polymerase engineering: towards the encoded synthesis of ... · polymerase engineering: towards the...

13
Polymerase engineering: towards the encoded synthesis of unnatural biopolymers David Loakes and Philipp Holliger* Received (in Cambridge, UK) 18th February 2009, Accepted 6th May 2009 First published as an Advance Article on the web 6th July 2009 DOI: 10.1039/b903307f DNA is not only a repository of genetic information for life, it is also a unique polymer with remarkable properties: it associates according to well-defined rules, it can be assembled into diverse nanostructures of defined geometry, it can be evolved to bind ligands and catalyse chemical reactions and it can serve as a supramolecular scaffold to arrange chemical groups in space. However, its chemical makeup is rather uniform and the physicochemical properties of the four canonical bases only span a narrow range. Much wider chemical diversity is accessible through solid-phase synthesis but oligomers are limited to o100 nucleotides and variations in chemistry can usually not be replicated and thus are not amenable to evolution. Recent advances in nucleic acid chemistry and polymerase engineering promise to bring the synthesis, replication and ultimately evolution of nucleic acid polymers with greatly expanded chemical diversity within our reach. 1. Polymerase mechanisms of discrimination against non-cognate substrates DNA polymerases perform an astonishing feat of molecular recognition by incorporating the correct complementary base opposite its template counterpart, with error rates as low as 10 4 in the case of Thermus aquaticus DNA polymerase I (Taq), even in the absence of exonucleolytic proofreading. 1 The availability of high-resolution structures for different stages of the polymerase cycle together with elegant work using synthetic base analogues has begun to illuminate the subtle interplay of molecular forces and structural determi- nants on which the correct readout depends. Elegant work by Kool and others has shown that formation of the nascent base pair is highly sensitive to geometric aberrations but surprisingly less dependent on Watson–Crick hydrogen bonding (provided other forces, e.g. stacking, can compensate for the loss of binding energy). 2 The next step, extension of the incorporated base, appears to depend not only on a correct geometric fit and the formation of a conformationally stable 3 0 -pair but on the presence of cognate minor groove hydrogen bond donor and acceptor groups. Furthermore, distortions of cognate duplex geometry by the newly formed pair cause significant stalling, not just at the primer 3 0 -end, 3 but up to four bases post-extension. 4 Polymerase substrate specificity thus involves passage through three Medical Research Council, Laboratory of Molecular Biology, Hills Road, Cambridge, Cambridgeshire, UK CB2 0QH. E-mail: [email protected] David Loakes David Loakes was born in London, England. He received his GRSC graduate degree whilst working full-time for Wellcome, before moving to Sussex University (UK) for his DPhil under the super- vision of Prof. Douglas Young. There he continued his interest in medicinal chemistry working on the synthesis of cephalosporin analogues. Following this he carried out postdoctoral research with Dr Daniel M. Brown at the MRC Laboratory of Molecular Biology (Cambridge, UK), working on mutagenic nucleoside analogues. Since being appointed full-time at the MRC research interests involve nucleic acid chemistry, polymerase function and synthetic biology. Philipp Holliger Philipp Holliger was born in Berne, Switzerland, in 1966. He graduated in Bioorganic Chemistry and Molecular Biology with Prof. S. A. Benner at the ETH Zu ¨rich before moving to the CPE (Cambridge Centre for Protein Engineering) for his PhD and postdoctoral fellowship with Dr G. Winter, FRS. In 2000, he joined the MRC Laboratory of Molecular Biology and was appointed program leader in 2005. His research interests span the fields of chemical biology, synthetic biology, protein engineering and directed evolution. This journal is c The Royal Society of Chemistry 2009 Chem. Commun., 2009, 4619–4631 | 4619 FEATURE ARTICLE www.rsc.org/chemcomm | ChemComm Downloaded by Portland State University on 04 February 2012 Published on 06 July 2009 on http://pubs.rsc.org | doi:10.1039/B903307F View Online / Journal Homepage / Table of Contents for this issue

Upload: others

Post on 11-Jun-2020

11 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Polymerase engineering: towards the encoded synthesis of ... · Polymerase engineering: towards the encoded synthesis of unnatural biopolymers ... This journal is c The Royal Society

Polymerase engineering: towards the encoded synthesis

of unnatural biopolymers

David Loakes and Philipp Holliger*

Received (in Cambridge, UK) 18th February 2009, Accepted 6th May 2009

First published as an Advance Article on the web 6th July 2009

DOI: 10.1039/b903307f

DNA is not only a repository of genetic information for life, it is also a unique polymer with

remarkable properties: it associates according to well-defined rules, it can be assembled into

diverse nanostructures of defined geometry, it can be evolved to bind ligands and catalyse

chemical reactions and it can serve as a supramolecular scaffold to arrange chemical groups in

space. However, its chemical makeup is rather uniform and the physicochemical properties of the

four canonical bases only span a narrow range. Much wider chemical diversity is accessible

through solid-phase synthesis but oligomers are limited to o100 nucleotides and variations in

chemistry can usually not be replicated and thus are not amenable to evolution. Recent advances

in nucleic acid chemistry and polymerase engineering promise to bring the synthesis, replication

and ultimately evolution of nucleic acid polymers with greatly expanded chemical diversity within

our reach.

1. Polymerase mechanisms of discrimination

against non-cognate substrates

DNA polymerases perform an astonishing feat of molecular

recognition by incorporating the correct complementary base

opposite its template counterpart, with error rates as low as

10�4 in the case of Thermus aquaticus DNA polymerase I

(Taq), even in the absence of exonucleolytic proofreading.1

The availability of high-resolution structures for different

stages of the polymerase cycle together with elegant work

using synthetic base analogues has begun to illuminate the

subtle interplay of molecular forces and structural determi-

nants on which the correct readout depends.

Elegant work by Kool and others has shown that formation

of the nascent base pair is highly sensitive to geometric

aberrations but surprisingly less dependent on Watson–Crick

hydrogen bonding (provided other forces, e.g. stacking, can

compensate for the loss of binding energy).2 The next step,

extension of the incorporated base, appears to depend not

only on a correct geometric fit and the formation of a

conformationally stable 30-pair but on the presence of cognate

minor groove hydrogen bond donor and acceptor groups.

Furthermore, distortions of cognate duplex geometry by the

newly formed pair cause significant stalling, not just at the

primer 30-end,3 but up to four bases post-extension.4 Polymerase

substrate specificity thus involves passage through three

Medical Research Council, Laboratory of Molecular Biology,Hills Road, Cambridge, Cambridgeshire, UK CB2 0QH.E-mail: [email protected]

David Loakes

David Loakes was born inLondon, England. He receivedhis GRSC graduate degreewhilst working full-time forWellcome, before moving toSussex University (UK) forhis DPhil under the super-vision of Prof. Douglas Young.There he continued his interestin medicinal chemistry workingon the synthesis of cephalosporinanalogues. Following this hecarried out postdoctoral researchwith Dr Daniel M. Brown atthe MRC Laboratory ofMolecular Biology (Cambridge,

UK), working on mutagenic nucleoside analogues. Since beingappointed full-time at the MRC research interests involvenucleic acid chemistry, polymerase function and syntheticbiology.

Philipp Holliger

Philipp Holliger was born inBerne, Switzerland, in 1966.He graduated in BioorganicChemistry and MolecularBiology with Prof. S. A. Bennerat the ETH Zurich beforemoving to the CPE (CambridgeCentre for Protein Engineering)for his PhD and postdoctoralfellowship with Dr G. Winter,FRS. In 2000, he joined theMRC Laboratory of MolecularBiology and was appointedprogram leader in 2005. Hisresearch interests span thefields of chemical biology,synthetic biology, proteinengineering and directedevolution.

This journal is �c The Royal Society of Chemistry 2009 Chem. Commun., 2009, 4619–4631 | 4619

FEATURE ARTICLE www.rsc.org/chemcomm | ChemComm

Dow

nloa

ded

by P

ortla

nd S

tate

Uni

vers

ity o

n 04

Feb

ruar

y 20

12Pu

blis

hed

on 0

6 Ju

ly 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B90

3307

FView Online / Journal Homepage / Table of Contents for this issue

Page 2: Polymerase engineering: towards the encoded synthesis of ... · Polymerase engineering: towards the encoded synthesis of unnatural biopolymers ... This journal is c The Royal Society

different ‘‘checkpoints’’, involving a readout of distinct

molecular parameters, each of which differentially affect

incorporation, extension and replication.

This ‘‘triple sieve’’ mechanism poses stringent requirements on

the properties of unnatural nucleotide analogues, e.g. hydro-

phobic base analogues (HBAs) or sugar analogues. While many

can be incorporated with good efficiency and specificity due to

energetically favourable interactions, the resulting unnatural base

pairs are often only very inefficiently extended when at the primer

30-end or replicated when acting as a templating base.

While the general features of the specificity mechanisms

described above apply to most polymerases, the relative

importance of the different molecular parameters can vary

significantly between different polymerase families and even

among members of the same family. For example, poly-

merases of the polA family, notably Klenow, are able to

efficiently form base pairs without Watson–Crick hydrogen

bonding (e.g. between difluorotoluene (1) and 4-methyl-

benzimidazole (2)) provided the nascent base pair displays

correct geometry, while the same unnatural bases are utilised

with much reduced efficiency by members of the polY family

of translesion polymerases.5

A striking example of the sophistication of DNA polymerase

molecular discrimination against non-cognate substrates is the

exclusion of non-cognate sugar structures. Ribonucleotides differ

from deoxyribonucleotides only by a 20-hydroxyl group on the

ribose sugar. The addition of an additional hydroxyl group on

the ribose sugar not only renders the product susceptible to

degradation by RNases but also alters the structure of DNA,

which may interfere with the activity of DNA modifying

enzymes. Since the concentration of ribonucleotide triphosphates

(NTPs) within the cell (Escherichia coli) is at about 10-fold higher

than that of deoxyribonucleotide triphosphates (dNTPs),6 DNA

polymerases must be effective at excluding NTPs in order to

maintain the integrity of the DNA genome. Indeed, in E. coli

DNA polymerase I NTPs are excluded from incorporation by a

factor of 104–105-fold, and this stringent discrimination is mainly

encoded in a single residue.7 This ‘‘steric gate’’ residue is

positioned such that it occupies the space where the 20-OH of

an incoming NTP would be placed, thereby sterically excluding

NTPs from the active site.

2. Incorporation and replication of modified

nucleic acids

Great strides have been made in the enzymatic synthesis and

replication of nucleic acids comprising unnatural chemical

modifications simply through elegant chemistry, judicious

use of polymerase and optimisation of reaction conditions

and additives. In general, the incorporation and replication of

modified nucleotide substrates by polymerases (natural or

engineered) comprises a vast body of literature. The work

discussed here therefore represents a blend of the personal

preferences and interests of the authors and is not meant to be

exhaustive.

2.1 Functionalised DNA

The most widely studied group of nucleoside analogues are

those comprising base modifications. Substitutions at the C5

position of either dU or dC, placing the substituent in the

major groove of the double helix, are especially well tolerated

by polymerases and a great variety of C5 substitutions have

been explored and found to be compatible with at least

sporadic incorporation.

Typically, pyrimidines are modified at the C5 position, and

purines at the C7 position (see in particular the extensive

works of Seela et al. for the latter). The number of such

triphosphate derivatives that have been described are too vast

to detail here, but many are dealt with in recent reviews.8–10

The triphosphate derivatives of C5-modified pyrimidines and

C7-modified purines are generally very well tolerated by a

variety of polymerases (primarily DNA polymerases and

reverse transcriptases), thus allowing for the introduction of

additional functional groups into nucleic acids for use in, for

example, aptamers11,12 and detection.13,14 Purines have also

been modified at the C8 position, a modification that generally

induces a syn conformation into the nucleoside which has

consequences for base pairing. As a result, not all C8-modified

purines are good polymerase substrates.15,16

Of particular interest in this group of compounds are

fluorescent dye-labelled nucleotides, as they are an essential

component of several core technologies of modern biology

including sequencing, microarrays and fluorescent in situ

hybridisation (FISH). These and other substituted nucleotides

provide the opportunity to utilise DNA as a molecular

scaffold, i.e. to synthesise nucleic acid polymers displaying

substituents at defined positions along the double helix.

However, despite systematic efforts to optimise substituent

and linker chemistry, many C5-substituted nucleotides, in

particular those bearing large hydrophobic dye molecules,

remain poor polymerase substrates, presumably due to steric

clashes between the large substituents and the polymerase.

Efficient replication of such polymers presents further

challenges as the bulky substituents in the template strand

cause pausing or abortive synthesis.17 Nevertheless, several

groups have reported some success with synthesising nucleic

acid polymers with several or all of the bases bearing

substituents.18–22 The most advanced of these systems was

described by Jager and Famulok, who achieved successful

PCR amplification of short (o100 bp) fragments with a

number of nucleotide analogues functionalised with small

non-fluorescent groups (functionalised DNA (fDNA)) using

Pwo polymerase. Such fDNA was found to stain poorly with

ethidium bromide and displayed a notable shift in the CD

spectrum suggestive of Z-DNA conformation.23,24 Some

reports also briefly commented on the poor solubility in

aqueous solvents of such polymers.18,23 While optimisation

of buffer conditions allowed significant improvements in the

synthesis of fDNA, even greater improvements should be

achievable through the engineering or evolution of polymerases

tailor-made for the synthesis of DNA-polymers bearing

substituents.

2.2 Hydrophobic base analogues

A decade ago Kool and co-workers turned the long-held belief

of nucleoside–polymerase recognition on its head by

demonstrating that certain nucleoside isosteres were

4620 | Chem. Commun., 2009, 4619–4631 This journal is �c The Royal Society of Chemistry 2009

Dow

nloa

ded

by P

ortla

nd S

tate

Uni

vers

ity o

n 04

Feb

ruar

y 20

12Pu

blis

hed

on 0

6 Ju

ly 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B90

3307

F

View Online

Page 3: Polymerase engineering: towards the encoded synthesis of ... · Polymerase engineering: towards the encoded synthesis of unnatural biopolymers ... This journal is c The Royal Society

specifically and efficiently incorporated into duplex DNA

without Watson–Crick hydrogen bonding.25,26 Working with

the thymine isostere difluorotoluene (1, X = F) it was shown

that E. coli DNA polymerase I (Klenow fragment, KF)

inserted the 50-triphosphate of (1, X = F) onto the growing

primer strand specifically opposite dA, and with an efficiency

approaching that of dTTP. Kool argued that hydrogen bond

recognition within the polymerase active site was not essential

provided there was a good geometric fit for the incoming

nucleotide. Kool and co-workers went on to describe the

adenine mimics (2, X = CH)27 and (2, X = N).28 The

triphosphates of both of these analogues are inserted opposite

dT or (1) by KF specifically and with good efficiency. When

these analogues are used as a triphosphate they extend the

primer by one nucleotide, whereas when present in the template

full-length product can be obtained. The analogue (2, X = N)

was used to assess the requirements for minor groove inter-

actions, the nitrogen being equivalent to N3 of adenine. When

(2, X = N) is present in the template full-length product is

readily achieved with KF, whereas (2, X = CH) is much less

efficiently extended, thus demonstrating that minor groove

H-bonding interactions do play an important role in

polymerase extension.

These analogues also offered an opportunity to explore

differences in minor groove H-bonding requirements among

different polymerases. Among the polymerases assayed for their

ability to bypass these non-polar isosteres of dA and dT,

KF(exo–), Taq and HIV-RT were all able to extend the

(1):(2, X = N) base pair, but not the (1):(2, X = CH) base pair,

indicating the requirement for minor groove hydrogen bond

recognition. However, Pola, Polb and T7 DNA polymerases did

not extend the non-hydrogen-bonding analogues.29,30 NMR

studies using 13C-methionine-labelled Polb suggested that the

thymine isostere (1) disrupts closure of the polymerase ternary

complex, which presumably accounts for why this polymerase is

unable to utilise the isostere.31

Single-turnover kinetic analysis with KF(exo–) found that the

presence of (1) decreased catalytic efficiency 30-fold, whilst when

paired with adenine isosteres catalytic efficiency is decreased

103-fold.32 As is the case with most unnatural base analogues,

incorporation and extension is dependent on polymerases lacking

proofreading ability (exo–). When KF(exo+) polymerase is

used, it will remove a terminal dT:(2, X = N) base pair

with the same efficiency as a mismatch, but interestingly the

(1):(2, X = CH) pair is edited much more slowly.33 The

analogues (1) and (2, X = N) when introduced into E. coli as

part of a modified phage genome are bypassed with moderate

efficiency, and with very high efficiency under SOS-induction.

Remarkably, each of the isosteres were read as their equivalent

natural base.34

Kool et al. have further explored the geometric fit of

thymine isosteres by assessing the polymerase specificity of

size analogues of (1) using each of the analogues where X is H,

F, Cl, Br and I and their ability to be incorporated opposite

dA by KF(exo–).35,36 These analogues present a series in which

the base size increases incrementally over the range of 1 A.

Surprisingly, it was the dichloro analogue, which is larger than

the natural thymine base, that was incorporated with the

greatest efficiency. Similar results were observed with the

translesion polymerases Dpo437 and E. coli polymerases II

and IV,38 though in the case of PolIV analogue incorporation

efficiency was much reduced. Using both the deoxyribosyl and

ribosyl analogues of (1, X = H, F, Cl, Br, I) the DNA- and

RNA-dependent synthesis by HIV-RT was explored. DNA

synthesis follows the same pattern as described above, whilst

with the RNA analogues HIV-RT only discriminates against

the smaller analogues.39,40

Isosteric base analogues have also been used to detect the

presence of DNA lesions such as abasic sites, which are

amongst the most common forms of DNA damage. Matray

and Kool et al. demonstrated that the pyrene C-nucleoside (3)

will form a specific ‘‘base pair’’ with an abasic site (4) in duplex

DNA.41,42 The pyrene base occupies a similar space to an entire

Watson–Crick base pair (i.e. is an isostere of the missing base

pair), as shown by an NMR structure of a DNA duplex

containing the (3):(4) pair.43 The 50-triphosphate derivative of

(3) is specifically incorporated opposite to an abasic site with

efficiency similar to that of a natural dNTP by both KF(exo–),

and yeast polZ.44 In the latter case, the efficiency of incorpora-

tion of pyrene exceeds that of the natural dNTP. It is suggested

that the preference for incorporation of pyrene is due to the

highly favourable stacking energy of the large aromatic pyrene

ring system. However, once pyrene has been incorporated into

the primer strand further extension is very inefficient. The large

pyrene ring has also been used as a mechanistic probe for DNA

polymerase bypass of other DNA lesions. Incorporation of

pyrene opposite various template thymine-dimer analogues

occurred preferentially opposite the 30-T of the dimer (except

in the case of trans, syn-photodimer) suggesting that, in this

case, the 30-thymine residue is in a conformation that is

unfavourable to Watson–Crick base pairing.45

In addition to pyrene, other large aromatic base analogues

have been shown to be efficiently incorporated opposite abasic

sites by various DNA polymerases. Berdis and co-workers

have investigated a number of indole nucleoside derivatives

and shown them to be efficiently and specifically incorporated

opposite an abasic site by T4 DNA polymerase and KF with

up B100-fold higher efficiency than dAMP.46–49 Once again,

once incorporated further extension beyond the indole is not

This journal is �c The Royal Society of Chemistry 2009 Chem. Commun., 2009, 4619–4631 | 4621

Dow

nloa

ded

by P

ortla

nd S

tate

Uni

vers

ity o

n 04

Feb

ruar

y 20

12Pu

blis

hed

on 0

6 Ju

ly 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B90

3307

F

View Online

Page 4: Polymerase engineering: towards the encoded synthesis of ... · Polymerase engineering: towards the encoded synthesis of unnatural biopolymers ... This journal is c The Royal Society

observed. The most efficient analogues were found to be

5-nitro- and 5-phenyl-indole derivatives,50,51 whilst larger52

and less planar53 derivatives were incorporated less efficiently,

as were analogues where p-electron density or p-electronsurface area are reduced compared to 5-nitroindole.54 All of

these support the notion that the energy gain from favourable

stacking interactions with a large p-electron surface area can

compensate for imperfect shape complementarity in the

incorporation step.

2.3 New base pairs

In recent years there has been much interest in identifying

alternative base-pairing systems. A potential advantage of

such a system is that it may allow an expansion of the genetic

code, creating new specific codons for the incorporation of

unnatural amino acids into proteins. Many different

base-pairing systems have been examined. Work in this area

was pioneered by Benner et al., who first examined base pairs

with unnatural Watson–Crick hydrogen-bonding patterns

such as isoguanine (isodG) and isocytosine (isodC).55 This

modified base pair does not impair the duplex structure, and

the nucleotides are substrates for DNA polymerases.

However, isodG does exist as tautomers and as such can form

base pairs with thymine. The use of 7-deazaisodG constrains

the analogue into a single tautomeric state and thus improves

specificity.56 Alternatively, use of 2-thiothymidine, which pairs

only inefficiently with the enol tautomer of isodG, leads to

efficient use of isodC and isodG in PCR reactions, with 98%

retention of the isodC:isodG base pair per PCR cycle.57 The

analogues isodC and isodG have found application in PCR58

and sequencing reactions,59 where they are good polymerase

substrates, though at slightly reduced efficiency compared with

the native nucleotides.

Apart from isodG:isodC Benner has described a family of

nucleoside analogues covering many possible permutations of

the Watson–Crick hydrogen-bonding patterns, some of which

may be suitable as alternative base pairs.60 For example, the

purine analogue dP (5) forms a specific base pair with the

pyrimidine analogue dZ (6), designed to present electron

density in the minor groove to facilitate specific polymerase

recognition. Each of the triphosphates of dP and dZ are

substrates for both A and B family DNA polymerases,

forming specific base pairs with better than 95% retention of

the dZ:dP base pair per cycle of PCR.61

Hirao et al. have designed a number of unnatural

base-pairing systems based on a combination of hydrogen

bond recognition and shape complementarity. The base pairs

consist of a modified adenine and a pseudobase similar in size

to a pyrimidine base. Each base pair occupies about the same

space as a cognate base pair, but they form specific base pairs

with each other while discriminating against pairing with any

of the natural bases due to steric exclusion. For example, the

base pair between the 2-thienylpurine (7, R = NH2, X1 = CH)

and the oxo(1H)pyridine (8, X2 = H) forms specifically as

there would be a steric clash between either the O4 of thymine

or N4 of cytosine and the thiophene ring. In addition, the base

pair is further stabilised by the presence of two hydrogen

bonds, which also enhance selectivity. These nucleobases

form a specific base pair and are a substrate pair for DNA

polymerases.62 Imidazolin-2-one also serves as a very good

partner, with the 2-thienylpurine (7, R = NH2, X1 = CH)

being a better DNA polymerase substrate than the nucleoside

derivative of pyridin-2-one.63 An alternative base pair is one in

which the thienyl ring is replaced by a methyl group (7, R = H)

that forms a specific base pair with difluorotoluene (1) but

forms a superior base pair, both in terms of oligonucleotide

hybridisation and polymerase specificity with the nucleoside

derivative of pyrrole-2-carbaldehyde.64,65 A further base pair,

the imidazo[4,5-b]pyridine derivative of (7, R=NH2, X1 = CH),

forms a specific base pair with the nucleoside derivative of

2-nitropyrrole. PCR reactions using the triphosphate

derivatives of these analogues resulted in B1% of the modified

base pair incorporated into the PCR product after just

20 cycles,66 and even better efficiency obtained using a

pyridoimidazole pairing with a nitropyrrole (see Fig. 2).67

These analogues are also good substrates for RNA

polymerases. Using T7 RNA polymerase the triphosphate

derivative of the ribonucleoside of (8, X2 = H or I or

arylalkynyl) is efficiently incorporated opposite the thienyl-

purine derivatives (7) in a DNA template.68,69 Furthermore,

the RNA transcripts were substrates for translation using an

E. coli cell free system to produce proteins containing

3-chlorotyrosine.70,71 Analogues of the ribonucleotide of (8)

have been incorporated into RNA transcripts bearing

fluorophores,72,73 biotin74 and iodine useful for photo-

crosslinking to proteins.75

Rappaport described an alternative three-base pair set of

nucleotides suitable for replication by T7 DNA polymerase.76

The three base pairs used are A:T, hypoxanthine:cytosine and

6-thiopurine:5-methyl-2-pyrimidinone. Error frequency rates

for these three base pairs were of the order 10�4–10�6, with

the major mismatches occurring between hypoxanthine and

thymine.

Another strategy for orthogonal base pairing through steric

exclusion has been explored by Kool et al. They devised a set

of nucleotides in which a benzene ring has been inserted

between the sugar and the nucleobase, termed xDNA77,78

4622 | Chem. Commun., 2009, 4619–4631 This journal is �c The Royal Society of Chemistry 2009

Dow

nloa

ded

by P

ortla

nd S

tate

Uni

vers

ity o

n 04

Feb

ruar

y 20

12Pu

blis

hed

on 0

6 Ju

ly 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B90

3307

F

View Online

Page 5: Polymerase engineering: towards the encoded synthesis of ... · Polymerase engineering: towards the encoded synthesis of unnatural biopolymers ... This journal is c The Royal Society

and yDNA79 (depending on the position of the phenyl ring),

and the effect of the additional benzene ring is to widen the

duplex structure. Initial studies to determine whether yDNA

can transmit genetic information showed that DNA

polymerases can insert the correct nucleotide, but at a lower

specificity than for normal DNA.80

The group of Romesberg have investigated a very large

range of chemical space in the search for a third base pair

based on hydrophobic base analogues. A large body of

analogues have been investigated to systematically probe

the steric requirements for polymerase recognition, and for

formation of homo- and hetero-pairs by the hydrophobic base

analogues. The range of chemical space investigated

includes substituted phenyl,81,82 substituted pyridyl83,84 and

isocarbostyril,85–87 as well as pyridones,88 azaindoles89 and

other heterocycles.90 More recently Romesberg et al. have

developed a novel base pair (see Fig. 2) which is replicated by

Klenow fragment (exo–) DNA polymerases with an efficiency

approaching that of a native base pair.91

2.4 Universal bases

Universal base analogues92,93 are generally characterised as

non-hydrogen bonding, hydrophobic base analogues (HBAs)

that pair indiscriminately with all the native nucleobases.94

Generally they are destabilising in a duplex due to loss of

hydrogen bonding and solvation, though this is compensated

for by stacking and hydrophobic interactions, the strength of

which is related to the size of the HBA. Largely due to their

lack of hydrogen bond functionality (but see isosteres above)

universal bases are very poorly recognised by polymerases.95,96

However, Romesberg et al. have shown that various isocarbo-

styril analogues, such as MICS, behave as efficient universal

base analogues, not only in hybridisation terms but also being

recognised indiscriminately by KF, albeit at reduced catalytic

efficiency.85 The universal bases benzimidazole, 5- and

6-nitrobenzimidazole and 5-nitroindole, are incorporated as

their 50-triphosphates opposite native DNA bases by KF and

human Pola 4000-fold more efficiently than a mismatch, with

Pola preferentially incorporating the universal bases opposite

pyrimidines and KF opposite purines.97 However, whilst these

analogues were shown to bind to Moloney murine leukaemia

virus (MMLV)-reverse transcriptase (RT) they were not

incorporated by it. Many HBAs can be incorporated into

DNA under rather forcing conditions, but once incorporated

they are blocking lesions. When present in a template strand

dAMP is preferentially incorporated opposite the universal

base, with polymerases presumably following the A-rule.98

The exception to this is when incorporated opposite a template

lesion, such as an abasic site (see section on isosteres above).

The analogue 3-nitropyrrole has also been shown to be a

substrate for the poliovirus RNA polymerase, preferentially

incorporating it opposite A and U, though at 100-fold reduced

efficiency compared to cognate pairs.99

2.5 Modified sugar–phosphate backbone

While there is a growing body of circumstantial evidence to

support the idea of a primordial RNA-based genetic system, a

plausible and efficient route for the prebiotic synthesis of

ribose has been lacking. This has spurred various groups to

investigate whether simpler sugar, or sugar-substitutes, can be

used to encode genetic information.

Meggers and co-workers have shown that the deoxyribose

ring may be substituted for a simple glycerol unit,100 GNA (9),

which forms stable duplexes with either GNA or (S)-, but not

(R)-GNA, with RNA,101 and does not form a stable duplex

with DNA. Szostack et al. have examined a variety of A and B

family DNA polymerases as well as reverse transcriptases to

determine whether GNA could be replicated as a genetic

system. Using (S)-glycerol NTPs it was shown that

Therminator DNA polymerase was able to perform +1

extension of a DNA primer–template duplex with an efficiency

comparable to incorporation of a native dNTP.102 Furthermore,

Bst DNA polymerase was able to fully extend across a short

region of template GNA, as was an engineered MMLV-RT,

albeit to a lesser extent.103 The fact that Bst DNA polymerase

is able to synthesise DNA from a GNA template suggests that

the polymerase is able to sufficiently stabilise the transient

DNA–GNA duplex for sequence readout from GNA to be

possible.

Another pre-RNA candidate that has been proposed is the

mixed acetal aminal of formyl glycerol, FNA, (10). The

triphosphate derivatives of FNA have been examined as

substrates by DNA polymerases and reverse transcriptases

on a DNA template. Unlike GNA, both enantiomers were

found to be substrates for a variety of polymerases, notably

those with lower fidelity, though the (R)-enantiomer was

preferentially incorporated over the ‘natural’ (S)-isomer.104

One of the most intriguing pre-RNA nucleic acids investi-

gated to date involves an aldo sugar containing only four

carbon atoms, (L)-a-threofuranosyl nucleic acid (TNA), (11),

in which the phosphodiester bonds are attached to the 20- and

30-positions of the threofuranose ring (30 - 20).105 TNA

duplexes show similar thermal stabilities to that of DNA

and RNA, and TNA can also form cross-pairs with either

DNA or RNA. X-Ray crystallography has shown that TNA

can be readily accommodated into B-form duplex DNA,106

but that it is a better mimic for A-form nucleic acids and

therefore forms better cross-pairs with RNA.107 Cytosine

TNA templates specifically direct the non-enzymatic

incorporation of activated rGMP leading to RNA products.108

Szostak and co-workers have examined in some detail the

templating properties of TNA and the incorporation of TNA

triphosphates by a variety of DNA polymerases and reverse

transcriptases. They have shown that certain polymerases are

capable of copying limited regions of template TNA, despite

the differences in the sugar–phosphate backbone between

DNA and TNA.109 Significant improvements in templating

ability were observed in the presence of Mn(II) ions. In order

to assess the effect of Mn(II) ions on fidelity, polymerase

reaction products were sequenced and the error rate compared

with template DNA and template TNA. It was shown that

using the DNA polymerase Sequenase the error rate for TNA

was 1.5-fold higher than for DNA, whilst with the MMLV-RT

Superscript II the TNA error rate was 3-fold increased. The

authors have suggested that evolution of a TNA-dependent

DNA polymerase could lead to improved efficiency, and thus

This journal is �c The Royal Society of Chemistry 2009 Chem. Commun., 2009, 4619–4631 | 4623

Dow

nloa

ded

by P

ortla

nd S

tate

Uni

vers

ity o

n 04

Feb

ruar

y 20

12Pu

blis

hed

on 0

6 Ju

ly 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B90

3307

F

View Online

Page 6: Polymerase engineering: towards the encoded synthesis of ... · Polymerase engineering: towards the encoded synthesis of unnatural biopolymers ... This journal is c The Royal Society

lead to applications in which TNA was used as genetic

information.

The kinetics of incorporation of TNA triphosphates by

DeepVent(exo–) and Bst and HIV-RT have been reported.

Whilst the kcat and Km (or Vmax/Km) values differ, the trend

is the same, which is that the Km increases with TNA

triphosphates, whilst the kcat decreases.110,111 Interestingly,

the kinetics of incorporation of a second TNA triphosphate

shows a further increase in Km and decrease in Vmax/Km.111

However, the polymerases retain fidelity for the incorporation

of TNA triphosphates.

The polymerase that has been most widely studied for the

incorporation of modified nucleotides and oligonucleotides is

Therminator DNA polymerase, an exo–variant of 91N DNA

polymerase, containing an A485L mutation.112,113 It has also

been shown to be one of the most effective DNA polymerases

for the incorporation of TNA triphosphates. The kinetics of

incorporation are much the same as described above for

DeepVent exo–DNA polymerase, in that the Km is higher

and the kcat reduced compared to dNTPs. Furthermore, when

the primer strand is modified to contain five TNA residues at

its 30-end there is no further significant decrease in efficiency of

incorporation of TNA triphosphates.114 Addition of Mn(II)

ions further improves the kinetics of incorporation as might be

anticipated. Therminator DNA polymerase has been used to

generate TNA sequences as long as 80 nt long. Error rates

were raised compared to the use of normal dNTPs, but low

enough, in principle, for use in SELEX.115,116 The authors

suggest that in vitro evolution could be used to identify a

mutant polymerase capable of faithful production of TNA

oligomers.

A class of modified sugars with interesting properties

that has been widely used for SELEX applications are those

with 20-modifications. Among these 20-F, 20-NH2 and

especially 20-OCH3 have found applications in the aptamer

field as they are reasonably well accepted by T7 RNA

polymerase and by reverse transcriptases and greatly stabilise

aptamers to nuclease degradation.117,118 Indeed, using an

optimised protocol of transcription using the Y693F/H784A

double mutant of T7 RNA polymerase, Burmeister et al.

described the selection of an anti-VEGF aptamer composed

entirely of 20-OCH3 nucleotides, which was stable in plasma

for 496 h and survived autoclaving at 125 1C.119

Locked nucleic acids (LNA) were first described by

Imanishi120 and Wengel,121 and contain a methylene bridge

between 20-O and C40 of the ribose sugar. The presence of the

bridging methylene group locks the sugar into a 30-endo

conformation, reducing the conformational flexibility of the

ring. LNA exhibits enhanced binding towards RNA and

DNA, and incorporation of LNA into DNA induces A-like

conformations. Various DNA polymerases have been assessed

for their ability to either bypass LNA within a DNA template,

or to incorporate LNA triphosphates. When present in a DNA

template LNA is bypassed provided the LNA residues are

distributed along the strand.122 The most efficient polymerases

were found to be KOD(exo–) and Vent(exo–), though

efficiency of replication was reduced compared to native

DNA. LNA triphosphates may be incorporated into DNA

using Phusion and Therminator DNA polymerases,123–125

again providing there are not too many consecutive

substitutions. Pfu exo–DNA polymerase will incorporate

LNA triphosphates, but is unable to further extend the

product.

Another class of modified sugars that have been studied are

those bearing modification at C40, which have been extensively

examined by Marx and co-workers. Such analogues were

examined to investigate steric effects in the polymerase

active site using various DNA polymerases and reverse

transcriptases.126 The effect of a bulky group at C40 is different

between polymerases, and is dependent to some extent on the

sequence context. The C40-methyl derivative is accepted the

most readily, although kinetics of incorporation with HIV-RT

showed reduced rate of incorporation.127 However, increasing

the size of the alkyl substituent to larger than methyl does

cause the rate of incorporation to reduce markedly, but

increases specificity.128,129

The final class of sugar-modified nucleoside to be considered

here are those based on pyrofuranose. A study of base-pairing

properties of a variety of six-membered carbohydrate mimics

with RNA showed that RNA can cross-pair with a broad

range of nucleic acid structures.130 Amongst those analogues

studied anhydrohexitol nucleic acid (HNA) (12), a homologue

of DNA, was found to be the most stable, forming the most

stable homo duplex in a polyA–polyT system, as well as

forming the most stable cross-pair with RNA (more stable

than RNA:RNA). Crystal structures of HNA:HNA131 and of

HNA:RNA132 reveal that it adopts an A-form duplex, though

with more pronounced major groove, and a more reinforced

backbone hydration, which may account for the enhanced

stability of such duplexes. HNA oligonucleotides have been

used in aptamers,133 and in antisense,134 but are not substrates

for RNase H. Various DNA polymerases were able to catalyse

the incorporation of a single HNA triphosphate, but only

DeepVent(exo–) DNA polymerase could catalyse incorpora-

tion of more than one substitution.135 Up to six substitutions

could be achieved using DeepVent at high enzyme concentra-

tion, though under such conditions fidelity of incorporation

was reduced when more than two HNA triphosphates were

included in the reaction. The HNA triphosphates have also

been tested as substrates for various reverse transcriptases

(RTs).136 Whilst all RTs tested were able to incorporate a

single HNA triphosphate, only the dimeric RTs RAV2 and

HIV-RT were able to incorporate more than one. A series of

mutant HIV-RTs were assayed and the M184V mutant of

HIV-RT was found to be the most efficient at making multiple

incorporations of HNA. HNA incorporated into the AUG

start codon and the UUC codon in mRNA did not prevent

translation.137 Another interesting six-membered ring

replacement for ribose is the cyclohexene nucleoside (CeNA),

(13), which like HNA forms exceptionally stable duplexes with

4624 | Chem. Commun., 2009, 4619–4631 This journal is �c The Royal Society of Chemistry 2009

Dow

nloa

ded

by P

ortla

nd S

tate

Uni

vers

ity o

n 04

Feb

ruar

y 20

12Pu

blis

hed

on 0

6 Ju

ly 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B90

3307

F

View Online

Page 7: Polymerase engineering: towards the encoded synthesis of ... · Polymerase engineering: towards the encoded synthesis of unnatural biopolymers ... This journal is c The Royal Society

RNA. CeNA has also been used in antisense, but unlike HNA,

CeNA oligomers are substrates for RNase H.138,139

3. Polymerase engineering and evolution

As detailed in the previous section, some mutant polymerases

show dramatic improvements in their capacity to incorporate,

extend and replicate a range of unnatural nucleotide analogues.

However, in many cases further improvements would be

desirable to enable processive synthesis, replication and evolution

of modified nucleic acids. This has provided a strong incentive

to engineer polymerase function by design, screening and

directed evolution.

Polymerase engineering started with simple deletion of

accessory domains of the wild-type polymerase, which sometimes

yielded polymerases with improved properties. For example,

variants of E. coli DNA polymerase I (Klenow) and Taq

polymerase (Stoffel fragment and Klentaq) have been generated

by full or partial deletion of the N-terminal 50–30 exonuclease

domain and show improved stability and fidelity. Mutation of

two essential aspartates in the 30–50 exonuclease domain is

commonly used to generate proofreading defective variants

(e.g.KOD(exo–)). These exo-variants are often the most useful

for the incorporation of unnatural nucleotide substrates, as the

proofreading 30–50 exonuclease will efficiently remove most

unnatural nucleotides. A particularly interesting mutation turned

out to be A485L in the 91N exo–polymerase (Therminator).

Originally designed to allow efficient incorporation of acyclic

nucleotides for terminator sequencing,112,113 it has turned out

to improve the incorporation of a wide spectrum of unnatural

nucleotide substrates (e.g. TNA, see above).

In some cases careful inspection of high-resolution

structures has allowed the rational design of mutants with

improved properties. Examples include mutants of Taq

polymerase with improved properties of dideoxynucleotide

incorporation,140 an increased capability to incorporate

ribonucleotides,141,142 reduced pausing143 as well as altered

fidelity.144 Grafting of new domains has been used to modify

polymerase properties, notably processivity. Insertion of the

thioredoxin binding loops of T7 or T3 DNA polymerase into

the E. coli Pol I Klenow fragment or Taq polymerase,

respectively, leads to an increase in processivity.145,146 Fusion

of the Sulfolobus solfataricus DNA binding protein Sso7d to

Taq and Pfu DNA polymerases also improved processivity in

both enzymes.147 Insertion of AviTagt peptide loops either

side of the primer–template duplex binding cleft of 91N DNA

polymerase allowed enzymatic biotinylation and assembly

with streptavidin into an artificial ‘‘processivity complex’’.

These complexes tethered to a solid surface lead to an

impressive gain in processivity from ca. 20 nt in the unmodified

polymerase to several thousand in the modified polymerase.148

Loeb et al. pioneered the use of in vivo screening by genetic

complementation of E. coli (polA12ts, recA718) (Fig. 1a) with

Taq polymerase, HIV reverse transcriptase (RT) and human

Polb.149–151 Complementation at non-permissive temperatures

was used to probe inter alia the mutability of the polymerase

active site152 and screening of the selected mutants has yielded

polymerase variants with a range of properties including

an increased capability to incorporate ribonucleotides141

or increased fidelity.153 Genetic complementation of

Saccharomyces cervisiae RAD30, RAD52 has been used to

screen variants of human polZ and yielded a polZmutant with

increased activity.154

In vitro screening for polymerase function has long been

used, for example, in drug discovery notably of antiviral

compounds such as nucleotide and non-nucleotide inhibitors

of HIV-1 RT.155 A number of clever high-throughput assay

formats have developed both as semi-quantitative continuous

and end-point assays. For example, several groups have used a

scintillation proximity assay in which a biotinylated primer is

extended with tritium-labelled nucleotide triphosphates and

immobilized on streptavidin coated scintillant beads. Other

assays take advantage of the increased fluorescence intensity

of some intercalating dyes, when bound to the double-

stranded DNA (dsDNA) synthesised by an active polymerase

Fig. 1 Strategies for the selection of polymerase repertoires. (a) Complementation of an E. coli strain with a temperature-sensitive mutant of

DNA polymerase I (polAts). Only active polymerases (Pol1) support colony growth at the non-permissive temperature (37 1C). (b) Proximal display

of polymerase and primer–template duplex substrate on filamentous phage. Primer extension leads to incorporation of a biotinylated nucleotide

which tags the phage displaying the active polymerase (Pol1) for selection. (c) Selection of polymerase function by compartmentalised self-

replication within the aqueous droplets of a water-in-oil emulsion (CSR). Only polymerases active under the selection conditions (Pol1) can

replicate their own encoding gene and generate ‘‘offspring’’, while those that are inactive (Pol2) cannot self-replicate and disappear from the gene

pool.

This journal is �c The Royal Society of Chemistry 2009 Chem. Commun., 2009, 4619–4631 | 4625

Dow

nloa

ded

by P

ortla

nd S

tate

Uni

vers

ity o

n 04

Feb

ruar

y 20

12Pu

blis

hed

on 0

6 Ju

ly 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B90

3307

F

View Online

Page 8: Polymerase engineering: towards the encoded synthesis of ... · Polymerase engineering: towards the encoded synthesis of unnatural biopolymers ... This journal is c The Royal Society

or detect the incorporation a tagged nucleotide substrate in an

ELISA format.156 Elegant continuous assays have also been

described either based on the change in fluorescence of

single-stranded DNA binding displaced upon primer extension

or on fluorescence resonance energy transfer (FRET) between

acceptor and donor fluorophores located on primer and

template strand or as part of a template hairpin structure.157

Marx et al. have adapted high-throughput liquid-handling to

perform in vitro screening of polymerase function by assaying

duplex DNA synthesis by this hairpin assay157 or simply by

monitoring the fluorescence increase of the intercalating dye

SYBR Green (in primer extension reactions or qPCR). These

in vitro screening methods were used to identify variants of

Taq polymerase with improved mismatch discrimination,158

variants of KlenTaq polymerase with substantial RT

activity159 or with an ability to amplify DNA damaged by

UV radiation.160

While in vivo screening by genetic complementation is a

highly effective and convenient way to identify active variants

from a large library of polymerase mutants, its scope is limited

by the fact that the only property that is screened for is basal

polymerase activity. In vitro screening is potentially much

more flexible as, in principle, any polymerase property can

be optimised, provided it is amenable to a high-throughput

assay. On the other hand, most of the assays rely on the

detection of double-stranded DNA or on the disruption of

template secondary structure by the strand-displacement

activity of the polymerase. In many ways, such assays are

not ideal for the discovery of polymerases with an enhanced

ability to incorporate and replicate unnatural nucleotide

substrates as polymerase activity may be too weak for

significant amounts of dsDNA to be synthesised or the

chemistry of the modified substrates may interfere with dye

binding or fluorescence. Other pitfalls include false positives

e.g. polymerases with an enhanced ability to synthesise

dsDNA from just three nucleotides160 (e.g. avoiding the

unnatural substrate), to preferentially misincorporate tagged

nucleotides or with an enhanced ability to melt template

hairpins. Finally, in vitro screening methods are currently

limited to polymerase libraries an order of magnitude smaller

than those used in genetic complementation.

Larger polymerase repertoires can, in principle, be processed

by the use of selection technologies. One of the most

productive selection methods has been phage display, in

particular for the identification of specific peptide and protein

interactions. Both Jestin and Romesberg have adapted phage

display for the selection of polymerase activity by proximal

display of both primer–template substrate and polymerase on

the surface of the phage particle (Fig. 1b).161,162 By selection

for in cis incorporation of a biotin-tagged nucleotide tri-

phosphate into the tethered template–primer duplex Jestin

has isolated variants of Taq polymerase with substantial RT

activity.163 Using a similar approach, Romesberg et al.

have selected variants of the Stoffel fragment of Taq poly-

merase with greatly improved incorporation of ribonucleotide

triphosphates,162 20-OCH3 ribonucleotide triphosphates164 and

a polymerase with 30-fold improved extension of the unnatural

PICS:PICS (14) self-pair.165 The phage selection approach

should in principle be ideal for the identification of polymerases

with enhanced activity with unnatural substrates. The only

caveat being that assay conditions must be compatible with

phage viability, which limits the selection of polymerases from

thermophilic organisms under realistic conditions.

Furthermore, the intramolecular tethering of the substrate

and polymerase may favour the selection of polymerase

variants with low affinity for template–primer duplex and/or

poor processivity. Finally, polymerases are usually not

efficiently secreted and displayed on the phage surface,

although this can be improved by optimisation of the signal

peptide.166 On the other hand, the phage method should be

extremely sensitive and in principle able to detect a single

incorporation event.

We have developed an alternative selection strategy for

the evolution of polymerases, called ‘‘compartmentalised

self-replication’’ (CSR).167 CSR is based on a simple feedback

loop, in which a polymerase replicates only its own encoding

gene with compartmentalisation into the aqueous compartments

of a water-in-oil (w/o) emulsion168 serving to isolate individual

self-replication reactions from each other (Fig. 1c). Thus each

polymerase replicates only its own encoding gene to the

exclusion of those in other compartments (i.e. self-replicates).

In such a system adaptive gains directly (and proportionally)

translate into genetic amplification of the encoding gene.

Therefore, the copy number of a polymerase gene after one

round of CSR is correlated to the catalytic activity of the

encoded polymerase under the selection conditions, with

polymerase genes encoding the most active polymerases best

adapted to the selection conditions dominating the population.

Segregation of self-replication into discrete, physically

separate compartments is critical to ensure linkage of pheno-

type and genotype during CSR. We developed w/o emulsions

that are stable for prolonged periods at temperatures

exceeding 90 1C and allowing selection of polymerases under

a wide range of experimental conditions including PCR

thermocycling.167 CSR has proved to be a productive method

for the selection of polymerase function yielding variants of

Taq polymerase with 410-fold increased thermostability,167

4130-fold increased resistance to heparin167 or a generically

enhanced substrate spectrum.169 Molecular breeding of

polymerase genes from the genus Thermus and CSR selection

also allowed the isolation of polymerases with a striking ability

PCR-amplify damaged DNA, which allowed the increased

recovery of DNA sequences from ice-age specimens.170 CSR

selections from the same polymerase library also yielded a

polymerase with a broad resistance to common environmental

PCR inhibitors or a generic ability to replicate large

hydrophobic base analogues (C. Baar, DL, PH, unpublished

results).

4626 | Chem. Commun., 2009, 4619–4631 This journal is �c The Royal Society of Chemistry 2009

Dow

nloa

ded

by P

ortla

nd S

tate

Uni

vers

ity o

n 04

Feb

ruar

y 20

12Pu

blis

hed

on 0

6 Ju

ly 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B90

3307

F

View Online

Page 9: Polymerase engineering: towards the encoded synthesis of ... · Polymerase engineering: towards the encoded synthesis of unnatural biopolymers ... This journal is c The Royal Society

CSR makes stringent demands on the catalytic efficiency

and processivity of selected polymerases (requiring replication

of the 42 kb polymerase gene). In order to reduce the

adaptive burden and increase the sensitivity of the method,

we devised short-patch CSR (spCSR), in which only a short,

defined segment of the polymerase gene is replicated and

evolved. spCSR has allowed the isolation of variants of Taq

with an expanded substrate range allowing enhanced

incorporation of 20-substituted nucleotides including the

generation of mixed RNA–DNA, 20-F-DNA amplification

products in PCR156 as well as greatly improved incorporation

of fluorescent dye-labelled nucleotides171 (N. Ramsay,

PH unpublished results).

4. Unnatural biopolymer synthesis

An improved understanding of the molecular mechanisms of

polymerase selectivity together with advances in nucleic acid

chemistry and polymerase engineering and evolution have

brought the synthesis of unnatural nucleic acid polymers

within our grasp.

Different types of polymers pose different challenges to

substrate design and polymerase engineering.

Nucleic acids can be viewed as repeats of tripartite nucleo-

tide building blocks comprising the nucleobase (A), the sugar

scaffold (B) and the backbone linkage (C). Different types of

polymers can comprise alternative compounds replacing or

modifying either of these components.

To simplify discussion, we propose a classification of

unnatural nucleic acid polymers based on this tripartite

structure, Fig. 2. Group A polymers DNA (or RNA) polymers

comprising standard ribofuranose sugars and phosphoester

backbone linkages with modifications to one or all of the four

canonical bases. These can be substituents, for example, to the

50-position of pyrimidines or substitutions of the natural bases

with non-canonical ones (as discussed in 2.1–2.4). Group B

polymers are DNA (or RNA) polymers comprising either

partial or complete replacement of standard ribofuranose

sugars by non-standard structures but retaining canonical

phosphoester backbone linkages (e.g. GNA, TNA) (as

discussed in 2.5). Group C polymers comprise polymers those

where the phosphodiester backbone linkages are replaced by

alternative structures.

There are many examples of partial group A polymers

(in which a natural base is partially replaced by another)

synthesised in the laboratory, notably the incorporation of

unnatural base pairs.57,62,85–87 Other partial group A polymers

are of biotechnological importance as DNA probes used in

fluorescent in situ hybridisation (FISH) or in microarrays are

generated using fluorescent dye substituted nucleotides. While

short probes can be synthesised by a number of polymerases

(e.g. Klenow for random prime dye labelling), longer group A

polymers can be difficult to synthesise in good yield especially

for larger substituents like fluorescent dyes. However, directed

evolution by CSR has yielded polymerases that are able to

completely replace one of the bases with their fluorescent

dye-labelled equivalent (e.g. dA with FITC-12-dA) in PCR.169

Fig. 2 The three panels on the left illustrate the tripartite nucleotide structure, comprising the nucleobase (red), ribofuranose sugar (blue) and

phosphoester backbone linkage (green). Nucleobases are either purine bases (A, G) (left) or pyrimidine bases (T, C) (right) and can bear

substitutions R1 (yellow) at the X (X = C) or exocyclic amino position (purines) or C5 position (pyrimidines). Panels on the right detail

modifications to the canonical structures that are compatible with enzymatic incorporation, such as base substitutions (yellow) (from left to right):

alkyne, enamine, alcohol, imidazole, guanidine and pyridyl groups; alternative base pairing systems (red) (from left to right) such as those

described by Benner (5:6),61 Yokoyama and Hirao67 and Romesberg;91 sugar modifications (blue) (from left to right), such as hexose nucleic acids

(HNA), threose nucleic acids (TNA), glycol nucleic acids (GNA) and 20-modifications (20-F, 20-OCH3) and Locked nucleic acids (LNA);

alternative internucleotide linkages (green) (from left to right): methylphosphonates, phosphorothioates, phosphoramidates, boranophosphates

and selenophosphates. We propose a classification of nucleic acid polymers, whereby group A polymers comprise nucleic acids with modifications

to the base (yellow / red panel), group B polymers comprise modifications to the sugar (blue panel) and group C comprise alternative backbones

(green panel).

This journal is �c The Royal Society of Chemistry 2009 Chem. Commun., 2009, 4619–4631 | 4627

Dow

nloa

ded

by P

ortla

nd S

tate

Uni

vers

ity o

n 04

Feb

ruar

y 20

12Pu

blis

hed

on 0

6 Ju

ly 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B90

3307

F

View Online

Page 10: Polymerase engineering: towards the encoded synthesis of ... · Polymerase engineering: towards the encoded synthesis of unnatural biopolymers ... This journal is c The Royal Society

Complete group A polymers, in which all of the four bases

bear a substituent have been described by Famulok for a

number of small substituent groups (fDNA).23 Several other

groups have also reported synthesis of group A polymers in

which all four bases are substituted with either biotin or a

fluorescent dye but with poor yields.18–22 Such polymers have

been proposed as the basis of a next-generation sequencing

method, termed exonuclease sequencing. To be practical more

quantitiative synthesis of these types of polymersis needed and

this is likely to require careful design of substituents to

maintain polymer solubility and substantial engineering of

the polymerase primer-template binding interface.

Given the stringency of molecular discrimination against

non-cognate chemistry by DNA polymerases, one would

expect the chemical make-up of natural DNA to be completely

uniform. However, the modifications to the cognate DNA

chemistry found in nature are diverse and surprisingly

wide-spread. They overwhelmingly comprise modifications to

the nucleobases (group A polymers) including, apart from the

well-known post-replicative epigenetic modifications, the

replacement of one of the canonical bases with an unnatural

analogue at the level of replication such as the complete

replacement of dA by 2,6-diaminopurine in the S. elongatus

phage S-2L. These modifications are thought to protect the

viral genome from cellular nucleases and promote preferential

replication of modified phage DNA by the cognate phage

replicases.172

Partial and complete group B polymers have so far not been

found in nature. However, there are several examples of

partial or evn complete group B polymers synthesized in the

laboratory using unnatural substitutes for ribose (e.g. GNA

and TNA) (as discussed in 2.5).101,103,109

Only one natural example of a group C polymer has been

found, that is the aS-modification in Streptomyces.173 However,

completely aS-modified DNA (in which every a oxygen is

replaced by sulfur) has been synthesized in the laboratory.169

The aS-modification is of course a modest alteration of the

standard phosphoester backbone but aS-DNA and mixed

aS-containing aptamers are nevertheless of interest due to their

resistance to many nucleases and their increased binding to

proteins.174 A number of other modifications to the phosphoester

backbone linkage have been described and found to be

compatible with enzymatic synthesis and replication. These

include boranophosphate, selenophosphate, phosphoramidate

or phosphonoester linkages. Boranophosphates are good poly-

merase substrates and have been successfullly used in

SELEX,175 and such aptamers may find clinical application

in the neutron capture therapy of tumors.

More drastic changes to the nucleic acid structure backbone,

in which both phosphoester linkage as well as sugar scaffold are

replaced by alternative chemical structures have been

attempted176 but few are satisfactory replacements. A notable

exception is PNA in which bases are displayed on a aminoethyl-

glycine backbone. PNA displays highly specific hybridization

properties and can direct non-enzymatic replication of

complementary DNA.177 However, enzymatic replication of

PNA would require extensive modification to the polymerase

active site. It is not inconceivable that polymers such as PNAs

could be synthesized on an engineered ribosome.178

5. Conclusion

Advances in polymerase engineering and evolution together

with sophisticated chemistry are poised to enable the enzymatic

synthesis, replication and evolution of unnatural nucleic acid

polymers with an ever widening array of chemical substitutions.

The new sequence and chemical space opened up by these

advances is likely to be a rich source of novel nucleic acid

therapeutics, aptamers and enzymes with useful applications

in medicine, biotechnology, nanotechnology and material

science. These novel polymers will provide insights into the

parameters of the molecular encoding of information leading

to artificial genetic systems based on unnatural chemistry.

Notes and references

1 K. R. Tindall and T. A. Kunkel, Biochemistry, 1988, 27,6008–6013.

2 E. T. Kool, Annu. Rev. Biophys. Biomol. Struct., 2001, 30,1–22.

3 M.-M. Huang, N. Arnheim and M. F. Goodman, Nucleic AcidsRes., 1992, 20, 4567–4573.

4 H. Miller and A. P. Grollman, Biochemistry, 1997, 36,15336–15342.

5 O. Potapova, C. Chan, A. M. DeLucia, S. A. Helquist,E. T. Kool, N. D. Grindley and C. M. Joyce, Biochemistry,2006, 45, 890–898.

6 A. Kornberg and T. A. Baker, DNA Replication, Freeman,San Francisco, 1992.

7 M. Astatke, N. D. Grindley and C. M. Joyce, J. Mol. Biol., 1998,278, 147–165.

8 Y. Brudno and D. R. Liu, Chem. Biol., 2009, 16, 265–276.9 M. Hocek and M. Fojta,Org. Biomol. Chem., 2008, 6, 2233–2241.10 S. H. Weisbrod and A. Marx, Chem. Commun., 2008, 5675–5685.11 M. Hollenstein, C. J. Hipolito, C. H. Lam and D. M. Perrin,

Nucleic Acids Res., 2009, 37, 1638–1649.12 P. Capek, H. Cahova, R. Pohl, M. Hocek, C. Gloeckner and

A. Marx, Chem.–Eur. J., 2007, 13, 6196–6203.13 H. Cahova, L. Havran, P. Brazdilova, H. Pivonkova, R. Pohl,

M. Fojta and M. Hocek, Angew. Chem., Int. Ed., 2008, 47,2059–2062.

14 M. Vrabel, P. Horakova, H. Pivonkova, L. Kalachova,H. Cernocka, H. Cahova, R. Pohl, P. Sebest, L. Havran,M. Hocek and M. Fojta, Chem.–Eur. J., 2009, 15, 1144–1154.

15 C. Lam, C. Hipolito and D. M. Perrin, Eur. J. Org. Chem., 2008,4915–4923.

16 H. Cahova, R. Pohl, L. Bednarova, K. Novakova, J. Cvacka andM. Hocek, Org. Biomol. Chem., 2008, 6, 3657–3660.

17 Z. Zhu, J. Chao, H. Yu and A. S. Waggoner, Nucleic Acids Res.,1994, 22, 3418–3422.

18 S. Brakmann and P. Nieckchen, ChemBioChem, 2001, 2, 773–777.19 Z. Foldes-Papp, B. Angerer, W. Ankenbauer and R. Rigler,

J. Biotechnol., 2001, 86, 237–253.20 Z. Foldes-Papp, B. Angerer, P. Thyberg, M. Hinz, S. Wennmalm,

W. Ankenbauer, H. Seliger, A. Holmgren and R. Rigler,J. Biotechnol., 2001, 86, 203–224.

21 G. Giller, T. Tasara, B. Angerer, K. Muhlegger, M. Amacker andH. Winter, Nucleic Acids Res., 2003, 31, 2630–2635.

22 T. Tasara, B. Angerer, M. Damond, H. Winter, S. Dorhofer,U. Hubscher and M. Amacker, Nucleic Acids Res., 2003, 31,2636–2646.

23 S. Jager and M. Famulok, Angew. Chem., Int. Ed., 2004, 43,3337–3340.

24 S. Jager, G. Rasched, H. Kornreich-Leshem, M. Engeser,O. Thum and M. Famulok, J. Am. Chem. Soc., 2005, 127,15071–15082.

25 S. Moran, R. X.-F. Ren and E. T. Kool, Proc. Natl. Acad. Sci.U. S. A., 1997, 94, 10506–10511.

26 S. Moran, R. X.-F. Ren, S. Rumney and E. T. Kool, J. Am.Chem. Soc., 1997, 119, 2056–2057.

4628 | Chem. Commun., 2009, 4619–4631 This journal is �c The Royal Society of Chemistry 2009

Dow

nloa

ded

by P

ortla

nd S

tate

Uni

vers

ity o

n 04

Feb

ruar

y 20

12Pu

blis

hed

on 0

6 Ju

ly 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B90

3307

F

View Online

Page 11: Polymerase engineering: towards the encoded synthesis of ... · Polymerase engineering: towards the encoded synthesis of unnatural biopolymers ... This journal is c The Royal Society

27 K. M. Guckian, J. C. Morales and E. T. Kool, J. Org. Chem.,1998, 63, 9652–9656.

28 J. C. Morales and E. T. Kool, J. Am. Chem. Soc., 1999,121, 2323–2324.

29 J. C. Morales and E. T. Kool, J. Am. Chem. Soc., 2000, 122,1001–1007.

30 J. C. Morales and E. T. Kool, Biochemistry, 2000, 39,12979–12988.

31 T. W. Kirby, E. F. DeRose, W. A. Beard, S. H. Wilson andR. E. London, Biochemistry, 2005, 44, 15230–15237.

32 O. Potapova, C. Chan, A. M. DeLucia, S. A. Helquist,E. T. Kool, N. D. Grindley and C. M. Joyce, Biochemistry,2006, 45, 890–898.

33 J. C. Morales and E. T. Kool, Biochemistry, 2000, 39, 2626–2632.34 J. C. Delaney, P. T. Henderson, S. A. Helquist, J. C. Morales,

J. M. Essigmann and E. T. Kool, Proc. Natl. Acad. Sci. U. S. A.,2003, 100, 4469–4473.

35 T. W. Kim and E. T. Kool, J. Org. Chem., 2005, 70, 2048–2053.36 T. W. Kim, J. C. Delaney, J. M. Essigmann and E. T. Kool, Proc.

Natl. Acad. Sci. U. S. A., 2005, 44, 15803–15808.37 S. Mizukami, T. W. Kim, S. A. Helquist and E. T. Kool,

Biochemistry, 2006, 45, 2772–2778.38 A. P. Silverman, Q. Jiang, M. F. Goodman and E. T. Kool,

Biochemistry, 2007, 46, 13874–13881.39 A. P. Silverman and E. T. Kool, J. Am. Chem. Soc., 2007, 129,

10626–10627.40 A. P. Silverman, S. J. Garforth, V. R. Prasad and E. T. Kool,

Biochemistry, 2008, 47, 4800–4807.41 T. J. Matray and E. T. Kool, Nature, 1999, 399, 704.42 L. Dzantiev, Y. O. Alekseyev, J. C. Morales, E. T. Kool and

L. J. Romano, Biochemistry, 2001, 40, 3215–3221.43 S. Smirnov, T. J. Matray, E. T. Kool and C. de los Santos,

Nucleic Acids Res., 2002, 30, 5561–5569.44 H. Hwang and J.-S. Taylor, Biochemistry, 2004, 43, 14612–14623.45 L. Sun, M. Wang, E. T. Kool and J.-S. Taylor, Biochemistry,

2000, 39, 14603–14610.46 A. Sheriff, E. Motea, I. Lee and A. J. Berdis, Biochemistry, 2008,

47, 8527–8537.47 X. Zhang, I. Lee and A. J. Berdis, Org. Biomol. Chem., 2004, 2,

1703–1711.48 E. Z. Reineks and A. J. Berdis, Biochemistry, 2004, 43,

393–404.49 D. Vineyard, X. Zhang, A. Donnelly, I. Lee and A. J. Berdis,Org.

Biomol. Chem., 2007, 5, 3623–3630.50 X. Zhang, I. Lee and A. J. Berdis, Biochemistry, 2005, 44,

13111–13121.51 I. Lee and A. J. Berdis, ChemBioChem, 2006, 7, 1990–1997.52 X. Zhang, A. Donnelly, I. Lee and A. J. Berdis, Biochemistry,

2006, 45, 13293–13303.53 X. Zhang, I. Lee, X. Zhou and A. J. Berdis, J. Am. Chem. Soc.,

2006, 128, 143–149.54 X. Zhang, I. Lee and A. J. Berdis, Biochemistry, 2005, 44,

13101–13110.55 C. R. Geyer, T. R. Battersby and S. A. Benner, Structure

(London), 2003, 11, 1485–1498.56 F. Seela and C. Wei, Helv. Chim. Acta, 1999, 82, 726.57 A. M. Sismour and S. A. Benner, Nucleic Acids Res., 2005, 33,

5640–5646.58 M. L. Abraham, M. Albalos, T. Guettouche, M. J. Friesenhahn

and T. R. Battersby, BioTechniques, 2007, 43, 617.59 J. D. Ahle, S. Barr, A. M. Chin and T. R. Battersby,Nucleic Acids

Res., 2005, 33, 3176–3184.60 S. A. Benner, Acc. Chem. Res., 2004, 37, 784–797.61 Z. Yang, A. M. Sismour, P. Sheng, N. L. Puskar and

S. A. Benner, Nucleic Acids Res., 2007, 35, 4238–4249.62 I. Hirao, T. Fujiwara, M. Kimoto and S. Yokoyama, Bioorg.

Med. Chem. Lett., 2004, 14, 4887–4890.63 I. Hirao, Y. Harada, M. Kimoto, T. Mitsui, T. Fujiwara and

S. Yokoyama, J. Am. Chem. Soc., 2004, 126, 13298–13305.64 T. Mitsui, A. Kitamura, M. Kimoto, T. To, A. Sato, I. Hirao and

S. Yokoyama, J. Am. Chem. Soc., 2003, 125, 5298–5307.65 T. Mitsui, M. Kimoto, A. Sato, S. Yokoyama and I. Hirao,

Bioorg. Med. Chem. Lett., 2003, 13, 4515–4518.66 I. Hirao, T. Mitsui, M. Kimoto and S. Yokoyama, J. Am. Chem.

Soc., 2007, 129, 15549–15555.

67 M. Kimoto, R. Kawai, T. Mitsui, S. Yokoyama and I. Hirao,Nucleic Acids Res., 2009, 37, e14.

68 M. Endo, T. Mitsui, T. Okuni, M. Kimoto, I. Hirao andS. Yokoyama, Bioorg. Med. Chem. Lett., 2004, 14, 2593–2596.

69 T. Mitsui, M. Kimoto, Y. Harada, S. Yokoyama and I. Hirao,J. Am. Chem. Soc., 2005, 127, 8652–8658.

70 I. Hirao, T. Ohtsuki, T. Fujiwara, T. Mitsui, T. Yokogawa,T. Okuni, H. Nakayama, K. Takio, T. Yabuki, T. Kigawa,K. Kodama, T. Yokogawa, K. Nishikawa and S. Yokoyama,Nat. Biotechnol., 2002, 20, 177–182.

71 I. Hirao, M. Kimoto, T. Mitsui, T. Fujiwara, R. Kawai, A. Sato,Y. Harada and S. Yokoyama, Nat. Methods, 2006, 3, 729.

72 R. Kawai, M. Kimoto, S. Ikeda, T. Mitsui, M. Endo,S. Yokoyama and I. Hirao, J. Am. Chem. Soc., 2005, 127,17286–17295.

73 M. Kimoto, T. Mitsui, Y. Harada, A. Sato, S. Yokoyama andI. Hirao, Nucleic Acids Res., 2007, 35, 5360–5369.

74 K. Moriyama, M. Kimoto, T. Mitsui, S. Yokoyama and I. Hirao,Nucleic Acids Res., 2005, 33, e129.

75 M. Kimoto, M. Endo, T. Mitsui, T. Okuni, I. Hirao andS. Yokoyama, Chem. Biol., 2004, 11, 47–55.

76 H. P. Rappaport, Biochem. J., 2004, 381, 709–717.77 H. B. Liu, J. M. Gao, L. Maynard, Y. D. Saito and E. T. Kool,

J. Am. Chem. Soc., 2004, 126, 1102–1109.78 H. B. Liu, J. M. Gao and E. T. Kool, J. Org. Chem., 2005, 70,

639–647.79 A. H. F. Lee and E. T. Kool, J. Am. Chem. Soc., 2005, 127,

3332–3338.80 J. Chelliserrykattil, H. Lu, A. H. F. Lee and E. T. Kool,

ChemBioChem, 2008, 9, 2976–2980.81 G. T. Hwang and F. E. Romesberg, Nucleic Acids Res., 2006, 34,

2037–2045.82 S. Matsuda, A. A. Henry and F. E. Romesberg, J. Am. Chem.

Soc., 2006, 128, 6369–6375.83 Y. Kim, A. M. Leconte, Y. Hari and F. E. Romesberg, Angew.

Chem., Int. Ed., 2006, 45, 7809–7812.84 Y. Hari, G. T. Hwang, A. M. Leconte, N. Joubert, M. Hocek and

F. E. Romesberg, ChemBioChem, 2008, 9, 2796–2799.85 M. Berger, Y. Wu, A. K. Ogawa, D. L. McMinn, P. G. Schultz

and F. E. Romesberg, Nucleic Acids Res., 2000, 28, 2911–2914.86 S. Matsuda and F. E. Romesberg, J. Am. Chem. Soc., 2004, 126,

14419–14427.87 A. M. Leconte, G. T. Hwang, S. Matsuda, P. Capek, Y. Hari and

F. E. Romesberg, J. Am. Chem. Soc., 2008, 130, 2336–2343.88 A. M. Leconte, S. Matsuda, G. T. Hwang and F. E. Romesberg,

Angew. Chem., Int. Ed., 2006, 45, 4326–4329.89 E. L. Tae, Y. Wu, G. Xia, P. G. Schultz and F. E. Romesberg,

J. Am. Chem. Soc., 2001, 123, 7439–7440.90 A. M. Leconte, S. Matsuda and F. E. Romesberg, J. Am. Chem.

Soc., 2006, 128, 6780–6781.91 Y. J. Seo, G. T. Hwang, P. Ordoukhanian and F. E. Romesberg,

J. Am. Chem. Soc., 2009, 131, 3246–3252.92 R. Nichols, P. C. Andrews, P. Zhang and D. E. Bergstrom,

Nature, 1994, 369, 492–493.93 D. Loakes and D. M. Brown, Nucleic Acids Res., 1994, 22,

4039–4043.94 D. Loakes, Nucleic Acids Res., 2001, 29, 2437–2447.95 T. H. Smith, J. Latour, C. Leo, S. Muthini, P. Siebert and

P. S. Nelson, Clin. Chem. (Washington, D. C.), 1995, 41, 10.96 C. L. Smith, A. C. Simmonds, I. R. Felix, A. L. Hamilton,

S. Kumar, S. Nampalli, D. Loakes, F. Hill and D. M. Brown,Nucleosides Nucleotides, 1998, 17, 541–554.

97 M. Chiaramonte, C. L. Moore, K. Kincaid and R. D. Kuchta,Biochemistry, 2003, 42, 10472–10481.

98 C. Crey-Desbiolles, N. Berther, M. Kotera and P. Dumy, NucleicAcids Res., 2005, 33, 1532–1543.

99 D. A. Harki, J. D. Graci, V. S. Korneeva, S. K. B. Ghosh,Z. Hong, C. E. Cameron and B. R. Peterson, Biochemistry, 2002,41, 9026–9033.

100 L. Zhang and E. Meggers, J. Am. Chem. Soc., 2005, 127,74–75.

101 M. K. Schlegel, A. E. Peritz, K. Kittigowittana, L. Zhang andE. Meggers, ChemBioChem, 2007, 8, 927–932.

102 A. T. Horhota, J. W. Szostak and L. W. McLaughlin, Org. Lett.,2006, 8, 5345–5347.

This journal is �c The Royal Society of Chemistry 2009 Chem. Commun., 2009, 4619–4631 | 4629

Dow

nloa

ded

by P

ortla

nd S

tate

Uni

vers

ity o

n 04

Feb

ruar

y 20

12Pu

blis

hed

on 0

6 Ju

ly 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B90

3307

F

View Online

Page 12: Polymerase engineering: towards the encoded synthesis of ... · Polymerase engineering: towards the encoded synthesis of unnatural biopolymers ... This journal is c The Royal Society

103 C.-H. Tsai, J. Chen and J. W. Szostak, Proc. Natl. Acad. Sci.U. S. A., 2007, 104, 14598–14603.

104 B. D. Heuberger and C. Switzer, J. Am. Chem. Soc., 2008, 130,412–413.

105 K. U. Schoning, P. Scholz, S. Guntha, X. Wu, R. Krishnamurthyand A. Eschenmoser, Science, 2001, 290, 1347.

106 C. J. Wilds, Z. Wawrzak, R. Krishnamurthy, A. Eschenmoserand M. Egli, J. Am. Chem. Soc., 2002, 124, 13716–13721.

107 P. S. Pallan, C. J. Wilds, Z. Wawrzak, R. Krishnamurthy,A. Eschenmoser and M. Egli, Angew. Chem., Int. Ed., 2003, 42,5893–5895.

108 B. D. Heuberger and C. Switzer, Org. Lett., 2006, 8, 5809–5811.109 J. C. Chaput, J. K. Ichida and J. W. Szostak, J. Am. Chem. Soc.,

2003, 125, 856–857.110 V. Kempeneers, K. Vastmans, J. Rozenski and P. Herdewijn,

Nucleic Acids Res., 2003, 31, 6221–6226.111 J. C. Chaput and J. W. Szostak, J. Am. Chem. Soc., 2003, 125,

9274–9275.112 A. F. Gardner and W. E. Jack, Nucleic Acids Res., 1999, 27,

2545–2553.113 A. F. Gardner and W. E. Jack, Nucleic Acids Res., 2002, 30,

605–613.114 A. Horhota, K. Zou, J. K. Ichida, B. Yu, L. W. McLaughlin,

J. W. Szostak and J. C. Chaput, J. Am. Chem. Soc., 2005, 127,7427–7434.

115 J. K. Ichida, A. Horhota, K. Zou, L. W. McLaughlin andJ. W. Szostak, Nucleic Acids Res., 2005, 33, 5219–5225.

116 J. K. Ichida, K. Zou, A. Horhota, B. Yu, L. W. McLaughlin andJ. W. Szostak, J. Am. Chem. Soc., 2005, 127, 2802–2803.

117 C. Wilson and A. D. Keefe, Curr. Opin. Chem. Biol., 2006, 10,607–614.

118 A. D. Keefe and S. T. Cload, Curr. Opin. Chem. Biol., 2008, 12,448–456.

119 P. E. Burmeister, S. D. Lewis, R. F. Silva, J. R. Preiss,L. R. Horwitz, P. S. Pendergrast, T. G. McCauley, J. C. Kurz,D. M. Epstein, C. Wilson and A. D. Keefe, Chem. Biol., 2005, 12,25–33.

120 S. Obika, M. Dorio, D. Nanbu and T. Imanishi, Chem. Commun.,1997, 1643.

121 S. K. Singh, P. Nielsen, A. A. Koshkin and J. Wengel, Chem.Commun., 1998, 455.

122 M. Kuwahara, S. Obika, J.-I. Nagashima, Y. Ohta, Y. Suto,H. Ozaki, H. Sawai and T. Imanishi, Nucleic Acids Res., 2008, 36,4257–4265.

123 R. N. Veedu, B. Vester and J. Wengel, ChemBioChem, 2007, 8,490–492.

124 R. N. Veedu, B. Vester and J. Wengel, Nucleosides, NucleotidesNucleic Acids, 2007, 26, 1207–1210.

125 R. N. Veedu, B. Vester and J. Wengel, J. Am. Chem. Soc., 2008,130, 8124–8125.

126 F. Di Pasquale, D. Fischer, D. Grohmann, T. Restle, A. Geyerand A. Marx, J. Am. Chem. Soc., 2008, 130, 10748–10757.

127 M. Strerath, J. Cramer, T. Restle and A. Marx, J. Am. Chem.Soc., 2002, 124, 11230–11231.

128 D. Summerer and A. Marx, Angew. Chem., Int. Ed., 2001, 40,3693–3695.

129 D. Summerer and A. Marx, J. Am. Chem. Soc., 2002, 124,910–911.

130 L. Kerremans, G. Schepers, J. Rozenski, R. Busson, A. VanAerschot and P. Herdewijn, Org. Lett., 2001, 3, 4129–4132.

131 R. Declercq, A. Van Aerschot, R. J. Read, P. Herdewijn andL. Van Meervelt, J. Am. Chem. Soc., 2002, 124, 928.

132 T. Maier, I. Przylas, N. Strater, P. Herdewijn and W. Saenger,J. Am. Chem. Soc., 2005, 127, 2937–2943.

133 G. Kolb, S. Reigadas, C. Boiziau, A. Van Aerschot,A. Arzumanov, M. J. Gait, P. Herdewijn and J.-J. Toulme,Biochemistry, 2005, 44, 2926–2933.

134 H. Kang, M. H. Fisher, D. Xu, Y. J. Miyamoto, A. Marchand,A. Van Aerschot, P. Herdewijn and R. L. Juliano, Nucleic AcidsRes., 2004, 32, 4411–4419.

135 K. Vastmans, S. Pochet, A. Peys, L. Kerremans, A. Van Aerschot,C. Hendrix, P. Marliere and P. Herdewijn, Biochemistry, 2000, 39,12757–12765.

136 K. Vastmans, M. Froeyen, L. Kerremans, S. Pochet andP. Herdewijn, Nucleic Acids Res., 2001, 29, 3154–3163.

137 I. N. Lavrik, O. N. Avdeeva, O. A. Dontsova, M. Froeyen andP. Herdewijn, Biochemistry, 2001, 40, 11777–11784.

138 J. Wang, B. Verbeure, I. Luyten, M. Froeyen, C. Hendrix,H. Rosemeyer, F. Seela, A. Van Aerschot and P. Herdewijn,Nucleosides, Nucleotides Nucleic Acids, 2001, 20, 785–788.

139 B. Verbeure, E. Lescrinier, J. Wang and P. Herdewijn, NucleicAcids Res., 2001, 29, 4941–4947.

140 Y. Li, V. Mitaxov and G. Waksman, Proc. Natl. Acad. Sci.U. S. A., 1999, 96, 9491–9496.

141 P. H. Patel and L. A. Loeb, J. Biol. Chem., 2000, 275,40266–40272.

142 M. Astatke, K. Ng, N. D. Grindley and C. M. Joyce, Proc. Natl.Acad. Sci. U. S. A., 1998, 95, 3402–3407.

143 K. B. Ignatov, A. A. Bashirova, A. I. Miroshnikov andV. M. Kramarov, FEBS Lett., 1999, 448, 145–148.

144 T. A. Kunkel and K. Bebenek, Annu. Rev. Biochem., 2000, 69,497–529.

145 E. Bedford, S. Tabor and C. C. Richardson, Proc. Natl. Acad. Sci.U. S. A., 1997, 94, 479–484.

146 J. F. Davidson, R. Fox, D. D. Harris, S. Lyons-Abbott andL. A. Loeb, Nucleic Acids Res., 2003, 31, 4702–4709.

147 Y. Wang, D. E. Prosen, L. Mei, J. C. Sullivan, M. Finney andP. B. Vander Horn, Nucleic Acids Res., 2004, 32, 1197–1207.

148 J. G. Williams, D. L. Steffens, J. P. Anderson, T. M. Urlacher,D. T. Lamb, D. L. Grone and J. C. Egelhoff, Nucleic Acids Res.,2008, 36, e121.

149 M. Suzuki, D. Baskin, L. Hood and L. A. Loeb, Proc. Natl. Acad.Sci. U. S. A., 1996, 93, 9670–9675.

150 B. Kim, T. R. Hathaway and L. A. Loeb, J. Biol. Chem., 1996,271, 4872–4878.

151 J. B. Sweasy and L. A. Loeb, Proc. Natl. Acad. Sci. U. S. A., 1993,90, 4626–4630.

152 E. Loh and L. A. Loeb, DNA Repair, 2005, 4, 1390–1398.153 E. Loh, J. Choe and L. A. Loeb, J. Biol. Chem., 2007, 282,

12201–12209.154 E. Glick, K. L. Vigna and L. A. Loeb, EMBO J., 2001, 20,

7303–7312.155 S. G. Sarafianos, B. Marchand, K. Das, D. M. Himmel,

M. A. Parniak, S. H. Hughes and E. Arnold, J. Mol. Biol.,2009, 385, 693–713.

156 J. L. Ong, D. Loakes, S. Jaroslawski, K. Too and P. Holliger,J. Mol. Biol., 2006, 361, 537–550.

157 D. Summerer and A. Marx, Angew. Chem., Int. Ed., 2002, 41,3620–3622; D. Summerer and A. Marx, Angew. Chem., Int. Ed.,2002, 41, 3516.

158 D. Summerer, N. Z. Rudinger, I. Detmer and A. Marx, Angew.Chem., Int. Ed., 2005, 44, 4712–4715.

159 K. B. Sauter and A. Marx, Angew. Chem., Int. Ed., 2006, 45,7633–7635.

160 C. Gloeckner, K. B. Sauter and A. Marx, Angew. Chem., Int. Ed.,2007, 46, 3115–3117.

161 J. L. Jestin, P. Kristensen and G. Winter, Angew. Chem., Int. Ed.,1999, 38, 1124–1127.

162 G. Xia, L. Chen, T. Sera, M. Fa, P. G. Schultz andF. E. Romesberg, Proc. Natl. Acad. Sci. U. S. A., 2002, 99,6597–6602.

163 S. Vichier-Guerre, S. Ferris, N. Auberger, K. Mahiddine andJ. L. Jestin, Angew. Chem., Int. Ed., 2006, 45, 6133–6137.

164 M. Fa, A. Radeghieri, A. A. Henry and F. E. Romesberg, J. Am.Chem. Soc., 2004, 126, 1748–1754.

165 A. M. Leconte, L. Chen and F. E. Romesberg, J. Am. Chem. Soc.,2005, 127, 12470–12471.

166 J. L. Jestin, G. Volioti and G. Winter, Res. Microbiol., 2001, 152,187–191.

167 F. J. Ghadessy, J. L. Ong and P. Holliger, Proc. Natl. Acad. Sci.U. S. A., 2001, 98, 4552–4557.

168 D. S. Tawfik and A. D. Griffiths, Nat. Biotechnol., 1998, 16,652–656.

169 F. J. Ghadessy, N. Ramsay, F. Boudsocq, D. Loakes, A. Brown,S. Iwai, A. Vaisman, R. Woodgate and P. Holliger,Nat. Biotechnol.,2004, 22, 755–759.

170 M. d’Abbadie, M. Hofreiter, A. Vaisman, D. Loakes,D. Gasparutto, J. Cadet, R. Woodgate, S. Paabo andP. Holliger, Nat. Biotechnol., 2007, 25, 939–943.

171 J. Ong, PhD Thesis, Cambridge University, 2004.

4630 | Chem. Commun., 2009, 4619–4631 This journal is �c The Royal Society of Chemistry 2009

Dow

nloa

ded

by P

ortla

nd S

tate

Uni

vers

ity o

n 04

Feb

ruar

y 20

12Pu

blis

hed

on 0

6 Ju

ly 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B90

3307

F

View Online

Page 13: Polymerase engineering: towards the encoded synthesis of ... · Polymerase engineering: towards the encoded synthesis of unnatural biopolymers ... This journal is c The Royal Society

172 J. H. Gommers-Ampt and P. Borst, FASEB J., 1995, 9,1034–1042.

173 L. Wang, S. Chen, T. Xu, K. Taghizadeh, J. S. Wishnok,X. Zhou, D. You, Z. Deng and P. C. Dedon, Nat. Chem. Biol.,2007, 3, 709–710.

174 A. Somasunderam, M. R. Ferguson, D. R. Rojo,V. Thiviyanathan, X. Li, W. A. O’Brien and D. G. Gorenstein,Biochemistry, 2005, 44, 10388–10395.

175 S. M. Lato, N. D. Ozerova, K. He, Z. Sergueeva, B. R. Shaw andD. H. Burke, Nucleic Acids Res., 2002, 30, 1401–1407.

176 P. E. Nielsen, Annu. Rev. Biophys. Biomol. Struct., 1995, 24,167–183.

177 J. G. Schmidt, L. Christensen, P. E. Nielsen and L. E. Orgel,Nucleic Acids Res., 1997, 25, 4792–4796.

178 O. Rackham and J. W. Chin, Nat. Chem. Biol., 2005, 1,159–166.

This journal is �c The Royal Society of Chemistry 2009 Chem. Commun., 2009, 4619–4631 | 4631

Dow

nloa

ded

by P

ortla

nd S

tate

Uni

vers

ity o

n 04

Feb

ruar

y 20

12Pu

blis

hed

on 0

6 Ju

ly 2

009

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/B90

3307

F

View Online