.-C.; Record, M. T., Jr. Biochemistry 1995,
Biochem. 1993, 218, 1005-1012. 34, 15624-15632.
(75) Volbeda, A.; Lahm, A.; Sakiyama, F.; Suck, D. EMBO J. 1991, (120) Suh, W. C.; Leirmo, S.; Record, M. T., Jr. Biochemistry 1992,
10, 1607-1618. 31, 7815-1725.
(76) Pingoud, A.; Jeltsch, A. Eur. J. Biochem. 1997, 246, 1-22. (121) Suh, W. C.; Ross, W.; Record, M. T., Jr. Science (Washington,
(77) Jeltsch, A.; Maschke, H.; Selent, U.; Wenz, C.; Kohler, E.; D.C.) 1993, 259, 358-361.
Connolly, B. A.; Thorogood, H.; Pingoud, A. Biochemistry 1995, (122) Kuwabara, M. D.; Sigman, D. S. Biochemistry 1987, 26, 7234-
34, 6239-46. 7238.
(78) Thielking, V.; Selent, U.; Kohler, E.; Landgraf, A.; Wolfes, H.; (123) Griep, M. A. Indian J. Biochem. Biophys. 1995, 32, 171-178.
Alves, J.; Pingoud, A. Biochemistry 1992, 31, 3727-3732. (124) Sousa, R.; Chung, Y. J.; Rose, J. P.; Wang, B.-C. Nature 1993,
(79) Piccirilli, J. A.; Vyle, J. S.; Caruthers, M. H.; Cech, T. R. Nature 364, 593-599.
1993, 361, 85. (125) Woody, A.-Y. M.; Eaton, S. S.; Osumi-Davis, P. A.; Woody, R.
(80) Moon, B. J.; Vipond, I. B.; Halford, S. E. J. Biochem. Mol. Biol. W. Biochemistry 1996, 35, 144-152.
1996, 29, 17-21. (126) Joyce, C. M.; Grindley, N. D. F. Proc. Natl. Acad. Sci. U.S.A.
(81) Vipond, I. B.; Moon, B.-J.; Halford, S. E. Biochemistry 1996, 35, 1983, 80, 1830-1834.
1712-1721. (127) Mullen, G. P.; Serpersu, E. H.; Ferrin, L. J.; Loeb, L. A.; Mildvan,
(82) Rosenberg, J. M. Curr. Opin. Struct. Biol. 1991, 1, 104-113. A. S. J. Biol. Chem. 1990, 265, 14327-14334.
(83) Alves, J.; Urbanke, C.; Fliess, A.; Maass, G.; Pingoud, A. (128) Han, H.; Rifkind, J. M.; Mildvan, A. S. Biochemistry 1991, 30,
Biochemistry 1989, 28, 7879-7888. 11104-11108.
(84) Newman, M.; Strzelecka, T.; Dorner, L. F.; Schildkraut, I.; (129) Blasco, M. A.; Lazaro, J. M.; Bernad, A.; Blanco, L.; Salas, M. J.
Aggarwal, A. K. Science 1995, 269, 656-663. Biol. Chem. 1992, 267, 19427-19434.
Enzymes in Nucleic Acid Biochemistry Chemical Reviews, 1998, Vol. 98, No. 3 1087
(130) Blasco, M. A.; Lazaro, J. M.; Blanco, L.; Salas, M. J. Biol. Chem. (151) Tan, C.-K.; Zhang, J.; Li, Z.-Y.; Tarpley, W. G.; Downey, K. M.;
1993, 268, 24106-24113. So, A. G. Biochemistry 1991, 30, 2651-2655.
(131) Pelletier, H.; Sawaya, M. R.; Kumar, A.; Wilson, S. H.; Kraut, (152) Aoyama, H. Biochem. Int. 1989, 19, 67-76.
J. Science 1994, 264, 1891-1903. (153) Kankare, J.; Salminen, T.; Lahti, R.; Cooperman, B. S.; Baykov,
(132) Sawaya, M. R.; Pelletier, H.; Kumar, A.; Wilson, S. H.; Kraut, A. A.; Goldman, A. Biochemistry 1996, 35, 4670-4677.
J. Science 1994, 264, 1930-1935. (154) Harutyunyan, E. H.; Kuranova, I. P.; Vainshtein, B. K.; Hohne,
(133) Lowenstein, J. M.; Schatz, M. N. J. Biol. Chem. 1961, 236, 305- W. E.; Lamzin, V. S.; Dauter, Z.; Teplyakov, A. V.; Wilson, K. S.
307. Eur. J. Biochem. 1996, 239, 220-228.
(134) Lowenstein, J. M. Biochem. J. 1958, 70, 222-230. (155) Janeway, C. M. L.; Xu, X.; Murphy, J. E.; Chaidaroglou, A.;
(135) Miller, D. L.; Ukena, T. J. Am. Chem. Soc. 1969, 91, 3050-3053. Kantrowitz, E. R. Biochemistry 1993, 32, 1601-1609.
(136) Cowan, J. A. Inorg. Chem. 1991, 30, 2740-2747. (156) Katz, R. A.; Skalka, A. M. Annu. Rev. Biochem. 1994, 133-173.
(137) Pelletier, H. Science 1994, 266, 2025-2026. (157) Katz, R. A.; Skalka, A. M. Annu. Rev. Genet. 1992, 26, 527-
(138) Steitz, T. A.; Smerdon, S. J.; Jager, J.; Joyce, C. M. Science 1994, 544.
266, 2022-2025. (158) Rice, P.; Mizuuchi, K. Cell 1995, 82, 209-220.
(139) Van de Sande, J. H.; Loewen, P. C.; Khorana, H. G. J. Biol. (159) Ariyoshi, M.; Vassylyev, D. G.; Iwasaki, H.; Nakamura, H.;
Chem. 1972, 247, 6140. Shinagawa, H.; Morikawa, K. Cell 1994, 78, 1063-1072.
(140) Davies, J. F.; Almassay, R. J.; Hostomska, Z.; Ferre, R. A.; (160) Asante-Appiah, E.; Skalka, A. M. J. Biol. Chem. 1997, 272,
Hostomsky, Z. Cell 1994, 76, 1123-1133. 16196-16205.
(141) Blain, S. W.; Goff, S. P. J. Biol. Chem. 1996, 271, 1448-1454. (161) Quigley, G. J.; Teeter, M. M.; Rich, A. Proc. Natl. Acad. Sci.
(142) Cirino, N. M.; Cameron, C. E.; Smith, J. S.; Rausch, J. W.; Roth, U.S.A. 1978, 75, 64.
M. J.; Benkovic, S. J.; Le Grice, S. F. J. Biochemistry 1995, 34, (162) Gessner, R. V.; Quigley, G. J.; Wang, A. H.-J.; van der Marel,
9936-43. G. A.; van Boom, J. H.; Rich, A. Biochemistry 1985, 24, 237-
(143) Sarafianos, S. G.; Pandey, V. N.; Kaushik, N.; Modak, M. J. 240.
Biochemistry 1995, 34, 7207-7216. (163) Weber, D. J.; Libson, A. M.; Gittis, A. G.; Lebowitz, M. S.;
(144) Kew, Y.; Song, Q.; Prasad, V. R. J. Biol. Chem. 1994, 269, Mildvan, A. S. Biochemistry 1994, 33, 8017-8028.
15331-15336. (164) Harutyunyan, E. H.; Oganessyan, V. Y.; Oganessyan, N. N.;
(145) Zhan, X.; Tan, C.-K.; Scott, W. A.; Mian, A. M.; Downey, K. M.; Terzyan, S. S.; Popov, A. N.; Rubinskiy, S. V.; Vainshtein, B.
So, A. G. Biochemistry 1994, 33, 1366-1372. K.; Nazarova, T. I.; Kurilova, S. I.; Vorobyova, N. N.; Avaeva, S.
(146) Tan, C.-K.; Cival, R.; Mian, A. M.; So, A. G.; Downey, K. M. M. Crystallographia Rep. 1996, 41, 84-96 [English translation
Biochemistry 1991, 30, 4831-4834. from Krystallografiya (in Russian)].
(147) Kanda, T.; Saigo, K. Biochim. Biophys. Acta 1993, 1163, 223- (165) Burgess, J. Metal Ions in Solution; Halsted Press: New York,
226. 1993.
(148) De Stefano, J. J.; Bambara, R. A.; Fay, P. J. Biochemistry 1993, (166) Lingner, J.; Hughes, T. R.; Shevchenko, A.; Mann, M.; Lundblad,
32, 6908-6915. V.; Cech, T. R. Science 1997, 276, 561-567.
(149) Kruhoffer, M.; Urbanke, C.; Grosse, F. Nucleic Acids Res. 1993, (167) Cotton, F. A.; Hazen, E. E.; Legg, M. J. Proc. Natl. Acad. Sci.
21, 3943-3949. U.S.A. 1979, 76, 2551-2555.
(150) Hsieh, J. C.; Zinnen, S.; Modrich, P. J. Biol. Chem. 1993, 268,
24607-24613. CR960436Q