Serine Hydrolase Selectivity - DiVA portal

70
KTH Biotechnology Serine Hydrolase Selectivity Kinetics and applications in organic and analytical chemistry Doctoral thesis by Anders Hamberg Department of Biochemistry School of Biotechnology Royal Institute of Technology (KTH) Stockholm 2010

Transcript of Serine Hydrolase Selectivity - DiVA portal

Page 1: Serine Hydrolase Selectivity - DiVA portal

KTH Biotechnology

Serine Hydrolase Selectivity

Kinetics and applications in organic and analytical

chemistry

Doctoral thesis by

Anders Hamberg

Department of Biochemistry School of Biotechnology

Royal Institute of Technology (KTH) Stockholm 2010

Page 2: Serine Hydrolase Selectivity - DiVA portal

©Anders Hamberg School of Biotechnology Royal Institute of Technology (KTH) AlbaNova University Centre 106 91 Stockholm ISBN 978-91-7415-663-8 TRITA-BIO-Report 2010:12 ISSN 1654-2312 Cover: GC-chromatogram of mono- and diacetylated 1,2-ethanediol obtained from reactions catalyzed by wild-type (red) and Ala281Val CALB. The amino acid residues are Ser105, His 224 and Val281 in the active site of the mutated variants.

ii

Page 3: Serine Hydrolase Selectivity - DiVA portal

Abstract The substrate selectivities for different serine hydrolases were utilized in various applications, presented in papers I-VI. The articles are discussed in the thesis in view of the kinetics of the enzyme catalysis involved. In paper I the enantioselectivities towards a range of secondary alcohols were reversed for Candida antarctica lipase B by site directed mutagenesis. The thermodynamic components of the enantioselectivity were determined for the mutated variant of the lipase. In papers II-III Candida antarctica lipase B was engineered for selective monoacylation using two different approaches. A variant of the lipase created for substrate assisted catalysis (paper II) and three different variants with mutations which decreased the volume of the active site (paper III) were evaluated. Enzyme kinetics for the different variants were measured and translated into activation energies for comparison of the approaches. In papers IV and V three different enzymes were used for rapid analysis of enantiomeric excess and conversion of O-acylated cyanohydrins synthesized by a defined protocol. Horse liver alcohol dehydrogenase, Candida antarctica lipase B and pig liver esterase were sequentially added to a solution containing the O-acylated cyanohydrin. Each enzyme caused a drop in absorbance from oxidation of NADH to NAD+. The product yield and enantiomeric excess was calculated from the relative differences in absorbance. In paper VI a method for C-terminal peptide sequencing was developed based on conventional Carboxypeptidase Y digestion combined with matrix assisted laser desorption/ionization mass spectrometry. An alternative nucleophile was used to obtain a stable peptide ladder and improve sequence coverage.

iii

Page 4: Serine Hydrolase Selectivity - DiVA portal

Sammanfattning Substratselektiviteter för olika serinhydrolaser har använts i olika sammanhang som presenteras i artikel I-VI. I avhandlingen diskuteras kinetiken för de enzymer som använts i arbetet som beskrivs i artiklarna. I artikel I så har sätesriktad mutagenes resulterat i omvänd enantioselektiviteten för lipas B från Candida antarctica mot ett antal sekundära alkoholer. De termodynamiska komponenterna för enantioselektiviteten bestämdes för den muterade varianten av lipaset. I artikel II-III så har två olika tillvägagångssätt använts för omarrangering av lipase B från Candida antarctica för selektiv monoacylering av dioler. En variant av lipaset tillverkad för substratassisterad katalys och tre olika varianter med mutationer som minskade aktiva sätets volym utvärderades. Enzymkinetiken för de olika varianterna mättes och omvandlades till aktiveringsenergier för jämförelse av tillvägagångssätten. I artikel IV och V beskrivs en metod för snabbanalys av enantiomert överskott och omsättning hos O-acylerade cyanohydriner. Tre olika enzymer, alkoholdehydrogenas från hästlever, lipas B från Candida antarctica och grisleveresteras sattes i nämnd ordning en lösning innehållande den O-acylerade cyanohydrinen. Varje orsakade en minskning i absorbans genom oxidering av NADH till NAD+. Utbyte och enantiomert överskott beräknades från de relativa absorbansskillnaderna. I artikel VI beskrivs en metod för C-terminalsekvensering av peptider genom konventionell trunkering med hjälp av karboxypeptidas Y. Fragmenterade peptider analyserades med matrix assisted laser desorption/ionization-masspektrometri. För att skapa en stabil peptidstege och därmed åstadkomma en förbättrad sekvenstäckning så tillsattes en alternativ nukleofil till trunkeringslösningen.

iv

Page 5: Serine Hydrolase Selectivity - DiVA portal

List of publications I Magnusson A., Takwa M., Hamberg A., Hult K., An S-

selective lipase was created by rational redesign and the enantioselectivity increased with temperature, Angewandte Chemie-International Edition, 2005, 44, 4582-4585.

II Hamberg A., Magnusson A., Hu F. J., Hult K., , Selective

monoacylation of diols by substrate assisted catalysis in T40A CALB, Submitted manuscript, 2010.

III Hamberg A., Maurer S., Hult K., Rational engineering of

CALB for selective monoacylation of diols, Manuscript, 2010.

IV Hamberg A., S. Lundgren, M. Penhoat, C. Moberg, K.

Hult, High-throughput method for enantiomeric excess determination of O-acetylated cyanohydrins, Journal of the American Chemical Society, 2006, 128, 2234-2235.

V Hamberg A, S. Lundgren, E. Wingstrand, C. Moberg, K.

Hult, High Throughput Synthesis and Analysis of Acylated Cyanohydrins, Chemistry: a European Journal, 2007, 13, 4334-4341.

VI Hamberg A, M. Kempka, J. Sjödal, J. Roerade, K. Hult, C-

terminal ladder sequencing of peptides using an alternative nucleophile in carboxypeptidase Y digests, Analytical Biochemistry, 2006, 357, 167-172.

v

Page 6: Serine Hydrolase Selectivity - DiVA portal

vi

Page 7: Serine Hydrolase Selectivity - DiVA portal

Table of contents 1 Introduction ...................................................................................... 1

1.1 Biocatalysis ............................................................................. 1 1.2 Substrate specificity................................................................. 2 1.3 Serine hydrolases .................................................................... 5

2 General methods ............................................................................... 8 2.1 Molecular modelling ............................................................... 8 2.2 Protein production and purification ........................................ 8 2.3 Enzyme-catalyzed reactions .................................................... 9

3 Enantioselectivity engineering.......................................................... 11 3.1 Activation energies................................................................ 12 3.2 Thermodynamic components ............................................... 15

4 Selective monoacylation of diols ...................................................... 17 4.1 Substrate assisted catalysis..................................................... 18 4.2 Restricting the size of the active site...................................... 21 4.3 Comparison of methods ....................................................... 24

5 Determination of enantiomeric excess ............................................. 28 5.1 Screening for enantioselectivity ............................................. 28 5.2 Enzymatic determination of enantiomeric excess .................. 30 5.3 Enzymatic analysis of O-acylated cyanohydrins..................... 31 5.4 Substrate specificity............................................................... 33 5.5 Results and discussion........................................................... 35

6 Amino acid sequence determination ................................................ 41 6.1 Peptide sequencing ............................................................... 41 6.2 Carboxypeptidase mediated sequencing and competing nucleophiles................................................................................ 44 6.3 Substrate specificity............................................................... 45 6.4 Identification frequency ........................................................ 47 6.5 Results .................................................................................. 48

Concluding remarks............................................................................ 52 Acknowledgements ............................................................................ 54 List of abbreviations ........................................................................... 56 References .......................................................................................... 57

vii

Page 8: Serine Hydrolase Selectivity - DiVA portal

viii

Page 9: Serine Hydrolase Selectivity - DiVA portal

Introduction

1 Introduction Enzymes are proteins developed in nature for catalyzing reactions required for survival of all organisms. The high effectiveness and specificity of enzymes have made them suitable for use in several different areas, such as organic synthesis and analytical chemistry. The use of enzymes as catalysts in organic synthesis has increased over the last decades, making them an important part of modern chemistry.[1-3] In analytical chemistry, the use of enzymes is extensive. Enzymatic methods in analytical chemistry have been described in several books and reviews.[4-6]

1.1 Biocatalysis A catalyst is generally defined as a substance which speeds up a chemical reaction without being consumed in the process. In turn, the definition of an enzyme is a biological catalyst. An example commonly used to illustrate the catalytic power of enzymes is the decarboxylation of orotidine 5’-phosphate.[7] Spontaneous decarboxylation of the substance occurs at a rate giving a half-time of 78 million years. When the reaction is catalyzed by the enzyme orotidine 5’-phosphate decarboxylase from yeast the rate is enhanced by a factor of 2×1023 M-1 making it one of the most proficient enzymes known. Catalysis is achieved by lowering the energy required for a reaction to take place, the activation energy. Enzymes reduce the activation energy by stabilizing the transition-state after binding a substrate (figure 1.1).[8]

1

Page 10: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

G

Reaction Coordinate

E + SES

ES‡

E + P

E + S ES ES‡ E + P Figure 1.1 The energy profile diagram for an enzyme-catalyzed reaction. The substrate and enzyme complex (ES) is formed from the free components (E and S). The complex then passes a transition-state (ES‡) before releasing the product (P). The efficiency of catalysis for an enzyme-catalyzed reaction is determined by catalytic constants kcat and KM. The constant kcat is the turnover number of an enzymatically catalyzed reaction and is proportional to the maximum reaction rate, Vmax, for a specific enzyme concentration, obtained at substrate saturation of the enzyme. KM is the Michaelis constant, which is defined as the substrate concentration needed to obtain ½ Vmax. KM is often described as a measure of the stability of the enzyme-substrate complex (ES, figure 1.1), which is true for single substrate kinetics when product formation is much slower than the dissociation of the substrate from the substrate-enzyme complex. For the reaction to take place, the ES complex has to pass transition-state (ES‡, figure 1.1), which occur at a rate determined by kcat.

1.2 Substrate specificity A single enzyme may catalyze conversion of various substrates at different rates. For example, isomerisation of glucose into fructose can be catalyzed by the enzyme xylose isomerase. However, the natural

2

Page 11: Serine Hydrolase Selectivity - DiVA portal

Introduction

reaction for xylose isomerase is isomerisation of D-xylose into D-xylulose.[9] Still, the activity towards glucose is sufficient for industrial production of high-fructose corn syrup, which is one of the largest industrial processes involving enzymes.[10] The efficiency of an enzyme-catalyzed reaction is quantified by the substrate specificity, defined by the specificity constant, kcat/KM, for an enzyme towards a substrate. Variations in substrate specificities for a single enzyme towards different substrates result in selectivity for an enzyme towards a substrate over another. The selectivity is calculated from the ratio of the kcat/KM-values towards the different substrates. As an example, in transacylation reactions catalyzed by lipases, alcohols are preferred over amines as nucleophiles.[11] Another example is that proteases can be highly specific depending on the protein sequence.[12, 13] An enzyme selectivity based on sometimes even more subtle differences between the substrates is enantioselectivity, which plays an important part in nature as well as in modern chemistry and is perhaps the most discussed selectivity for various enzymes. Enantioselectivity is derived from differences in substrate specificity towards the enantiomers of a certain substrate. This means that one enantiomer reacts faster than the other when a racemic mixture is subjected to an enantioselective enzyme. The enantiomeric ratio, E, is a measure of the enantioselectivity for an enzyme towards a substrate. E is defined by equation 1.1 where A denotes the fast reacting enantiomer and B denotes the slow reacting enantiomer.

BM

cat

AM

cat

Kk

Kk

E⎟⎠⎞⎜

⎝⎛

⎟⎠⎞⎜

⎝⎛

= (equation 1.1)

3

Page 12: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

The product in an enantioselectively catalyzed reaction will exhibit an enantiomeric excess, ee, defined by equation 1.2, where A and B represents the major and minor enantiomer, respectively.

[ ] [ ][ ] [ ]BA

BAee+−

= (equation 1.2)

In kinetic resolution reactions, E can be calculated from equation 1.3, ee-values are determined for both substrate and product, denoted as eeS and eeP, respectively.[14, 15]

⎟⎟⎟⎟

⎜⎜⎜⎜

+

+

⎟⎟⎟⎟

⎜⎜⎜⎜

+

=

P

S

S

P

S

S

eeee

ee

eeee

ee

E

1

1ln

1

1ln

(equation 1.3)

Substrate specificity can be described in terms of energy differences. The energy required to go from the ground state to transition state (from E + S to ES‡, figure 1.1) is defined as the activation energy, which can be calculated by equation 1.4.[8] Hence, the substrate specificity for an enzyme towards a substrate can be correlated to activation energy.

hTkRT

Kk

RTG B

M

cat‡ lnln +−=Δ (equation 1.4)

The activation energy can be broken down in its intrinsic components of enthalpy, ΔH‡, and entropy, ΔS‡, as defined by equation 1.5.[8] The enthalpy and entropy contribution is based on binding and degrees of freedom, respectively, in transition state compared to the ground state.

4

Page 13: Serine Hydrolase Selectivity - DiVA portal

Introduction

‡‡‡ STHG Δ−Δ=Δ (equation 1.5)

When selectivities are discussed, equations 1.1 and 1.4 can be rewritten into equation 1.6, which describes the difference in activation energies between the reactions of two substrates, A and B.

BM

cat

M

cat

Kk

Kk

RTG⎟⎠⎞⎜

⎝⎛

⎟⎠⎞⎜

⎝⎛

−=ΔΔ Aln (equation 1.6)

The equation can for example be used for calculating energy differences in transition state between enantiomers. It can also be used to determine the difference in activation energy when different catalysts are used for the same reaction. In the example mentioned in section 1.1 considering decarboxylation of orotidine 5’-phosphate, the spontaneous versus the enzymatically catalyzed reaction yields a difference in activation energy of more than 130 kJ/mol.[7]

1.3 Serine hydrolases Presented in the following chapters of this thesis are enzyme based methods for organic synthesis and analytical chemistry. The specificities of the serine hydrolases Candida antarctica lipase B, CALB, (also known as Pseudozyma antarctica lipase B)[16] and carboxypeptidase Y, CPY, have been utilized. An example of the reaction mechanism for a transacylation reaction catalyzed by CALB is illustrated in figure 1.2. The mechanism of serine hydrolases is divided in acylation and deacylation. The tetrahedral intermediates passed in each step are regarded as stabilized transition states.

5

Page 14: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

NN

His224

HOO

Asp187

OH

Ser105

NN

His224

HOO

Asp187

O

Ser105

HO

R

R'

O

RO R'

O

NN

His224

HOO

Asp187

O

Ser105

R'

O

HO R

Oxyanion hole

NN

His224

HOO

Asp187

O

Ser105

HO

R''

R'

O

Oxyanion hole

HO R''

R''O R'

O

A) B)

C)D)

Figure 1.2. Reaction mechanism of transacylation and ester hydrolysis catalyzed by CALB. An ester binds to the free enzyme, A, and a tetrahedral intermediate, B, is formed. The alcohol part of the ester leaves and an acyl enzyme is formed, C. A second tetrahedral intermediate, D, is formed after nucleophilic attack by a second alcohol (or water if the R’’ group is hydrogen). The newly formed ester (or carboxylic acid) leaves, completing the catalytic cycle. Chapters 3 and 4 describe work where CALB was engineered to alter the selectivity of the enzyme (papers I-III). The example in chapter 3 (paper I) explains the reversed enantioselectivity for CALB towards a range of secondary alcohols as a result of a single point mutation. In the following chapter (papers II and III), CALB is rationally engineered for increased selectivity towards diol over monoester as substrates in transacylation reactions. Chapters 5 and 6 describe the development of two different analysis methods mediated by serine hydrolases. Substrate specificities for the involved enzymes plays an important part in both cases. Chapter 5 describes the development of a high-throughput

6

Page 15: Serine Hydrolase Selectivity - DiVA portal

Introduction

7

screening method for library derived cyanohydrins (based on papers IV and V). In this case differences in substrate specificity, namely high enantioselectivity, were needed for determination enantiomeric excess. In chapter 6 (based on paper VI), a peptide sequencing method was developed. Differences in substrate specificity towards various substrates posed a problem in this case, which was decreased by adding a competing nucleophile.

Page 16: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

2 General methods The choice of methods when working in biochemistry is essential for making accurate scientific evaluations. Additionally, the ability to make predictions and outline future work are processes dependent on available methods. Below is a discussion of the methods used for the work presented in this thesis, full description of the methods can be found in papers I-VI.

2.1 Molecular modelling A valuable tool in the work discussed in section 4.2 was the computer modelling of tetrahedral intermediates, corresponding to transition states in the active site of CALB (paper III). Visual analysis of the tetrahedral intermediates obtained after a series of energy minimizations and molecular dynamics was sufficient to predict a change in selectivity towards a range of diols over monoesters. The software used for modelling was YASARA[17] (Yet Another Scientific Artificial Reality Application), which is a comprehensible tool for modelling and evaluation of protein structures.

2.2 Protein production and purification In the work presented in chapters 5 and 6 (papers IV-VI) most of the enzymes were purchased in a ready to use form. However, all CALB variants that were used for the work presented in chapters 3-5 (papers

I-V) were from in-house productions obtained by the use of different expression systems. The Trp104Ala variant was expressed using an alcohol oxidase promoter, AOX1, in Picchia pastoris as explained in earlier work.[18] A high protein yield (g/l) was obtained by fed-batch temperature limited fermention in a bioreactor. Working with the AOX1-expression system is generally considered a laborious task since the amount of methanol has to be under strict control in order not to harm the cells. For this reason, CALB has been introduced into another

8

Page 17: Serine Hydrolase Selectivity - DiVA portal

General methods

expression system, using a glyceraldehyde-3-phosphate promoter, GAP, in P. pastoris.[19] The method has significantly facilitated small-scale cultivations in shake flasks and was used for the production of Thr40Val CALB. The use of P. pastoris as a host organism is excellent for large scale production of CALB in cultivations with high cell density. For creating and evaluating new CALB variants, relatively small amounts of the enzyme is required, why an expression system using Escherichia coli as host organism has been developed.[20] The method provides a simple route for mutating the enzyme and was used for the production of Ala281Glu and Ala281Val CALB. An additional advantage with the expression of CALB in E. coli, is that a His-tag is attached to the protein, which is useful for purification. When expressing CALB in P. pastoris, the protein is normally purified using hydrophobic interaction chromatography, HIC. If the AOX1 expression system is used, size exclusion chromatography, SEC, is often needed as an additional purification step to remove the alcohol oxidase. In attempts to purify Trp104Ala, a problem was encountered with the stability of the protein. The large cavity in the interior of the enzyme caused by the mutation made it sensitive to high ionic strength and the activity was lost during HIC purifications. Instead of purifying the protein it was immobilized on carrier beads directly from the media.

2.3 Enzyme-catalyzed reactions The reactions used for evaluations in chapters 3 and 4 were carried out in organic solvent, which is enabled by immobilisation of CALB (papers I-III). Furthermore, the use of immobilized enzyme allowed measurements of enantiomeric ratios at high temperatures (70° C) necessary for the work discussed in chapter 3 (paper I). The use of organic solvent as reaction media also had the advantage of the ability to perform analysis by gas chromatography, GC, which was especially important for the work discussed in chapter 3 (paper I). Measurements of the enantiomeric excess of the substrate and product from the

9

Page 18: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

10

transacylation reaction of the secondary alcohols enabled straightforward E-determinations according to equation 1.3. However, determination of E for wild-type CALB towards 1-phenylethanol was done by measuring the kcat/KM-values towards each enantiomer separately. In the work presented in chapters 5-6, the reactions were performed in aqueous buffer and analysed spectrophotometrically (papers IV-V) or by MALDI-MS (paper VI). The amount of enzyme loaded on the carrier beads for the different CALB variants discussed in chapters 3-4 were verified by active site titration. This was required for accurate determination of the presented kcat- and kcat/KM-values. Total protein concentrations can be measured by for example absorbance, but the risk of contaminations or partially inactive enzyme may cause deviations from the actual amount of active enzyme. By measuring the amount of immobilized active enzyme directly on the beads these sources of variation could be circumvented.

Page 19: Serine Hydrolase Selectivity - DiVA portal

Enginering enantioselectivity

3 Engineering enantioselectivity Molecular chirality plays an important role when designing compounds for use in various fields such as drug development and agricultural chemistry. Kinetic resolution is a well established method for obtaining enantiopure compounds.[21, 22] The enantiomers in a racemic mixture are separated in an enantioselectively catalyzed reaction. In dynamic kinetic resolution the method is integrated with a racemisation of the substrate, improving the reaction yield.[23, 24] The access to enantioselective catalysts is crucial in order to accomplish enantiopurity in kinetic resolution methods, and can often be accomplished by the use of biocatalysts.[25-27] The enantioselectivity for CALB towards secondary alcohols along with its ability to withstand high temperatures in immobilized form have made it suitable for use in dynamic kinetic resolution.[28] A problem when using lipases for resolution of secondary alcohols is their preference only towards the (R)-enantiomer, formulated as Kazlauskas rule.[29] Subtilisins have shown functional as catalysts in dynamic resolution of the (S)-secondary alcohols with some limitations regarding reaction temperatures.[27, 30] An engineered lipase with (S)-selectivity towards secondary alcohols would be of great value. Several examples can be found in the literature regarding engineering of enantioselectivity for various enzymes.[2, 31, 32] The degree of rationality involved in enzyme engineering may vary from totally random to site specific.[33-36] Lately, semi-rational approaches have been demonstrated successful for improving and altering enantioselectivity for various enzymes.[37, 38] In paper I, enantioselectivities for CALB towards a range of secondary alcohols were reversed from (R)- to (S)-preference by rational design. Substitutions of the Trp104 residue in CALB had previously shown increased activity towards large symmetric secondary alcohols.[18] The Trp104 residue is important for limiting the space in the stereo-specificity pocket. When the Trp104 is mutated into alanine, a

11

Page 20: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

cavity is created and the pocket is enlarged. Large substituents which would normally not fit in the stereo-specificity pocket of wild-type CALB can be accommodated in the enlarged space created by the Trp104Ala mutation. Recent studies have demonstrated that CALB Trp104Ala can be a powerful tool in kinetic resolution of bulky phenyl allyl secondary alcohols.[39]

3.1 Activation energies The reversed enantioselectivity caused by the Trp104Ala mutation in CALB is evident in figure 3.1. Only the (R)-1-phenylethanol is acylated when wild-type CALB was used as catalyst. When the Trp104Ala variant catalyzes the acylation reaction, the (S)-enantiomer reacts faster.

Figure 3.1. GC chromatogram showing (S)- and (R)-1-phenylethanol butyrate, from right to left, obtained from transacylation reactions of catalyzed by wild-type and Trp104Ala CALB, denoted as dashed and solid lines, respectively. Individual kcat- and KM-values were determined for wild-type and Trp104Ala CALB towards (R)- and (S)-1-phenylethanol (paper I). The resulting kcat/KM-values can be used to calculate the activation energies, ΔG‡, individually for the acylations of each enantiomer catalyzed by both of the CALB variants (figure 3.2). A big difference is observed

12

Page 21: Serine Hydrolase Selectivity - DiVA portal

Enginering enantioselectivity

between the activation energies for the acylations of (R)- and (S)-1-phenylethanol catalyzed by wild-type CALB, which is derived from its extreme enantioselectivity towards the substrate (E=1 300 000). When Trp104Ala CALB catalyzes the reaction, the enantioselectivity is reversed (E=6.6).

0

10

20

30

40

50

60

70

80

90

100

wild type Trp104Ala

ΔG

‡(k

J/m

ol)

Figure 3.2. The transition state energies for acylation of (R)- and (S)-1-phenylethanol, shown in black and white, respectively. The reactions are catalyzed by wild-type and Trp104Ala CALB. The difference between activation energies for the acylations of the two enantiomers is shifted, favouring (S)-1-phenylethanol as substrate. The reversed enantioselectivity in the Trp104Ala variant is caused by changes in activation energies for both enantiomers of 1-phenylethanol, increasing by 10 kJ/mol when (R)-1-phenylethanol and decreasing by 30 kJ/mol when (S)-1-phenylethanol acted as substrates. Thereby, the reduced activation energy for the reaction of (S)-1-phenylethanol can be seen as the predominant cause of the reversed enantioselectivity for Trp104Ala compared to wild-type CALB.

13

Page 22: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

-20

0

20

40

60

80

100

E+S

ES

ES‡

E+P

ΔG

‡(k

J/m

ol)

Figure 3.3. Energy profile diagram for the acylation of (S)-1-phenylethanol catalyzed by wild-type and Trp104Ala CALB represented in the graph by drawn and dashed lines, respectively. The E+S, ES and ES‡ represent the acyl enzyme and alcohol in dissociated, bound and bound in transition state forms, respectively. The E+P represent the enzyme and ester product, dissociated. The change in kcat/KM towards (S)-1-phenylethanol caused by the Trp104Ala mutation originates from changes in both KM and kcat (paper

I). Decreased KM together with increased kcat in Trp104Ala compared to wild-type CALB contributes to an increased kcat/KM. Translated into energies the change in KM reduces the activation energy by 1.9 kJ/mol, while the change in kcat contributes by 28 kJ/mol (figure 3.3). Thus, lowering of the energy required to reach transition state when the substrate is bound by the enzyme is the major contributor to the decreased activation energy for acylation of (S)-1-phenylethanol caused by the Trp104Ala mutation. The drastic change in enantioselectivity towards 1-phenylethanol (8 300 000-fold) caused by the Trp104Ala mutation in CALB can be summarized as mainly an effect of increased kcat (64 000-fold).

14

Page 23: Serine Hydrolase Selectivity - DiVA portal

Enginering enantioselectivity

3.2 Thermodynamic components The activation energies, ΔG‡-values, for the different enantiomers of 1-phenylethanol were discussed in relation to enantioselectivity in the previous section. Enantioselectivities for Trp104Ala CALB towards various secondary alcohols were used for calculating differences in transition-state energies, ΔΔG‡, according to equation 1.6 (paper I). By measuring enantioselectivities at varying temperatures, the ΔΔG‡-values were broken down into intrinsic thermodynamic components, enthalpy and entropy, ΔΔH‡ and ΔΔS‡, by using equation 1.5. Interestingly, the enantioselectivity for Trp104Ala CALB towards some of the tested secondary alcohols increases with temperature, suggesting that entropy contributes positively to the enantioselectivity. The enantioselectivity towards 1-phenylethanol, 1-phenylpropanol and 3-hexanol, preferring the (S)-enantiomer, was shown to be not only assisted but driven by entropy. The (R)-enantiomer is favoured by enthalpy, while the entropy contribution causes the enantioselectivity to switch towards the (S)-enantiomers. The observation indicates an increased degree of freedom in transition state as driving force for the enantioselectivity towards the substrates. When the (S)-enantiomer is acylated by Trp104Ala CALB, the degree of freedom in transition state is higher than in the acylation of the (R)-enantiomer catalyzed by the same enzyme. The enantioselectivity for Trp104Ala CALB towards 1-phenylethanol and its thermodynamic components were further evaluated in different solvents. Measurements of the temperature dependence of the enantioselectivity were done using acetonitrile, cyclohexane and cis-decaline as solvents. A trend can be seen for the various solvents, causing the enantioselectivity to increase with the size of the solvent molecules. The largest difference in enantioselectivity between the reactions run in various solvents was observed at the highest

15

Page 24: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

16

temperatures used in the assay as the enantioselectivity was driven by entropy, in resemblance with the reactions described above. A possible explanation for entropy as the driving force for enantioselectivity could be found in the release of solvent. The enlarged stereospecificity pocket in Trp104Ala CALB has to accommodate the large substituent in order for the (S)-enantiomer to react. Consequently, the solvent molecule(s) which would otherwise occupy the pocket has to be released, leading to an increased degree of freedom in the system. A small solvent molecule may be accommodated in the stereospecificity pocket in a more relaxed state and more freedom may be gained by releasing a larger solvent molecule.

Page 25: Serine Hydrolase Selectivity - DiVA portal

Selective monoacylation of diols

4 Selective monoacylation of diols Diols can easily be acylated using a wide variety of common methods. Enzymatic or more traditional chemical methods can be used. However, a challenge arises when a high monoester yield is desired. The monoester formed from acylation of a diol can easily be further acylated into a diester. An efficient way to improve the monoester yield is to use an excess of diol. However, many diols are expensive and downstream processing of monoester preparations with low product purity is laborious. Therefore, selective methods for high monoester yield obtained at high diol conversion are of great value in preparative monoester synthesis. The product purity in acylation of diols can be measured in monoester excess (equation 4.1).

[ ] [ ][ ] [ ]diestermonoester

diestermonoesterexcessmonoester+−

= (equation 4.1)

The selectivity for CALB towards diol over monoester in transacylation reactions was increased by site directed mutagenesis. Two different approaches to improve the selectivity for CALB are presented (papers

II and III). The methods vary by the residues in CALB that were targeted for mutation. Additionally, the produced CALB variants utilize different chemical properties to discriminate between diol and monoester as substrates. In the first example, a CALB variant with a mutation in the oxyanion hole was used as catalyst (paper II). The variant made use of the free hydroxyl group, present in addition to the nucleophilic hydroxyl group in the diol. Increased selectivity towards diol over monoester was observed as a result of substrate assisted catalysis. In the second example, steric hindrances towards large substrates were introduced by mutations restricting the size of the active site of CALB (paper III). Monoesters were discriminated as substrates in the transacylation reaction as a result of the mutations.

17

Page 26: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

4.1 Substrate assisted catalysis A hydrogen bond was removed in the oxyanion hole in the active site of CALB by the Thr40Ala mutation (paper II). The hydrogen bond provided by Thr40 is important for stabilizing the transition state (figure 4.1 A). The mutation resulted in increased transition state energy, causing the activity to drop severely. Previous studies of Thr40Ala CALB have shown that the lost hydrogen bond in transition state can be replaced by a hydroxyl group on the acyl part of the substrate.[40] Some of the lost stability in transition state is thereby recovered by substrate assisted catalysis. If no hydroxyl group or other hydrogen bond donor is present on the substrate, the activity of Thr40Ala remains low. A) B) C)

O O

O OH

Ser 105

N H

NH

Thr40Ala

Gln106

O O

O

Ser 105

N H

NH

Thr40

Gln106

OH

OH

O O

O

Ser 105

N H

NH

Thr40Ala

Gln106

O

O

Figure 4.1. Tetrahedral intermediates in acetylations of 1,2-ethanediol (A and B) and 1,2-ethanediol acetate (C). The reaction is catalyzed by wild-type (A) and Thr40Ala CALB (B and C). The free hydroxyl group of the diol may interact with the oxyanion when the reaction is catalyzed by Thr40Ala CALB (B). Substrate assisted catalysis could also be accomplished by using a diol as substrate in a transacylation reaction. A free hydroxyl group is provided in transition state when the diol acts as a nucleophile. Interactions between the free hydroxyl group and the oxyanion may restore some of the transition state stability lost in Thr40Ala CALB (figure 4.1 B). On the other hand, when the formed monoester acts as nucleophile the hydroxyl group is protected from interactions with the oxyanion, and the destabilisation of the transition state will remain (figure 4.1 C). Thereby, kcat/KM towards the diol will be higher than towards the monoester.

18

Page 27: Serine Hydrolase Selectivity - DiVA portal

Selective monoacylation of diols

Measurements of kcat/KM for wild-type and Thr40Ala CALB towards a range of different diols and their corresponding monoesters in transacylation reactions were carried out (paper II). As expected, the increased energies in transition state caused by the Thr40Ala mutation could partially be recovered when diols acted as nucleophiles in the reactions, resulting in improved monoester excess at a given diol conversion As an example, the monoester excess at around 33% diol conversions were 41% and 97% for the wild-type and Thr40Ala CALB catalyzed reactions, respectively (figure 4.2).

min12.4 12.6 12.8 13 13.2 13.4

counts

4000

6000

8000

10000

12000

FID1 A, (080108\CT40A128.D) FID1 A, (080108\C_VT_2.D)

Figure 4.2. GC chromatogram displaying mono- and diacetylated 1,2-ethanediol, from left to right, formed in acylation reactions catalyzed by wild-type and Thr40Ala CALB, represented by dashed and drawn lines, respectively. The product measurements were done at 32% and 33% conversion for wild-type and Thr40Ala catalyzed reactions, respectively. The selectivities increased around 2-4 times towards diols over their corresponding monoester as a result of the mutation. Similar selectivities towards the diols over their corresponding monoesters were also detected for Thr40Val CALB, which was produced as a compliment to the Thr40Ala variant. However, the expressed amounts of enzyme were

19

Page 28: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

low and no quantitative measurements of the selectivity were conducted. The selectivities for Thr40Ala and wild-type CALB varied depending on the diol acting as a nucleophile, which is discussed in more detail below. Furthermore, the selectivities towards different diols were tested over 1-butanol, acting as a competing substrate in transacylation reactions. Also in this investigation the Thr40Ala mutation was shown to cause an increase in selectivity towards the diols over the competing 1-butanol. The competition experiment was repeated with a changed acyl donor and an alternative solvent. The selectivities towards diols over 1-butanol were influenced by the variation of acyl donor and solvent. However, when comparing ratios of selectivities for Thr40Ala over wild-type CALB, no essential variation was found depending on acyl donor and solvent. The effect of the Thr40Ala mutation depends mainly on the diol rather than the surrounding reaction conditions. Variation of the tested diols showed similar effects on the selectivity over the resultant monoester as well as over 1-butanol as a competing substrate. The increase in selectivity caused by the Thr40Ala mutation in CALB was predominant towards shorter rather than longer straight chained primary diols. Substrate assisted catalysis may be complicated as a result of increased length of the diol. Different conformations in the transition state of the various diols have to be formed for interactions between the free hydroxyl group and the oxyanion to occur. For example, the length of the diol affects the energy in the ring structure, which has to be formed for substrate assisted catalysis (figure 4.1 B). Substrate assisted catalysis as a plausible reason for the increased selectivity towards diol over monoester in Thr40Ala compared to wild-type CALB was further evaluated by computer modelling. The difference in activation energies between mono- and diester formation was found when 1,2-ethanediol was acylated using Thr40Ala CALB as catalyst. Hence, the transition state of the monoacetylation of 1,2-ethanediol was chosen for evaluation in the computer simulations. The

20

Page 29: Serine Hydrolase Selectivity - DiVA portal

Selective monoacylation of diols

tetrahedral intermediate of deacylation with 1,2-ethanediol acting as nucleophile was modelled in the active site of Thr40Ala CALB. The results show that the free hydroxyl group is able to hydrogen bond to the oxyanion in the transition state. Thereby, support is found to substrate assisted catalysis as a plausible explanation for the increased selectivity towards diol over monoester caused by the Thr40Ala mutant. The increased energy in transition state in CALB caused by the Thr40Ala mutation cannot be completely recovered by using a diol as substrate (table 2, paper II). Substrate assisted catalysis can only occur when the diol acts as nucleophile, which is in the deacylation step of the reaction mechanism (figure 1.2). In the acylation part of the reaction mechanism no free hydroxyl group is present to interact with the oxyanion in transition state. When vinyl butyrate was used instead of ethyl acetate as acyl donor the reaction rates were improved. The activation of the acyl donor allowed more of the activity lost as a result of the Thr40Ala mutation in CALB to be restored by substrate assisted catalysis. Using vinyl esters may be useful in monoester preparations catalyzed by Thr40Ala CALB. Still, even when vinyl esters are used in acylation reactions with diols the activity of Thr40Ala remain at around 10% compared to wild-type CALB. Moreover, if other acyl donors are required to obtain a specific product, the decreased activity limits the usefulness of the method. A method where the selectivity towards diol over monoester increases without affecting the overall activity would be valuable.

4.2 Restricting the size of the active site Since the method for selective monoacylation of diols presented in paper II had drawbacks concerning loss of activity, an alternative method was developed. In this method the oxyanion hole in CALB was maintained intact (paper III). Rational design was applied to CALB in order to increase the selectivity towards diol over monoester. Site directed mutagenesis was performed on residues suggested from

21

Page 30: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

computer modelling. The created mutants were evaluated by determining kcat/KM-values towards diols and the monoesters produced in transacylation reactions. Tetrahedral intermediates representing mono- and diester formation of 1,2-ethanediol and 1,4-butanediol using acetate and butyrate acyl donors were modelled in wild-type CALB. Two residues were found interesting as targets for mutagenesis from visual analysis. Mutations of Ala281 and Ala282 into larger residues should cause steric hindrance in transition state of the acylation of monoester into diester (figure 4.3). The activities towards the diols were expected to be intact since diols are smaller than monoesters. By selectively discriminating the monoester as nucleophile by introducing steric hindrance, the selectivity towards diol over monoester would increase.

Ala281 Ala281

Ala282 Ala282

A) B)

Figure 4.3. Tetrahedral intermediates demonstrating the transition states in acylation of 1,2-ethanediol using a butyrate acyl donor. The free diol and the formed monoester act as nucleophiles, shown respectively in (A) and (B). A mutated variant of CALB, Ala282Leu, was available in the lab from previous work.[41] The variant was suitable for evaluation of effects on selectivity towards diol over monoester caused by mutations of Ala282. Additionally, two variants were produced, Ala281Val and Ala281Glu, which were expected to possess selectivities different from Ala282Leu CALB. The mutations of Ala281 were done to distinguish between

22

Page 31: Serine Hydrolase Selectivity - DiVA portal

Selective monoacylation of diols

smaller substrates, increasing the selectivity towards 1,2-ethanediol over 1,2-ethanediol monoacetate. The Ala282Leu should result in increased selectivity towards larger diols, such as 1,4-butanediol, over the monoester formed in the reaction. Transacylation reactions of 1,2-ethanediol and 1,4-butanediol with vinyl acetate and vinyl butyrate as acyl donors were performed for evaluation of the selectivities towards diol over monoester for the produced mutants. Individual kcat/KM-values towards the diols and the formed monoesters were obtained from the reaction data. For all reactions, the selectivity increased as a result of the mutations. As expected from modelling, kcat/KM towards the monoesters produced in the reactions decreased as result of the mutations. Moreover, in several cases kcat/KM towards the diols increased in the mutants compared to the wild type. The biggest change in kcat/KM caused by the mutations was observed towards 1,2-ethanediol when vinyl butyrate was used as acyl donor, with a 5- and 6-fold increase compared to wild-type for Ala282Leu and Ala281Glu CALB, respectively. A better fit of the diol in the active site caused by the mutations may be attributed to the increased kcat/KM. Interactions in the active site which favour binding of the substrate and/or stabilizes transition state may be improved by the mutation. For the Ala281Glu variant, a charged or polar group is introduced, which may contribute to catalysis by electrostatic interactions. The free hydroxyl group in the diol may interact with the carboxylic acid group of the side chain of the glutamate introduced by the mutation, which could promote substrate binding and stabilize transition state. The acyl donor had a considerable effect on the kcat/KM towards many of the diols. For Ala281Glu, kcat/KM towards 1,2-ethanediol increased almost 2-fold when vinyl butyrate was used as acyl donor instead of vinyl acetate. The amount of acylated enzyme can differ depending on the acyl donor, which may explain the variations in

23

Page 32: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

kcat/KM. A complete table of the kcat/KM towards diols and the formed monoesters is presented in paper III.

4.3 Comparison of methods Two methods for selective monoacylation of diols were developed by mutating CALB in order to increase the selectivity towards diol over monoester. Different characteristics of the substrates were utilized in the CALB variants which were produced to increase the selectivity. In the first case the enzyme was engineered for substrate assisted catalysis. The free hydroxyl group available when a diol acts as nucleophile was used to increase the selectivity towards diol over monoester. In the second example, the difference in size between diol and monoester was exploited to increase the selectivity by designing new CALB variants. Both methods were successful in respect of increasing the selectivity towards diol over monoester in the mutated variants compared to wild-type CALB. However, the two approaches resulted in different activities towards the diols tested in the assay and their corresponding monoesters. The Thr40Ala variant showed a severe decrease in activity compared to wild-type CALB whereas the mutations of Ala281 and Ala282 increased the activity in some cases. In paper II, the kcat/KM-values for mutated and wild-type CALB towards a common substrate were recalculated into differences in activation energy, ΔΔG‡. The ΔΔG‡ is calculated using the ratio of two kcat/KM-values towards a substrate (e.g. 1,2-ethanediol) for two different enzyme variants (e.g. Thr40Ala and wild-type CALB) inserted in equation 1.6. Calculation of the difference in transition state energy between reaction catalyzed by mutated and wild-type CALB enables comparison between the different CALB variants (figure 4.4). The Thr40Ala mutation cause increased activation energies for both mono- and diester formation. The activation energies for monoester formation do not increase as much as those for diester formation, which result in increased selectivities towards diol over monoester. In the Ala281Glu,

24

Page 33: Serine Hydrolase Selectivity - DiVA portal

Selective monoacylation of diols

Ala281Val and Ala282Leu variants of CALB, the transition state energies are mostly decreased for monoester formation and increased for diester formation. An exception is the acylation of 1,4-butanediol catalyzed by Ala281Val CALB, where the mutation causes a slight increase in transition state energy also in monoester formation. The steric hindrance introduced to discriminate the monoester also has a negative effect on the diol as nucleophile.

-4

-2

0

2

4

6

8

10

12

14

T40A

A281

E

A281

V

A282

L

T40A

A281

E

A281

V

A282

L

1,2-ethanediol 1,4-butanediol

diolmonoester

ΔΔ

G‡

(kJ/

mol

)

Figure 4.4. Diagram of activation energies for mutated variants related to wild-type CALB. Transition state energies for acetylation of free diol (1,2-ethanediol and 1,4-butanediol) and the corresponding monoesters are shown in black and white, respectively. The ΔΔG‡-values for Thr40Ala CALB were obtained using ethyl acetate as acyl donor and solvent. The ΔΔG‡-values for Ala282Leu, Ala281Glu and Ala281Val CALB were obtained using vinyl acetate as acyl donor and MTBE as solvent. The results in figure 4.4 are sufficient for displaying the discussed differences between the presented methods (papers II and III). However, some caution is required when interpreting the results since the reaction conditions differ when determining the kcat/KM-values for

25

Page 34: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

the CALB variants. The ΔΔG‡-values for Thr40Ala in relation to wild-type CALB (figure 4.4) are calculated from kcat/KM-values determined using ethyl acetate as acyl donor and solvent. In the corresponding measurements for Ala281Glu, Ala281Val and Ala282Leu compared to wild-type CALB, vinyl acetate was used as acyl donor and MTBE as solvent. As discussed in section 4.1, the activity lost as a result of the Thr40Ala mutation in CALB could be recovered by substrate assisted catalysis to a higher extent when vinyl butyrate than when ethyl acetate was used as acyl donor. It is possible that the difference in activation energy for Thr40Ala compared to wild-type CALB would be lower than presented in figure 4.4 if vinyl acetate had been used as acyl donor. Nevertheless, when vinyl butyrate was used as acyl donor in diol acylation, the Thr40Ala mutation caused reduced activity, but not as severe as with ethyl acetate. As the activity towards the diol in many of the cases are improved by the Ala281Glu, Ala281Val or Ala282Leu mutation in CALB, any of these variants would be preferred over Thr40Ala for preparative monoester synthesis. Which of the variants created by mutating positions Ala281 and Ala282 is more suitable for monoester production depend on the diol and acyl donor. The Ala281Glu and Ala281Val variants were able to discriminate between the shortest substrates in the assay, exhibiting high selectivity free over monoacetylated 1,2-ethanediol. The Ala282Leu variant is more suitable for synthesis of the longest monoesters in the assay, 1,4-butanediol butyrate. An interesting aspect of the presented method would be to combine the mutation of Thr40 with a mutation of Ala281 or Ala282. The selectivity enhancements in these two methods largely depend on different properties of the substrates. The increased selectivity towards diol over monoester caused by substrate-assisted catalysis in Thr40Ala originates from interactions between the free hydroxyl group and the oxyanion in transition state. In the CALB variants with mutations of

26

Page 35: Serine Hydrolase Selectivity - DiVA portal

Selective monoacylation of diols

27

Ala281 and Ala282 the selectivity increased as a result of restrictions of the active site volume. An exception is the Ala281Glu, where electrostatic interactions between the free hydroxyl group and the substituted residue may have added to the selectivity. If combined, the Thr40 and Ala281 or Ala282 mutations in CALB could result in further increased selectivity towards diol over monoester.

Page 36: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

5 Determination of enantiomeric excess As mentioned earlier, production of enantioselective catalysts is a central field of research, not only in enzymatic synthesis, but in modern organic chemistry at large. In order to meet a great demand of enantiomerically pure or enriched compounds, several efficient methods for enantioselective synthesis are necessary. High-throughput methods have frequently been used for finding enantioselective catalysts for various reactions. These methods involve generation and evaluation of large libraries of diverse catalysts, rather than single-sample experiments. Catalyst libraries can be generated by combinatorial methods, for example are based on the use of metal complexes with chiral ligands and activators.[42, 43] The various combinations of ligands and activators induce diversity in the library which may lead to enantiomeric enrichment of the synthesized compound. An alternative to combinatorial chemistry is directed evolution for an enzyme-catalyzed reaction.[44] Various approaches for inducing enzyme diversity have been described.[2, 32]

5.1 Screening for enantioselectivity A challenge associated with production of large compound libraries is the demand of increased throughput in analysis. Screening for enantiomeric excess, ee, of compounds created by a catalyst library has become a research field of great interest reviewed in several publications.[44-46] As no single screening method has yet proven to be universal, new methods have to be specifically developed for the synthesized substance when creating a catalyst library. Hence, several methods for rapid ee-determinations based on various techniques have been developed. Conventional methods for ee-determination have been modified and optimized for increased throughput. Parallel apparatus for gas chromatography (GC) equipped with chiral columns were optimized for

28

Page 37: Serine Hydrolase Selectivity - DiVA portal

Determination of enantiomeric excess

rapid ee-determination of the substance 2-phenylpropanol.[44, 47] Studies using modified capillary electrophoresis (CE) for high throughput screening of secondary amines have shown promising results.[44, 48] Capillary array electrophoresis (CAE), originally designed for DNA separations, and CE on micro-chips have a potential to increase throughput significantly compared to conventional CE. Circular dichroism has been successfully employed in combination with high performance liquid chromatography (HPLC) for ee-determinations of 1-phenylethanol. The HPLC is used for separation of reaction product from the reactants but no chromatographic separation of the enantiomers is necessary, which increases throughput significantly.[44, 49] Both chromatography and capillary electrophoresis are universal methods in the sense that almost any compound can be analyzed by separation of the enantiomers. However, the retention times needed for separation are dependent on the compound and may not be feasible when catalyst libraries grow. Mass spectrometry has been used for ee-determinations of pseudo-enantiomers differing in absolute configuration and mass.[44, 50-52] A mass ratio obtained by different amounts of the pseudo-enantiomers can be correlated to the ee of the product.[44, 50, 51] Analysis has also been carried out directly on pseudo-enantiomeric reaction intermediates associated with the catalyst.[52] Fourier transform infrared spectroscopy (FTIR) and flow-through nuclear magnetic resonance spectroscopy (NMR) are other suitable methods for determining enantiomeric excess of 13C-labeled pseudo-enantiomers as reviewed by Reetz et al.[44] Using MS, FTIR or NMR for ee-determinations in catalysis requires pure forms of mass- or isotope-tagged pseudo-enantiomers either as a reactant or product in the catalyzed reaction. These methods are therefore limited to ee-determinations in kinetic resolution of racemates and desymmetrization of prochiral compounds bearing enantiotopic groups.

29

Page 38: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

Numerous methods have been developed for enabling ee-determinations by conventional spectrophotometry. Many of these methods are based on using molecules which can give a change in fluorescence,[53, 54] color[55-57] or UV-absorbance[58] by enantiospecific interactions with a target substance. The ultra high-throughput potential of fluorescence was demonstrated by ee-determinations of various amino acids in microarray format.[44, 59] Primarily, interaction induced spectrophotometric ee-determinations can be done on chiral amines, amino alcohols, amino acids, α-hydroxycarboxylic acids and monosaccarides. Compounds such as simple alcohols, esters and ketones may require alternative methods.[54] Highly substance-specific methods for ee-determinations have been developed by utilizing various biochemical techniques such as immunochemistry, protein-substance interactions and enantioselective enzyme catalysis. Immunochemical methods are based on antibodies with affinity towards the synthesized product.[60, 61] An attractive system is to use competitive enzyme immunoassays for ee-determination,[60] which requires enantiospecific antibodies against the target substance. Theoretically, the immune system can raise antibodies against any compound. However, production of antibodies with the required affinity is often a laborious task. Recently, enantiospecific interaction with an engineered transmembrane protein (i.e. an α-hemolysin variant) was discovered useful for ee-determinations.[62] Enantiomeric excess was determined for the pharmaceuticals ibuprofen and thalidomide. Enzymatic methods for determining enantiomeric excess (EMDee) have been developed for a few substances.[63-66] The principels of EMDee are described in detail in the following section (5.2).

5.2 Enzymatic determination of enantiomeric excess Enantioselective enzymes can be used for determining enantiomeric excess of a substance.[63-67] The enzyme catalyzes conversion of the

30

Page 39: Serine Hydrolase Selectivity - DiVA portal

Determination of enantiomeric excess

substance into a chemically different species, enabling quantification of the different enantiomers by absorbance measurement. The reaction rate at high substrate concentrations (Vmax) can be directly correlated to the ee-value if the enantiomer which is not favored in catalysis by the enzyme inhibits the enzyme [63]. Otherwise, knowledge of the total concentration is required for calculation of the ee-value from the reaction rate.[64, 66] Alternatively, two different enzymes catalyzing the same reaction but with different enantiopreference can be used for ee-determinations by separate measurements without knowledge of the substrate concentration.[65, 67] A change in absorbance can be obtained by a pH-indicator,[64] a reaction product[63, 65, 67] or the yield of a coupled reaction.[66] The method presented herein relies on using relative endpoints rather than relative reaction rates for determining enantiomeric excess. This gives the advantage that the method becomes less sensitive to variations in concentration of substrate and added enzymes. Consequently, the yield from the synthesis does not have to be known before screening of the product. Furthermore, the system becomes insensitive to fluctuations in enzyme concentrations and is therefore suitable for miniaturization. For successful ee determination using endpoints it is important that the enantioselectivity for the enzyme is sufficiently high (i.e. E>100) towards the screened substrate, enabling depletion of one of the enantiomers. In this way a stable signal can be obtained over time without any interference from background reaction of the remaining enantiomer, which can be quantified by using a complementary enzyme. Thereby the total concentration of the synthesized product can be determined from the total change.

5.3 Enzymatic analysis of O-acylated cyanohydrins The enzymatic method for determination of enantiomeric excess relying on relative endpoints was implemented on O-acylated cyanohydrins (papers IV, V), synthesized according to a previously defined

31

Page 40: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

method.[68] The analysis of ee and conversion is carried out by measuring changes in absorbance caused by three consecutive reaction steps. Each step is carried out by addition of an enzyme exhibiting a desired specificity (figure 5.1).

time

abso

rban

ce

1. HLADHUnreactedsubstrate

2. CALB(S)-enantiomer

3. PLE(R)-enantiomer

R

O

R

OH

NADH + H+ NAD+

HLADH

R CN

O

O

R'

R CN

OH

H2O

HO

O

R'

R

O

R

OH

NADH + H+ NAD+HCN

CALB Spontanous HLADH

R CN

O

O

R'

R CN

OH

H2O

HO

O

R'

R

O

R

OH

NADH + H+ NAD+HCN

PLE Spontanous HLADH

Figure 5.1 Three different enzymes are added in successive steps: 1) Reduction of aldehyde to alcohol catalyzed by HLADH. 2) Selective ester hydrolysis of the (S)-enantiomer of the product. 3) Unselective ester hydrolysis of the remaining enantiomer, (R), of the product. Oxidation of NADH to NAD+ corresponding to the yield in each reaction step enables determination of ee and conversion by absorbance measurement. The first enzyme added is horse liver alcohol dehydrogenase, HLADH. By adding HLADH reduction of aldehyde to alcohol is catalyzed yielding NAD+ from NADH (nicotinamide adenine dinucleotide), which causes a drop in absorbance (step 1, figure 5.1). The next step of the analysis is carried out by adding the enzyme Candida antarctica Lipase B, CALB. Thereby, ester hydrolysis of the (S)-enantiomer of the

32

Page 41: Serine Hydrolase Selectivity - DiVA portal

Determination of enantiomeric excess

acylated cyanohydrin is catalyzed (step 2, figure 5.1). After depletion of the (S)-enantiomer pig liver esterase, PLE, is added for hydrolysis of the (R)-enantiomer (step 3, figure 5.1). The products from hydrolysis of both enantiomers are coupled to the oxidation of NADH to NAD+ and the decrease in absorbance can be used for determining ee. Additionally, conversion can be calculated using the decrease in absorbance of all three reaction steps. The product from the ester hydrolysis is in spontaneous equilibrium with the corresponding aldehyde, which is used in synthesis. Hydrolysis of acylated cyanohydrins can thereby be quantified from the reduction of aldehyde, coupled to oxidation of NAD+ from NADH (step 2 and 3, figure 5.1). The reaction is catalyzed by HLADH which is already prevalent in the reaction buffer as it is the first enzyme added (step 1, figure 5.1). This color reaction enables determination of ee from absorbance measurements. Additionally, the amount of aldehyde remaining from the synthesis of the acylated cyanohydrin is assessed and can be used to calculate the conversion. Again, absolute concentrations do not have to be known as relative measurements are used. CALB is known to exhibit high enantioselectivity towards secondary alcohols and esters and is therefore a suitable catalyst for depletion of the (S)-acylated cyanohydrins. PLE catalyzes hydrolysis of both (R)- and (S)-enantiomers of the acylated cyanohydrins. However, the amount of the (R)-enantiomer can be determined by assuming that the remaining substrate is in this configuration when the enzyme is added.

5.4 Substrate specificity An advantage with using enzymes for analysis is the possibility to use the same method for screening of a relatively wide variety of substances. Still, highly selective conversion of specific components in a reaction mixture can be obtained. In order to achieve accurate determinations of ee and conversion using the presented method certain requirements on

33

Page 42: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

the substrate specificity for the involved enzymes towards their substrates have to be met. As discussed earlier in section 5.2, high enantioselectivity for the enzyme used for selective depletion of one enantiomer is crucial for successful determination of ee in the presented method. In agreement with Kazlauska’s rule,[29] the enantioselectivity for a lipase towards the O-acylated cyanohydrins should be dependent on the difference in size between the R- and CN-group of the cyanohydrin (figure 5.1). The enantioselectivity for CALB towards secondary alcohols and esters is derived from a stereospecificity pocket in the active site,[69] which should fit the cyano group of the cyanohydrin. Therefore, successful ee-determination using CALB in the presented method may be possible for O-acylated cyanohydrins with substituents at the R-position large enough to obtain the required enantioselectivity. Since the R-group points out towards the entrance of the active site, its size should not be limited by the enzyme. The active site of CALB also have a pocket for the acyl part of the substrate (R’-group, figure 5.1).[69] Linear aliphatic R’-substituents of various lengths should not have any problems to fit in this pocket.[70] However, too large hydrophobic substituents on either the R- or the R’-group of the substrate may affect the solubility in aqueous buffer of both substrate and product and thereby prevent analysis. Some nonlinear acyl chains may also fit into this pocket, since it is not as confined as the stereospecificity pocket. The dynamic range in absorbance measurements is limited due to Lambet-Beer’s law (0.2-0.8 absorbance units). The enzymes in the assay therefore have to have kcat/KM-values towards the tested substances which enable catalysis at low substrate concentrations (10-100 µM). As an example the kcat/KM for HLADH towards benzaldehyde is around 200 s-1mM-1,[71] which was sufficient for the analysis of mandelonitrile acetate with the presented method. In contrast, the specificity for

34

Page 43: Serine Hydrolase Selectivity - DiVA portal

Determination of enantiomeric excess

HLADH towards smaller substrates such as acetaldehyde is considerably lower, making the enzyme inappropriate for analysis of lactonitrile acetate. As mentioned earlier (section 5.3), the enzyme used for complementary ester hydrolysis does not have to be enantioselctive towards the analyzed substance. Commercially available PLE is often a mixture of isozymes which have low enantioselectivity and may thereby be a good catalyst for both enantiomers.[72] In most type of chemical analysis, impurities such as byproducts or remaining reactants and reagents from the synthesis may affect the analysis. Using enzymes with adequate substrate specificities towards their substrates can prevent such effects without purification of the analyte.

5.5 Results and discussion The screening method described in section 5.4 was evaluated by enzymatic determination of various ee of mandelonitrile acetate, published in paper IV. This initial study demonstrated that the design of the method was adequate and applicable in both cuvette- and microtiter plate format. However, a more extensive evaluation was needed to define the scope and limitations of the method concerning substrate diversity, robustness and applicability in high-throughput screening. In paper V, the method was hence tested using O-acylated cyanohydrins with various substitutions at the chiral center of the cyanohydrin and at the acyl part of the compound (respectively as defined in figure 5.1). Additionally, paper V includes a study using the method for high-throughput screening of a combinatorial catalyst library model.

35

Page 44: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

Table 5.1 Substances subjected to enzymatic analysis Substancenumber*

Substituents** R R’

Enzymatic determination ee conversion

1a C6H5 CH3 yes yes 1b 4-CH3-C6H5 CH3 yes yes 1c 4-CH3-O-C6H5 CH3 yes yes 1d 4-Cl-C6H5 CH3 yes yes 1e 2-Furyl CH3 yes yes 1f 3-(PhO)C6H5 CH3 no yes 1g 3-Pyridyl CH3 yes yes 1h 2-Pyridyl CH3 no yes 1i (E)-CH=CHPh CH3 yes yes 1j C(CH3)3 CH3 no no 1k (CH2)4CH3 CH3 no no 1l C6H5 CH2CH3 yes yes 1m C6H5 (CH2)2CH3 yes yes 1n C6H5 (CH2)3CH3 no no 1o C6H5 CH(CH3)2 yes yes 1p C6H5 C(CH3)3 no no 1q C6H5 (E)-CH=CHPh no no 1r C6H5 C6H5 no no

O*Structures of substance 1a-1r can be found in paper V. **Substituents on an acylated cyanohydrin as defined: The substrates listed in table 5.1 were tested using crude samples of the reaction mixture. A racemic sample and samples containing an enantiomeric excess of the (S)- as well as the (R)-enantiomer were used for assessment. Verification was done by plotting the enzymatically determined ee to their corresponding value determined by GC (figure 5.2). Furthermore, each sample was analyzed at two substrate concentrations differing by a factor of 2. Enzymatic determination of ee was achieved for 10 and conversion was enzymatically determined for 12 of 18 tested substances (tab. 5.1). Possible explanations for the reason why some compounds could not be enzymatically analyzed are

R CN

O R'

36

Page 45: Serine Hydrolase Selectivity - DiVA portal

Determination of enantiomeric excess

presented in paper V. High accuracy of the enzymatically determined ee-values and conversion was obtained regardless the variations in substrate concentrations.

-100

-50

0

50

100

-100 -50 0 50 100

GC (% ee )

Enzy

mat

ic (%

ee

)

1d1l

Figure 5.2. Example of enzymatic determination of ee plotted as a function of the corresponding values determined by GC for 1d ( ) and 1l ( ). Positive ee values correspond to an excess of the (R)-enantiomer, negative to an excess of the (S)-enantiomer. The dotted line corresponds to EMD value = GC value. The continuous line is a linear regression of results from 1d (EMDee = 0.97 GCee -2.8, r2 = 0.9992). The dashed line is a linear regression of results from 1l (EMDee = 0.82GCee – 0.66, r2 = 0.996). A small catalyst library was created for synthesis of mandelonitrile acetate in a micro reactor. A salen-Ti complex were used as a catalyst activated by various Lewis bases and samples were taken at two different reaction times to obtain samples with different ee and conversion (table 5.2). Enzymatic determination of ee and conversion was applied as previously described in 96-well microtiter plate format for screening the library. No purification of the samples was done prior to the screening. Quadruplicate measurements were made for each sample to evaluate the reproducibility of the method.

37

Page 46: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

Table 5.2 Catalyst library created from various Lewis bases, catalysts and reaction times*.

Lewis base Catalyst Reaction time(min)

ee (%)

conversion (%)

Et3N salen-Ti2 20 -49 30 Et3N salen-Ti2 40 -62 70 DBU salen-Ti2 20 -20 96 DBU salen-Ti2 40 -7 97 cinchonidine salen-Ti2 20 -63 22 cinchonidine salen-Ti2 40 -74 66 quinine salen-Ti2 20 -42 23 quinine salen-Ti2 40 -61 64 DIPEA salen-Ti2 20 -59 68 DIPEA salen-Ti2 40 -55 73 DMAP salen-Ti2 20 -75 45 DMAP salen-Ti2 40 -75 74 DEA salen-Ti2 20 -59 24 DEA salen-Ti2 40 -64 65 cinchonidine ent-salen-Ti2 20 46 42 cinchonidine ent-salen-Ti2 40 87 80 quinine ent-salen-Ti2 20 75 94 quinine ent-salen-Ti2 40 83 94 DABCO salen-Ti2 20 -35 12 DABCO salen-Ti2 40 -79 63

*All reactions were carried out in dichloromethane at room temperature using 2 equivalents of acetyl cyanide, 5 mol% catalyst and 10 mol% Lewis base. Both ee and conversion were determined by GC. Positive ee denote an excess of the (R)-enantiomer whereas negative ee denote an excess of the (S)-enantiomer.

H

O

CN

O

CN

O

O

CatalystLewis base

The enzymatically determined values for ee and conversion were successfully expressed as a function of the values obtained from GC

38

Page 47: Serine Hydrolase Selectivity - DiVA portal

Determination of enantiomeric excess

(paper V). The experiments showed that the different catalytic complexes and Lewis bases used for synthesis had no apparent effect on the analysis. Standard deviations of the measured ee and conversion differed between the analyzed samples. This can be explained by variations in dynamic range of absorbance caused by differences in conversion and concentration. In an early stage of the development of the enzymatic method for determination of enantiomeric excess para-nitrophenol was used as a pH indicator to yield a color reaction for absorbance measurement. The enzymes CALB and PLE were used as described in section 5.3. The method was evaluated by spectrophotometric measurements in cuvettes using purified samples of O-acetylated mandelonitrile of various ee (figure 5.3). A linear relationship between ee-values determined enzymatically and by GC was obtained (paper V).

Figure 5.3. Enzymatic determination of ee for purified samples of mandelonitrile by sequential addition of CALB and PLE. The concentration of each enantiomer was quantified using para-nitrophenol as ph-indicator.

39

Page 48: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

40

The use of a pH-indicator broadens the scope of ee-determinations by using CALB and PLE for enantioselective hydrolysis of ester bonds. However, a limitation is that a weak buffer had to be used in order to obtain a measurable change in pH, making the method sensitive to contaminations such as atmospheric carbon dioxide. Using crude reaction samples for ee-determinations was for example not possible and the reproducibility in micro-titer plates was low. Still, the pH-indicator was a valuable tool for evaluating CALB and PLE as suitable enzymes for the assay. If these enzymes would be used in similar methods for determination of ee for other types of secondary esters, the pH-indicator could be useful. An alternative colour reaction could open up possibilities to use CALB and PLE in a robust analysis method for measuring enantioselectivities in ester synthesis protocols other than the synthesis of cyanohydrins evaluated herein.

Page 49: Serine Hydrolase Selectivity - DiVA portal

Amino acid sequence determination

6 Amino acid sequence determination Protein science includes research fields of great interest. Proteomics can provide basic knowledge for understanding of living organisms. Synthesis of novel proteins is attractive in drug development. An important part in such research is amino acid determination of expressed proteins and peptides.[73-77] Amino acid sequence determination of a full length protein was first accomplished in 1953, revealing the sequence of bovine hormone insuline.[78] Today, most sequencing is performed on peptides or protein fragments afforded by site specific proteases as peptides and protein fragments are easier to handle than full length proteins. Tools for amino acid sequence determination of peptides are therefore valuable in both proteomics and protein synthesis.

6.1 Peptide sequencing Modern amino acid sequence determination of peptides or proteins often involves ladder sequencing in mass spectrometry (MS).[73-75] Mass analysis of a fragmented peptide allows sequence determination from the mass differences between peptide fragments detected as peaks in mass spectra with specific mass to charge ratios (m/z) as illustrated in figure 6.1. The fragmentation can be accomplished by sequential truncations or by random peptide cleavage; examples of both approaches are presented in the following text. Due to limitations in mass accuracy when using MS for large macromolecules, ladder sequencing is mainly applicable to sequencing peptides and small proteins.[73] If a large protein is to be sequenced, it has to be digested into smaller fragments prior to sequencing.

41

Page 50: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

A-B – C – D – E – F

A-B – C – D – E

A-B – C – D

A-B – C

m/z

A – B – C – D – E – F

A – B – C – D – E F

A – B – C – D E – F

A – B – C D – E – F

A – B C – D – E – F

A B – C – D – E – F

m/z

A B

Figure 6.1 Ladder sequencing of a peptide with the sequence ABCDEF. Peptide fragments are displayed as peaks in a mass spectrum. The amino acid sequence can be elucidated from mass differences between peaks in the mass spectrum. In A the ladder is created from truncation of the peptide from the C-terminus. The colours match C-terminal amino acids of the peptide fragment with the peaks in the corresponding mass spectrum. In B two ladders, C-terminal and the N-terminal, are created by random cleavage of the peptide bonds. The colour of the peaks in the mass spectrum is correlated to the colour of the amino acid adjacent to the cleavage site. Both the C- and N-terminal amino acids in the starting peptide have the same colour seen as a single peak with the highest m/z in the mass spectrum.

42

Page 51: Serine Hydrolase Selectivity - DiVA portal

Amino acid sequence determination

The most commonly used method for ladder sequencing is tandem-mass spectrometry (MS/MS).[73-75] Peptide sequencing by MS/MS can be divided in two steps; the first step involves isolation of a peptide of a certain mass, which is then fragmented in the second step.[79-81] The sequence obtained in MS/MS can be read from both the N- and C-terminal of the peptide, which is mostly an advantage. On the other hand, the double set of ladders displayed in a mass spectrum may give peptide fragments with overlapping masses.[82] Furthermore, depending on the fragmentation technique, all peptide bonds do not have the same tendency to dissociate in MS/MS.[82, 83] Incomplete fragmentation combined with the risk of obtaining peptide fragments with overlapping masses often lead to incomplete sequence readout. Complementary methods are therefore often used to obtain sufficient sequence coverage. An alternative to MS/MS-sequencing is microwave assisted digestion of proteins followed by mass analysis.[84] Fragmentation is accomplished by subjecting the peptide to acid hydrolysis assisted by microwave irradiation prior to analysis. The amino acid sequence can then be determined by simple MS. The microwave based methods have the same advantage as MS/MS-sequencing concerning the ability to determine the amino acid sequence from both ends of the peptide. This also results in the same risk as in MS/MS-sequencing to obtain overlapping masses. Selective N-terminal sequencing of peptides and proteins is routinely performed by Edman degradation.[85] Peptide degradation is carried out in cycles using phenyl isothiocyanate for selective cleavage of the N-terminal amino acid, which can then be analyzed by chromatography or capillary electrophoresis. As analysis is carried out on the amino acids one by one, Edman degradation can be applied to full length proteins as well as peptides. Over the decades, Edman degradation has matured into a reliable and automated method for peptide- and protein

43

Page 52: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

sequencing. However, Edman degradation can be regarded as a laborious and slow technique exhibiting poor sensitivity. Nonetheless, it is often used as a complement to MS/MS-sequencing.[75] A protocol for peptide ladder sequencing resembling that of Edman degradation is also commonly used as a complement to MS/MS-sequencing.[86, 87] Phenyl isothiocyanate is used together with a terminating agent to prepare a peptide ladder which can be analyzed in MS. As in Edman degradation, the amino acid sequence is determined from the N-terminus. If the amino acid sequence information is desired from the C-terminus, other methods have to be employed. C-terminal specific degradation of peptides carried out in cycles can be accomplished by alkylation chemistry.[88, 89] However, low yields in the degradation reactions often make this method insufficient for sequencing, especially when peptide samples are limited. Another C-terminal specific alternative is carboxypeptidase mediated peptide ladder sequencing.[90-96] Mass analysis of a carboxypeptidase digestion of a peptide can provide information about the C-terminal amino acid sequence. Carboxypeptidase mediated amino acid determination is described in detail in the following section (6.2).

6.2 Carboxypeptidase mediated sequencing and competing nucleophiles Carboxypeptidases are proteolytic enzymes with selectivity towards the C-terminal amino acid,[13] and are frequently used for C-terminal specific peptide sequencing.[90-96] Aliquots are taken from a carboxypeptidase digest over a period of time. C-terminally truncated peptide fragments are derived as digestion proceeds, and a peptide ladder is generated. The C-terminal amino acid sequence can thereby be determined by mass spectrometry. However, variations in substrate specificity combined with the generation of numerous peptide fragments may cause insufficient sequence readout (explained in detail in section 6.3).

44

Page 53: Serine Hydrolase Selectivity - DiVA portal

Amino acid sequence determination

A commonly used enzyme for amino acid sequence determination of peptides is Carboxypeptidase Y (CPY).[90, 93-96] It has been used both as a single enzyme and together with another carboxypeptidase exhibiting complementary substrate specificity. CPY have also been studied for the use as a catalyst in transpeptidation reactions to make alterations at the C-terminus of peptides.[97-100] In this case, an amine is used as an alternative nucleophile which competes with water in a CPY digest. The result is that a fraction of the digested peptides undergoes aminolysis while the rest of the peptides are hydrolysed. An alternative nucleophile could be introduced to compete with water in CPY mediated peptide sequencing (figure 1, paper VI). The altered end group could slow down further degradation and thereby stabilize the peptide ladder. The stabilized peptide fragments obtained from aminolysis should give an increased number of peptide fragments of different lengths. Another effect of the presence of a competing nucleophile is that amine-capped peptide fragments can function as a buffer to refill the amount of peptide fragments which are rapidly digested by the enzyme (figure 3, paper VI). Altogether these effects may contribute to increased sequence readout. This method would resemble the DNA sequencing method developed by Sanger, where ddNTP competes with ordinary NTP in PCR (polymerase chain reaction) [78]. The random incorporation of ddNTP stops further elongation and DNA fragments of different lengths are created.

6.3 Substrate specificity Peptides bind to CPY by interactions to six different subsites named S1-S5 and S1’[101]. Each site binds an amino acid in the peptide analogously named P1-P5 and P1’. P1’ is the C-terminal amino acid, which is hydrolyzed off by the enzyme after cleavage between P1’ and P1. The C-terminus of a peptide is recognized by interactions between the carboxylic end and protein residues in the S1’-pocket of the enzyme [98,

102, 103]. The C-terminal amino acid is thereby selectively cleaved off. The

45

Page 54: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

newly formed peptide fragment can undergo further truncations at its C-terminus. In this way new substrates are created as degradation of the peptide proceeds. The subsites of CPY vary in specificity. For example, the S1 preferentially accommodates large hydrophobic amino acids [104]. As a result, the kcat/KM for CPY towards a peptide is dependent of its amino acid sequence, and may vary more than 1000 times solely by variations in the P1’- and P1-positions.[13] This corresponds to possible differences in activation energy larger than 16 kJ/mol (room temperature). The large variation in substrate specificity dependent on the amino acid sequence of the digested peptide pose a problem in carboxypeptidase mediated ladder sequencing. As fragments with various C-terminal amino acid sequences are generated throughout the carboxypeptidase digestion, concentrations of the different peptide fragments may not always be sufficient for analysis. The aim of incorporating an alternative nucleophile at the C-terminus of a fraction of the digested peptides is to cancel out some of the variations in substrate specificity. An ultimate goal would be to obtain totally stable peptide fragments from the alternative reaction. Although, such goal demands that the big differences in substrate specificity towards the variety of peptide fragments derived in the digest have to be overcome. In order to effectively incorporate an alternative nucleophile at the C-terminus of the digested peptide it should fit in the S1’-subsite. This also means that the enzyme will have affinity towards the peptide with a bound alternative end group and will subject it to further degradation. However, KM towards the free nucleophile does not have to be in the same range as towards the digested peptide if the frequency of incorporation can be adjusted by varying the concentration of free nucleophile. In the experiments presented in paper VI the concentration of the free alternative nucleophile is 4500 times higher than the starting concentration of the peptide. The large difference in

46

Page 55: Serine Hydrolase Selectivity - DiVA portal

Amino acid sequence determination

concentrations is combined with the possibility for the alternative nucleophile to bind in any of the different subsites. Thus, the concentration of the alternative nucleophile has to be balanced to avoid competitive inhibition from the alternative nucleophile.

6.4 Identification frequency Carboxypeptidase mediated peptide ladder sequencing protocols can be compared using identification frequency as a parameter of efficiency (paper VI). The identification frequency is based on the amount of information per mass spectrum of aliquots taken from digests of different peptides at set reaction times. The full procedure for calculation of identification frequencies is defined in paper VI. Determining identification frequencies can be advantageous when comparing sequencing results over a defined range of sequences obtained by different digest methods. The use of various peptides and several aliquots for determination of an identification frequency can cancel out accidental deviations between measurements. Variations in the success rate of a sequence experiment may be caused by many factors in the sequencing method other than the enzymatic step; such as sample handling or discriminations between different fragments in MS. The identification frequency can also reveal trends derived from different digest protocols for ladder sequencing. For example, the probability to identify an amino acid at a certain position in a peptide is dependent on the reaction times used for taking aliquots from a carboxypeptidase digest. Substrate specificity of the carboxypeptidase and concentrations of enzyme and substrate can also cause variations which will be displayed in the identification frequency. An alternative to identification frequency is sequence coverage, which also can be calculated per aliquot. A difference is that the sequence coverage is peptide specific whereas results from sequencing of more than one peptide can be included when determining the identification frequency. The sequence coverage is related to the identification

47

Page 56: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

frequencies by their corresponding averages over a given sequence interval for one or several different peptides. An average of the sequence coverage per aliquot taken from digests of various peptides will be identical to the average identification frequency calculated from the same experiments. However, the average of either coverage or identification frequency may not be satisfactory when comparing two different digest methods as it will not show any trends dependent on the amino acid sequence. Such trends can be displayed by calculating the identification frequency for each amino acid position in the sequence interval.

6.5 Results The identification frequencies in carboxypeptidase Y digests of six different peptides were determined for the amino acids at positions 1-10 from the C-terminus (paper VI). Mass spectra of aliquots taken at 1, 10, 100 and 1000 minutes of reaction time were used in the calculations. The experiments were carried out using a buffer containing 2-pyridylmethylamine as an alternative nucleophile. A digest carried out in a buffer without any alternative nucleophile was used as a reference. The average detection frequency in the interval was 29% when 2-pyridylmethylamine was present compared to 19% in the reference. An improvement of the identification frequency of around 50% was thereby accomplished. As previously discussed (section 6.5) these values are identical to the average sequence coverage per aliquot, calculated for the interval 1-10 from the C-terminus. Initially, the identification frequency for each position of the amino acid sequence was calculated (paper VI, figure 5). Variations in identification frequencies dependent on the amino acid position were revealed for both the 2-pyridylmethylamine containing digest and the reference. Both digest methods resulted in similar trends with the highest identification frequency around position 7, and higher values for the digest containing 2-pyridylmethylamine. These trends can be caused

48

Page 57: Serine Hydrolase Selectivity - DiVA portal

Amino acid sequence determination

by differences in substrate specificities combined with a relatively limited number of peptides. Another reason for the observed trend is the reaction times. Changing the time protocol for aliquot sampling could result in a shifted peak of identification frequency and a different trend. It is apparent that both time protocol and variations in substrate specificities have the same effect regardless of the presence of 2-pyridylmethylamine. The amine-capping by 2-pyridylmethylamine stabilize the peptide fragments and enhance the probability to identify amino acids in CPY mediated peptide sequencing. Consequently, a smaller amount of aliquots and less peptide specific optimization of the digest protocol is necessary to obtain a sufficient sequence coverage. Both time and valuable peptide samples can thereby be saved by using the presented method. A comparison between mass spectra of 100 minutes aliquots taken from CPY digests of bombesin with and without the alternative nucleophile 2-pyridylmethylamine is done in paper VI, figure 3. A similar example with mass spectra of 10 min. aliquots taken from CPY digests of N-acetyl-renin-tetradecapeptide shows an increased number identified amino acids when the alternative nucleophile 2-pyridylmethylamine is used (figure 6.2). The example with bombesin discussed in paper VI suggests that an improvement in identification frequency can be accomplished by a buffering effect from the amine-capped peptides. The C-terminal bound 2-pyridylmethylamine can be hydrolyzed off to yield peptide fragments which are by large extent degraded by the enzyme (paper VI, figure 4).

49

Page 58: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

ValLeuLeuHisPhe

Phe His Leu

Phe His Leu

Tyr + Ser

-90 Da

* **

**

*

**

** *

*

A

B

* *

*

m/z

Arb

itrar

y un

its

m/z

Arb

itrar

y un

its

Figure 6.2 Mass spectra of two CPY digest of N-acetyl-renin-tetradecapeptide after 10 min reaction time. The upper mass spectrum (A) corresponds to CPY digest without any alternative nucleophile. The lower mass spectrum (B) corresponds to CPY digest with 2-pyridylmethylamine present as an alternative nucleophile. The amino acids identified from hydrolysis fragment series are indicated with solid double-point arrows and amino acids identified from amine-capped fragments are indicated with dashed double-point arrows. A dotted double-point arrow in spectrum B indicates a difference in the peptide ladder corresponding to the additive mass of tyrosine and serine minus 90 Da, which is the mass of a peptide-bound 2-pyridylmethylamine. A signal to noise ratio of 3 was used as a limit for identification of peptide fragments. Peaks marked with an asterisk (*) correspond to Na or K adducts. The example in figure 6.2 shows that an improvement can be caused by amine-capped peptide fragments remaining in the digest solution after hydrolysis of the peptide fragments. A double set of peptide fragments created from hydrolysis and aminolysis are displayed as two ladders in mass spectra, with a shift corresponding to the mass of a peptide bound

50

Page 59: Serine Hydrolase Selectivity - DiVA portal

Amino acid sequence determination

51

2-pyridylmethylamine. A gap in the sequence corresponding to the mass of tyrosine and serine after addition of the mass of peptide-bound 2-pyridylmethylamine (90 Da) provides information regarding the amino acid composition without revealing the internal sequence. The double identifications of amino acids from the double set of peptide fragments and mass differences correlated to more than one amino acid were not included in the identification frequencies discussed above.

Page 60: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

Concluding remarks A broad spectrum of applications has been demonstrated where the substrate selectivities for various serine hydrolases have been exploited. Methods for organic synthesis as well as for analytical purposes have been developed. Kinetic properties of the involved enzymes have been evaluated in relation to the presented methods. The mutation of Trp104Ala in CALB reversed the enantioselectivity towards a range of secondary alcohols is discussed in chapter 3 (paper

I). The kinetic constants for Trp104Ala and wild-type were determined towards each enantiomer of 1-phenylethanol and compared in terms of activation energies. An increased kcat towards the slow (S)-1-phenylethanol was revealed as the main cause for the reversed enantioselectivity. Thermodynamic analysis of the Trp104Ala variant showed that the reversed enantioselectivities were driven by entropy. Two different approaches for increasing the selectivity for CALB in favour of selective monoacylation of diols are presented in chapter 4 (paper II and III). In the first example, the selectivity towards a range of diols over their corresponding monoesters increased as a result of the Thr40Ala mutation CALB (paper II). Substrate assisted catalysis was evaluated as reason for the change in selectivity. In the second example, visual analysis of structures acquired by computer modelling was sufficient to suggest mutations in CALB for increasing the selectivity towards a range of diols over their monoesters (paper III). The residues Ala281 and Ala282 were mutated into larger residues resulting in three different CALB variants, Ala281Gly, Ala281Val and Ala282Leu. Both cases using CALB variants engineered for substrate assisted catalysis and restricted size of the active site showed increased selectivities towards the tested diols over their corresponding monoester (papers II and III). However, the latter example was showed to be superior to the former by comparison in terms of activation energies. The substrate specificity

52

Page 61: Serine Hydrolase Selectivity - DiVA portal

Concluding remarks

53

towards some of the diols increased significantly as a result of the mutations of residues Ala281 and Ala282 in CALB. Chapter 5 (based on paper IV and V) presents a method for determining enantiomeric excess and yield. The method was applied for evaluation of a protocol for combinatorial synthesis of acylated cyanohydrins. The substrate specificity two different serine hydrolases, CALB and PLE, were utilized to accomplish measurable changes in absorbance. The high enantioselectivity for CALB towards acylated secondary alcohols enabled determinations of enantiomeric excess from relative endpoints. Evaluations of the method showed an expected substrate diversity and promising high-throughput applicability showed good and results. The robustness of the method was improved by shifting from quantification using a pH-indicator to the oxidation of NADH to NAD+ catalyzed by HLADH. In the work presented in chapter 6 (based on paper VI), an alternative nucleophile was used for improving the sequence coverage in carboxypeptidase Y mediated peptide sequencing. The developed method was proven to be useful for C-terminal sequencing of peptides. The method can be directly applied on any peptide sequencing experiment where carboxypeptidase Y truncation is considered.

Page 62: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

Acknowledgements Many persons have contributed to the work presented in this thesis. I am especially grateful to the following persons. I would like to express my sincere gratitude to my supervisor, Professor Kalle Hult. Thank you for accepting me as a PhD student and all the guidance and mentorship you have shown me along the way. I have learnt a lot from you and you have supported my work and helped me to develop as a scientist. The chemical company, BASF, Ludwigshafen is acknowledged for financially supporting the work presented in papers II-III. Additionally, I would like to thank Steffen Maurer (BASF) for scientific contributions in paper III, and Bernard Hauer and Cecilia Branneby (former BASF) for valuable discussions during meetings. The Nanochemistry program, funded by the Swedish Foundation for Strategic Research is acknowledged for financing the work presented in papers IV-VI. I would like to thank present and former people at Organic Chemistry, KTH School of Chemical Science and Engineering, Professor Christina Moberg, Dr. Stina Lundgren, Dr. Erica Wingstrand and Dr. Maël Penhoat, for the excellent collaboration with the work presented in papers IV-V. The present and former people at Analytical Chemistry, KTH School of Chemical Science and Engineering, Professor Johan Roerade, Martin Kempka and Dr. Johan Sjödahl, for your contributions to the work presented in paper VI. I learned a lot working with you, and I would like to thank you for confiding in me as a biochemistry “rookie”. I would like to thank all former and present members of the Biocat group who contributed to my work: Dr. Anders Magnusson, thank you for your letting me work on your project and supervising me as a diploma student (paper I), thank you also for your contribution to the

54

Page 63: Serine Hydrolase Selectivity - DiVA portal

Acknowledgements

55

work presented in paper II, Cecilia Branneby for recommending our group to BASF, Mohammad Takwa, thank you for contributing to paper I, and for accompanying me from the diploma-work corner days and onwards. Marianne Wittrup Larsen, thank you for sharing some of your microbiological jedi skills, and comments on (or trying to read) my thesis, Linda, thank you for reading the thesis and contributing with valuable comments, my “room mates”, Karim Cassimjee, Cecilia Hedfors, Michaela Vallin and Magnus Eriksson for your company and answers to FAQ, Magnus also for introducing me to the art of being an “evil but righteous” course lab assistant, Per-Olof Syrén, for giving me the left over Ala282Leu, Maria Svedendahl, for hints on the way to my dissertation, senior scientists Dr. Mats Martinelle and Professor Per Berglund for your scientific contributions to the group, all the previous group members who I did not get to know for building up the “group knowledge” that I have made use of in my work. Administration staff, Ela, Marlene, and Lotta thank you for running floor 2. Kaj Kauko and Erik Björkvall thank you for IT-support My former diploma workers, Saied Ghods, Karin Engström and Francis Jinxin Hu (paper II), gets thanked for contributing to my work. The following are acknowledged for travelling opportunities: The Nanochemistry program, funded by the Swedish Foundation for Strategic Research, Peter Klasons stipendiefond, COST D25 and Anders Lindstedts stiftelse. Ett stort tack Percy, min pappa för visat stöd genom åren. Tack Sonja, min mamma och Arnulf, för välbehövd barnpassning. Många pussar och kramar till mina söner, Bruno och Robin, världens finaste killar. Ni finns alltid i mina tankar även när jag inte är i närheten. Tack till min fru Newroz, för att ditt sälskap och kärlek, du har verkligen ställt upp för mig under mitt arbete med avhandlingen.

Page 64: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

List of abbreviations CAE Capillary array electrophoresis CALB Candida antarctica Lipase B (also known as Pseudozyma) CE Capillary electrophoresis CPY Carboxypeptidase Y E Enantiomeric ratio ee Enantiomeric excess EMDee Enzymatic method for determining enantiomeric excess FTIR Fourier transform infrared spectroscopy GC Gas chromatography HIC Hydrophobic interaction chromatography HLADH Horse liver alcohol dehydrogenase HPLC High Performance Liquid Chromatography kcat Turnover number for an enzyme towards a substrate KM Michaelis-Menten constant MS Mass spectrometry MS/MS Tandem mass spectrometry m/z Mass to charge ratio NAD+ Nicotinamide adenine dinucleotide, oxidized form NADH Nicotinamide adenine dinucleotide, reduced form NMR Nuclear magnetic resonance spectroscopy PCR Polymerase chain reaction Pn Amino acid in a digested peptide or protein at position n

on N-terminal side from the cleavage site Pn’ Amino acid in a digested peptide or protein at position n

on C-terminal side from the cleavage site PLE Pig liver esterase Sn Subsite in a peptidase or protease which accommodates

the amino acid n in a peptide or protein SEC Size exclusion chromatography UV Ultra violet Vmax Maximum reaction rate for an enzyme-catalyzed reaction

56

Page 65: Serine Hydrolase Selectivity - DiVA portal

References

References [1] U. T. Bornscheuer, R. J. Kazlauskas, M. T. Reetz, Protein

Engineering Handbook, Vol. 1-2, Wiley-VCH, 2009. [2] L. G. Otten, F. Hollmann, I. W. C. E. Arends, Trends in

Biotechnology 2010, 28, 46-54. [3] K. Faber, Biotransformations in Organic Chemistry (fourth edition), 4

ed., Springer-Verlag, 2000. [4] H. Bergmeyer, Methods of Enzymatic Analysis, 3rd Edition, Vol 1-4,

1983. [5] T. M. Devlin, Analytical Chemistry 1959, 31, 977-981. [6] M. M. Fishman, Analytical Chemistry 1980, 52, R185-R199. [7] A. Radzicka, R. Wolfenden, Science 1995, 267, 90-93. [8] A. R. Fersht, Structure and Mechanism in Protein Science, First ed., W.

H. Freeman, 1999. [9] G. Danno, Agricultural and Biological Chemistry 1970, 34, 10. [10] S. H. Bhosale, M. B. Rao, V. V. Deshpande, Microbiological Reviews

1996, 60, 280-&. [11] M. Cammenberg, K. Hult, S. Park, Chembiochem 2006, 7, 1745-

1749. [12] J. J. Perona, C. S. Craik, Protein Science 1995, 4, 337-360. [13] R. Hayashi, Y. Bai, T. Hata, Journal of Biochemistry 1975, 77, 69-79. [14] C. S. Chen, Y. Fujimoto, G. Girdaukas, C. J. Sih, Journal of the

American Chemical Society 1982, 104, 7294-7299. [15] J. L. L. Rakels, A. J. J. Straathof, J. J. Heijnen, Enzyme and

Microbial Technology 1993, 15, 1051-1056. [16] T. Boekhout, Journal of General and Applied Microbiology 1995, 41,

359-366. [17] E. Krieger, G. Koraimann, G. Vriend, Proteins-Structure Function

and Genetics 2002, 47, 393-402. [18] A. O. Magnusson, J. C. Rotticci-Mulder, A. Santagostino, K.

Hult, Chembiochem 2005, 6, 1051-1056. [19] M. W. Larsen, Doctoral thesis, Royal Institute of Technology

(KTH) (Stockholm), 2009. [20] M. W. Larsen, U. T. Bornscheuer, K. Hult, Protein Expression and

Purification 2008, 62, 90-97. [21] K. Naemura, R. Fukuda, M. Murata, M. Konishi, K. Hirose, Y.

Tobe, Tetrahedron-Asymmetry 1995, 6, 2385-2394.

57

Page 66: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

[22] K. Burgess, L. D. Jennings, Journal of the American Chemical Society 1991, 113, 6129-6139.

[23] J. Steinreiber, K. Faber, H. Griengl, Chemistry-a European Journal 2008, 14, 8060-8072.

[24] B. Martin-Matute, J. E. Backvall, Current Opinion in Chemical Biology 2007, 11, 226-232.

[25] J. V. Allen, J. M. J. Williams, Tetrahedron Letters 1996, 37, 1859-1862.

[26] A. L. E. Larsson, B. A. Persson, J. E. Backvall, Angewandte Chemie-International Edition in English 1997, 36, 1211-1212.

[27] M. J. Kim, Y. I. Chung, Y. K. Choi, H. K. Lee, D. Kim, J. Park, Journal of the American Chemical Society 2003, 125, 11494-11495.

[28] B. A. Persson, A. L. E. Larsson, M. Le Ray, J. E. Backvall, Journal of the American Chemical Society 1999, 121, 1645-1650.

[29] R. J. Kazlauskas, A. N. E. Weissfloch, A. T. Rappaport, L. A. Cuccia, Journal of Organic Chemistry 1991, 56, 2656-2665.

[30] L. Boren, B. Martin-Matute, Y. M. Xu, A. Cordova, J. E. Backvall, Chemistry-a European Journal 2006, 12, 225-232.

[31] P. F. Mugford, U. G. Wagner, Y. Jiang, K. Faber, R. J. Kazlauskas, Angewandte Chemie-International Edition 2008, 47, 8782-8793.

[32] M. T. Reetz, D. Kahakeaw, J. Sanchis, Molecular Biosystems 2009, 5, 115-122.

[33] E. Henke, U. T. Bornscheuer, Biological Chemistry 1999, 380, 1029-1033.

[34] M. T. Reetz, K. E. Jaeger, Chemistry-a European Journal 2000, 6, 407-412.

[35] B. Heinze, R. Kourist, L. Fransson, K. Hult, U. T. Bornscheuer, Protein Engineering Design & Selection 2007, 20, 125-131.

[36] H. Raj, B. Weiner, V. P. Veetil, C. R. Reis, W. J. Quax, D. B. Janssen, B. L. Feringa, G. J. Poelarends, Chembiochem 2009, 10, 2236-2245.

[37] S. Bartsch, R. Kourist, U. T. Bornscheuer, Angewandte Chemie-International Edition 2008, 47, 1508-1511.

[38] M. T. Reetz, L. W. Wang, M. Bocola, Angewandte Chemie-International Edition 2006, 45, 1236-1241.

[39] M. Vallin, P. O. Syren, K. Hult, Chembiochem 2010, 11, 411-416. [40] A. Magnusson, K. Hult, M. Holmquist, Journal of the American

Chemical Society 2001, 123, 4354-4355.

58

Page 67: Serine Hydrolase Selectivity - DiVA portal

References

[41] Z. Marton, V. Léonard-Nevers, P.-O. Syrén, C. Bauer, S. Lamare, K. Hult, V. Tranc, M. Graber, Journal of Molecular Catalysis B: Enzymatic 2010, in press.

[42] K. L. Ding, H. F. Du, Y. Yuan, J. Long, Chemistry-a European Journal 2004, 10, 2872-2884.

[43] K. D. Shimizu, M. L. Snapper, A. H. Hoveyda, Chemistry-a European Journal 1998, 4, 1885-1889.

[44] M. T. Reetz, Advances in Catalysis 2006, 49, 1-69. [45] J. F. Traverse, M. L. Snapper, Drug Discovery Today 2002, 7, 1002-

1012. [46] M. Tsukamoto, H. B. Kagan, Advanced Synthesis & Catalysis 2002,

344, 453-463. [47] M. T. Reetz, K. M. Kuhling, S. Wilensek, H. Husmann, U. W.

Hausig, M. Hermes, Catalysis Today 2001, 67, 389-396. [48] M. T. Reetz, K. M. Kuhling, A. Deege, H. Hinrichs, D. Belder,

Angewandte Chemie-International Edition 2000, 39, 3891-3893. [49] M. T. Reetz, K. M. Kuhling, H. Hinrichs, A. Deege, Chirality

2000, 12, 479-482. [50] J. H. Guo, J. Y. Wu, G. Siuzdak, M. G. Finn, Angewandte Chemie-

International Edition 1999, 38, 1755-1758. [51] M. T. Reetz, M. H. Becker, H. W. Klein, D. Stockigt, Angewandte

Chemie-International Edition 1999, 38, 1758-1761. [52] C. Markert, A. Pfaltz, Angewandte Chemie-International Edition 2004,

43, 2498-2500. [53] Z. B. Li, J. Lin, L. Pu, Angewandte Chemie-International Edition

2005, 44, 1690-1693. [54] L. Pu, Chemical Reviews 2004, 104, 1687-1716. [55] R. Eelkema, R. A. van Delden, B. L. Feringa, Angewandte Chemie-

International Edition 2004, 43, 5013-5016. [56] L. Zhu, E. V. Anslyn, Journal of the American Chemical Society 2004,

126, 3676-3677. [57] J. F. Folmer-Andersen, V. M. Lynch, E. V. Anslyn, Journal of the

American Chemical Society 2005, 127, 7986-7987. [58] X. F. Mei, C. Wolf, Journal of the American Chemical Society 2006,

128, 13326-13327. [59] G. A. Korbel, G. Lalic, M. D. Shair, Journal of the American

Chemical Society 2001, 123, 361-362.

59

Page 68: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

[60] F. Taran, C. Gauchet, B. Mohar, S. Meunier, A. Valleix, P. Y. Renard, C. Creminon, J. Grassi, A. Wagner, C. Mioskowski, Angewandte Chemie-International Edition 2002, 41, 124-127.

[61] M. Matsushita, K. Yoshida, N. Yamamoto, P. Wirsching, R. A. Lerner, K. D. Janda, Angewandte Chemie-International Edition 2003, 42, 5984-5987.

[62] X. F. Kang, S. Cheley, X. Y. Guan, H. Bayley, Journal of the American Chemical Society 2006, 128, 10684-10685.

[63] P. Abato, C. T. Seto, Journal of the American Chemical Society 2001, 123, 9206-9207.

[64] M. B. Onaran, C. T. Seto, Journal of Organic Chemistry 2003, 68, 8136-8141.

[65] Z. Li, L. Butikofer, B. Witholt, Angewandte Chemie-International Edition 2004, 43, 1698-1702.

[66] M. Baumann, R. Sturmer, U. T. Bornscheuer, Angewandte Chemie-International Edition 2001, 40, 4201-4204.

[67] I. Bustos-Jaimes, W. Hummel, T. Eggert, E. Bogo, M. Puls, A. Weckbecker, K. E. Jaeger, Chemcatchem 2009, 1, 445-448.

[68] S. Lundgren, E. Wingstrand, M. Penhoat, C. Moberg, Journal of the American Chemical Society 2005, 127, 11592-11593.

[69] J. Uppenberg, N. Ohrner, M. Norin, K. Hult, G. J. Kleywegt, S. Patkar, V. Waagen, T. Anthonsen, T. A. Jones, Biochemistry 1995, 34, 16838-16851.

[70] J. Ottosson, K. Hult, Journal of Molecular Catalysis B-Enzymatic 2001, 11, 1025-1028.

[71] C. N. Ryzewski, R. Pietruszko, Archives of Biochemistry and Biophysics 1977, 183, 73-82.

[72] A. Musidlowska-Persson, U. T. Bornscheuer, Journal of Molecular Catalysis B-Enzymatic 2002, 19, 129-133.

[73] H. Steen, M. Mann, Nature Reviews Molecular Cell Biology 2004, 5, 699-711.

[74] S. D. Patterson, R. H. Aebersold, Nature Genetics 2003, 33, 311-323.

[75] G. K. Agrawal, R. Rakwal, Mass Spectrometry Reviews 2006, 25, 1-53.

[76] P. Wang, G. Arabaci, D. Pei, Journal of Combinatorial Chemistry 2001, 3, 251-254.

[77] S. H. Joo, Q. Xiao, Y. Ling, B. Gopishetty, D. H. Pei, Journal of the American Chemical Society 2006, 128, 13000-13009.

60

Page 69: Serine Hydrolase Selectivity - DiVA portal

References

[78] F. Sanger, Annual Review of Biochemistry 1988, 57, 1-28. [79] D. F. Hunt, J. R. Yates, J. Shabanowitz, S. Winston, C. R. Hauer,

Proceedings of the National Academy of Sciences of the United States of America 1986, 83, 6233-6237.

[80] K. Biemann, H. A. Scoble, Science 1987, 237, 992-998. [81] M. Wilm, A. Shevchenko, T. Houthaeve, S. Breit, L.

Schweigerer, T. Fotsis, M. Mann, Nature 1996, 379, 466-469. [82] B. A. Budnik, M. L. Nielsen, J. V. Olsen, K. F. Haselmann, P.

Horth, W. Haehnel, R. A. Zubarev, International Journal of Mass Spectrometry 2002, 219, 283-294.

[83] D. M. Horn, R. A. Zubarev, F. W. McLafferty, Proceedings of the National Academy of Sciences of the United States of America 2000, 97, 10313-10317.

[84] H. Y. Zhong, Y. Zhang, Z. H. Wen, L. Li, Nature Biotechnology 2004, 22, 1291-1296.

[85] P. Edman, G. Begg, European Journal of Biochemistry 1967, 1, 80-91. [86] B. T. Chait, R. Wang, R. C. Beavis, S. B. H. Kent, Science 1993,

262, 89-92. [87] Y. Suzuki, M. Suzuki, Y. Nakahara, Y. Ito, E. Ito, N. Goto, K.

Miseki, J. Iida, A. Suzuki, Analytical Chemistry 2006, 78, 2239-2243.

[88] V. L. Boyd, M. Bozzini, G. Zon, R. L. Noble, R. J. Mattaliano, Analytical Biochemistry 1992, 206, 344-352.

[89] B. Samyn, K. Hardeman, J. Van der Eycken, J. Van Beeumen, Analytical Chemistry 2000, 72, 1389-1399.

[90] D. H. Patterson, G. E. Tarr, F. E. Regnier, S. A. Martin, Analytical Chemistry 1995, 67, 3971-3978.

[91] K. Klarskov, K. Breddam, P. Roepstorff, Analytical Biochemistry 1989, 180, 28-37.

[92] F. Daldegan, B. Ribadeaudumas, K. Breddam, Applied and Environmental Microbiology 1992, 58, 2144-2152.

[93] B. Thiede, B. Wittmannliebold, M. Bienert, E. Krause, Febs Letters 1995, 357, 65-69.

[94] D. R. Cool, A. Hardiman, Biotechniques 2004, 36, 32-34. [95] V. Bonetto, A. C. Bergman, H. Jornvall, R. Sillard, Analytical

Chemistry 1997, 69, 1315-1319. [96] R. H. Huang, Y. Zhang, D. C. Wang, Rapid Communications in

Mass Spectrometry 2003, 17, 903-908.

61

Page 70: Serine Hydrolase Selectivity - DiVA portal

Serine Hydrolase Selectivity, Anders Hamberg

62

[97] U. Christensen, H. B. Drohse, L. Molgaard, European Journal of Biochemistry 1992, 210, 467-473.

[98] U. H. Mortensen, H. R. Stennicke, M. Raaschounielsen, K. Breddam, Journal of the American Chemical Society 1994, 116, 34-41.

[99] H. R. Stennicke, K. Olesen, S. B. Sorensen, K. Breddam, Analytical Biochemistry 1997, 248, 141-148.

[100] S. B. Sorensen, M. Raaschounielsen, U. H. Mortensen, S. J. Remington, K. Breddam, Journal of the American Chemical Society 1995, 117, 5944-5950.

[101] H. Nakase, S. Murata, H. Ueno, R. Hayashi, Bioscience Biotechnology and Biochemistry 2001, 65, 2465-2471.

[102] U. H. Mortensen, S. J. Remington, K. Breddam, Biochemistry 1994, 33, 508-517.

[103] T. L. Bullock, B. Branchaud, S. J. Remington, Biochemistry 1994, 33, 11127-11134.

[104] K. Olesen, U. H. Mortensen, S. Aasmulolsen, M. C. Kiellandbrandt, S. J. Remington, K. Breddam, Biochemistry 1994, 33, 11121-11126.