Download Moving one DNA double helix through another by a type II DNA

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Zinc finger nuclease wikipedia , lookup

Deoxyribozyme wikipedia , lookup

Transcript
Quarterly Reviews of Biophysics ,  (), pp. –.
Printed in the United Kingdom

#  Cambridge University Press
Moving one DNA double helix through
another by a type II DNA topoisomerase : the
story of a simple molecular machine
J A M E S C. W A N G
Department of Molecular and Cellular Biology, Harvard University, Cambridge,
Massachusetts €‚ƒˆ, USA
.             
. The double helix structure of DNA and its topological ramifications 
. Entering the DNA topoisomerases 
. Type II DNA topoisomerases 
.                                   
. Transporting one DNA double helix through another 
. The clamp model : a type II DNA topoisomerase as an ATP-modulated
protein clamp 
. The number of gates in the protein clamp 
. Three-dimensional structures of type II DNA topoisomerase fragments

. How are the various subfragments connected ? 
. A molecular model of DNA transport by a type II DNA topoisomerase

.                                              
      
. ATP utilization 
. How tight is the coupling between DNA transport and ATP binding and
hydrolysis ? 
.. Dependence of the coupling efficiency on the degree of supercoiling of
the DNA substrate 
.. Coupling efficiency and DNA binding 
. DNA relaxation by a type II DNA topoisomerase : how is the high
efficiency of coupling achieved ? 
.. ATP hydrolysis and the closure of the enzyme clamp 
.. Coupling of ATPase activity to DNA binding 
.. Structural changes in the enzyme upon closure of the N-gate 
.. ATP binding}hydrolysis and DNA cleavage 
.. Entrapment of the T-segment and opening of the DNA gate 
.. Closure of the DNA gate and exit of the T-segment 

J. C. Wang
. Directionality of DNA transport : why is bacterial gyrase unique ? 
.. Structural basis of the ability of bacterial DNA gyrase to catalyse
the ATP-dependent negative supercoiling of DNA 
.. The ATP-independent DNA relaxation activity of gyrase 
.                 
.                
.          



.            
.. The double helix structure of DNA and its topological ramifications
The discovery of the double helix structure of DNA led immediately to questions
on the mechanics of unravelling its intertwined strands during replication. If a
parental DNA is to be duplicated into two progeny molecules by separating its two
strands and copying each, then the strands must untwine rapidly during
replication (Watson & Crick, ).
That DNA indeed replicates in such a semiconservative fashion was soon
demonstrated by the Meselson–Stahl experiment (). At first, it appeared that
the unravelling of the intertwined strands should not pose an insurmountable
mechanical problem. The two strands at one end of a linear DNA, for example,
can be pulled apart with concomitant rotation of the double-stranded portion of
the molecule around its helical axis. If the strands of a DNA double helix are to
separate at an estimated replication rate of   base pairs (bp) per minute, then
the speed of this rotation would be   revolutions per minute from the  bp
per turn helical geometry of the double helix. This speed, though impressive,
seemed reasonable : owing to the slender rod-like shape of the double helix, the
estimated viscous drag for this rotational motion is actually rather modest
(Meselson, ).
Two findings in the s, however, heightened interests in the DNA uncoiling
problem. First, evidence began to accumulate that the lengths of DNA molecules
from natural sources are much longer than previously realized. Once it was found
that mechanical shear during DNA preparation can cause extensive breakage of
DNA, new records of longer and longer DNA molecules were set. It soon became
clear that an entire chromosome containing tens of millions of base pairs is
consisted of a single DNA molecule, with a length of several millimetres along its
contour (Kavenoff et al. ). For such a long DNA, the simple idea of
unravelling its intertwined strands by rotating the entire thread-like molecule
around its helical axis is untenable. Second, the s also saw the discovery of
various ring-shaped or ‘ circular ’ DNA molecules. Of particular importance was
the finding that the small DNA of polyoma virus is a double-stranded ring with
intact strands (Dulbecco & Vogt,  ; Weil & Vinograd, ). It was shown that
for such a ‘ covalently closed ’ DNA, the two complementary strands are
topologically linked, as predicted by the double-helix structure : a pair of
intertwined single-stranded DNA rings in a double-stranded DNA ring can not

A simple molecular machine
(a)
(b)
A
B
C
Fig. . Schematic drawings illustrating two topological problems of DNA during
semiconservative replication. In (a), the general problem of separating the two parental
strands is illustrated for a double-stranded DNA ring. In this two-line representation of the
duplex DNA ring, each of the circular DNA strands is represented by a closed line. Because
of the double helix geometry of DNA, the two strands are topologically linked ; the degree
of linkage between them must gradually reduce as DNA replication proceeds, and the two
strands must become completely unlinked at the end of a round of replication. For the two
closed lines drawn in the illustration, one way of unlinking them is to carry out multiple
cycles of a simple operation : in each cycle, one line is transiently broken, the other line is
then passed through the break once before the break is resealed. The linking number (Lk)
between the two closed strands is the minimal number of such cycles that is required to
unlink them. Inspection of the particular drawing shows that all crossovers are of the same
sign and each strand passage event removes two crossovers ; Lk is thus equal to one half of
the number of crossovers. For other ways of defining Lk, see Crick (), Cozzarelli et al.
(), Wang (). In (b), the conversion of twists between the strands of an unreplicated
DNA segment to intertwines between two fully replicated progeny DNA molecules is
illustrated. Two replication forks are shown to approach each other near the end of a round
of DNA replication (drawing A). If copying of the unreplicated portion of the parental
DNA segment is faster than unlinking the parental strands coiled around each other, two
intertwined progeny molecules are formed (C). Under such conditions, the replication
product of a circular DNA would be a pair of multiply linked rings. Illustration taken from
Varshavsky et al. ().
be separated without breaking at least one of the strands. A parameter defining the
degree of topological linkage between the two strands of a duplex DNA ring is
termed the linking number (Lk). As this quantity will pop up again in subsequent
discussions, an operational definition of Lk is illustrated in Fig.  a.
Whereas the problem of unravelling the complementary strands of a long linear
DNA during replication might appear to be kinematic in nature, the same
problem for a covalently closed DNA ring is surely a topological one. Actually,
because a linear chromosome inside a cell is organized in a compact form with

J. C. Wang
multiple loops, the problem of separating its strands is not all that different from
that for a covalently closed DNA ring. Therefore, it would appear that during
evolution, as DNA became very long or circular, a solution must be found for the
disentanglement of its topologically intertwined strands during replication.
The separation of two parental DNA strands during replication is the best
known topological manifestation of the double helix structure. It is not the only
one, however, and a variation of the theme is illustrated in Fig.  b. In the top
drawing (A), a DNA molecule near the end of its replication is shown, with a short
stretch of the molecule remaining unreplicated. If unravelling of the parental
strands in this unreplicated segment is incomplete before the strands are
completely unpaired and copied, then the residual intertwists between the
parental strands would be converted to intertwines between the newly replicated
progeny molecules (B and C) (Sundin & Varshavsky, ,  ; Varshavsky et
al. ). These interwoven DNA duplexes must be resolved if the progeny
molecules are to be segregated into two newly divided cells.
Historically, replication had inspired much of the earlier considerations on the
topological problems of DNA. It is now well-known that the necessity of
disentangling DNA strands or duplexes also arises in nearly all other cellular
transactions of DNA, including transcription, chromosome condensation and
decondensation, and recombination. These aspects have been reviewed elsewhere
(see Wang, , and references therein). As illustrated by the two examples
summarized above, many of the topological problems of DNA are deeply rooted
in its double-helix structure.
. Entering the DNA topoisomerases
Nature invented a family of enzymes to solve the topological problems of DNA.
These enzymes, termed the DNA topoisomerases (Wang & Liu, ), catalyse
the interpenetration of DNA strands or double helices. In their presence, DNA
strands and double helices can go through one another as if there were no physical
boundaries in between. Since the discovery of the first member of this family of
enzymes in the bacterium Escherichia coli nearly three decades ago (Wang, ),
all organisms have been found to possess several of these enzymes. The limited
scope of this review does not permit even a cursory coverage of the topoisomerase
literature ; references on earlier studies can be found in a number of monographs
(Cozzarelli & Wang,  ; Bates & Maxwell,  ; Liu,  a, b).
The DNA topoisomerases perform their magic through transient breakage of
DNA strands. They can be divided into two types : the type I enzymes break one
DNA strand at a time and the type II enzymes both strands of a double helix in
concert (Brown & Cozzarelli,  ; Liu et al. ). The type I enzymes can be
further divided into two subfamilies : the IA subfamily represented by bacterial
DNA topoisomerases I and III and eukaryotic DNA topoisomerase III, and the
type IB subfamily by eukaryotic DNA topoisomerase I and pox virus DNA
topoisomerases. Members of the same subfamily are closely related in their amino
acid sequences and reaction characteristics, and members of different subfamilies
A simple molecular machine

share little sequence homology and are mechanistically distinct (for reviews, see
Wang, , and references therein). The type II DNA topoisomerases are
thought to form a single subfamily, but recent studies of an enzyme DNA
topoisomerase VI from archaeal hyperthermophiles suggest that the type II
enzymes may also be divided into two subfamilies IIA and IIB (Bergerat et al.
) : subfamily IIB is represented by archaeal DNA topoisomerase VI, and
subfamily IIA by all other type II topoisomerases, including bacterial gyrase
(DNA topoisomerase II), bacterial DNA topoisomerase IV, yeast and Drosophila
DNA topoisomerase II, mammalian DNA topoisomerases IIα and IIβ, and Teven phage DNA topoisomerases. Because mechanistic analysis of archaeal DNA
topoisomerase VI is still in its infancy, this fascinating enzyme will not be
discussed in this review. It is plausible, however, that the archaeal enzyme may
share many of the mechanistic features to be described and discussed below.
All DNA topoisomerases catalyse transient DNA strand breakage by
transesterification, a mechanism first postulated in  (Wang, ). For the
type I enzymes, the phenolic oxygen of an active-site tyrosyl residue undergoes
transesterification with a phosphoryl group in a DNA strand, breaking the DNA
phosphodiester bond and forming a phosphotyrosine linkage (Tse et al.  ;
Champoux, ). Rejoining of the DNA strand occurs through an apparent
reversal of the DNA breakage reaction. Because the enzyme–DNA complex is
likely to undergo large conformation changes between the DNA breakage and
rejoining steps (Lima et al. ), the two steps are not necessarily the exact
microscopic reversal of each other (Stivers et al. ). For DNA topoisomerases
other than the type IB enzyme, the tyrosyl residue becomes linked to a DNA «phosphoryl group in the formation of the covalent intermediate (Fig. ) ; for the
type IB enzymes, it becomes linked to a «-phosphoryl group. The type I enzymes
act as monomers in their breakage and rejoining of DNA strands one at a time ; the
type II enzymes are dimeric, and a pair of active-site tyrosyl residues undergo
transesterification with a pair of phosphoryl groups in the two strands of a duplex
DNA (Morrison and Cozzarelli,  ; Sander & Hsieh,  ; Liu et al. ).
. Type II DNA topoisomerases
In this review, the focus is on the reaction mechanism of the type II DNA
topoisomerases. The type II enzymes are of special interest in a number of ways.
They unlink DNA catenanes and resolve intertwined chromosome pairs during
mitosis, and in their absence cells die (reviewed in Yanagida & Sternglanz, ).
These enzymes have been identified as the targets of a large number of natural
toxins, antimicrobial agents, and anti-tumour therapeutics (Liu,  b ; Maxwell,
). They are also DNA-dependent ATPases, and how they couple their
manipulation of DNA to ATP binding and hydrolysis raises fascinating
mechanistic questions.
The first members of the type II subfamily of DNA topoisomerases, bacterial
DNA gyrase (DNA topoisomerase II), was discovered by Gellert et al. () as

J. C. Wang
5′ DNA
O
O
P
O
CH2
DNA 3′
O
OH
5′ DNA
3′ OH
O
CH2
O
P
O
CH2
DNA 3′
O
CH2
Fig. . Formation of a covalent intermediate between a DNA and a DNA topoisomerase.
Nucleophilic attack of an enzyme tyrosyl group on a DNA backbone phosphorus leads to
the breakage of the DNA strand and the simultaneous formation of a phosphotyrosine bond
between the enzyme and the DNA. In the illustration shown, the phenolic oxygen of the
tyrosyl group is attacking from the opposite side of a «-oxygen, and the enzyme tyrosyl
becomes linked to a DNA «-phosphoryl group in the covalent intermediate. Such a
covalent intermediate has been identified in reactions catalyzed by the type IA and type II
DNA topoisomerases ; for the type IB DNA topoisomerases, the enzyme tyrosyl group
attacks from the opposite site of a «-oxygen and becomes linked to a DNA «-phosphoryl
group in the covalent intermediate.
an ATP-dependent DNA negative supercoiling activity. The term negative
supercoiling refers to reducing the linking number between the two strands of a
duplex DNA ring. If a DNA ring is in its most stable structure (termed a relaxed
DNA), its linking number Lk° at a temperature of – °C and in a dilute
aqueous buffer around neutral pH is readily estimated to be N}±, where N is
the size of the DNA ring in base pairs, and ± is the helical periodicity of a
typical DNA under these conditions (Wang,  ; Rhodes & Klug, ). If Lk
of a DNA ring deviates significantly from Lk°, torsional and flexural strains are
introduced into the molecule, and the DNA ring becomes contorted in space
much like the deformation of a torsionally unbalanced rope. It is this contorted
A simple molecular machine
(A)
409
1
(B)
679
ATP
(C)
1201
Y*

1428
(dimer)
Yeast DNA topoisomerase II
1
804 1
875
GyrB
39
GyrA
60
52
Phage T4
Fig. . A schematic alignment of the amino acid sequences of three representative type II
DNA topoisomerases. Yeast DNA topoisomerase II is composed of a single subunit, E. coli
DNA gyrase of two (GyrA and GyrB), and phage T DNA topoisomerase of three (encoded
by genes , , and ). A, B, and C along the yeast polypeptide denote three proteasesensitive sites. For the yeast enzyme, the region between sites A and B, corresponding to the
C-terminal half of the E. coli GyrB protein, is termed the B« subfragment ; the region
between B and C, corresponding to the N-terminal two thirds of the E. coli GyrA protein,
is termed the A« subfragment. Thick lines represent regions with significant homology and
thin lines regions with marginal homology. The approximate locations of the ATPase site,
the active-site tyrosine, and residues that form the primary dimer interface in the yeast
enzyme are indicated by ATP, Y*, and (dimer) above the line representing the enzyme.
Illustration drawn after fig.  in Berger & Wang ().
shape that inspired the term supercoiling, supertwisting, or superhelix formation.
By definition, a negatively supercoiled DNA is one with Lk ! Lk°, and a
positively supercoiled DNA one with Lk " Lk°. DNA gyrase can apparently
utilize the chemical energy of ATP hydrolysis to drive a DNA ring to an
energetically unfavourable state with an Lk ! Lk°.
The discovery of bacterial DNA gyrase was soon followed by the discovery of
phage T DNA topoisomerase (Liu et al.  ; Stetler et al. ) and eukaryotic
DNA topoisomerase II (Baldi et al.  ; Hsieh & Brutlag,  ; Miller et al.
). The phage and eukaryotic enzymes differ from bacterial DNA gyrase in
that they catalyse the relaxation of a positively or negatively supercoiled DNA in
the presence of ATP, but not DNA supercoiling. These enzymes also differ in
their quaternary structures. Bacterial gyrases are comprised of two subunits
(Gellert et al. ), the eukaryotic enzymes one (Miller et al.  ; Sander &
Hsieh,  ; Goto & Wang, ), and the phage T enzyme three (Liu et al.
 ; Stetler et al. ). The amino acid sequences of the enzymes showed
clearly, however, that the enzymes are closely related (see for example, the
compilation of Caron & Wang, ) ; a schematic alignment of the amino acid
sequences of three type II DNA topoisomerases of different quaternary structures
is illustrated in Fig. .

J. C. Wang
.                                  
. Transporting one DNA double helix through another
A breakthrough in understanding the mechanism of the type II DNA
topoisomerases came around . In earlier years, it was assumed that in
reactions catalysed by all topoisomerases only one of the two strands of a DNA
double helix would be transiently broken so that the intact strand could hold the
broken ends close to each other for their subsequent rejoining. Several lines of
evidence showed, however, that bacterial DNA gyrase and phage T DNA
topoisomerase do not act in this fashion. First, it was observed in  that the
addition of a protein denaturant to a plasmid-bound gyrase would linearize rather
than nick the plasmid in the presence of nalidixate, an antibiotic that targets
bacterial gyrase (Sugino et al.  ; Gellert et al. ). Furthermore, a protein
moiety was found to be covalently attached to each of the two « ends of the linear
DNA produced in this reaction (Morrison & Cozzarelli, ), and this protein
moiety was shown to be the A-subunit of gyrase (Tse et al.  ; Sugino et al.
). These findings demonstrated that the double-stranded breakage of DNA
by gyrase is a consequence of covalent adduct formation between the enzyme and
both strands of a duplex DNA. Second, it was found that the bacterial and phage
enzymes can interconvert several novel topological forms of duplex DNA rings,
for examples the unlinking of DNA catenanes or the knotting of unknotted rings
(Liu et al.  ; Kreuzer & Cozzarelli,  ; Mizuuchi et al. ). These
topological transformations strongly suggested that the bacterial and phage
enzymes can catalyse the passage of one DNA double helix through a transient
double-stranded break in another. Third, quantitative measurements of linking
number changes of DNA rings by DNA gyrase or phage T DNA topoisomerase
showed that these enzymes alter Lk in units of two (Brown & Cozzarelli,  ; Liu
et al.  ; Mizuuchi et al. ). Such an even-number change in Lk can best be
interpreted in terms of a mechanism in which an enzyme mediates the passage of
one DNA segment through another of the same DNA ring during each round of
reaction (Brown & Cozzarelli,  ; Liu et al.  ; Mizuuchi et al. ). It
turned out that the dyadic change in Lk by such an operation was foretold in a
mathematical analysis of the linking number between the two edges of a closed
ribbon (Fuller, ), and could also be inferred from earlier examples illustrating
the topological properties of a DNA ring (Crick, ). Studies of eukaryotic type
II DNA topoisomerases also led to the same conclusion that these enzymes act by
a double-stranded DNA breakage, passage, and rejoining mechanism (Baldi et al.
 ; Hsieh & Brutlag, ).
Once it was recognized that the type II enzymes transiently cleave doublestranded DNA, it was straightforward to determine the relative positions of the
pair of cleavage sites such an enzyme would make in a DNA duplex. From these
measurements, DNA breakage by a type II topoisomerase was shown to involve
transesterification between the active-site tyrosyl residues of the enzyme and a pair
of staggered phosphoryl groups four base pairs apart (Morrison & Cozzarelli,
A simple molecular machine

 ; Sander & Hsieh,  ; Liu et al. ). In the case of bacterial DNA gyrase,
the A-subunit is often referred to as the ‘ DNA breakage-and-rejoining subunit ’
because of the presence of an active-site tyrosyl group in each A-subunit (Tse et
al.  ; Sugino et al.  ; Horowitz & Wang, ). However, essential
residues from both A- and B-subunits are most likely present in each catalytic
pocket for transesterification. Whereas purified DNA gyrase A-subunit has no
detectable DNA cleavage}rejoining activity (Higgins et al. ), the combination
of the A-subunit and a C-terminal fragment of the B-subunit does (Brown et al.
 ; Gellert et al. ). Similarly, yeast DNA topoisomerse II lacking the
‘ GyrB ’ portion of the polypeptide has no DNA cleavage activity, but a fragment
comprised of amino acid residues  to about , corresponding to the Cterminal half of the gyrase B-subunit plus the N-terminal two thirds of the gyrase
A-subunit, is capable of forming enzyme–DNA covalent adduct (Berger et al.
). Site-directed mutagenesis studies of yeast DNA topoisomerase II further
confirmed the requirement of residues in both A« and B« domains of the enzyme
(Q. Liu and J. Wang, to be published ; see the legend to Fig.  for the boundaries
of the A« and B« domains).
. The clamp model : a type II DNA topoisomerase as an ATP-modulated
protein clamp
From the above description, a type II DNA topoisomerase presumably binds to
a double-stranded DNA segment, to be termed the gate-segment or G-segment,
with the pair of active-site tyrosyl residues of the enzyme positioned near a pair
of DNA phosphorus atoms for creating an opening or gate in the DNA double
helix. The DNA-bound enzyme must then capture a second double-stranded
DNA segment, termed the transported-segment or T-segment, in order to move
it through the G-segment. How does this occur ?
Studies carried out between the mid s and early s led to the proposal
of the clamp model, in which a type II DNA topoisomerase is assumed to act as
an ATP-modulated protein clamp (Roca & Wang, ). According to this model,
the enzyme possesses a pair of jaws, which come into contact to close a molecular
gate upon binding of ATP, and come apart to re-open the gate upon ATP
hydrolysis and product release. The closing and opening of the jaws can occur
either in a free enzyme, or in one bound to a G-segment. The ATP-modulated
cycling time is faster when the enzyme is DNA-bound, and the magnitude of this
increase in rate can be deduced from the DNA dependence of the rate of ATP
hydrolysis. In the case of yeast DNA topoisomerase II, the Michaelis constant KM
and the apparent turnover number kcat were measured to be ± m and ± s−" per
dimeric enzyme in the absence of DNA (at  °C in a medium containing  m
Tris acetate, pH ±,  m potassium acetate,  m magnesium acetate, and
 m -mercaptoethanol). In the presence of DNA, positive cooperativity
between the two ATPase sites was observed. The concentration of ATP at halfmaximal velocity drops to about ± m, and kcat is increased by about -fold
(Lindsley & Wang,  a). Thus at saturating ATP concentrations the motion of

J. C. Wang
G
G
+ ATP
T
Fig. . The protein clamp model of a type II DNA topoisomerase. An enzyme molecule is
depicted as a dimeric protein clamp bound to a DNA segment termed the gate or Gsegment. According to this model, closing and opening of the pair of jaws of the protein
clamp are modulated by ATP binding and hydrolysis and release of the hydrolytic products.
A second DNA segment, termed the transported or T-segment, can enter the DNA-bound
protein clamp ; ATP-mediated closure of the jaws captures the T-segment and drives it
through the transiently opened DNA gate in the G-segment. Drawing taken from Roca &
Wang ().
clamp opening and closing is probably accelerated by about -fold when the yeast
enzyme is bound to DNA. For E. coli DNA gyrase, binding to DNA was also
found to stimulate the ATPase activity of the enzyme (Sugino et al.  ; Sugino
& Cozzarelli,  ; Staudenbauer & Orr,  ; Maxwell & Gellert, ).
The ATP-modulated opening and closing of the enzyme clamp is assumed to
be directly related to the type II topoisomerase-mediated transport of one DNA
double helix through another. When a DNA-bound protein clamp is in the open
state, a T-segment can enter through the open jaws into the enzyme interior. This
entrance is presumably facilitated by weak DNA–protein interactions and is
also influenced by the topology of the DNA substrate. Closing of the jaws by
ATP-binding then captures the T-segment and drives it through the G-segment
(Fig. ).
The clamp model was first hinted by images of DNA gyrase molecules in
electron micrographs (Kirschhausen et al. ). Experimental evidence
supporting the model came initially from studies of DNA binding by yeast
DNA topoisomerase II in the absence and presence of «-adenylyl-β,γimidodiphosphate (ADPNP), a non-hydrolysable β,γ-imido analogue of ATP. In
the absence of the nucleotide, the yeast enzyme binds readily to linear or circular
DNA, with a slight preference for supercoiled DNA (by about a factor of ). Preincubation of the enzyme with ADPNP, however, converts it to a form capable of
binding linear DNA but not any form of DNA rings (Roca & Wang, ). The
specific binding of linear DNA to the ADPNP–enzyme complex is not caused by
interactions between the DNA ends and the enzyme, as the ends of an enzymebound linear DNA can be readily joined by DNA ligase. Thus it appeared that the
A simple molecular machine

binding of ADPNP to the enzyme converts it to an annular form with a hole
sufficiently large for a linear DNA to thread through (Roca & Wang, ).
Results of similar experiments with E. coli DNA gyrase were, however, not as
clear cut (A. Maxwell and M. Gellert, personal communications) ; it is plausible
that measurements with the bacterial enzyme are subject to complications arising
from the association}dissociation of the DNA gyrase subunits. In a later section,
mechanistic differences between DNA gyrase and other type II DNA
topoisomerases will be discussed.
The clamp model explains readily an earlier observation with Drosophila DNA
topoisomerase II that addition of ADPNP makes the Drosophila enzyme resistant
to dissociation by salt if it is bound to a DNA ring, but not if it is bound to a linear
DNA (Osheroff, ). For a type II topoisomerase bound to a DNA ring, closing
the protein clamp by ADPNP would introduce a salt-resistant topological link
between the protein and DNA (Pommier et al.  ; Roca & Wang, ). The
same interpretation explains why a linear DNA bound to the enzyme–ADPNP
complex is readily dissociated from the complex upon exposure to high salt, but
ligation of the ends of the linear DNA before salt addition prevents its separation
from the protein (Roca & Wang, ).
According to the clamp model, the jaws of a type II DNA topoisomerase bound
to a DNA G-segment must close after the entrance of the T-segment in order to
effect the transport of the T-segment through the G-segment. This notion was
supported by an order-of-mixing experiment in which two different DNA rings
and ADPNP were added sequentially to yeast DNA topoisomerase II. Under
conditions such that an enzyme-bound DNA could not be displaced by a
subsequently added second DNA, it was shown that the G-segment always resides
in the first DNA ring added, and catenation between the two sequentially added
DNA rings is possible only if the second ring is added before the addition of
ADPNP (Roca & Wang, ).
. The number of gates in the protein clamp
A protein clamp has at least one molecular gate which closes when the jaws of the
clamp come into contact and opens when they come apart. The question was
raised, however, about whether there are separate gates for the entrance and exit
of the T-segment, or there is only a single gate through which the T-segment
would first enter and later exit (Roca & Wang, ). Fig.  illustrates the two
situations. In each case, an ATP-modulated enzyme clamp bound to a G-segment
is shown, and a T-segment can enter the clamp in its open state as described in the
section above. In the two-gate model (a), closing the jaws denoted by N drives the
T-segment through the transiently opened G-segment and out of a second protein
gate on the opposite side of the entrance gate [right-side drawing in (a)]. In the
one-gate model (b), closing of the protein clamp drives the T-segment through the
G-segment into a holding dock ; following ATP-hydrolysis, the clamp re-opens
for the exit of the T-segment through the same gate it had entered earlier. As
illustrated in the right-side drawing in (b), in the one-gate model at least a part of

J. C. Wang
(a)
T
G
G
T
N
N
N
N
T
(b)
T
G
N
N
T
Fig. . The two-gate (a) and one-gate (b) model of DNA transport by a type II DNA
topoisomerase. See the text for explanation. Illustration taken from Roca & Wang ().
the G-segment must dissociate from the enzyme prior to the exit of the Tsegment, because backtracking of the T-segment through the G-segment would
mean no net transport.
In the two-gate model, exist of the T-segment requires transient disruption of
protein–protein interaction between the pair of jaws constituting the exit gate ; in
the one-gate model, transient disruption of DNA-protein interaction is involved
instead. The two-gate model was first proposed in  for DNA gyrase
(Mizuuchi et al.  ; Wang et al. ). Because gyrase interacts with about
 bp of the DNA G-segment (Liu & Wang,  a, b ; Klevan & Wang,  ;
Kirkegaard & Wang,  ; Morrison & Cozzarelli,  ; Fisher et al.  ; Rau
et al.  ; Orphanides & Maxwell, ), in the absence of experimental data it
appeared that separating two interacting protein domains would be no more
difficult than dissociating a long stretch of DNA from the protein surface.
The first attempt to distinguish the two models was made more recently
(Roca & Wang, ). A series of experiments were carried out to examine the
unlinking of two singly linked DNA rings by yeast DNA topoisomerase II upon

A simple molecular machine
G
ATP
T
2
N
N
3
T
T
1
N
G
N
N
G
N
N
N
T
4
ADP + Pi
5
G
N
N
Fig. . Reaction steps in the two-gate model of ATP-dependent DNA transport by a type II
DNA topoisomerase. An enzyme molecule in its open-clamp conformation can bind a Gsegment, and the jaws labelled N in the drawings can be in the open or closed state
depending on the absence or presence of enzyme-bound ATP. A T-segment can enter the
enzyme when it is in the open-clamp conformation. The closure of the enzyme clamp upon
ATP-binding traps the T-segment and forces it to first go through the DNA gate and then
an exit gate on the opposite side of the entrance gate (the N-gate) formed by the pair of jaws
labelled N.
the addition of ADPNP. Because the non-hydrolysable nucleotide can trigger the
closure of the N-gate but does not allow its re-opening, the one-gate model would
predict that both of the unlinked DNA rings would be topologically linked with
the annular enzyme. The two-gate model, on the other hand, would predict that
only the DNA ring containing the G-segment would be bound in such fashion, as
the ring containing the T-segment would be expelled from the enzyme through
the exit gate. The results of the experiments are in agreement with the two-gate
model : decatenation of the singly linked rings was found to occur readily upon the
addition of ADPNP, and analysis of the products showed that the ring in which
the T-segment resided was free in solution and the ring in which the G-segment
resided remained topologically linked to the annular enzyme (Roca & Wang,
). Incorporation of this finding into the ATP-modulated protein clamp model
of type II DNA topoisomerases led to the model depicted in Fig. .
. Three-dimensional structures of type II DNA topoisomerase fragments
The structures of several type II DNA topoisomerase fragments have been solved
by X-ray crystallography in recent years (Wigley et al.  ; Berger et al.  ;
Lewis et al.  ; Morais Cabral et al. ). Fig.  depicts a ± AI structure of

J. C. Wang
a -kDa N-terminal fragment of the B-subunit of E. coli DNA gyrase (Wigley et
al. ). This fragment, comprised of amino acid residues –, contains the
ATPase domain of the enzyme. The crystal was obtained in the presence of
ADPNP, and there is one bond ADPNP molecule in each half of the dimeric
protein molecule ; therefore the structure is likely to resemble the ATP-bound
form of the protein fragment. Two lysyl side-chains lining the nucleotide-binding
pocket, Lys- and Lys-, were previously identified to form covalent links
with an ATP affinity analogue pyridoxol «-diphospho-«-adenosine (Tamura &
Gellert, ).
Several studies showed that the -kDa E. coli GyrB fragment is monomeric in
solution but dimerizes upon binding of ADPNP, but not ADP (Ali et al. ,
). Measurements of ATP hydrolysis by E. coli DNA gyrase (Maxwell et al.
) and yeast DNA topoisomerase II (Lindsley & Wang,  a), as well as
measurements of ADPNP-binding by DNA gyrase (Tamura et al. ), also
suggest a cooperative interaction between the two nucleotide binding sites. Taken
together, these results indicate that the binding of an ATP to one of the ATPase
catalytic pockets of a type II DNA topoisomerase triggers a conformation change
of the ATPase domain, which in turn leads to intramolecular dimerization of the
domains and facilitates the binding of a second ATP. An important conclusion
that can be drawn from the crystallographic and biochemical experiments is that
ATP hydrolysis can not occur without dimerization of the ATPase domains of the
enzyme (Wigley et al.  ; Tamura et al.  ; Ali et al. ). This point will
be elaborated upon in a later section on the coupling between ATP usage and
DNA transport by the type II DNA topoisomerases.
Because ADPNP binding is known to trigger the closure of the enzyme clamp
in the case of yeast DNA topoisomerase II, from the above discussion it can be
deduced that the pair of ATPase domains in a type II DNA topoisomerase
constitute the entrance gate ; this gate has been termed the N-gate (Figs  and )
because in a single-polypeptide eukaryotic enzyme the ATPase domain is close to
the N-terminus.
In the structure shown in Fig.  (Wigley et al. ), an N-terminal arm (amino
acid residues –) of the DNA gyrase B-subunit extends from the ATPase
domain surface to form dimer contacts with the other monomer. The C-terminus
proximal helix (residues –) also extends from the core of the C-terminal
subdomain of the structure to contact the symmetry-related helix of the other
monomer. This C-terminal interface is likely to represent a crystallographic rather
than native protein contact, however.
Fig.  depicts a ± AI crystal structure of a -kDa fragment of yeast DNA
topoisomerase II spanning residues – (Berger et al. ). This fragment
lacks the ATPase subfragment as well as  amino acid residues at the Cterminus of the intact enzyme, but it is capable of covalent adduct formation with
double-stranded DNA. From the alignment shown in Fig. , the –
fragment can be subdivided into two parts, termed the B« subfragment (residues
 to about ), which corresponds to the C-terminal half of the gyrase Bsubunit, and the A« subfragment, which corresponds to the gyrase A-subunit
A simple molecular machine

lacking the -residue C-terminal fragment. In the cases of E. coli DNA gyrase
and yeast and Drosophila DNA topoisomerase II, this missing C-terminal
segment is known to be dispensable for catalysis of DNA transport (Reece &
Maxwell,  a ; Shiozaki & Yanagida,  ; Crenshaw & Hsieh,  ; Caron et
al.  ; Kampranis & Maxwell, ). The dispensability of the C-terminal
segment is also consistent with its absence in phage T and T DNA
topoisomerase (see Fig. ). Therefore, in combination, the structures of the gyrase
ATPase subfragment and the yeast DNA topoisomerase II B«A« fragment provide
a D sketch of a functional type II DNA topoisomerase.
In the heart-shaped (B«A«) structure of the yeast enzyme fragment, the two B«
#
subfragments contact each other to form an arch, which caps the V-shaped A«–A«
dimer to enclose a large hole  AI wide at its base. The two A« polypeptides
contact at their C-terminal region to form the A«–A« dimer interface. This
interface represents the major contact between the two B«A« halves and it buries
over  AI # of surface. Three of the four known yeast top‚ mutations that result
in cold-sensitivity of the mutant cells were mapped to this interface (Thomas et
al. ). The B«–B« contacts are less extensive but appear significant (Berger et
al. ).
A pair of semicircular grooves formed by residues from both the A« and B«
subfragments are present near the top of the V-shaped A«–A« dimer. These
grooves, with an approximately  AI opening, are decorated with positive surface
charges and have been implicated in the binding of the G-segment (Berger et al.
). A double-stranded DNA segment with a -nucleotide «-overhang can be
modelled into each groove in the protein structure, with the single-stranded
nucleotides extending through a narrow tunnel and joining to the active-site
tyrosyl residue Tyr- (Fig. ). Thus the -kDa structure is believed to
correspond closely to the conformation of the enzyme after it has cleaved the Gsegment and pulled it apart (Berger et al. ). In the proposed model, a part of
the DNA-binding surface is formed by a domain comprised of residues –.
This domain has the same fold as the DNA binding domain of the E. coli
catabolite activator protein (CAP) and histone H, and has been termed the CAP
domain (Berger et al. ). Recent lysine footprinting data are consistent with the
assignment of the G-segment binding surface described above (Li & Wang, ).
Fig.  illustrates a ± AI structure of a -kDa fragment (residues –) of E.
coli DNA GyrA protein (Morais Cabral et al. ). As expected from amino acid
sequence comparisons (Fig. ), the overall architecture of this fragment is very
similar to that of the A« subfragment of yeast DNA topoisomerase II (Fig. ).
There are significant differences between the relative positions of the various
domains within the E. coli and the yeast enzyme fragment, however. Whereas the
pair of CAP domains in the yeast enzyme are well-separated, they are rotated from
their positions in the yeast enzyme and brought into contact to form the ‘ head
dimer interface ’ in the gyrase structure shown in Fig.  (Morais Cabral et al.
). The three long helices (α, α, and α) connecting the N-proximal core
and the primary dimer interface also adopt different conformations in the yeast
and E. coli fragment structures, which will be discussed in a later section.

J. C. Wang
. How are the various subfragments connected ?
No high-resolution structure of a type II DNA topoisomerase capable of moving
one DNA through another is presently available. Electron microscopic
examination of E. coli gyrase and its subunits (Kirschhausen et al. ), human
DNA topoisomerase II (Schultz et al. ), and intact and fragments of yeast
DNA topoisomerase II (Benedetti et al. ) has provided low-resolution
information on the organization of the subdomains. Visualization of negatively
stained human enzyme by scanning transmission electron microscopy, for
example, showed a V-shaped tripartite structure with a dense globular domain,
about  AI in diameter, flanked by two smaller spheres about – AI in diameter
(Schultz et al. ). The pair of smaller spheres at the opening of the V-shaped
molecule appeared to be connected to the larger globular body by linkers – AI
in length that were barely visible in the micrographs, and the angle extended by
the smaller spheres was typically in the range of –°. A similar tripartite
structure was seen by transmission microscopy with metal-shadowed yeast DNA
topoisomerase II (Benedetti et al. ). Deletion of the N-terminal  residues
of the yeast enzyme resulted in a large decrease in the average size of the smaller
spheres in the tripartite structures (Benedetti et al. ), indicating that these
spheres represent the N-terminal or the GyrB-halves of the dimeric enzyme.
Interestingly, with either the human or the yeast enzyme, preincubation of the
enzyme with ADPNP caused a major change in the appearance of the images ; in
the majority of molecules, the two smaller spheres appeared to be in contact with
each other following incubation with the ATP analogue (Schultz et al.  ;
Benedetti et al. ). Micrographs of the human enzyme showed a preponderance
of bipartite structures in the presence of ADPNP, with a larger globular structure
of about  AI in diameter and a smaller one of about  AI in diameter (Schultz et
al. ).
In combination, the X-ray crystallography and electron microscopy results
suggest that each polypeptide of eukaryotic DNA topoisomerase II is organized
into two modules A and B, corresponding approximately to the A- and B-subunits
of gyrase. The pair of A modules in a dimeric enzyme are normally in contact with
each other in the presence or absence of bound ATP. The B modules, on the other
hand, form strong contacts only upon binding of ADPNP (and by implication
ATP) ; in the absence of bound ATP, the pair of B modules often stay apart.
Remarkably, there seem to be few protein–protein contacts between the A and B
modules in the absence of DNA, as suggested by the separation of the two
modules in electron micrographs of the molecules (Schultz et al.  ; Benedetti
et al. ), and by the results of protein footprinting experiments designed to test
whether any lysyl residue in the GyrA or GyrB half of yeast DNA topoisomerase
II might be protected against citraconylation by the presence of the other half of
the polypeptide (Li & Wang, ).
The structural studies, especially the crystal structures of the -kDa yeast
DNA topoisomerase II fragment and the -kDa E. coli DNA gyrase A subunit
fragment, also reveal a high degree of flexibility in the detailed positioning of the
A simple molecular machine

various domains of the type II DNA toposiomerases. This structural variability
is presumably related to the ability of a type II DNA topoisomerase to assume
very different conformations during its catalytic steps. At the same time, this
structural flexibility increases the difficulty of reconstructing the various
conformational states of a functional enzyme from the structures of its parts.
. A molecular model of DNA transport by a type II DNA topoisomerase
The structural data summarized above provided additional support of the model
depicted in Fig.  and added much molecular details to the model. Fig. 
illustrates a refined model combining the known biochemical and structural data
(Berger et al. ). The particular enzyme in this illustration is that of yeast DNA
topoisomerase II lacking the C-terminal amino acid residues that are dispensable
for its catalytic actions. It is most likely, however, that the basic features of the
model are shared by all type II DNA topoisomerases. In a later section, the
plausible difference between bacterial DNA gyrase and type II enzymes
exemplified by yeast DNA topoisomerase II will be discussed.
The structure denoted by  in Fig.  is a representation of the free enzyme in
an open-clamp conformation. The purple (A«) structure in  is the same as that
#
seen in the -kDa crystal structure except that the pair of B« subfragments with
the ATPase subfragments tagged on, are not in contact. Such an open
conformation with detached B« subfragments is presumably one of the major
conformations of the free enzyme, and this open conformation is necessary for the
binding of the DNA G-segment (shown as a rod above the enzyme).
Upon binding of the DNA G-segment, there is a significant change in the
conformation of the enzyme (). The two enzyme halves move toward each other
by a large distance, estimated to be – AI (Berger et al. ), and the pair of
active-site tyrosyl residues are now positioned near the scissile phosphates for
nucleophilic attack. The structure of the purple (A«) dimer depicted in 
#
resembles more the one shown in Fig.  for the A« dimer of E. coli gyrase (Morais
Cabral et al. ) than the one shown in Fig.  for the -kDa fragment of the
yeast enzyme. The precise positions of the B« subfragments in  are uncertain. It
has been suggested (Berger et al., ) that these subfragments may be oriented
very differently from their positions in the -kDa fragment structure of Berger
et al. (). The pair of ATPase subfragments in  can combine or separate
depending on the state of ATP binding, and a T-segment can enter the enzyme
clamp when these jaws are apart. Closure of the jaws upon the binding of ATP
would then trap the T-segment and initiate a conformational cascade in the
enzyme-DNA complex : the DNA G-segment is transiently broken and splayed
apart, and the T-segment is forced through the DNA gate into the large cavity
below (illustrations  to  in Fig. ).
The conformation of the (B«A«) part of the enzyme shown in  of Fig.  is that
#
seen in the crystal structure of the -kDa fragment of the yeast enzyme. When
the two enzyme halves retract toward each other following the passage of the Tsegment through the DNA gate, the size of the cavity containing the T-segment

J. C. Wang
is reduced and the T-segment exits the enzyme through the A«–A« dimer interface
(illustration  of Fig. ). The A«–A« interface has been termed the C-gate because
its constituents are close to the carboxyl termini of the A« polypeptides. The open
state of the C-gate is probably a transient one. Following the exit of the T-segment
and closure of the C-gate, the N-gate can re-open after ATP hydrolysis and the
release of the hydrolytic products ( in Fig. ), and the enzyme is posed for
another cycle of DNA transport.
The availability of detailed structural information had also made it possible to
test more directly a key postulate of the two-gate protein clamp model, namely the
role of the C-gate for the exit of the T-segment (Roca et al. ). Based on the
crystal structure of the -kDa fragment of the yeast enzyme, site-directed
mutagenesis was carried out to replace Lys- and Asn- by cysteines. In
the mutant enzyme, a pair of disulphide bonds can form across the A«–A« dimer
interface to lock the C-gate. When the single-step decatenation experiment of
Roca & Wang () was repeated with the mutant enzyme, with its C-gate locked
by the formation of disulphide bonds across the A«–A« interface, the unlinked
DNA rings following one round of DNA transport were found to remain
topologically linked to the enzyme–ADPNP complex (Roca et al. ).
Furthermore, exit of the DNA ring containing the T-segment was observed upon
subsequent addition of -mercaptoethanol, which reduces the disulphide bonds
and hence unlocks the C-gate, to the reaction products (Roca et al. ). These
and other data, while providing strong evidence in support of the two-gate
mechanism, do not preclude the possibility that under certain conditions (for
example when the G-segment can dissociate readily from the enzyme) a type II
DNA topoisomerase could promote DNA transport by a one-gate mechanism
(Lindsley, ). A special one-gate mechanism to account for the ATPindependent relaxation of negative supercoiled DNAs by bacterial DNA gyrase
will be discussed in Section ...
.                                              
     
A type II DNA topoisomerase can be viewed as a molecular machine that utilizes
ATP as the fuel in the transport of one DNA double helix through another. From
a thermodynamic point of view, only DNA supercoiling by bacterial DNA gyrase,
and not the removal of supercoils by the other type II DNA topoisomerases, is
expected to require a high-energy cofactor like ATP. It is clear, however, that
kinetically the transport of DNA by a type II DNA topoisomerase is strongly
dependent on ATP. Removal of DNA supercoils by yeast DNA topoisomerase II
is undetectable in the absence of ATP, and mutating a critical ATPase domain
residue Gly- of the enzyme to alanine also abolishes its DNA relaxation
activity in the presence of ATP (Linsley & Wang,  a).
A low level of DNA transport activity in the absence of ATP is present,
however, in several type II DNA topoisomerases including E. coli DNA gyrase
(Gellert et al. ,  ; Sugino et al.  ; Higgins et al.  ; Kreuzer &
A simple molecular machine

Cozzarelli,  ; Mizuuchi et al.  ; Krasnow & Cozzarelli,  ; Marians,
), phage T DNA topoisomerase (Liu et al. , ) and Drosophila DNA
topoisomerase II (Osheroff et al., ). A derivative of gyrase lacking the ATPase
domain was also found to relax supercoiled DNA in the presence or absence of
ATP (Brown et al.  ; Gellert et al. ). In the sections below, the available
information on how a type II DNA topoisomerase couples ATP usage to DNA
transport is summarized. Attempts are made, at the risk of oversimplification, to
interpret this intricate coupling in terms of the known or predicted structural
features of the various enzyme–DNA complexes. The low level ATP-independent
DNA transport activity of DNA gyrase will be discussed in a separate section.
. ATP utilization
Because the binding of the non-hydrolysable ATP analogue ADPNP to a type II
topoisomerase can effect one round of DNA transport, it is generally assumed that
DNA transport by the enzyme is driven by ATP binding, and ATP hydrolysis and
product release are only necessary for enzyme turnover (Sugino et al.  ;
Peebles et al. ). It is plausible, however, that ATP hydrolysis may further
accelerate the coupled process. An example is provided by the recent study of the
roles of nucleotide cofactors in the reaction steps catalysed by the elongation factor
G, a GTPase involved in the transloction of bacterial ribosomes along messenger
RNA (Rodnina et al. ). It is known that non-hydrolysable GTP analogues can
effect one round of translocation. With GTP itself, however, rapid hydrolysis of
the nucleotide was found to precede translocation. These observations suggest
that although the binding of a non-hydrolysable GTP analogue can effect
translocation, GTP hydrolysis may further accelerate the rearrangements of the
ribosome that drive translocation (Rodnina et al. ).
The intricacy of ATP utilization in the transport of DNA by the type II DNA
topoisomerases is illustrated by the recent pre-steady state kinetic analysis of ATP
hydrolysis by yeast DNA topoisomerase II (Harkins & Lindsley,  a, b). As
expected, the DNA-bound enzyme consumes two ATP per enzyme turnover
event. These bound ATP molecules are hydrolysed sequentially, however. Rapid
hydrolysis of one ATP and the release of the hydrolytic products appear to occur
before the second ATP is hydrolysed. Deducing the structural changes
accompanying this stepwise hydrolysis of bound ATP should provide further
insight on the intricate coupling of ATP usage to DNA transport.
. How tight is the coupling between DNA transport and ATP binding and
hydrolysis ?
.. Dependence of the coupling efficiency on the degree of supercoiling of the
DNA substrate
Measurements with yeast DNA topoisomerase II indicate that the coupling
efficiency can be very high under optimal conditions. Quantitative analysis of
linking number changes of DNA rings showed that for a moderately supercoiled

J. C. Wang
DNA,  % of all bound enzyme molecules are able to promote one round of
DNA transport upon the addition of excess ADPNP (Roca & Wang, ).
Because a fraction of the bound enzyme molecules was probably not active, the
actual coupling efficiency could be even higher and close to the theoretical limit of
 %. The same study showed that with a completely relaxed DNA ring the
efficiency is considerably lower : about one out of four bound enzyme molecules
would complete a DNA transport event upon ADPNP addition.
Two sets of measurements of the coupling efficiency in the presence of ATP
itself rather than its non-hydrolysable analogue were also carried out with the
yeast enzyme (Lindsley & Wang,  a). In one, parallel measurements of the
rate of ATP hydrolysis and the rate of removal of DNA supercoils were carried
out at a saturating ATP concentration of  m. The reaction temperature was set
below ± °C so that the ATPase and DNA transport rates were slowed down
sufficiently to permit manual quenching of the reactions. From these
measurements it was calculated that about ±³ ATP were hydrolysed per DNA
transport event. When the same measurements were carried out at  °C and
lower ATP concentrations, however, the number of ATP hydrolysed per DNA
transport event was measured to be as low as ±, with an average of ±³±.
These results suggest that at a high ATP concentration when ATP binding is
rapid, closure of the enzyme clamp often occurs without trapping a T-segment ;
at a low ATP concentration, at least in the case of a supercoiled DNA the rates of
ATP binding and the closure of the enzyme clamp are both slow relative to the
rate of entrance of the T-segment into an enzyme in its open-clamp conformation,
leading to a very high coupling efficiency.
In terms of the number of ATP required for one round of DNA transport by
a type II DNA topoisomerase, the theoretical limit is probably close to one ATP
per DNA transport event at very low ATP concentrations : the binding of an ATP
to one half of the dimeric enzyme could be sufficient in triggering dimerization of
the N-terminal domain and the conformational cascade that follows, leading to the
transport of a T-segment through the enzyme-bound G-segment. Studies with a
heterodimeric yeast enzyme with one wild-type polypeptide and one mutant
polypeptide with a mutation in the ATPase domain suggest that the binding of a
single ATP to a dimeric enzyme is sufficient to trigger the closure of the enzyme
clamp (Lindsley & Wang,  b). At high ATP concentrations, however, the
binding of ATP to wild-type yeast DNA topoisomerase II is cooperative, and the
binding of an ATP to one protomer of the dimeric enzyme is likely to be followed
rapidly by the closure of the N-gate and the binding of a second ATP to the other
protomer (Lindsley & Wang,  a). Therefore at high ATP concentrations the
theoretical limit of the coupling efficiency is expected to be two ATP for each
DNA transport event.
For E. coli DNA gyrase, the apparent coupling efficiency was also found to be
strongly dependent on DNA topology. With a positively supercoiled DNA, the
coupling efficiency approaches  %, and a linking number reduction of two per
(BA) molecule was observed upon the addition of ADPNP (Bates et al. ).
#
The coupling efficiency decreases steadily with DNA substrates of increasing
A simple molecular machine

degrees of negative supercoiling, and the ADPNP-mediated reduction of Lk
becomes undetectable for a moderately negatively supercoiled DNA with a
specific linking difference of ®± [Bates et al.  ; the specific linking
difference is defined as the fractional deviation of Lk from Lk° or (Lk®Lk°)}Lk°].
Earlier measurements also showed that the addition of the non-hydrolysable
nucleotide to E. coli DNA gyrase bound to a relaxed DNA reduces Lk by about
± per A protomer or ± per dimeric enzyme molecule (Sugino et al. ),
corresponding to an efficiency of about  %. With ATP as the cofactor, earlier
measurements of negative supercoiling of a relaxed DNA by gyrase suggested that
the apparent coupling efficiency is high during the early phase of the reaction, but
drops to zero when the specific linking difference of the DNA approaches ®±
(reviewed in Reece & Maxwell,  b).
The results summarised above show that in terms of the relaxation of a
positively supercoiled DNA, the coupling efficiency for bacterial DNA gyrase is
not very different from that for yeast DNA topoisomerase II. Because the yeast
enzyme does not catalyse DNA negative supercoiling, it is not possible to compare
the efficiencies of the two enzymes in this reaction. There are two complications
in the interpretation of the coupling efficiencies of DNA negative supercoiling by
gyrase. First, in nearly all measurements with the bacterial enzyme, only the net
change in the average linking number of a DNA ring was examined. For a DNA
gyrase bound to a negatively supercoiled DNA, the formation of a positive-noded
configuration between the G-segment and the incoming T-segment is favoured by
the DNA-enzyme binding energy, but disfavoured by the free energy of DNA
supercoiling (see Section ..). Therefore the number of DNA transport events
can be equated to one half of the net decrease in Lk only if the DNA–enzyme
binding energy is the dominant term, otherwise ATP-triggered inversion of both
positive and negative nodes may occur in a population of negatively supercoiled
DNA molecules. Second, DNA gyrase is known to possess an ATP-independent
DNA relaxation activity that is specific for negatively supercoiled DNA (Sugino
et al.  ; Gellert et al. ,  ; Higgins et al. ). This uncoupled
pathway may significantly reduce the apparent coupling efficiency, especially if a
highly negatively supercoiled DNA substrate is used. In experiments on the
coupling efficiency of ADPNP triggered DNA transport in a positively supercoiled
DNA, these complications are largely absent (the first complication can also be
eliminated entirely by the use of a DNA substrate consisting of a single linking
number isomer, so that the inversion of both positive and negative nodes can be
measured directly from an analysis of the distribution of the linking numbers of
the products ; see Roca & Wang, ).
.. Coupling efficiency and DNA binding
In terms of the two-gate protein clamp model, the above data can be understood
by assuming that normally the binding of ADPNP or ATP to an enzyme bound
to a G-segment always leads to the closure of the N-gate and the sequential
transport of a T-segment, if one is present inside the enzyme, through the DNA
gate and the protein exit gate. This interpretation is supported by two

J. C. Wang
observations. First, as shown by the single-step decatenation experiments of Roca
& Wang (), the DNA ring containing the T-segment always exits the enzyme
clamp following its passage through the DNA gate. Second, through the use of a
mutant enzyme YF, which is incapable of opening the DNA gate owing to the
replacement of the pair of active site tyrosyl residues by phenylalanines, it was
found that the binding of ADPNP to a G-segment bound mutant enzyme can
trigger the closure of the N-gate, but a second DNA ring can not be captured
(J. Roca, W. Li, and J. Wang, unpublished). Together, these two observations
demonstrate that the probability of capturing a T-segment without its being
transported through the DNA gate and the exit gate of the enzyme is low.
If the binding of ADPNP to a G-segment-bound enzyme is always followed by
the transport of a T-segment that has entered the enzyme clamp, then the
dependence of the overall coupling efficiency on the topology of the DNA
substrate must reflect differences in the probability of finding a T-segment within
an open enzyme clamp with DNA substrates supercoiled to different extents. For
type II DNA topoisomerases resembling yeast DNA topoisomerase II, the
coupling efficiencies measured in experiments using ADPNP indicate that for a
supercoiled DNA substrate the probability of finding a T-segment within an open
enzyme clamp is very high, and the same is true for bacterial DNA gyrase bound
to a positively supercoiled DNA substrate. The situation with DNA gyrase bound
to a negatively supercoiled DNA is probably rather different, however, as
discussed above. In experiments using ATP as the cofactor, the coupling
efficiencies are likely to be determined by the relative rates of T-segment entrance
and clamp closure, rather than by the probability of finding a T-segment in an
equilibrium population of enzyme molecules.
For yeast DNA topoisomerase II bound to a relaxed DNA, the measured
coupling efficiency of DNA transport upon addition of ADPNP is around  %
(Roca & Wang, ). This result suggests that the binding of a T-segment to the
enzyme in its open-clamp conformation is rather weak under the experimental
conditions (in  m Tris HCl, pH ,  m EDTA,  m KCl,  m MgCl
#
and  m -mercaptoethanol, at  °C). The T-segment presumably goes in and
out of the enzyme clamp in the absence of the nucleotide, and is not in contact with
the enzyme most of the time.
Stable binding of eukaryotic DNA topoisomerase II to DNA crossovers has
been implicated by results obtained by electron microscopy (Zechiedrich &
Osheroff, ), by the formation of DNA knots (Hsieh,  ; Wassermann &
Cozzarelli,  ; Roca et al. ), and by experiments measuring the residual
number of supercoils upon relaxation of a supercoiled DNA in the presence of
varying amounts of yeast DNA topoisomerase II (Roca et al.  ; relaxation was
carried out with a type IB DNA topoisomerase in the absence of ATP so that the
DNA transport activity of the type II enzyme was not manifested). It is uncertain
in these measurements, however, whether a DNA crossover bound by an enzyme
molecule represents a pair of G- and T-segments. Under conditions that favour
crossover binding, the efficiency of DNA transport upon the addition of ADPNP
is low (Roca et al. ). Furthermore, in the ADPNP-triggered decatenation of
A simple molecular machine

a supercoiled DNA ring singly linked to a relaxed DNA ring, yeast DNA
topoisomerase II was found to preferentially bind to the supercoiled ring, which
is consistent with the binding of the enzyme to a DNA crossover ; this one-step
DNA transport event was found to efficiently unlink the pair of rings, however,
indicating that the T-segment is from the relaxed DNA ring rather than a
crossover in the supercoiled ring (Roca & Wang, ). The plausible binding of
type II DNA topoisomerases to crossovers that do not include a T-segment was
also invoked to account for a remarkable property of these enzymes that was
revealed by recent experiments (Rybenkov et al. ). It was observed that in the
presence of ATP, a variety of the type II DNA topoisomerases can reduce the
steady state fraction of knotted or catenated DNA molecules to levels up to two
orders of magnitude below those at thermodynamic equilibrium.
. DNA relaxation by a type II DNA topoisomerase : how is the high efficiency
of coupling achieved ?
In essence, the available data for the type II DNA topoisomerases suggest that in
the relaxation of a positively supercoiled DNA by either yeast DNA topoisomerase
II or E. coli DNA gyrase, the overall high efficiency of coupling of DNA transport
to ATP usage is achieved by the high efficiencies of two consecutive reactions.
First, the binding of two ATP molecules to a DNA-bound dimeric enzyme at a
high ATP concentration almost always leads to the closure of the N-gate. Second,
provided that a T-segment has already entered the enzyme clamp, the closure of
the N-gate almost always leads to the transport of the trapped T-segment, first
through the DNA gate and then the exit gate. The details of these and additional
factors involved in the coupling of ATP usage to DNA transport are discussed in
the sections below.
.. ATP hydrolysis and the closure of the enzyme clamp
A key step in the ATP-triggered conformational cascade of a type II DNA
topoisomerase is the dimerization of the ATPase domains of the enzyme following
ATP binding. Therefore a high efficiency of coupling requires that dimerization
must precede ATP hydrolysis. This is nicely accomplished by the participation of
amino acid residues of both protomers in each of the ATPase site of the dimeric
enzyme : ATP hydrolysis can not occur unless the N-gate is closed to form the
ATPase catalytic pocket (Wigley et al.  ; Tamura et al.  ; Ali et al.  ;
O’Dea et al. ).
.. Coupling of ATPase activity to DNA binding
In the presence of a high concentration of DNA, the ATPase activity of yeast
DNA topoisomerase II is stimulated by about -fold (Lindsley & Wang,  a).
For the  residue N-terminal fragment of the yeast enzyme, the ATPase activity
is only marginally stimulated, by about  %, by the presence of excess DNA (S.
Olland and J. Wang, unpublished). Similarly, for purified E. coli DNA GyrB

J. C. Wang
protein, the ATPase activity is not significantly stimulated by DNA in the absence
of the GyrA protein (Sugino et al.  ; Sugino & Cozzarelli,  ; Staudenbauer
& Orr,  ; Maxwell & Gellert, ). Because the binding of a DNA Gsegment is likely to involve both the GyrA and GyrB regions of a type II DNA
topoisomerase (see Section .), these results are consistent with the notion that
stimulation of the ATPase activity of a type II DNA topoisomerase by DNA is
largely through the binding of the enzyme to the G-segment. The binding of the
two halves of a dimeric enzyme molecule to a contiguous DNA segment may
facilitate dimerization of the N-gate jaws to form the active form for ATP
hydrolysis ; allosteric changes within the ATPase domain, triggered by the
binding of a dimeric enzyme to DNA, may further enhance the ATPase activity
(see the section below).
Whether the DNA dependence of the ATPase activity involves interaction
between the enzyme and the T-segment as well is less clear. Studies with E. coli
DNA gyrase showed that the stimulation of its ATPase activity by DNA is
dependent on the length of the DNA ; DNA shorter than  bp can stimulate the
ATPase only at a high concentration, suggesting that DNA must bind to two or
more sites of the enzyme to have an effect on the ATPase activity (Maxwell &
Gellert, , ). These data can not distinguish, however, between a model
in which multiple sites are involved in the binding of the G-segment, and one in
which the multiple sites represent both G- and T-segment binding sites (Maxwell
& Gellert, , ).
Following the solution of the three-dimensional structure of the -kDa E. coli
GyrB fragment (Wigley et al. ), site-directed mutagenesis within this
fragment was carried out to replace Arg- by a glutamine (Tingey & Maxwell,
). Arg- is situated at a constriction of the channel formed by the closing
of the N-gate (Fig. ), and was postulated to be a part of a binding surface for the
T-segment (Wigley et al. ). Interestingly, the RQ mutant enzyme retains
the intrinsic ATPase activity of the wild-type enzyme but the activity is not
stimulated by DNA (Tingey & Maxwell, ). These observations led to the
suggestion that the binding of the T-segment within the enzyme clamp may be
required for stimulation of the ATPase activity. The effect of the RQ mutation
on the DNA dependence of the ATPase activity could also be attributed to,
however, an allosteric change within the ATPase clamp of the enzyme when a wild
type but not an RQ mutant enzyme binds to a G-segment (see Section ..).
It might seem ideal for a very high efficiency of coupling between ATP usage
and DNA transport if ATP hydrolysis requires the entrance of a T-segment into
a G-segment bound enzyme. This is not so. As described earlier, it appears that
at a high concentration of ATP a G-segment-bound enzyme clamp may often
close without capturing a T-segment. In such a situation, if the binding of a
T-segment is indeed a prerequisite for ATP hydrolysis, an enzyme bound to a
G-segment would get stuck in the closed clamp conformation whenever it fails to
capture a T-segment upon closure of the N-gate : it would have to rely on its
DNA-independent ATPase activity to free itself from the closed-clamp
conformation.
A simple molecular machine

.. Structural changes in the enzyme upon closure of the N-gate
Following ATP binding and the closure of the N-gate, a conformational cascade
ensues to effect the transport of the T-segment through the G-segment. Several
structural changes in the enzyme have been detected upon the binding of ADPNP,
the most conspicuous of which being the switch of an SV endoproteinase
cleavage site in yeast DNA topoisomerase II (Lindsley & Wang, ). In the
absence of ADPNP, this proteinase cleaves the yeast enzyme after Glu- (site
A in Fig. ) and around residue  (site C in Fig. ) ; in the presence of ADPNP,
cleavage occurs after Glu- (site B in Fig. ) and site C. Limited proteolysis of
the -kDa N-terminal E. coli gyrase B-subunit also showed that in the presence
of ADPNP, cleavage by trypsin occurs after Lys-, but the -kDa N-terminal
product is resistant to further cleavage that would occur in the absence of the
nucleotide (Ali et al. ).
Probing the chemical reactivity of individual yeast DNA topoisomerase II lysyl
residues showed that ADPNP binding affects lysine citraconylation at at least six
sites. Three of these sites, around Lys-}, }, and }, showed a
reduced level of citraconylation in the presence of ADPNP (Li & Wang, ).
Based on the crystal structure of the ATPase domain of E. coli gyrase and amino
acid sequence alignment of type II DNA topoisomerases, it was suggested that the
Lys-} and } sites are probably contacted directly by a bound
ADPNP, and reactivity at the Lys-} site is probably affected by ADPNP
because of its location at the N-gate dimerization interface. Citraconylation at
Lys-, , and , on the other hand, was found to be enhanced by ADPNP
binding, suggesting that these lysyl residues become more exposed to the solvent
through allosteric conformational changes induced by the binding of the
nucleotide to the ATPase catalytic sites. ADPNP binding also affects reactivity of
one or more lysyl residues around positions  and  in a less direct way : these
positions are protected against citraconylation by DNA binding, and this DNAmediated protection appears to be reduced in the presence of the non-hydrolysable
ATP analogue (Li & Wang, ).
.. ATP binding}hydrolysis and DNA cleavage
There are conflicting hints on the relation between the binding and hydrolysis of
ATP by a type II DNA topoisomerase and the cleavage and rejoining of the Gsegment. Two lines of evidence suggest that ATP utilization affects transient
cleavage of DNA mediated by type II DNA topoisomerases and vice versa. First,
covalent adduct formation between DNA and type II DNA topoisomerases, as
revealed by the addition of a protein denaturant to the DNA–enzyme complexes,
is often enhanced by the presence of ATP or ADPNP (Sugino et al.  ; Peebles
et al.  ; Fisher et al.  ; Sander & Hsieh,  ; Kreuzer & Alberts,  ;
Osheroff, ). Second, when the active site tyrosine Tyr- of yeast DNA
topoisomerase II is mutated to phenylalanine, and thus abolishing the DNA
cleavage activity of the enzyme, the DNA-dependent ATPase activity of the
enzyme was found to be largely abolished (W. Li, J. Roca, and J. Wang,

J. C. Wang
unpublished ; J. Lindsley, personal communication ; S. Gasser, personal
communication ; A. Maxwell, personal communication). In contrast, there have
also been ample examples that the type II DNA topoisomerases and their
truncation derivatives can cleave DNA in the absence of ATP (see for examples,
Sander & Hsieh,  ; Osheroff,  ; Reece & Maxwell,  a). Furthermore,
enzymes lacking the ATPase activity because of the presence of point mutations
or deletions were nevertheless found to have DNA cleavage activities similar to
that of the wild-type enzyme (Jackson & Maxwell,  ; Lindsley & Wang,
 a ; Berger et al.  ; O’Dea et al.  ; Tingey & Maxwell, ).
These seemingly conflicting observations are yet to be placed in the same
mechanistic framework. It is noteworthy, however, that the experimentally
measured efficiency of DNA cleavage by a type II DNA topoisomerase probably
represents a composite of contributions from different conformations of the
enzyme. The effect of ATP or ADPNP binding on DNA cleavage by the enzyme
is probably realized only if the enzyme is in the proper closed-clamp conformation.
In terms of the coupling between ATP usage and DNA transport, the effect of the
nucleotide cofactor on the cleavage of the G-segment by the enzyme in its closedclamp conformation is likely to be the important one, as the DNA transport event
presumably occurs only inside the enzyme in its closed-clamp conformation.
Judging from the large drop in the ATPase activity of yeast DNA topoisomerase
II upon mutating the active site tyrosyl residue Tyr- to phenylalanine, the
binding and}or hydrolysis of ATP is likely to have a significant effect on the
cleavage of the G-segment in the closed-clamp conformation of the enzyme.
.. Entrapment of the T-segment and opening of the DNA gate
Closely related to the question on the effects of ATP binding and hydrolysis on the
cleavage of the G-segment is whether the cleavage reaction involves the Tsegment. Studies with a  bp DNA showed that the efficiency of Drosophila
DNA topoisomerase II-mediated cleavage of this short DNA has a sigmoidal
dependence on the concentration of the DNA (Corbett et al. ). Based on this
observation, the authors suggested that the enzyme might interact with more than
one -mer in its cleavage of one of them, and that the presence of a T-segment
might be required for the cleavage of a G-segment (Corbett et al. ).
Studies of DNA cleavage by the closed clamp form of yeast DNA topoisomerase
II show however, that the transient breakage of the G-segment does not
require the entrapment of a T-segment (Roca & Wang, ). In one experiment,
yeast DNA topoisomerase II was first converted to the closed clamp conformation
by the addition of ADPNP, and a linear DNA was allowed to thread through the
enzyme annulus to provide a G-segment. When etoposide and SDS were
sequentially added to the enzyme–DNA mixture, it was found that the G-segment
was efficiently cleaved by the enzyme. Because only one DNA double helix can be
threaded through an enzyme in its closed-clamp conformation, the finding of this
experiment demonstrates that the entrapment of a T-segment is not required for
the transient cleavage of the G-segment (Roca & Wang, ).
The opening of the DNA gate by a type II DNA topoisomerase, however,
A simple molecular machine

requires not only breakage of the DNA strands through transesterification
between the enzyme and DNA, but also the movement of the enzyme-linked
DNA ends away from each other to permit the passage of the T-segment. These
two steps are akin to unlocking a gate and opening it. Whereas the T-segment is
not required for unlocking the DNA gate, its entrapment may be strongly coupled
to opening it. From the crystal structure of the ATPase subfragment of bacterial
gyrase B-subunit, the dimension of the archway formed by the dimerized
fragment is not sufficiently large to accommodate a duplex DNA (Tingey &
Maxwell, ). Therefore steric repulsion is likely to accompany the trapping of
a T-segment, and this repulsion may force the widening of the DNA gate (Berger
et al. ).
In one experiment with the YF mutant of yeast DNA topoisomerase II, it
was found that closure of the N-gate of a G-segment-bound mutant enzyme
would occur normally upon the addition of ADPNP, but this closure could not
trap a second DNA ring present at a high concentration (J. Roca, W. Li, and J.
Wang, unpublished). This finding suggests that the widening of the DNA gate
may be coupled to the trapping of the T-segment. For the mutant enzyme,
unlocking and opening the DNA gate are not possible, and hence the G-segmentbound enzyme can not assume the closed clamp conformation even if a T-segment
has entered the enzyme clamp. In a functional enzyme, on the other hand,
widening the DNA gate upon entrapment of a T-segment allows the T-segment
to move through the DNA gate into the large cavity on the other side of the Gsegment. The positive surface charges lining this cavity is likely to provide further
attraction for the T-segment to move into it.
.. Closure of the DNA gate and exit of the T-segment
Once the T-segment has passed the G-segment and entered the large hole lined
with positive surface charges (structure  in Fig. ), it would seem to be in a
favourable micro-environment. The question is then what drives its subsequent
exit ?
The answer to the above question is likely to lie in the energetics of the enzymeDNA complex in which the G-segment is cleaved and the DNA gate is wide open.
This state is most likely thermodynamically unstable ; the enzyme–DNA complex
is probably driven into this state only when the T-segment is trapped in the closed
N-terminal jaws of the protein. Therefore once the T-segment has passed through
the open DNA gene into the large hole, the DNA–enzyme complex would resume
its more stable conformation by closing the DNA gate. The movement of the two
halves of the enzyme–DNA complex toward each other would reduce the size of
the large hole, which could in turn force the exit of the T-segment (Berger et al.
).
From the above discussion, it is clear that the stepwise movement of a Tsegment through the multiple ports and locks of the canal in a G-segment-bound
enzyme requires the existence of energetically well-balanced conformational states
of the various enzyme–DNA complexes. The existence of these conformational
states is in turn dependent on the presence of intricately balanced structural

J. C. Wang
elements or hinges in the enzyme–DNA complexes. For example, based on the
crystal structures of the -kDa yeast DNA topoisomerase II fragment and the
-kDa bacterial gyrase fragment, it was suggested that conformational changes at
a pair of hinge structures in a type II DNA topoisomerase may lead to the opening
of the C-gate for the exit of the T-segment, as illustrated in Fig.  (Morais Cabral
et al. ).
. Directionality of DNA transport : why is bacterial gyrase unique ?
Among the type II DNA topoisomerases, bacterial DNA gyrase is unique in two
respects. First, whereas all type II DNA topoisomerases are ATP dependent, only
DNA gyrase is capable of catalysing the negative supercoiling of DNA. Second,
DNA gyrase and a derivative of it lacking the N-terminal half of the B-subunit
possess an ATP-independent DNA relaxation activity, which differs from the
ATP-dependent DNA transport activities of the other type II enzymes in that it
is specific for negatively supercoiled DNA (Sugino et al.  ; Gellert et al. ,
 ; Higgins et al. ). [An early report also ascribed to bacterial DNA gyrase
an ADPNP-dependent activity capable of removing positive supercoils
catalytically, but recent measurements showed that positive supercoil removal in
the presence of the non-hydrolysable analogue is stoichiometric and amounts to a
maximum of two supercoils per gyrase molecule (Bates et al. ).]
... Structural basis of the ability of bacterial DNA gyrase to catalyse the
ATP-dependent negative supercoiling of DNA
What might be the structural features that set gyrase apart from the other
members of the type II subfamily of DNA topoisomerases ? It was suggested
many years ago that the ability of DNA gyrase to catalyse DNA supercoiling is a
consequence of the unique way DNA gyrase binds to DNA (Wang, ). Only
in the case of DNA gyrase the G-segment is wrapped around the enzyme with a
right-handed writhe, and about  bp of the segment are in contact with the
enzyme (Liu & Wang,  a, b ; Klevan & Wang,  ; Fisher et al.  ;
Kirkegaard & Wang,  ; Morrison & Cozzarelli,  ; Rau et al.  ;
Orphanides & Maxwell, ) ; for the other type II DNA topoisomerases, a much
shorter region of the G-segment (about – bp) is contacting the enzyme
without a significant writhe of the segment (Spitzer & Muller,  ; Lee et al.
 ; Thomsen et al. ).
How can the wrapping of a DNA segment around a type II DNA toposiomerase
affect the outcome of the enzyme-mediated topological transformation of a DNA
ring ? When a T-segment enters the interior of an enzyme bound to a G-segment
on the same DNA ring, the two DNA segments can form a crossover with either
a plus sign or a minus sign, as illustrated in Fig. . The enzyme-mediated
transport of the T-segment through the G-segment always inverts this nodal sign.
Therefore, in order for a type II DNA topoisomerase to catalyse the negative
supercoiling of a DNA ring, the T- and G-segment must form a plus node before
the DNA transport event ; the linking number of the DNA ring is then reduced
A simple molecular machine

(a)
G
T
(b)
(c)
–
+
(d)
–
+
Fig. . The sign of the node between a pair of G and T-segments in a DNA ring before the
DNA transport event. The schematic in (a) illustrates a G-segment-bound enzyme (circle
with a cross inside) viewed down the molecular dyad from the N-gate to the C-gate. The Tsegment can enter the opened N-gate to form a positive node (b) or a negative node (c) with
the G-segment. If the enzyme does not impose a significant writhe in the G-segment, the
nodal sign between the incoming T-segment and the enzyme-bound G-segment must be
determined by the topology of the DNA ring : it would be positive in a positively
supercoiled DNA, negative in a negatively supercoiled one, and either positive or negative in
a relaxed DNA ring. In drawing (d ), the two crossing-segments are shown to be
perpendicular to each other ; the plus and minus mode drawn are structurally equivalent. In
the case of bacterial DNA gyrase, the enzyme imposes a right-handed writhe in the Gsegment. The nodal sign is then influenced by two terms : enzyme–DNA interaction favours
a positive node, and DNA conformation favours a particular nodal sign depending on the
sense of supercoiling. Illustration taken from Roca & Wang ().
by two by the inversion of the plus node to a minus one. The wrapping of a DNA
segment around gyrase apparently imposes a strong bias in favour of the formation
of a plus node between the two segments. For the other type II DNA
topoisomerases, wrapping of the DNA around the enzyme is not involved, and the
nodal sign an incoming T-segment makes with the G-segment is specified by the
conformation of the DNA : the segments would preferentially assume a plus sign

J. C. Wang
in a positively supercoiled DNA and a minus sign in a negatively supercoiled
DNA (Roca & Wang, ). Thus for type II DNA topoisomerases other than
gyrase, enzyme-mediated transport of a T-segment through a G-segment always
reduces the net number of supercoils in the DNA.
Recent studies have provided strong evidence in support of the above
interpretation. A C-terminal -kDa fragment of E. coli DNA gyrase A-subunit
(residues –) was shown to be involved in the right-handed wrapping of
DNA around the enzyme : mutant gyrase lacking this -kDa domain does not
impose a right-handed writhe in the DNA segment bound to it (Kampranis &
Maxwell, ), but the purified -kDa protein does by itself (Reece & Maxwell,
 a). In addition, deletion of this domain from DNA gyrase converts the
enzyme from a DNA-negative supercoiling activity to one that catalyses the ATPdependent removal of supercoils (Kampranis & Maxwell, ). Unlike DNA
gyrase but similar to the other type II DNA topoisomerases, this deletion mutant
of gyrase is also efficient in unlinking DNA catenanes in the presence of ATP
(Kampranis & Maxwell, ).
.. The ATP-independent DNA relaxation activity of gyrase
The ATP-independent DNA relaxation activity of gyrase is of particular
mechanistic interest in that it is specific for negative supercoils. In other words,
in the absence of ATP the enzyme retains a directionality in its transport of a
DNA segment through another segment of the same DNA ring, but this
directionality is the opposite of that in the presence of ATP.
One interpretation of the observed directionality of gyrase-mediated DNA
transport in the absence of ATP is that the ATP-independent reaction follows a
reverse pathway of the ATP-driven reaction : the T-segment enters through the
C-gate, passes through the G-segment, and exits the N-gate (Kampranis &
Maxwell,  ; Cullis et al.  ; Critchlow et al. ). This interpretation
implies that the C-gate of DNA gyrase might not be tightly closed and might often
pop open to allow a T-segment to slip through. There is no data to suggest,
however, that interaction across the A–A dimer interface in DNA gyrase is weaker
than that in yeast DNA topoisomerase II. The A subunit of E. coli as well as
Micrococcus luteus DNA gyrase exists as stable dimers in solution under a wide
range of conditions (see for examples, Klevan & Wang,  ; Kirschhausen et al.
), and the crystal structure of the -kDa gyrase A fragment shows an A–A
dimer interface burying an area of over  AI # (Morais Cabral et al. ).
A second problem with the reverse-path interpretation of the ATP-independent
DNA relaxation activity of DNA gyrase is that it provides no satisfactory
explanation for the specificity of the reaction, namely only negatively supercoils
are removed. This problem can be better seen by first considering a gyrase
molecule bound to a G-segment in a supercoiled DNA. As discussed before, the
wrapping of the G-segment around the enzyme positions a T-segment, T , which
"
makes a positive node with the G-segment, and is presumably located close to the
normal entrance gate of the enzyme. This means that the T-segment which would
follow the reverse path could not be this well-positioned T segment and must
"
A simple molecular machine

therefore be a different T-segment T . The specificity for the removal of negative
#
supercoils would then require that the enzyme could somehow position T and T
"
#
near the normal entrance and exit gate of the enzyme, respectively, and that the
nodal sign between the G-segment and T must be a negative one. There is
#
presently no structural hint of such an arrangement.
The above discussion demonstrates that alternatives to the reverse-path
interpretation should be considered. One model that is consistent with the
available data is that the ATP-independent removal of every two negative
supercoils involves the binding of a DNA gyrase molecule at a negative node in
a negatively supercoiled DNA, perhaps through the assembly of one GyrA dimer
and two GyrB monomers at that node, to form an enzyme–DNA complex similar
to the one which is formed in the normal ATP-driven pathway immediately after
the passage of the T-segment through the DNA gate. In this complex, one of the
arms of the negative DNA node forms the G-segment, and the other arm, which
would subsequently serve as the T-segment, is present inside the cavity bounded
by the A protomers and the G-segment. In the absence of the ATP dependent
closure of the entrance gate and the conformational cascade which would normally
follow, this pre-assembled T-segment could not pass through the normal exit gate
(the C-gate). It could escape through the DNA gate whenever the gate opens,
however, to form the more stable enzyme-DNA complex in which the T-segment
assumes a positive-noded orientation relative to the G-segment. This process
would remove two negative supercoils.
In the mechanism postulated above, thermodynamic considerations predict that
the formation of such a pre-assembled complex with a negative node between the
two DNA segments could occur only in a negatively supercoiled DNA. Assembly
of gyrase subunits at a positive node in a positively supercoiled DNA could occur,
of course, but the energetics would strongly favour the formation of the normal
enzyme-DNA complex before the transport of the T-segment, in which the DNA
wraps around the enzyme to position a T-segment in a positive-noded orientation
with respect to the G-segment. Such a thermodynamically stable complex can not
be relaxed further in the absence of ATP, which explains why the ATPindependent DNA relaxation activity of bacterial gyrase does not remove positive
supercoils.
.                 
Biochemical, genetic, and crystallographic studies of the type II DNA
topoisomerases in the past two decades have led to a detailed model on how such
an enzyme accomplishes its remarkable feat of moving one DNA duplex through
another. A type II DNA topoisomerase is in every way an ATP-fuelled molecular
machine. It is now well recognized that in the biological world most if not all
processes are carried out by intricate macromolecular machines, usually composed
of a complex assembly of components (see for examples, Alberts, ,  ;
Echols, ). The type II DNA topoisomerases are unusual in that they act as
single molecules : a single dimeric enzyme does all the work of binding to one

J. C. Wang
DNA double helix, capturing a second, opening a gate in the first DNA segment,
and transporting the captured segment through the transiently opened DNA gate,
all in a split second ! The beauty of this simple and yet efficient machine is
awesome. Because of their simplicity, the type II DNA topoisomerases also
provide opportunities for deeper understanding of coupled processes ; their
indispensability in cells and their importance as targets of antimicrobial and
anticancer therapeutics add yet another bonus to their studies.
.               
I am grateful to many who made the study of DNA topoisomerases a joyful
undertakening, to Qiyong Liu for his help in the preparation of several of the
figures, to Janet Lindsley and James Berger for sharing results prior to publication,
to Tao-shih Hsieh and Tony Maxwell for discussions, and to NIH for three
decades of support of the research of my laboratory on DNA topoisomerases. I
thank Richard Henderson and Kurt Wu$ thrich for their patience and prodding
during the writing of this review ; without patience they would have given up on
this review, and without their prodding I would have abandoned my charge.
.          
A, B. M. (). The DNA enzymology of protein machines. Cold Spring Harb.
Symp. Quant. Biol. , –.
A, B. (). The cell as a collection of protein machines : Preparing the next
generation of molecular biologists. Cell , –.
A, J. A., J, A. P., H, A. J. & M, A. (). The -kDa Nterminal fragment of the DNA gyrase B protein hydrolyses ATP and binds coumarin
drugs. Biochemistry , –.
A, J. A., O, G. & M, A. (). Nucleotide binding to the kilodalton N-terminal fragment of the DNA gyrase B protein. Biochemistry ,
–.
B, M. I., B, P., M, E. & T-V, G. P. (). In
vitro catenation and decatenation of DNA and a novel eukaryotic ATP dependent
topoisomerase. Cell , –.
B, A. D., O’D, M. H. & G, M. (). Energy coupling in Escherichia coli
DNA gyrase : the relationship between nucleotide binding, strand passage, and DNA
supercoiling. Biochemistry , –.
B, A. D. & M, A. (). DNA Topology. Oxford : IRL Press.
B, P., S, A., F, P. & W, J. C. (). Study of yeast DNA
topoisomerase II and its truncation derivatives by transmission electron microscopy.
J. Biol. Chem. , –.
B, J. M., F, D., W, J. C. & H, S. C. (). Structural similarities
between topoisomerases that cleave one or both DNA strands. Proc. natn. Acad. Sci.
U.S.A. , –.
B, J. M., G, S. J., H, S. C. & W, J. C. (). Structure and
mechanism of DNA topoisomerase II. Nature , –.
B, J. M. & W, J. C. (). Recent developments in DNA topoisomerase II
structure and mechanism. Curr. Op. Struct. Biol. , –.
A simple molecular machine

B, A.,  M, B., G, D., V, P. C., N, A. & F,
P. (). An atypical topoisomerase II from Archaea with implications for meiotic
recombination. Nature , –.
B, P. O. & C, N. R. (). A sign inversion mechanism for enzymatic
supercoiling of DNA. Science , –.
B, P. O., P, C. L. & C, N. R. (). A topoisomerase from
Escherichia coli related to DNA gyrase. Proc. natn Acad. Sci. U.S.A. , –.
C, P. R. & W, J. C. (). Alignment of primary sequences of DNA
topoisomerases. In DNA Topoisomerase and Their Applications in Pharmacology (ed.
L. Liu), pp. –. San Diego : Academic Press.
C, P. R., W, P.  W, J. C. (). The C-terminal domain of
Saccharomyces cerevisiae DNA toposiomerase II. J. Mol. Cell. Biol. , –.
C, M. (). Ribbons .. J. appl. Crystallogr. , –.
C, J. J. (). DNA is linked to the rat liver DNA nicking-closing enzyme by
a phosphodiester bond to tyrosine. J. biol. Chem. , –.
C, A. H., Z, E. L. & O, N. (). A role for the passage helix
in the DNA cleavage reaction of eukaryotic topoisomerase II. J. biol. Chem. ,
–.
C, N. R., B, T. C. & W, J. H. (). In DNA Topology and its
Biological Effects (ed. N. R. Cozzarelli and J. C. Wang), pp. –. Cold Spring
Harbor : Cold Spring Harbor Laboratory Press.
C, N. R. & W, J. C. (). DNA Topology and its Biological Effects. Cold
Spring Harbor : Cold Spring Harbor Laboratory Press.
C, D. G. & H, T. (). Function of the hydrophilic carboxyl terminus
of type II DNA topoisomerase from Drosophila melanogaster. I. In vitro studies. J.
biol. Chem. , –.
C, F. H. C. (). Linking numbers and nucleosomes. Proc. natn. Acad. Sci.
U.S.A. , –.
C, S. E., O’D, M. H., H, A. J., C, M., G, M. &
M, A. (). The interaction of the F plasmid killer protein, CcdB, with
DNA gyrase : induction of DNA cleavage and blocking of transcription. J. molec. Biol.
, –.
C, P. M., M, A. & W, D. P. (). Exploiting nucleotide
thiophosphates to probe mechanistic aspects of Escherichia coli DNA gyrase.
Biochemistry , –.
D, R. & V, M. (). Evidence for a ring structure of polyoma virus DNA.
Proc. natn. Acad. Sci. U.S.A. , –.
E, H. (). Nucleoprotein structures initiating DNA replication, transcription,
and site-specific recombination. J. biol. Chem. , –.
F, L. M., M, K., O’D, M. H., O, H. & G, M. (). Sitespecific interaction of DNA gyrase with DNA. Proc. natn. Acad. Sci. U.S.A. ,
–.
F, F. B. (). Decomposition of the linking number of a closed ribbon. Proc.
natn. Acad. Sci. U.S.A. , –.
G, M., F, L. M. & O’D, M. H. (). DNA gyrase : purification and
catalytic properties of a fragment of gyrase B protein. Proc. natn. Acad. Sci. U.S.A.
, –.
G, M., F, L. M. & O, H., O’D, M. H. & M, K. ().

J. C. Wang
DNA gyrase : site-specific interactions and transient double-stranded breakage of
DNA. Cold Spring Harbor Symp. Quant. Biol. , –.
G, M., M, K., O’D, M. H., I, T. & T, J.-I. ().
Nalidixic acid resistance : a second genetic character involved in DNA gyrase activity.
Proc. natn. Acad. Sci. U.S.A. , –.
G, M., M, K., O’D, M. H. & N, H. A. (). DNA gyrase : An
enzyme that introduces superhelical turns into DNA. Proc. natn. Acad. Sci. U.S.A.
, –.
G, T. & W, J. C. (). Yeast DNA topoisomerse II is encoded by a singlecopy, essential gene. Cell , –.
H, T. T. & L, J. E. ( a). Pre-steady state analysis of ATP hydrolysis
by Saccharomyces cerevisiae DNA topoisomerase II. . A DNA-dependent burst in
ATP hydrolysis. Biochemistry , –.
H, T. T., L, T. J. & L, J. E. ( b). Pre-steady state analysis of
ATP hydrolysis by Saccharomyces cerevisiae DNA topoisomerase II. . Kinetic
mechanism for the sequential hydrolysis of two ATP. Biochemistry , –.
H, N. P., P, C. L., S, A. & C, N. R. (). Purification of
subunits of Escherichia coli DNA gyrase and reconstitution of enzymatic activity. Proc.
natn. Acad. Sci. U.S.A. , –.
H, D. S. & W, J. C. (). Mapping the active site tyrosine of Escherichia
coli DNA gyrase. J. biol. Chem. , –.
H, T. (). Knotting of circular duplex DNA by type II DNA topoisomerase
from Drosophila melanogaster. J. biol. Chem. , –.
H, T. & B, D. (). ATP-dependent DNA topoisomerase from D.
melanogaster reversibly catenates duplex DNA rings. Cell , –.
J, A. P. & M, A. (). Identifying the catalytic residue of the ATPase
reaction of DNA gyrase. Proc. natn. Acad. Sci. U.S.A. , –.
K, S. C. & M, A. (). Conversion of DNA gyrase into a
conventional type II topoisomerase. Proc. natn. Acad. Sci. U.S.A. , –.
K, R., K, L. C. & Z, B. H. (). On the nature of chromosome-sized
DNA molecules. Cold Spring Harbor Symp. Quant. Biol. , –.
K, T., W, J. C. & H, S. C. (). DNA gyrase and its
complexes with DNA : direct observation by electron microscopy. Cell , –.
K, K. & W, J. C. (). Mapping the topography of DNA wrapped
around gyrase by nucleolytic and chemical probing of complexes of unique DNA
sequences. Cell , –.
K, L. & W, J. C. (). DNA gyrase-DNA complex containing  base
pairs of DNA and an α β protein core. Biochemistry , –.
# #
K, M. A. & C, N. R. (). Catenation of DNA rings by
topoisomerases. Mechanism of control by spermidine. J. biol. Chem. , –.
K, K. N. & A, B. M. (). Site-specific recognition of bacteriophage T
DNA by T type II DNA topoisomerase and Escherichia coli DNA gyrase. J. biol.
Chem. , –.
K, K. N. & C, N. R. (). Formation and resolution of DNA
catenanes by DNA gyrase. Cell , –.
L, M. P., S, M. & H, T. (). Nuclease protection by Drosophila DNA
topoisomerase II : enzyme}DNA contacts at the strong topoisomerase II cleavage
sites. J. biol. Chem. , –.
L, R. J., S, O. M., S, C. V., S, T., M, A., W,
A simple molecular machine

A. J. & W, D. B. (). The nature of inhibition of DNA gyrase by the
coumarins and the cyclothiolidines revealed by X-ray crystallography. EMBO J. ,
–.
L, W. & W, J. C. (). Footprinting of yeast DNA topoisomerase II lysyl side
chains involved in substrate binding and interdomainal interactions. J. biol. Chem.
, –.
L, C., W, J. C. & M! , A. (). Three dimensional structure of the
 K N-terminal fragment of E. coli DNA topoisomerase I. Nature , –.
L, J. E. (). Intradimerically tethered DNA topoisomerase II is catalytically
active in DNA transport. Proc. natn. Acad. Sci. U.S.A. , –.
L, J. E. & W, J. C. (). Proteolysis patterns of epitopically labeled yeast
DNA topoisomerase II suggest an allosteric transition in the enzyme induced by ATP
binding. Proc. natn. Acad. Sci. U.S.A. , –.
L, J. E. & W, J. C. ( a). On the coupling between ATP usage and DNA
transport by yeast DNA topoisomerase II. J. biol. Chem. , –.
L, J. E. & W, J. C. ( b). Study of allosteric communication between
protomers by immunotagging. Nature , –.
L, L. F. ( a). DNA Topoisomerases : Biochemistry and Molecular Biology. Boca
Raton : Academic Press.
L, L. F. ( b). DNA Topoisomerases and Their Applications in Pharmacology. Boca
Raton : Academic Press.
L, L. F., L, C. C. & A, B. M. (). T DNA topoisomerase : a new ATPdependent enzyme essential for initiation of T bacteriophage DNA replication.
Nature , –.
L, L. F., L, C. C. & A, B. M. (). Type II DNA topoisomerases :
Enzymes that can unknot a topologically knotted DNA molecule via a reversible
double-stranded break. Cell , –.
L, L. F., R, T. C., Y, L, T, K. M. & C, G. L. (). Cleavage of
DNA by mammalian DNA topoisomerase II. J. biol. Chem. , –.
L, L. F. & W, J. C. ( a). DNA–DNA gyrase complex : the wrapping of the
DNA duplex outside the enzyme. Cell , –.
L, L. F. & W, J. C. ( b). Micrococcus luteus DNA gyrase : active components
and a model for its supercoiling of DNA. Proc. natn. Acad. Sci. U.S.A. ,
–.
M, K. J. (). DNA gyrase-catalysed decatenation of multiply linked DNA
dimers. J. biol. Chem. , –.
M, A. (). DNA gyrase as a drug target. Trends Microbiol. , –.
M, A. & G, M. (). The DNA-dependence of the ATPase activity of
DNA gyrase. J. biol. Chem. , –.
M, A., R, D. C. & G, M. (). Mechanistic studies of DNA gyrase.
In Biomacromolecular Stereodynamics III : Proceedings of the Fourth Conversation in
the Discipline of Biomolecular Stereodynamics (ed. R. H. Sarma and M. H. Sarma),
pp. –. Albany, New York : Adenine Press.
M, M. (). Formation of hybrid DNA by rotary diffusion during genetic
recombination. J. molec. Biol. , –.
M, M. & S, F. W. (). The replication of DNA in E. coli. Proc. natn.
Acad. Sci. U.S.A. , –.
M, K. G., L, L. F. & E, P. T. (). A homogeneous type II DNA
topoisomerase from HeLa cell nuclei. J. biol. Chem. , –.

J. C. Wang
M, K., F, L. M., O’D, M. H. & G, M. (). DNA gyrase action
involves the introduction of transient double-stranded breaks in DNA. Proc. natn.
Acad. Sci. U.S.A. , –.
M C, J. H., J, A. P., S, C. V., S, N., M, A. &
L, R. C. (). Crystal structure of the breakage-reunion domain of DNA
gyrase. Nature , –.
M, A. & C, N. R. (). Site-specific cleavage of DNA by E. coli
DNA gyrase. Cell , –.
M, A. & C, N. R. (). Contacts between DNA gyrase and its
binding site on DNA : features of symmetry and asymmetry revealed by protection
from nucleases. Proc. natn. Acad. Sci. U.S.A. , –.
O’D, M. H., T, J. K. & G, M. (). Mutations in the B subunit of
Escherichia coli DNA gyrase that affect ATP-dependent reactions. J. biol. Chem. ,
–.
O, G. & M, A. (). Evidence for a conformational change in the
DNA gyrase–DNA complex from hydroxyl radical footprinting. Nucleic Acids Res.
, –.
O, N. (). Eucaryotic topoisomerase II. Characterization of enzyme
turnover. J. biol. Chem. , –.
O, N., S, E. R. & B, D. (). DNA topoisomerase II from
Drosophila melanogaster : Relaxation of supercoiled DNA. J. biol. Chem. ,
–.
P, C. L., H, N. P., K, K. N., M, A., B, P. O., S,
A. & C, N. R. (). Structure and activities of Escherichia coli DNA
gyrase. Cold Spring Harbor Symp. Quant. Biol. , –.
P, Y., K, D. & K, K. (). Topological complexes between DNA
and topoisomerase II and effects of polymines. Biochemistry , –.
R, D. C., G, M. & M, A. (). Structure of the DNA gyrase-DNA
complex as revealed by transient electric dichroism. J. molec. Biol. , –.
R, R. J. & M, A. ( a). Probing the limits of the DNA breakage-reunion
domain of the Escherichia coli DNA gyrase A protein. J. biol. Chem. , –.
R, R. J. & M, A. ( b). DNA gyrase : structure and function. CRC Critical
Rev. Biochem. Mol. Biol. , –.
R, D. & K, A. (). Helical periodicity of DNA determined by enzyme
digestion. Nature , –.
R, J., B, J. M., H, S. C. & W, J. C. (). DNA transport by a
type II topoisomerase : Direct evidence for a two-gate mechanism. Proc. natn. Acad.
Sci. U.S.A. , –.
R, J., B, J. M. & W, J. C. (). On the simultaneous binding of
eukaryotic DNA topoisomerase II to a pair of double-stranded DNA helices. J. biol.
Chem. , –.
R, J. & W, J. C. (). The capture of a DNA double-helix by an ATPdependent protein clamp : a key step in DNA transport by type II DNA
topoisomerases. Cell , –.
R, J. & W, J. C. (). DNA transport by a type II DNA topoisomerase :
evidence in favor of a two-gate mechanism. Cell , –.
R, J. & W, J. C. (). The probabilities of supercoil removal and decatenation
by yeast DNA topoisomerase II. Genes to Cells , –.
A simple molecular machine

R, M. V., S, A., K, V. I. & W, W. ().
Hydrolysis of GTP by elongation factor G drives tRNA movement on the ribosome.
Nature , –.
R, V. V., U, C., V, A. V. & C, N. R. ().
Simplification of DNA topology below equilibrium values by type II topoisomerases.
Science , –.
S, M. & H, T. (). Double strand DNA cleavage by type II DNA
topoisomerase from Drosophila melanogaster. J. biol. Chem. , –.
S, P., O, S., O, P. & H, R. (). Structure and
conformational changes of DNA topoisomerase II visualized by electron microscopy.
Proc. natn. Acad. Sci. U.S.A. , –.
S, K. & Y, M. (). Functional dissection of the phosphorylated
termini of fission yeast DNA topoisomerase II. Mol. Cell. Biol. , –.
S, J. R. & M, M. T. (). A consensus sequence for cleavage by
vertebrate topoisomerase II. Nucleic Acids Res. , –.
S, W. L. & O, E. (). DNA gyrase : affinity chromatography on
novobiocin-sepharose and catalytic properties. Nucleic Acids Res. , –.
S, G. L., K, G. J. & H, W. M. (). T DNA-delay proteins required
for specific DNA replication form a complex that has ATP-dependent DNA
topoisomerase activity. Proc. natn. Acad. Sci. U.S.A. , –.
S, J. T., S, S. & M, A. S. (). Vaccinia DNA topoisomerase, I :
Kinetic evidence for general acid-base catalysis and a conformational step.
Biochemistry , –.
S, A. & C, N. R. (). The intrinsic ATPase of DNA gyrase. J. biol.
Chem. , –.
S, A., H, N. P., B, P. O., P, C. L. & C, N. R. ().
Energy coupling in DNA gyrase and the mechanism of action of novobiocin. Proc.
natn. Acad. Sci. U.S.A. , –.
S, A., H, N. P. & C, N. R. (). DNA gyrase subunit
stoichiometry and the covalent attachment of subunit A to DNA during DNA
cleavage. Nucleic Acids Res. , –.
S, A., P, P. L., K, K. N. & C, N. R. (). Mechanism of
action of nalidixic acidL : purification of E. coli nalA gene product and its relationship
to DNA gyrase and a novel nicking-closing enzyme. Proc. natn. Acad. Sci. U.S.A. ,
–.
S, O. & V, A. (). Terminal stages of SV DNA replication
proceed via multiply intertwined catenated dimers. Cell , –.
S, O. & V, A. (). Arrest of segregation leads to accumulation of
highly intertwined catenated dimers : dissection of the final stages of SV DNA
replication. Cell , –.
T, J. K. & G, M. (). Characterization of the ATP binding site of
Escherichia coli DNA gyrase. Affinity labeling of Lys- and Lys- of the Bsubunit by pyridoxol «-diphospho-« adenosine. J. biol. Chem. , –.
T, J. K., B, A. D. & G, M. (). Slow interaction of «-adenylylbeta, gamma-immidodiphosphate with Escherichia coli DNA gyrase. Evidence for
cooperativity in nucleotide binding. J. biol. Chem. , –.
T, W., S, R. M., M, M. E. & H, C. (). Genetic analysis of the
gyrase A-like domain of DNA topoisomerase II of Saccharomyces cerevisiae. Genetics
, –.

J. C. Wang
T, A. P. & M, A. (). Probing the role of the ATP-operated clamp in
the strand-passage reaction of DNA gyrase. Nucleic Acids Res. , –.
T, B., B, C., L, K., A, A. H., S, B. S. &
W, O. (). Characterization of the interaction between topoisomerase
II and DNA by transcriptional footprinting. J. molec. Biol. , –.
T, Y.-C., K, K. & W, J. C. (). Covalent bonds between protein and
DNA. Formation of phosphotyrosine linkage between certain DNA topoisomerases
and DNA. J. biol. Chem. , –.
V, A., L, L., S, O., B, J., O, E., S, P.
& F, D. (). Cellular and SV chromatin : replication, segregation,
ubiquitination, nuclease-hypersensitivity sites, HMG-containing nucleosomes, and
heterochromatin-specific protein. Cold Spring Harbor Symp. Quant. Biol. , –.
W, J. C. (). Interaction between DNA and an Escherichia coli protein ω. J.
molec. Biol. , –.
W, J. C. (). Helical repeat of DNA in solution. Proc. natn. Acad. Sci. U.S.A.
, –.
W, J. C. (). DNA topoisomerases. In Nucleases (ed. R. Roberts and S. Linn),
pp. –. New York : Cold Spring Harbor Laboratory.
W, J. C. (). An introduction to DNA supercoiling and DNA topoisomerasecatalyzed linking number changes of supercoiled DNA. In DNA Topoisomerase and
Their Applications in Pharmacology, (ed. L. Liu), pp. –. San Diego : Academic
Press.
W, J. C. (). DNA topoisomerases. Ann. Rev. Biochem. , –.
W, J. C., G, R. I., J, K. J., K, K., K, L.,
K, M. L. & T, Y-C. (). DNA topoisomerases. In Mechanistic Studies
of DNA Replication and Genetic Recombination (ed. B. Alberts), pp. –. New
York : Academic Press.
W, J. C. & L, L. F. (). DNA topoisomerases : enzymes that catalyze the
concerted breaking and rejoining of DNA backbone bonds. In Molecular Genetics,
Part III, (ed. J. H. Taylor), pp. –. New York : Academic Press.
W, S. A. & C, N. R. (). Supercoiled DNA-directed knotting by
T topoisomerase. J. biol. Chem. , –.
W, J. D. & C, F. H. C. (). Genetic implications of the structure of
deoxyribonucleic acid. Nature , –.
W, R. & V, J. (). The cyclic coil helix and cyclic forms of polyoma viral
DNA. Proc. natn. Acad. Sci. U.S.A. , –.
W, D. B., D, G. J., D, E. J., M, A. & D, G. ().
Crystal structure of the N-terminal domain of DNA gyrase B protein. Nature ,
–.
Y, M. & S, R. (). Genetics of DNA topoisomerases. In DNA
Topology and its Biological Effects (ed. N. R. Cozzarelli and J. C. Wang), pp. –.
Cold Spring Harbor : Cold Spring Harbor Laboratory Press.
Z, E. L. & O, N. (). Eukaryotic topoisomerases recognize
nucleic acid topology by preferentially interacting with DNA crossovers. EMBO J. ,
–.