36
Lipase chemoselectivity – kinetics and applications Cecilia Hedfors Licentiate Thesis School of Biotechnology Royal Institute of Technology Stockholm 2009

Lipase chemoselectivity – kinetics and applications ...211890/FULLTEXT01.pdf · Lipase chemoselectivity – kinetics and applications . Cecilia Hedfors . ... can be used for the

  • Upload
    doliem

  • View
    240

  • Download
    0

Embed Size (px)

Citation preview

Lipase chemoselectivity – kinetics and applications

Cecilia Hedfors

Licentiate Thesis

School of Biotechnology Royal Institute of Technology

Stockholm 2009

© Cecilia Hedfors School of Biotechnology Royal Institute of Technology AlbaNova University Center 106 91 Stockholm Sweden ISBN 978-91-7415-275-3 TRITA-BIO Report 2009:7 ISSN 1654-2312 Printed in Stockholm, April 2009 UniversitetsService US AB Box 700 14 100 44 Stockholm Sweden

ABSTRACT

A chemoselective catalyst is preferred in a chemical reaction where protecting groups

otherwise are needed. The two lipases Candida antarctica lipase B and Rhizomucor miehei

lipase showed large chemoselectivity ratios, defined as (kcat/KM)OH / (kcat/KM)SH, in a

transacylation reaction with ethyl octanoate as acyl donor and hexanol or hexanethiol as acyl

acceptor (paper I). The chemoselectivity ratio of the uncatalyzed reaction was 120 in favour

of the alcohol. Compared to the uncatalyzed reaction, the chemoselectivity was 730 times

higher for Candida antarctica lipase B and ten times higher for Rhizomucor miehei lipase.

The KM towards the thiol was more than two orders of magnitude higher than the KM towards

the corresponding alcohol. This was the dominating contribution to the high chemoselectivity

displayed by the two lipases. In a novel approach, Candida antarctica lipase B was used as

catalyst for enzymatic synthesis of thiol-functionalized polyesters in a one-pot reaction

without using protecting groups (paper II). Poly(ε-caprolactone) with a free thiol at one of

the ends was synthesized in an enzymatic ring-opening polymerization initiated with

mercaptoethanol or terminated with either 3-mercaptopropionic acid or γ-thiobutyrolactone.

SAMMANFATTNING

En kemoselektiv katalysator är att föredra i en kemisk reaktion där skyddsgrupper annars är

nödvändiga. I en transacyleringsreaktion med etyloktanoat som acyldonator och hexanol eller

hexantiol som acylacceptor var den kemoselektiva kvoten, definierad som (kcat/KM)OH /

(kcat/KM)SH, stor för både Candida antarctica lipas B och Rhizomucor miehei lipas (artikel I).

Den kemoselektiva kvoten för den ickekatalyserade reaktionen var 120, till fördel för

alkoholen. Jämfört med den ickekatalyserade reaktionen var kemoselektiviteten 730 gånger

större för Candida antarctica lipas B och tio gånger större för Rhizomucor miehei lipas. KM

för tiolen var mer än två tiopotenser större jämfört med KM för motsvarande alkohol. Detta

var det dominerande bidraget till den stora kemoselektiviteten som de båda lipaserna

uppvisade. I artikel II användes Candida antarctica lipas B som katalysator för enzymatisk

syntes av tiol-funktionaliserade polyestrar i en enkärlsreaktion utan användande av

skyddsgrupper. Poly(ε-kaprolakton) med en fri tiolgrupp i en av ändarna syntetiserades med

enzymatisk ringöppningspolymerisation initierad med merkaptoetanol alternativt terminerad

med 3-merkaptopropionsyra eller γ-tiobutyrolakton.

LIST OF PUBLICATIONS

This thesis is based on the following publications which are referred to by their Roman numerals:

I Lipase chemoselectivity between thiol and alcohol acyl acceptors in

a transacylation reaction. C. Hedfors, K. Hult and M. Martinelle. Manuscript II Thiol end-functionalization of poly(ε-caprolactone), catalyzed by

Candida antarctica lipase B. C. Hedfors, E. Östmark, E. Malmström, K. Hult and M. Martinelle. Macromolecules 2005, 38, 647-649.

TABLE OF CONTENTS

1. INTRODUCTION 1

2. ENZYME SELECTIVITY 3

2.1 Regioselectivity 32.2 Enantioselectivity 4 2.3 Chemoselectivity 5 2.4 The Michaelis-Menten equation 5

3. TRIACYLGLYCEROL LIPASES 7

3.1 Reaction mechanism 8 3.2 Candida antarctica lipase B 10 3.3 Rhizomucor miehei lipase 10 3.4 Active site titration of immobilized lipase 10

4. LIPASE CATALYZED CHEMOSELECTIVE TRANSACYLATION REACTIONS 13

4.1 Chemoselectivity 14

5. LIPASES IN POLYESTER SYNTHESIS 17

5.1 Polycondensation 17 5.2 Ring-opening polymerization 18 5.3 Selective lipase polymerization 20 5.4 Chemoselective thiol end-functionalization of polyesters 20

ACKNOWLEDGMENTS 23

REFERENCES 25

1. INTRODUCTION

The existence of life is fascinating. How was it created and why does is exist at all?

Independently of how life began, one of its keys is enzymes. Enzymes are proteins catalyzing

chemical reactions by lowering the energy barrier when a substrate is transformed to a

product, increasing the probability of a reaction. The rate enhancement of an enzyme

catalyzed reaction, compared to the uncatalyzed reaction, is often in the magnitude of 108

[Wolfeneden, 2006], but the range is wide. The enzyme with the highest rate enhancement

known is orotidine monophosphate decarboxylase (EC 4.1.1.23), which converts orotidine

monophosphate to uridine monophosphate. The rate enhancement of this enzyme is 1017, or

with other words, the half-time of the reaction is 18 milliseconds with enzyme and 78 million

years without [Miller et al, 2002]. Enzymes are also important in another perspective as they

often display substrate selectivity. This allows them to control reaction pathways, prohibiting

side-reactions that could cause unwanted effects in a living organism. This selectivity is a

powerful tool and we learn more and more how to use it in vitro. This thesis will consider two

lipases, Candida antarctica lipase B and Rhizomucor miehei lipase, both capable to

distinguish between an alcohol and a thiol in transacylation reactions, and also, the use of

Candida antarctica lipase B as catalyst in thiol end-functionalized polyester synthesis.

Ever since the discovery of enzymes and their ability to catalyze chemical reactions, scientists

have tried to take advantage of their properties. With the technical developments in the

biotechnology area, it is possible to produce enzymes on a large scale (fermenter size > 10

m3) to a relatively low cost; enzyme concentrates are sold from 5 €/L [Bornscheuer and

Buchholz, 2005]. This opens the door for more industrial processes to use enzyme catalysts

and benefit from their high substrate selectivity.

- 1 -

The production of enzymes reached a value of 2 billion € in year 2007 with Novozymes A/S,

which held a market share of 47 percent, as the dominating manufacturer [Novozymes report

2007]. Enzymes can be used as final products, for example in detergents or in pharmaceutical

and animal feed industries, or as a processing aid like in textile, leather and sugar industries.

Enzymes are also used in food and beverage production; alcohol, baking and dairy industries.

The two dominating market areas for enzymes, with respect to volume and value of the

enzyme, are food processing (40-45%) and detergent industry (35-40%) [Bornscheuer and

Buchholz, 2005]. The single largest enzymatic industrial process is the production of high

fructose corn syrup from corn starch. The starch is broken down to oligosaccharides with the

enzyme α-amylase (EC 3.2.1.1), and further to glucose by glucoamylase (EC 3.2.1.3). Xylose

isomerase (EC 5.3.1.5) converts the glucose to fructose, yielding the syrup used as a substitute

to sugar [Cheetham, 2000].

- 2 -

2. ENZYME SELECTIVITY

One of the main advantages of using enzymes as catalysts is that they often display substrate

selectivity, i.e. the ability to favor one of many possible substrates. A chemically catalyzed

multiple step process, possibly needing protecting groups, was with use of enzymes made in a

one-step reaction with good yield (paper II). There are different kinds of substrate selectivity,

enzymes can for example be regioselective, enantioselective or chemoselective.

2.1 Regioselectivity If the production of one structural isomer is favored above others, the reaction is said to be

regioselective [Smith and March, 2001]. Oligosaccharides are formed from monosaccharides

which are linked together by glycosidic bonds (figure 1).

O

O OOH

OH

O

OOH

OH

O

O

OH

OH

O

OOH

OH

O

O OOH

OH

α1,4-glycosidic bond

α1,6-glycosidic bond

OH

OH OH

OH

Figure 1. The structure of glycogen with the α1,4-glycosidic and α1,6-glycosidic bonds marked.

Depending on the linkage, different kinds of oligosaccharides are formed. In glycogen, the

bonds between the glucose molecules are mainly α1,4-glycosidic bonds, but one in twelve is

- 3 -

an α1,6-glycosidic bond, which makes the polymer branched. Glycogen synthase (EC

2.4.1.11) can only form the α1,4-glycosidic bonds to yield α-amylose and are thus

regioselective. The branching of α-amylose to form glycogen is accomplished by amylo-

(1,4→1,6)-transglycosylase (EC 2.4.1.18) [Voet and Voet, 1995].

2.2 Enantioselectivity Enantiomers are a pair of molecular entities which are non-superposable mirror images of

each other [Moss, 1996] and enantioselectivity is the ability to distinguish between the two

enantiomers. Aspartame is a sweetener 200 times as sweet as sucrose. In the production of

aspartame, thermolysin (EC 3.4.24.27) can be used for the peptide bond formation between

the two precursors; blocked L-aspartic acid and L-phenylalanine methyl ester (figure 2). The

enantioselectivity of the enzyme makes it possible to use the cheaper racemic DL-

phenylalanine methyl ester, instead of the pure L-form. The unreacted D-form can be

racemised and reused in the synthesis, leading to high atom efficiency. Thermolysin is also

regioselective and reacts only with the α-carboxylate of the aspartic acid and not with the side

chain β-carboxylate, which would give rise to a bitter tasting β-aspartame [Cheetham, 2000].

L-Asp-L-Phe-OMe

L-Z-Asp + DL-Phe-OMe

L-Z-Asp-L-Phe-OMe D-Phe-OMe

L-Z-Asp-L-Phe-OMe

D-Phe-OMe

thermolysinracemerisation

HCl

deblocking

(aspartame)

aspartame

H3NNH

O

O

O

O

O

Figure 2. Production scheme of aspartame. Thermolysin is used in the first step as an enantio- and regioselective catalyst. The produced di-peptide forms a complex together with the unwanted form of phenylalanine (D-Phe-OMe) and precipitates. The D-Phe-OMe can be racemised and used as starting material again. Z = benzyloxycarbonyl protecting group on the amine group of aspartate.

- 4 -

2.3 Chemoselectivity Chemoselectivity is the existence of a preferential reaction of a chemical reagent with one

functional group in the presence of other functional groups [Smith and March, 2001]. The

lipase of Burkholderia cepacia (EC 3.1.1.3) catalyzes acylation of 1-phenylethanol hundred

times faster than of the corresponding amine using methyl butyrate as acyl donor

[Cammenberg et al, 2006].

O

O

NH2

OH

O

O

NH

O

100:1

+

+

+

Bcl

Figure 3. In a transacylation reaction Burkholderia cepacia lipase (Bcl) prefer the alcohol over the amine as acyl acceptor by a factor of 100.

As we have shown, Candida antarctica lipase B catalyzed the acylation of alcohol 88 000

times faster than the corresponding thiol in a transacylation reaction (paper I).

2.4 The Michaelis-Menten equation Most enzymes follow the kinetics proposed by Michaelis and Menten in year 1913. Enzyme

(E) and substrate (S) first form an enzyme-substrate complex (ES), called the Michaelis-

Menten complex, which is a non-covalently bound intermediate. The complex can then either

dissociate or cross the reaction energy barrier to form enzyme and product (P), (equation 1).

The reaction rate (v) of product formation is dependent on the catalytic constants kcat and KM,

the substrate concentration [S] and the enzyme concentration [E], according to the Michaelis-

Menten equation (equation 2). The specificity constant, kcat/KM, is an apparent second-order

rate constant of the enzyme reaction reflecting the specificity and efficiency of the reaction

(equation 3). Enzyme selectivity is the ratio of the specificity constants of the two reactions.

- 5 -

E + S ES E + PEP (equation 1)

[ ] [ ][ ]S

SE

M

0

+=

Kkv cat (equation 2)

[ ] [ ]SEM

freecat

Kkv = (equation 3)

- 6 -

3. TRIACYLGLYCEROL LIPASES

Triacylglycerol lipases (EC 3.1.1.3), in this thesis called lipases, belong to the family of

hydrolases acting on ester bonds. Lipases hydrolyze triglycerides to glycerol and free fatty

acids. They are found in animals, plants, bacteria and fungi, where they play an important role

in the metabolism. Lipases have an α/β-hydrolase fold [Ollis et al, 1992], a conserved

catalytic triad (Ser, His, Asp/Glu) [Brady et al, 1990] and an oxyanion hole which stabilizes

the oxyanion in the transition state [Brzozowski et al, 1991]. Most of them share a consensus

sequence of Gly/Thr-X-Ser-X-Gly making up the nucleophilic elbow containing the catalytic

serine [Wong and Schotz, 2002].

The definition of lipases and how they differ to esterases, also acting on ester bonds, have

been discussed over the years. Many lipases have a lid, an amphiphilic loop covering the

active site, which undergoes a structural change upon binding to a lipid surface [Brzozowski

et al, 1991]. Most of the lipases having a lid show interfacial activation, i.e. the lipase activity

increases at the critical micellar concentration. This was for long the definition of a lipase, but

since not all lipases have a pronounced lid or show interfacial activation, the definition was

changed in 1997 to “a carboxyl-esterase that hydrolyse long-chain acyl glycerols” [Verger,

1997].

Lipases have found many industrial applications, as reviewed by Ghanem, 2007 and Houde et

al, 2004. Detergent formulas use proteases, cellulases and lipases with broad substrate

selectivity and good activity in basic environments and at high temperatures. The most

common lipase used in detergents is Thermomyces lanugiosus. It is produced by Novozymes

A/S with trade names like Lipolase® Ultra, Lipo Prime™ and Lipex®. Kinetic resolution is

an additional large area for lipases. For example is BASF producing more than 2 500 tons

annually of enantiopure secondary amines using lipase as catalyst. Lipases are also widely

- 7 -

used in oleochemical transformations, for example to make cocoa butter from cheaper

glycerides and to produce isopropyl myristate for use in cosmetics.

3.1 Reaction mechanism The reaction mechanism of lipases involves the three catalytical residues; serine, histidine and

aspartate/glutamate, which are arranged to lower the pKa of the serine hydroxyl group and

thus enable it to do a nucleophilic attack on the carbonyl carbon of the substrate (scheme 1,

E). The first substrate, an acyl donor, finds its way into the active site of the enzyme and a

Michaelis-Menten complex is formed. The histidine acts as a general base, activating the

serine hydroxyl group, which attacks the carbonyl carbon on the substrate, forming a

tetrahedral intermediate (scheme 1, TI 1). The oxyanion charge is stabilized by hydrogen

bonds to the residues of the oxyanion hole, while the positive charge of the histidine is

stabilized by aspartate/glutamate. The tetrahedral intermediate collapses and by expulsion of

the first leaving group, an alcohol, the acyl enzyme intermediate is formed (scheme 1, AE).

Asp / Glu

O

O

His

H-N N

Ser

H-O

oxyanion hole

Asp / Glu

O

O

His

H-N N-H

Ser

O

oxyanion hole

Asp / Glu

O

O

His

H-N N

Ser

O

oxyanion hole

Asp / Glu

O

O

His

H-N N-H

Ser

O

oxyanion hole

OR1 O

R2

R2

OR2

OOH

R1O R2

O

OHR1

H2O

HO R2

O

acylation

deacylation

Free enzyme (E) Thetrahedral intermediate (TI 1)

Acyl enzyme (AE)Thetrahedral intermediate (TI 2)

Scheme 1. The enzyme mechanism of lipases is exemplified by a hydrolysis reaction. The first substrate, an ester, is attacked by the serine in the enzyme active site creating an acyl enzyme and the first product, an alcohol. The second substrate, a water molecule, attacks the acyl enzyme and a new ester is formed and released from the enzyme.

- 8 -

The tetrahedral intermediates are flanked on both sides by transition states, where bonds are

broken and formed. The transition states are the highest energy barriers the reaction has to

overcome. The deacylation step consists of the release of the substrate in the acyl enzyme.

The second substrate, an acyl acceptor, attacks the carbonyl carbon in the acyl enzyme

forming the second tetrahedral intermediate (scheme 1, TI 2). A proton is transferred, now

from the second substrate via the histidine to the serine oxygen. The second product is

released from the enzyme, which is then ready for the next cycle.

Depending on the substrates, different outcomes can be achieved. An ester as acyl donor will

create an alcohol when forming the acyl enzyme. If the acyl acceptor is water, the product

will be an acid and the overall reaction is hydrolysis of the ester (scheme 1). If the lipase is

used in solvents, an alcohol can be used as acyl acceptor and the product will be a new ester.

In the same way a thiol or amine as acyl acceptor will produce a thioester or amide as product

(scheme 2).

R1O R2

O

R3O R2

O

R3S R2

O

R3NH

R2

O

Acyl enzyme

R1 OH

R3 OH

R3 SH

R3 NH2

Ser OH

Free enzyme

Ser O R2

O

Ser OH

Free enzyme

Ser OH

Free enzyme

Ser OH

Free enzyme

Thioester

Amide

Ester

Scheme 2. Lipases can form esters, thioesters or amides depending on the acyl acceptor.

- 9 -

3.2 Candida antarctica lipase B The lipase B from the yeast Candida antarctica consists of 317 amino acids and weights 33

kDa. The active site residues are serine 105, histidine 224 and aspartate 187. The oxyanion

hole consists of backbone and side group of threonine 40 and the backbone of glutamine 106

[Uppenberg et al, 1994].

Candida antarctica lipase B (CalB) often shows high enantioselectivity and is a popular

catalyst for resolution of alcohols, amines, acids and esters as reviewed by Ghanem, 2007 and

Gotor et al, 2006. Resolution of secondary alcohols and amines are made at industrial scale by

BASF [Schmid et al, 2001]. In polyesterification reactions, CalB shows high performance

and its use have extension to functionalization of polymers as reviewed by Varma et al, 2005.

A number of engineering approaches have been applied on CalB to change the substrate

specificity, alter the enantioselectivity or to make non-conventional reactions like aldol

reactions and Michael-type additions as reviewed by Hult and Berglund 2003 and 2007. CalB

is commercially available from Novozymes A/S and Boehringer Mannheim. In this thesis the

preparation Novozym 435, which is CalB immobilized on an acryl resin, was investigated.

3.3 Rhizomucor miehei lipase Rhizomucor miehei lipase (Rml) is a fungal lipase, has 269 amino acids and weights 29 kDa.

The catalytic residues are serine 144, histidine 257 and aspartate 203 [Brady et al, 1990]. The

backbones of serine 82 and leucine 145 constitute the oxyanion hole [Norin et al, 1994]. Rml

has a pronounced interfacial activation and the lid region is constituted of the amino acid

residues 85 to 91 [Brzozowski et al, 1991].

Rml is frequently used as an industrial catalyst, for example in fat modifications for

cosmetics, bio-diesel and food industry [Houde et al, 2004]. Resolution of chiral compounds

using Rml as catalyst has been reviewed by Alcántara et al, 1998. Lipozyme®, from

Novozymes A/S, is Rml immobilized on ion-exchange resin and was the preparation

investigated in this thesis.

3.4 Active site titration of immobilized lipase The reaction rate is dependent on the amount of active enzyme. To measure the amount active

lipase, the enzyme preparation can be titrated with an active site inhibitor. The inhibitor has a

leaving group whose concentration can be measured by UV-Vis or fluorescence spectroscopy

[Rotticci et al, 2000; Dijkstra et al, 2008].

- 10 -

In paper I, the amount of active enzyme of two immobilized enzyme preparations, Novozym

435 (CalB) and Lipozyme® (Rml), was measured by active site titration. The active site

inhibitor methyl 4-methylumbelliferyl hexylphosphonate (figure 4) was used, and the amount

of 4-methylumbelliferone released was measured using a fluorometer [Fujii et al, 2003;

Magnusson et al, 2005].

O OOPO

OC6H13

Figure 4. The active site inhibitor methyl 4-methylumbelliferyl hexylphosphonate.

The amount of active lipase was measured to 3.3 weight percent on Novozym 435 and 0.14

weight percent on Lipozyme® (table 1). The inhibition of immobilized enzyme was slow

compared to lipase in solution. After four days incubation with the inhibitor, a plateau value

was reached. This may be compared with inhibition of CalB solubilized in an aqueous buffer,

where a plateau value was reached after 30 minutes using the same inhibitor [Hedfors,

unpublished data].

Table 1. Active site titration using methyl 4-methylumbelliferyl hexylphosphonate.

Lipase nmol lipase / g carrier weight percent

Novozym 435 1000 3.3

Lipozyme® 35 0.14

- 11 -

- 12 -

4. LIPASE CATALYZED CHEMOSELECTIVE TRANSACYLATION REACTIONS

In paper I, the chemoselectivity between alcohol and thiol was investigated for the two

lipases CalB and Rml. The transacylation of ethyl octanoate with hexanol and hexanethiol as

acyl acceptors was used (scheme 3). The acyl donor concentration was kept constant to

achieve pseudo one-substrate kinetics. Apparent kinetic constants kcatapp, KM

app and

(kcat/KM)app were determined (table 2).

O R

O

XH

X R

O

OH

lipase

R = C7H15X = S or O

Scheme 3. Transacylation reactions with ethyl octanoate as acyl donor and hexanol or hexanethiol as acyl acceptors.

Rml catalyzed the transacylation reaction with hexanol with a kcat ten times higher than CalB,

while the KM-values were almost the same (table 2). For the reaction with hexanethiol, neither

of the lipase preparations could be saturated with the thiol. The initial rates were directly

proportional to the hexanethiol concentration up to 1.8 M. Concentrations above 1.8 M were

not tested since that would have altered the reaction conditions severely. Only the specificity

ratio, (kcat/KM)app, from the Michaelis-Menten curve could be derived from these experiments.

Rml was 1 600 times more efficient in using hexanethiol as acyl acceptor than CalB (table 2).

In the hexanol reaction, Rml was 20 times more efficient.

- 13 -

Table 2. Apparent kinetic constants for the acyl transfer using ethyl octanoate as acyl donor and hexanol or hexanethiol as acyl acceptor in cyclohexane.

Lipase (kcat/KM)app (s-1M-1) kcatapp (s-1)[a] KM

app (M)[a]

CalB

hexanol 710 14 ± 9 0.019 ± 0.002

hexanethiol 0.0081[b] - >1.8

Rml

hexanol 16000 130 ± 10 0.0084 ± 0.002

hexanethiol 13[b] - >1.8

[a] Non-linear regression of Michaelis-Menten equation.

[b] Calculated from rates as a function of substrate concentrations far below KM.

In the literature, only one KMapp-value for a thiol in a transacylation reaction with a lipase can

be found. In a reaction with oleic acid and butanethiol using Lipozyme as catalyst, the KMapp

of butanethiol is reported to 1.85 M [Caussette et al, 1997]. Reports of transacylation

reactions with thiol substrates, where the two lipases CalB and Rml are compared with respect

to their activities have been made by Weber et al, 1999. It is concluded that CalB and Rml

have the same enzyme activity when compared per gram immobilized enzyme after 24 h

reaction. Using the data from Weber et al. 1999, and combining it with the data from the

active site titration in table 1 (paper I), showed that Rml is 300 times faster than CalB in

transacylation with thiol. This is in line with our results (table 2), where the ratio between Rml

and CalB for the thiol transacylation reaction was 1 600. Iglesias et al. published a paper on

the chemoselectivity between thiols and alcohols for lipases in 1996. They found that only O-

acylation and no S-acylation occurs for mercaptoalcohols in a transacylation reaction using

the lipase from procine pancreatic, Candida cylindracea or Rhitzomucor miehei as catalysts.

The low reactivity of thiols for CalB has been noticed by Öhrner et al. in 1996, investigating

the acylation of secondary alcohols and their amine and thiol analogues.

4.1 Chemoselectivity There was a large difference in specificity constants (kcat/KM) between the two acyl acceptors,

hexanol and hexanethiol, for both CalB and Rml (table 2). The ratios of the specificity of the

two reactions, (kcat/KM)OH / (kcat/KM)SH, gave the chemoselectivity (table 3). Both lipases

- 14 -

showed a much higher selectivity for the hexanol as compared to the hexanethiol, and this

was the major contribution to the great chemoselectivity displayed by the lipases.

Table 3. Chemoselectivity between hexanol and hexanethiol in the acyl transfer reaction with ethyloctanoate as acyldonor.

Catalyst chemoselectivity ratios relative to uncatalyzed

(kcat/KM)OH / (kcat/KM) SH

CalB 88000 730

Rml 1200 10

(knon)OH / (knon)SH

Uncatalyzed[a] 120 1

[a] Background reactions with no enzyme. Vinyl octanoate was used as acyl donor since no reaction was detected within ten days using ethyl octanoate.

To put the chemoselectivity displayed by the two lipases into perspective, the

chemoselectivity of the uncatalyzed reaction was investigated. Using the same reaction

conditions as for the enzymes, no product was detected within ten days of reaction.

Consequently, the activated ester of vinyl octanoate was used. The transacyaltion reaction was

run in a competitive manner, and the ratio of the reaction constants (knon)OH / (knon)SH was

measured to 120. The enzymatic contribution was derived by dividing the chemoselectivity of

the lipases with the chemoselectivity of the uncatalyzed reaction. The enzymatic contribution

of the chemoselectivity was 730 for CalB and ten for Rml (table 3).

The conclusions that can be drawn from table 2 and 3, and from the data in the literature, are

that alcohols react more readily than thiols in transacylation reactions with lipases, and that

the thiol substrates posses high KM-values. Is the high KM the reason to the low kcat/KM for

thiols, as compared to the kcat/KM for the corresponding alcohols? The KM for hexanethiol is

more than two orders of magnitude higher than the KM for hexanol for both CalB and Rml. It

seems that the KM-effect is the dominating, or maybe the only reason to the high

chemoselectivity.

- 15 -

- 16 -

5. LIPASES IN POLYESTER SYNTHESIS

The production of polymers is a big and growing market. Synthetic polymers are found in

many products; plastics, fabrics and building materials just to mention a few. Today polymers

are made by chemical catalysis, but during the last ten years the interest in enzyme-catalyzed

polymerization has grown. The advantages of using enzymes as catalyst are that they often

display high selectivity, are non-toxic and operate under mild reaction conditions; i.e.

temperature and pH [Varma et al, 2005]. The group of enzymes receiving most attention in

polyester synthesis is lipases. Their hydrolytic activity can in a low water environment be

turned to, for example, ester synthesis and produce polyesters when suitable monomers are

provided. Lipases can make polyesters by either polycondensation or ring-opening

polymerization.

5.1 Polycondensation Polyesters can be synthesized from diacids (or diesters) and diols (AA, BB-type monomers)

or from hydroxy acids (AB-type monomers) by polycondensation.

HO OH HO OH

O O

HO O O O OH

O O O O

HOOH HO

OO

O O

O

m

+mn

m m

n n

p

mm

OHO

m

p

lipase

lipase

A)

B)

Scheme 4. Schematic mechanism of polycondensation reaction. A) Example of an AA, BB-type reaction with diacid and diol as monomers. B) Example of AB-type reaction with a hydroxy acid monomer.

- 17 -

To receive polyesters of high molecular mass, water produced during the reaction has to be

removed to shift the equilibrium from hydrolysis to elongation of the polyester. The

conventional way of receiving polyesters by polycondensation polymerization, is to use an

acidic catalyst at high temperature (> 200◦C) and under reduced pressure. At this high

temperature, side reactions such as dehydration of diols can occur [Varma et al, 2005]. By

using a lipase as catalyst, the temperature can be lowered, leading to less side reactions. The

two pioneering works on enzyme catalyzed polycondensation of diacids and diols by

Okumura et al. in 1984 and of hydroxyl acids by Ajima et al. in 1985 have been followed by

many more as reviewed by Varma et al. 2005.

5.2 Ring-opening polymerization Polyesters can be made by a ring-opening polymerization (ROP) reaction if the monomer

consists of a ring-closed ester, for example lactones or lactides (figure 5).

OO

O

O

1 2 3

O O O

O O

m n

Figure 5. Lactones (1), cyclic carbonates (2) and lactides (3) can be used as monomers in ring-opening polymerization.

R OH R O

OO

O

m

OH

m+lipase

R O

OO

O

m

OH

m n+1n+lipase

R O

O

OH

m

Initiation

Propagation

Scheme 5. Schematic mechanism of a ring-opening polymerization (ROP) reaction. The reaction is initiated by an alcohol or water and then the opened monomer can be used for further reaction.

- 18 -

The advantage of using a ring-ester monomer is that no product is formed in the acylation

step. Thus, the atom efficiency is 100 percent and the equilibrium is shifted to elongation

since no by-product is formed which can hydrolyze/cleave the polyester. The ring-opening

polymerization requires an initiator to start the reaction. Thus, by defining the initiator to

monomer ratio the polymer length can be controlled.

Enzyme catalyzed ROP of lactones was first published independently by the groups of Knani

et al. and Uyama and Kobayashi, both in 1993. Organometallic initiators show high activity

towards small lactones (4-7 member rings) and the polymerizability decreases with the ring

size, since the driving force for ROP is the ring-strain [Duda et al, 2002]. Lipases, on the

other hand, show increasing polymerizability with increasing ring size [Duda et al, 2002; van

der Mee et al, 2006]. The reason for this was further investigated for CalB and it was shown

that when the lactone possessed a transoid conformation, which is only possible for larger

lactones, the reactivity in the enzyme increases [van Buijtenen et al, 2007]. With larger

lactones, the polymeric product becomes more like low density poly(ethylene) in its physical

properties, while the ester groups may still make it biodegradable [Focarete et al, 2001]

O

O

E-OH

O

OE-O

E-O

OOH

E-O

OOH

O

OO

nR

O

OOH

H n

n ≥ 1

O

OOH

Rn+1

R OH

A n = 0B n ≥ 1

A n = 0B n ≥ 1

A

B

Scheme 6. The mechanism of ring-opening polymerization; A initiation and B propagation. The first substrate, a ring-closed ester, is opened by the enzyme, creating the acyl enzyme. The second substrate, which is an alcohol or water in the initiation step (A), or the hydroxyl end of a polymer in the propagation step (B) attacks the acyl enzyme and the product is released.

- 19 -

5.3 Selective lipase polymerization One of the major benefits of using lipases as catalysts in polymer synthesis is the possibility

of easily achieving selective polymerization. Enantiopure polyesters can be synthesized by

using ω-methylated lactones and CalB as catalyst [van Buijtenen et al, 2007]. It has also been

shown that CalB can selectively grow polymers from glycopyranosides [Córdova, 1998]. The

selectivity is towards the primary alcohol, being much more reactive than the secondary

alcohols in the glycopyranoside. CalB can grow polyesters selectively from dendrimeric

initiators [Córdova, 2001]. Also the use of a selective lipase in grafting from polymers has

been investigated [Duxbury, 2007]. In both these cases the selectivity is mainly due to the

relative steric hindrance of the hydroxyl groups in a partly grafted polymer/dendrimer

compared to the hydroxyl groups at the ends of the growing polyesters. It has to be

remembered that the substrate must be able to reach the active site of the enzyme for

polymerization or hydrolysis to occur.

5.4 Chemoselective thiol end-functionalization of polyesters Polyesters with free thiol ends are of interest due to their ability to cross-link with enes to

form networks [Dondoni, 2008]. The network makes the polymer more rigid and the

technique is used for example in production of polymeric films used in paints, lacquer and

glue. Thiol end-functionalized poly(ε-caprolactone) has been made chemically by Trollsås et

al, using Al(iOPr)3 as catalyst and α-(2,4-dinitrophenylthio)ethanol as combined initiator and

protection group [Trollsås et al, 1998, Carrot et al, 1999].

In paper II, we showed a novel route to thiol end-functionalized polyesters in a one-pot

synthesis using the chemoselective CalB as catalyst. As backbone in the polymer chain, ε-

caprolactone was used. The polymer was either initiated with mercaptoethanol or terminated

with γ-thiobutyrolactone or 3-mercaptopropionic acid (scheme 7). The enzyme used was the

commercially available Novozym 435, which is CalB immobilized on acrylic resin. The

degree of thiol functionalization was 70-90 percent. Since we utilized the high

chemoselectivity displayed by the lipase, no protection of the thiol group was needed.

For the initiation reaction with mercaptoethanol (A, scheme 7), 70% of the polymers were

initiated with the alcohol moiety of mercaptoethanol, giving the desired thiol end-

functionalized polymer. Of the remaining polymers, 20% were water-initiated and 10% were

initiated with the thiol moiety of mercaptoethanol. Using 6-mercapto-hexanol as initiator, the

degree of thiol functional polymers can be increased to 97% [Takwa et al, 2006].

- 20 -

HSOH O

O

+ HSO

OO

OOH

n-1

O

O

HOO

O

OOH

n-1

+ S

O

HOO

SH

O

O

n

1 2 3

2 4

4

5 6

I)

II)

+ HOO SH

O

O

n4

7 8

HO SH

O

n

n

B)

A)

H2O +

Scheme 7. A) Mercaptoethanol (1) was used as initiatior in the polyesterification of ε-caprolactone (2) giving the thiol end-functionalized product (3). B) Poly(ε-caprolactone) was synthesized from ε-caprolactone (2) initiated with water. I) Poly(ε-caprolactone) (4) was end functionalized by γ-thiobutyrolactone (5) or as in II) with 3-mercaptopropionic acid (7).

Table 4. Analysis of the thiol functionalized polymer products.

Polymer product

I or T I:M T:M Time (h) solvent polymers with -SH (%)

Mn (Da)

Mw/Mn (Da)

Dp

3 1 1:30 24 bulk 70 2900 1.4 26

6 5 5:1 72 MTBE 90 2100 1.8 19

8 7 1:30 24 bulk 70 6700 2.7 59

The polymer product numbers correspond to the numbers in figure 7. I = initiator, T = terminator, M = monomer. The initiator to monomer, or terminator to monomer, ratio is specified. MTBE = methyl tertiary-butyl ether. Mn = number average molecular weight, Mw = weight average molecular weight, both measured by GPC (gel permeation chromatography). Dp = average polymer length.

- 21 -

For the two termination reactions, poly(ε-caprolactone) was first synthesized with water as

initiator. Termination with γ-thiobutyrolactone resulted in 90% thiol end-functionalized

polymers. The attractiveness of using γ-thiobutyrolactone, a ring-closed thioester, as

terminator is that no water by-product is formed, and thus the probability for hydrolysis of the

polymer is reduced. However, a large excess of terminator was needed since the thioester of

γ-thiobutyrolactone was not as good substrate for the enzyme as the esters in the poly-ε-

caprolactone. When 3-mercaptopropionic acid was used as terminator, 70% of the polymers

were thiol end-functionalized.

This new approach with thiol end-functionalization of polyesters using a chemoselective

lipase as catalyst has been followed up within the Biocatalysis group together with the

Coating technology group at KTH, by Takwa et al, giving macromonomers and materials

[Takwa et al, 2006 and 2008; Simpson 2008]. Kato et al. have made polyesters with free thiol

groups along the polyester chain by a polycondensation reaction of dimethyl 2-

mercaptosuccinate and hexane 1,6-diol using CalB as catalyst [Kato et al, 2009]. Also, Kerep

et al. reported that 2-mercaptoethanol initiated ring-opening polymerization of ε-caprolactone

with lipase and microwave irradiation gives a higher chemoselectivity than polymerization

with only lipase as catalyst [Kerep et al, 2007].

- 22 -

ACKNOWLEDGMENTS Jag vill rikta ett stort TACK till följande personer: Prof. Karl Hult. För att jag fick möjlighet att göra mitt examensarbete i biokatalysgruppen och fascineras av enzymer, samt för att jag fick fortsätta som doktorand. Ditt intresse för biokemi smittar av sig och ditt sätt att ifrågasätta allt är utvecklande. Dr. Mats Martinelle. För bra handledning och lärorika diskussioner. Du tar dig alltid tid och gör allt lika noggrant. Prof. Eva Malmström och Emma Östmark. För bra samarbete och intressanta diskussioner om polymerkemi. Alla nuvarande och tidigare medlemmar i biokatalysgruppen. För alla trevliga stunder! För allt jag har lärt mig och all hjälp jag fått med både stort och smått. Nu kan jag sluta isolera mig framför datorn och återgå till labbet och vara mer social. Alla ni andra på plan 2. För trevliga samtal under lunch och fika. Speciellt tack till Lotta och Ela. Till mina vänner. För alla trevliga stunder, i lekparken, över en fika eller en trevlig middag. Till min stora släkt. Mamma och pappa, mormor och morfar, farmor och farfar, alla kusiner, morbröder, mostrar, farbröder och fastrar – ni betyder alla mycket för mig. Även alla släktingar på Davids sida vill jag tacka, speciellt Anna och Thomas. Till min lilla familj. David, Astrid och Irma, ni är så värdefulla för mig. Tack för allt stöd och all er omtanke.

- 23 -

- 24 -

REFERENCES

Ajima A., Yoshimoto T., Takahashi K., Tamura Y., Saito Y., Inada Y., Polymerization of 10-hydroxydecanoic acid in benzene with polyethylene glycol-modified lipase. Biotechnol. Lett. 1985, 7, 303-306. Alcántara A. R., de Fuentes I. E., Sinisterra J. V., Rhizomucor miehei lipase as the catalyst in the resolution of chiral compounds: an overview. Chem. Phys. Lipids 1998, 93, 169-184. Review. Bornscheuer U. T., Buchholz K., Highlights in biocatalysis – Historical landmarks and current trends. Eng. Lipfe Sci. 2005, 5, 309-323. Review. Brady L., Brzozowski A. M., Derewenda Z. S., Dodson E., Dodson G., Tolley S., Turkenburg J. P., Christianses L., Huge-Jensen B., Norskov L., Thim L., Menge U., A serine protease triad forms the catalytic centre of a triacylglycerol lipase. Nature 1990, 343, 767-770. Brzozowski A. M., Derewenda U., Derewenda Z. S., Dodson G. G., Lawson D. M., Turkenburg J. P., Bjorkling F., Huge-Jensen B., Patkar S. A., Thim L., A model for interfacial activation in lipases from the structure of a fungal lipase-inhibitor complex. Nature 1991, 351, 491-494. Cammenberg M., Hult K., Park S., Molecular basis for the enhanced lipase-catalyzed N-acylation of 1-phenylethanamine with methoxyacetate. ChemBioChem 2006, 7, 1745-1749. Carrot G., Hillborn J., Hedrick J. L., Trollsås M., Two general methods for the synthesis of thiol-functional polycaprolactones. Macromolecules 1999, 32, 5171-5173. Caussette M., Marty A., Combes D., Enzymatic synthesis of thioesters in non-coventional solvents. J. Chem. Tech. Biotechnol. 1997, 68, 257-262. Cheetham P. S. J., Applied biocatalysis, 2nd ed. Harwood academic publisher, 2000. Córdova A., Iversen T., Hult K., Lipase-catalyzed synthesis of methyl 6-O-poly(ε-caprolactone)glycopyranosides. Macromolecules 1998, 31, 1040-1045. Córdova A., Synthesis of amphiphilic poly(ε-caprolactone) macromonomers by lipase catalysis. Biomacromolecules 2001, 2, 1347-1351.

- 25 -

Derewenda Z. S., Derewenda U., Dodson G. G., The crystal and molecular structure of the Rhizomucor miehei triacylglyceride lipase at 1.9 Å resolution. J. Mol. Biol. 1992, 227, 818-839. Dijkstra H. P., Sprong H., Aerts B. N. H., Kruithof C. A., Egmond M. R., Gebbink R. J. M. K., Selective and diagnostic labelling of serine hydrolases with reactive phosphonate inhibitors. Org. Biomol. Chem. 2008, 6, 523-531. Dondoni A., The emergence of thiol-ene coupling as a click process for materials and bioorganic chemistry. Angew. Chem. Int. Ed. 2008, 47, 8995-8997. Duda A., Kowalski A., Penczek S., Uyama H., Kobayashi S., Kinetics of the ring-opening polymerization of 6-, 7-, 9-. 12-, 13-, 16-, and 17-membered lactones. Comparison of chemical and enzymatic polymerizations. Macromolecules 2002, 35, 4266-4270. Duxbury C. J., Cummins D., Heise A., Selective enzymatic grafting by steric control. Macromol. Rapid Commun. 2007, 28, 235-240. Focarete M. L., Scandola M., Kumar A. and Gross R. A. J., Physical characterization of poly(omega-pentadecalactone) synthesized by lipase-catalyzed ring-opening polymerization. Polym. Sci. Part B: Polym. Phys. 2001, 39, 1721-1729. Fujii R., Utsunomiya Y., Hiratake J., Sogabe A., Sakata K., Highly sensitive active-site titration of lipase in microscale culture media using fluorescent organophosphorus ester. Biochim. Biophys. Acta 2003, 1631, 197-205. Ghanem A., Trends in lipase-catalyzed asymmetric access to enantiomerically pure/enriched compounds. Tetrahedron 2007, 63, 1721-1754. Review. Gotor-Fernandez V., Busto E., Gotor V., Candida antarctica lipase B: an ideal biocatalyst for the preparation of nitrogenated organic compounds. Adv. Synth. Catal. 2006, 348, 797-812. Houde A., Kademi A., Leblance D., Lipases and their industrial applications – an overview. Appl. Biochem. Biotech. 2004, 118, 155-170. Review. Hult K., Berglund P., Engineered enzymes for improved organic synthesis. Curr. Opin. Biotech. 2003, 14, 395-400. Review. Hult K., Berglund P., Enzyme promiscuity: mechanism and applications. Trends Biotechnol. 2007, 25, 231-238. Review. Iglesias L. E., Baldessari A., Gros E. G., Lipase-catalyzed chemospecific O-acylation of 3-mercapto-1-propanol and 4-mercapto-1-butanol. Bioorg. Med. Chem. Lett. 1996, 6, 853-856. Kato M., Toshima K., Matsumura S., Direct enzymatic synthesis of a polyester with free pendant mercapto groups. Biomacromolecules 2009, 10, 366-373. Kerep P., Ritter H., Chemoenzymatic synthesis of polycaprolactone-block-polystyrene via macromolecular chain transfer reagents. Macromol. Rapid Commun. 2007, 28, 759-766.

- 26 -

Knani D., Gutman A. L., Kohn D. H., Enzymatic polyesterification in organic media. Enzyme-catalyzed synthesis of linear polyesters. I. Condensation polymerization of linear hydroxyesters. II. Ring-opening polymerization of ε-caprolactone. J. Polym. Sci., Part A: Polym. Chem. 1993, 31, 1221. Magnusson A. O., Rotticci-Mulder J., Santagostino A., Hult K., Creating space for large secondary alcohols by rational redesign of Candida antarctica lipase B. ChemBioChem 2005, 6, 1-7.

Miller B. G. and Wolfenden R., Catalytic proficiency: The unusual case of OMP decarboxylase. Annu. Rev. Biochem. 2002, 71, 847-885.

Moss G.P., Basic terminology of stereochemistry. Pure Appl. Chem. 1996, 68, 2193-2222

Nardini M., Dijkstra D. W., α/β Hydrolase fold enzymes: the family keeps growing. Curr Opin. Struc. Biol. 1999, 9, 732-737. Norin M., Haeffner F., Achour A., Norin T., Hult K., Computer modelling of substrate binding to lipases from Rhizomucor miehei, Humicola lanuginosa and Candida rugosa. Protein Sci. 1994, 3, 1493-1503. Novozymes A/S, Annual report 2007. Öhrner N., Orrenius C., Mattson A., Norin T., Hult K., Kinetic resolutions of amine and thiol analogues of secondary alcohols catalyzed by the Candida antarctica lipase B. Enzyme. Microb. Tech. 1996, 19, 328-331. Okumura S., Iwai M., Tominaga T., Synthesis of ester oligomer by Aspargillus niger lipase. Agric. Biol. Chem. 1984, 48, 2805-2813. Ollis D. L., Cheah E., Cygler M., Dijkstra B., Frolow F., Franken S. M., Harel M., Remington S. J., Silman I., Schrag J., Sussman J. L., Verschueren K. H. G., Goldman A., The α/β hydrolase fold. Protein Eng. 1992, 5, 197-211. Rotticci D., Norin T., Hult K., Martinelle M., An active-site titration method for lipases. Biochim. Biophys. Acta 2000, 1483, 132-140. Schmid A., Dordick J. S., Hauer B., Kiner A., Wubbots M., Witholt B., Industrial biocatalysis today and tomorrow. Nature 2001, 409, 258-268. Review. Simpson N., Takwa M., Hult K., Johansson M., Martinelle M., Malmström E., Thiol-functionalized poly(ω-pentadecalactone) telechelics for semicrystalline polymer networks. Macromolecules 2008, 41, 3613-3619. Smith M. B. and March J., March´s advanced organic chemistry, 5th ed. John Wiley & Sons, Inc. 2001. Takwa M., Hult K., Martinelle M., Single-step, solvent-free enzymatic route to α,ω-functionalized polypentadecalactine macromonomers. Macromolecules 2008, 41, 5230-5236.

- 27 -

- 28 -

Takwa M., Simpson N., Malmström E., Hult K., Martinelle M., One-pot difunctionalization of poly(ω-pentadecalactone) with thiol-thiol or thiol-acrylate groups, catalyzed by Candida antarctica lipase B. Macromol. Rapid Commun. 2006, 27, 1932-1936. Trollsås M., Hawker C. J., Hedrick J. L., Carrot G., Hillborn J., A mild and versatile synthesis for the preparation of thiol-functionalized polymers. Macromolecules 1998, 31, 5960-5963. Uppenberg J., Hansen M. T., Patkar S., Jones T. A., The sequence, crystal structure determination and refinement of two crystal forms of lipase B from Candida antarctica. Structure 1994, 2, 293-308. Uyama H. and Kobayashi S., Enzymatic ring-opening polymerization of lactones catalyzed by lipase. Chem. Lett. 1993, 1149-1150. van Buijtenen J., van As B. A. C., Verbruggen M., Roumen L., Vekemans J. A. J. M., Pieterse K., Hilbers P. A. J., Hulshof L. A., Palmans A. R. A. and Maijer E. W., Switching from S- to R-selective in the Candida antarctica lipase B-catalyzed ring-opening of ω-methylated lactones: tuning polymerization by ring size. J. Am. Chem. Soc. 2007, 129, 7393-7398. van der Mee L., Helmich F., de Bruijn R., Vekemans J. A. J. M., Palmans A. R. A., Maijer E. W., Investigation of lipase-catalyzed ring-opening polymerizations of lactones with various ring sizes: kinetic evaluation. Macromolecules 2006, 39, 5021-5027. Varma I. K., Albertsson A-C., Rajkhowa R., Srivastava R. K., Enzyme catalyzed synthesis of polyesters. Prog. Polym. Sci. 2005, 30, 949-981. Review. Verger R., Interfacial activation of lipases: facts and artifacts. Trends Biotechnol. 1997, 15, 32-38. Review. Voet D., Voet J. G., Biochemistry. 2nd ed. John Wiley & Sons, Inc. 1995. Weber N., Klein E., Mukherjee K. D., Long-chain acyl thioesters prepared by solvent-free thioesterification and transthioesterification catalyzed by microbial lipases. Appl. Microbiol. Biotechnol. 1999, 51, 401-404. Wolfenden R., Degrees of difficulty of water-consuming reactions in the absence of enzymes. Chem. Rev. 2006, 106, 3379-3396. Wong H., Schotz M., The lipase gene family. J. Lipid Res. 2002, 43, 993-999. Review.