• No results found

Structure and Dynamics of Non Ribosomal Peptide Synthetases

N/A
N/A
Protected

Academic year: 2021

Share "Structure and Dynamics of Non Ribosomal Peptide Synthetases"

Copied!
13
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Non Ribosomal Peptide Synthetases

Bachelor thesis of M.E. Hooghwinkel

Student nr. s2367831

Supervision: A.J.M. Driessen

31-05-2017

(2)

SUMMARY

Secondary metabolites make up a large class of socially relevant compounds. Many of these, like antibi- otic Vancomycin and anticancer agent Bleomycin, are produced by Non-Ribosomal Peptide Synthetases (NRPSs). Named Non-Ribosomal Peptides (NRPs), these compounds often contain non-proteinogenic amino acids and often have other interesting features, like e.g. a cyclic architecture. NRPSs are large, modu- lar enzyme complexes, arranged in a beads-on-a-string-like manner, and work much like an assembly belt as each module activates a single substrate and adds it to the growing peptide product chain. Engineering these complexes to add non-native substrates to their product peptides opens up exciting possibilities, as this may be an easy way to produce new, custom peptides. Engineering NRPSs is difficult, however, due to their great size and the many interactions that take place in-protein; attempts to change substrate specificity often result in a drastically lowered turnover. Detailed structural data may help to solve issues contributed to the breaking of protein-protein interactions. Unfortunately, no complete structure of any NRPS is currently available. This is attributed again to the size of the proteins, but also the flexibility and the many conformational changes the enzyme undergoes in its catalytic cycle. Structures of dissected catalytic domains, didomains and modules are abundant, however, and may be combined to get a model of the full structure of a NRPS. Recently, this was done by Marahiel (2015), providing exciting new insights into the structure and dynamics of NRPSs, and bringing the possibility of functional engineered NRPSs closer than ever.

TABLE OF CONTENTS

§ Paragraph Page

1 Introduction 3

1.1 Non-Ribosomal Peptides 3

1.2 NRP synthesis 3

1.3 Engineering NRPSs 4

1.4 NRPS structure 4

2 Pcp-domain 5

3 A-domain 6

4 C-domain 8

5 Te-domain 9

6 MbtH-like proteins 9

7 Structures of entire modules and NRPS complexes 10

8 Conclusion and future prospects 11

9 References 13

Cover image: Model of a heptamodular NRPS.

Image derived from Marahiel et al., 2015

(3)

1. INTRODUCTION

1.1 Non-Ribosomal Peptides

Secondary metabolites make up an enormous di- versity of compounds, often characterized by their medical or antimicrobial activities. These compounds are interesting because of their great diversity and versatility: examples are the anticancer agent bleo- mycin, antifungal agent echinocandin, immunosup- pressant cyclosporine and antibiotics like penicillins, vancomycin and gramicidin (see figure 1A). Many of these compounds are widely used and therefore (often biochemically) mass-produced, showing the interest in acquiring knowledge about the synthesis process of these secondary metabolites. All mentioned com- pounds are part of a class called Non-Ribosomal Pep-

tides (NRPs), a particularly large class of secondary metabolites, which are synthesized independently of the mRNA/ribosomal synthesis mechanism by large enzymatic complexes called Non-Ribosomal Peptide Synthetases (NRPSs) (Marahiel et al., 1997). Remov- ing the constraints imposed by synthesis based on the ribosome contributes to the versatility of these com- pounds. More than 500 monomers have been found to be part of different NRPs, including non-proteinogen- ic amino acids and other building blocks not regularly found in peptides - increasing the diversity of possible NRPS products enormously (Walsh et al., 2013).

1.2 NRP synthesis

NRP synthesis takes place in an assembly line-like fashion. NRPSs are large, multi-modular enzymatic

Figure 1: A) Structures of commercially relevant secondary metabolites synthesised by NRPSs. B) Schematic overview of NRP synthesis. Peptide formation finds place in an assembly line-like fashion, where each step is a reaction catalysed by a succession of separate domains. The steps in peptide synthesis are explained in the text in paragraph 1.2. Image derived from Kittilä et al., 2016.

(4)

complexes, mostly with a linear architecture, wherein each module activates a substrate and incorporates it into the ever-growing product peptide chain. Each module generally has specificity for only one sub- strate, so the sequence of the modules in the NRPS reflects the primary structure of the product NRP (Marahiel et al, 1997).

NRPS modules comprise several distinct structural domains, each functioning separately.

The adenylation domain (A-domain) selects and acti- vates a substrate monomer, and binds it to the pep- tidyl carrier protein domain (pcp-domain, often re- ferred to as the thiolation or T-domain). This domain uses a highly mobile 4’-phosphopanteneine (ppant)- arm to move the substrate between domains and/or modules. At the Condensation domain (C-domain), a peptide bond is formed between two adjacent pcp- bound substrates. The pcp-domain then moves the intermediate to the next module, where the process is repeated, untill it encounters the thioesterase domain (Te-domain) which is usually present at the final mod- ule of a given NRPS. Here the product is released from the NRPS, by either hydrolyzation or cyclization.

A minimal module includes the A-, pcp- and C-do- main, except for the initiation module which lacks the C-domain since condensation finds place downstream of the activation of the first substrate. However, many more domains are found in NRPSs: many NRPSs utilize additional tailoring domains to achieve even more diversity in their products. Common domains are epimerization domains (E-domains), epimerizing activated L-amino acids into their D-enantiomer;

methylation, oxidation and formylation domains. For a schematic overview of NRP synthesis, see figure 1B.

The exact functioning, dynamics and structure of the aforementioned domains will be elucidated in their respective paragraphs further in this thesis.

1.3 Engineering NRPSs

The ultimate goal of most NRPS research is to be able to engineer the complexes, manually selecting the substrates activated by each module, to be able to create custom peptide products fast and easily. To get to this goal, large advances still need to be made in the knowledge about the enzymatic complexes. Research- ers have already tried out many strategies to obtain

a NRPS with altered specificity, with some strategies being more successful than others.

Initial strategies were focused around switching entire modules and/or domains from different NRPSs with each other, hoping to gain a hybrid product peptide (Winn et al., 2016). Unfortunately, the activity of the hybrid enzymes turned out to be very low, if at all present. This happened supposedly due to the altera- tions in structural integrity and other protein-protein interactions. To avoid this problem, researcher tried to rationally mutate the substrate-specificity determin- ing pocket of the adenylation domain, to alter just the specificity of the A-domain thus leaving other interac- tions intact. Only little success was achieved using this strategy (Winn et al, 2016). More recently, researchers have tried to switch parts of the domains, e.g. a small part of the N-terminal subdomain of the adenyla- tion domain comprising all specificity determining residues (Kries et al., 2015). This strategy turned out promising, though still the activity of engineered enzymes is significantly lower than their wild-type counterparts.

Kries et al., among others, point out that the structural knowledge of NRPS enzyme complexes is still not ad- vanced enough to predict why certain swapping strat- egies work and others do not. Little is known about the interactions in-protein between different modules, since still no structures of entire NRPSs have been solved. Recently though, Marahiel (2015) published a model of a heptamodular NRPS, possibly aiding in the understanding of the structure of these mega-en- zymes, which might in turn help us to produce viable engineered NRPSs, producing custom peptides.

1.4 NRPS structure

The structure of complete NRPS complexes have, as mentioned before, still not been solved. This is attrib- uted to their plasticity and flexibility, which results in large difficulties in obtaining crystallizable solutions of the enzymes. Also, the sheer size of the enzymes presents great challenges for NMR (Payne et al., 2017).

However, more and more structures of separate do- mains, domains coupled to other domains, and even complete modules have been solved in the past couple of years. Significant progress has also been made in understanding the separate processes each domain ca- talyses, based on structures solved in various phases of

(5)

substrate activation, peptide bond formation, thioes- terification and more.

In this thesis I will give an overview of the most recent insights in NRPS structure and dynamics per domain, to find out if it is possible to find a structure of an entire NRPS enzyme complex (based upon the currently available structures of separate domains, didomains and modules) and try apply this knowledge to the ultimate goal of engineering NRPSs to produce custom peptides.

2. PCP-DOMAIN

The small, flexible peptidyl carrier protein (pcp) domains are perhaps the most studied. Their highly mobile, 18Å long phosphopanteinyl cofactor is struc- turally of great interest since it has the most versatile role in the synthesis of the NRP. Therefore, many crystal structures of this domain have been solved, often in combination with other domains, to investi-

gate the cooperation of the domains. The domain itself is relatively small with at most ±100 residues, always comprising a conserved serine, to which the ppant cofactor is covalently attached to convert the apo- pcp-domain to its holo-state. This is done by a specific pptase, often co-overexpressed in recombinant organ- isms with the NRPS gene cluster (Beld et al., 2014).

The small domain mainly consists of a four-helix bundle and an extended loop region between the first two helices, not unlike other acyl carrier proteins. At the extent of the extended loop region the conserved serine is found. The loop is diverse in structure and sequence between different pcp-domains, while the rest of the structure remains somewhat similar (Miller

& Gulick, 2017). Three of the four helices (helices H1, H2 and H4) are long and mostly parallel, while the third helix runs almost perpendicular to the other three and is shorter (figure 2A).

It is not a surprise that the loop preceding helix H2 with the cofactor-binding serine and that helix itself are the main determinants for the interactions with Figure 2: Various structures of the pcp-domain. A) only subtle configurational changes are found in the four-helix bundle of the pcp-domains of teicoplanin biosynthesis and yerbaplanin biosynthesis enzymes, respectively, upon binding the ppant-cofactor or its substrate. B) Multi-domain structures of the pcp-domain together with other domains show that the majority of interactions finds place through the helices 2 and 4 and the linker region before helix 2 (which is where the ppant arm is bound). Image derived from Kittilä et al., 2016.

(6)

the other catalytic domains (figure 2B; Goodrich et al., 2015). However, shotgun mutagenesis of the EntB protein followed by screen-testing to assay its function in vivo indicated that regions of the H3 helix also play a role in intra-protein interactions: they form a part of the hydrophobic patch, governing interactions with the condensation domain situated downstream (Lai et al., 2006).

The structure of pcp-domains often undergoes large conformational changes, depending on the domain that it is interacting with (Miller & Gulick, 2017).

Many structures have been solved of the pcp-domain in complex with the A- and C- domain (see figure 2B), and also recently a structure of a complete module with the pcp-domian in different positions has been solved of the gramicidin synthase initiation module by Reimer et al., 2016. These cases are elucidated in the other paragraphs.

In engineering NRPSs, the pcp-domain is usual- ly left unchained from the native version since the domain itself does not share any interactions with the substrate and/or intermediate (only through the ppant-cofactor which covalently binds the sub- strate). However, it has been shown that sometimes Cytochrome P450 enzymes interact with pcp-bound intermediates to modify the loaded peptide (Uhlmann et al., 2013). Manipulating this process can be used to even further increase the amount of modifications available in engineered NRPSs (Payne et al., 2017).

3. A-DOMAIN

Arguably the most versatile domain in the NRPS assembly line, adenylation (A) domains have received a lot of attention in the past years. It is the first domain the substrates encounter before being added to the growing peptide product chain in the assembly line- like process. The A-domain is characterized by its two distinct subdomains: a large N-terminal subdomain (±430) and a smaller, flexible C-terminal subdo- main (±120 residues). The A-domain belongs to the ANL-superfamily of adenylating enzymes and works similar to other enzymes in the family, comprising acyl-CoA synthetases and luciferases as well (Gulick., 2009). It catalyzes two partial reactions. Firstly it lets a carboxylate (in this case often an amino acid) react

with ATP, catalyzing the formation of an acyl-adeny- late and inorganic PPi. Then, after the PPi leaves the active site, the thiol group of the ppant-arm of the pcp-domain attacks the carboxylate carbon and at- taches to it, releasing the AMP leaving group (Gulick, 2009).

Structural changes in this process have been thor- oughly investigated, among others by Reimer et al.

(2016), who recently solved crystal structures of vari- ous conformations of the first module of the gramici- din synthase complex, LgrA (figure 3A). They showed that a series of conformational changes takes place during the two partial reactions. The A-domain firstly has an open conformation, but when ATP and the substrate (Valine in this case) bind to it the C-terminal subdomain rotates some 30º and closes, subsequently catalyzing the formation of the acyl-adenylate. Then a major alternation takes place: the C-terminal sub- domain rotates another 140º to allow access for the ppant-arm of the pcp-domain to attack the acyl-ade- nylate. This domain alternation changes the active site, but does not move the substrate. The structure was obtained by attaching a non-hydrolysable analogue of the product to the pcp-domain. Upon solving the structure, they found the product (Val-NH-pcp) still attached to the active site after the rotation. The next solved structure showed a 75º rotation and 61Å trans- location of the pcp-domains ppant arm, which they dedicate to a full 180º rotation of the C-terminal sub- domain of the A-domain, moving the pcp-substrate to - in this case - a formylation domain. See figure 3A for an overview of A-domain dynamics in an initiation module.

Because of their essential role in substrate recognition and activation, early researchers focused their interest on this domain, hoping to gain knowledge about this process and thereafter possibly modify it to activate non-native substrates. Stachelhaus et al. (1999) deter- mined a binding pocket in the N-terminal subdomain of the A-domain, consisting of 10 residues, which determines the specificity of the domain. Besides these, a lysine in the C-terminal subdomain has been found to be required for the acyl-adenylate formation.

This residue is poised in the active site, interacting with the carboxylate as well as the phosphate group on the AMP, but moving away at least 45Å upon the first major structure alternation (Miller & Gulick, 2017).

(7)

It was speculated that rationally mutating the afore- mentioned 10 active site residues would alter the substrate specificity of the domain, but numerous studies have shown that just changing this pocket did not lead to a product being synthesized, or only at a very low rate (Winn et al., 2016). Different views have surfaced upon the question why: for instance, it is

speculated that other domains like the condensation domain also have a certain selectivity (Uguru et al., 2004; see paragraph 4). More recent studies suggest that the reaction coupling between the A-domain and the pcp-domain might form a bottleneck. In engi- neered NRPSs with an altered A-domain specificity, the rate of activation moght be reduced – resulting in

Figure 3: The A-domain makes large conformational rearrangements and guides the pcp-domain movement. A) the catalytic cycle of an initiation module, LgrA (Reimer et al., 2016). B) the catalytic cycle of an termination module (Tano- vic et al., 2008). Yellow: large subdomain of the A-domain (AL ), grey: small subunit of the A-domain (AS ), multicoloured cartoon/pink circle: pcp-domain, brown: F-domain, blue: C-domain. Image derived from Kittilä et al., 2016.

(8)

a slowed motion of the pcp-domains ppant arm, and subsequently enlarging the chance of hydrolysis of the peptide intermediates due to the possible interception by water (Kaljunen et al., 2015). They state that slow- ing down the motion of the pcp-domain might help to avoid this problem. However, more research is needed to confirm that hypothesis.

4. C-DOMAIN

In the large condensation domains (±450 residues), the formation of a peptide bond is catalyzed between two activated substrates. Two substrates of adjacent modules bound to the pcp-domain as thioesters are brought to two opposite sides of the domain. The C-domain has a distinct V-shaped structure, which consists of two similar subdomains, called a CAT- fold (Marahiel et al., 1997). The active site is built up around the catalytic histidine found in the middle of the fold, which can be reached from both sides. That is exactly what happens when the substrates are brought towards the active site by the pcp-domain (Stachel- haus et al., 1998).

A model from Bloudoff et al. (2013), based on their extensive structural research of the first C-domain of CDA synthase (CDA-C1), suggests that the V-shaped latch is actually in a ‘closed’ state in CDA-C1, and that the substrates arrive at the active site through a tunnel.

They state that the tunnel entrance is 30Å long and wide enough to accommodate substrates. Also the entrance of this tunnel corresponds to what is known about binding sites for the up- and downstream pcp-domains. The ppant-moieties fit nicely into a pocket in the N-terminal subdomain, whereas the ser- ine faces the C-terminal subdomain. Also, the reactive amino group is around 3Å away from the catalytic histidine, as would be expected, for the intermediate to be in range of it (Bloudoff et al., 2013).

The closed state of the CDA synthase is not observed in other C-domain structures. For instance, in the SrfC-A structure, the lid seems to be opened. This change in configuration does not seem to by dynamic, nor does it seem to be the result of crystal packing: in both the opn SrfA-C and the closed CDA-C1 cases, the condensation domain is biochemically active. Ini- tially it was speculated that the state of the C-domain

had to do with the size of the intermediate substrate it has to accommodate - the larger the substrate, the more open the configuration. However, this does not seem to be the case, since the condensation domain in the terminal module SrfA-C is more open than CDA-C1, in spite of serving earlier in the assembly line of the NRPS thus accommodating a smaller intermediate. Therefore, the exact function of differ- ent states of the C-domain is currently unknown, and more research has to be performed to shed a light on this issue.

The C-domain generally has a low binding affinity for analogs of their natural substrate (Kittilä et al., 2016).

Unfortunately, the structural characterisation of the C-domain has remained limited due to the lack of substrate bound structures. However, major scientific advances have been made recently this field. In 2016, Bloudoff et al. developed a chemical probe which resembled the natural substrate to bind the C-domain covalently, allowing them to determine the structure of the complex. They suggest that the catalytic histi- dine, which has been generally cited as the C-domains active residue but with an unclear exact role, plays a major part in substrate positioning and orienta- tion. This tells us more about the domain controls the stereochemistry of the peptide synthesis, and the extra layer of control it has over A-domain selectivity.

Also, they determined other residues interacting with the ppant arm of the pcp-domain, which can now be assessed via site-directed mutagenesis to determine their role in C-domain substrate selectivity (Bloudoff et al., 2016).

The C-domain does not only act as a gatekeeper for substrate selectivity, but are also selecting stereo con- figuration (Stachelhaus et al., 1998). It seems that the domain only selects (l) amino acids as acceptor, while the donor residues can be (l) or (d). (Hur et al., 2012) This suggests that the C-domain is very similar to the so-called E-domain (which epimerizes acceptor sub- strates from the (l) to the (d)-form). These epimeriza- tion domains are often situated next to the C-domain, which is useful since most A-domains only select and activate (d)-configured substrates. Little structural information about these domains is currently available (Miller & Gulick, 2017).

(9)

5. TE-DOMAIN

Since all intermediates in NRP synthesis are covalently bound to pcp-domains via a thioester bond, the final product needs to be liberated from its respective car- rier protein after the peptide product chain has been completed. Thioesterase (TE) domains work to this end. Typically found at the end of the NRPS chain, TE-domains do often more than just cleaving the sub- strate: they may form all sorts of cyclisation products that characterize many important NRPs (Bruner et al., 2002). The Te-domain utilizes a highly conserved cat- alytic triad oftentimes found in other serine proteases, comprising a serine, asparagine and histidine, where the serine covalently attaches to the substrate which is subsequently either hydrolyzed or released by cycli- zation (Miller & Gulick, 2017). To perform the latter, an internal nucleophile on the bound peptide attacks the ester bond of the intermediate. This could be an amine, resulting in an amide product, or an alcohol group, resulting in an ester product, respectively. The nucleophile is usually one of the side chains of the amino acid residues but can also be the amino termi- nus of the peptide (Kittilä et al., 2016).

Furthermore, Te-domains offer the possibility to mul- timerize their substrates. Presumably a first peptide product is loaded onto the Te-domains active serine, after which a following pcp-bound peptide is used to cleave the ester, resulting in dimerized or even trimer- ized cyclic peptides (Jaitzig et al., 2014).

Aside from their catalytic triad, an alpha/beta hy- drolase fold is also highly conserved in Te-domains.

However, little sequence similarity is further observed in the ca. 280aa that make up the domain. In some Te-domains, like in the EntF-YbdZ Te-domain, con- tain a lid consisting of a-helices that can be in either an open or closed conformation (Miller et al., 2016;

figure 4B). It is speculated that this conformational change may block out water as a hydrolyzing agent from the active site, promoting the cyclization by an internal nucleophile.

6. MBTH-LIKE PROTEINS

It has been found that many NRPSs work in con- vention with small partner proteins (±70 residues), resembling and thus named after the mycobacterium tuberculosis mycobactin MbtH protein. These MbtH- like proteins (MLPs) are speculated to function as activators of NRPS adenylation domains, activating acyl-adenylate formation through inducing a structur- al alternation in the A-domain (Zhang et al, 2010). A recent solved structure of the complete EntF-module, bound to a MLP called YbdZ, counteracts this hy- pothesis (Miller et al., 2016). They show that no differ- ent structure is found in EntF with or without YbdZ bound. However, they state that EntF is not dependent on YbdZ to function completely, but it merely en-

Figure 4: conformational rearrangements of the Te-domain, and interactions with the pcp-domain. A) The Te-domain is in an open configuration and close to the pcp-domain, which is undergoing interactions with the C-domain in a full-module structure (Drake et al., 2016). B) The lid of the Te-domain rotates and closes upon binding with the pcp-do- main in an overlay of the Te-domain form an pcp-Te-structure with the whole-module Te-domain (Liu et al., 2011). Image derived from Kittilä et al., 2016.

A. B.

(10)

hances the activity of the A-domain. Therefore, it is possible that in other proteins, completely relying on MLPs to function, these helper proteins may indeed have a conformation-altering function. To assess this, crystal structures of these NRPS’ A-domains need to be solved, bound to MLPs as well as without them.

Until then the exact function of MLPs remains specu- lation.

7. STRUCTURES OF ENTIRE MOD- ULES AND NRPS COMPLEXES

It is clear that many structures of domains are now known, as well as structures of multiple domains together. Information about structure and dynamics of the basic domains has almost become common knowledge. However, it is evident that the domains of an NRPS maintain many mutual interactions, contrib- uting to the structure of the protein. To explore the importance of these interactions and the collaboration of the domains, it was essential to obtain structures of complete NRPS modules.

The first structure comprising an entire NRPS module was that of SrfA-C, the terminal module of surfactin biosynthesis (Tanovic et al., 2008). It contains a C-A- pcp-Te architecture. The ppant-cofactor was missing in their structure, making it difficult to predict the vital interactions between this domain and the rest of the protein. They propose that the A- and C-do- main form the core of a NRPS module, due to the extensive amount of interactions they share. Also they conveniently showed that the folds of the domains in the module are the same as in their dissected counter- parts.

Recently, some structures of complete holo-NRPS modules have been solved. The recent crystal structure of the linear gramicidin synthase initiation module (LgrA) has not only given some insights in the inter- actions between the pcp-domain and other domains, but also shows the structure and dynamics of the formylation domain (Reimer et al., 2016), hopefully rendering formylation an accessible reaction in future NRPS engineering. Most importantly, they show the three roles of the small subunit of the adenylation do- main: it provides catalytic residues for the adenylation reaction, it positions the pcp-domain-domain for the

thiolation reaction, and bridges the distance between the actives sites visited by the pcp-domain (figure 2A).

Miller et al. solved the EntF protein in 2016, a sin- gle-module NRPS found in E. coli, bound to its MLP YbdZ. The structures are in thioester-forming con- formation, with the pcp-domain covalently trapped to the adenylation domain, due to an inhibitor they introduced. In comparison to the other NRPS mod- ule structures with the same architecture, SrfA-C (mentioned above) and AB3403, they noticed that the Te-domain was in all cases in a very different locations. The Te-domain never seems to be mak- ing significant contacts with the rest of the protein, out of which they conclude that the location of this Te-domain within its module appears to be highly variable and only loosely based on the location of the pcp-domain. Furthermore, the linker regions between the pcp-domain and adenylation and Te-domains, respectively, are disordered in the crystal structure.

This suggests that these regions do not have a single conformation, but are instead highly flexible in the thioester forming conformation.

The structure of holo-AB3403, a protein in human pathogen Acinetobacter baumannii, was recently solved by Drake et al. (2016). The structure was solved with the ppant-moiety inside the C-domain, indi- cating important residues in the hydrophobic tunnel which interact with the pcp-domain (the latch of the C-domain was closed, just as in the CDA-C1 from Bloudoff et al., 2013). Drake et al. also present the structure of holo-EntF, with the pcp-domains cofactor covalently trapped in thioester formation conforma- tion in the A-domain (figure 4A).

Based on on the then available structures, Marahiel proposed a model for NRPS complexes in 2015 (figure 5). By superposing a pcp-C-domain structure and C-A-pcp structure from Tanovic et al. (2008) they obtained a pcp-C-A-pcp structure. After successively superposing these, they proposed a helical model, in which each module makes a 120º turn relative to its neighbour, along the helical axis. In the model, most A-C-didomains are directed outside, protecting the pcp-domains from solvent to prevent the intermedi- ates from hydrolysation. Since the C-domains of the different modules are modelled to be in close proxim- ity of each other, the pcp-domain would need only a simple rotational motion to deliver a substrate from

(11)

one domain to the other. The active site serine resi- dues of the pcp-domains are about 45Å apart, which means that the pcp-domains have to make a substan- tial rearrangement to pass along their substrates.

8. CONCLUSION AND FUTURE PROSPECTS

Despite the grand amount of structures already avail- able, a full multi-modular NRPS crystal structure re- mains to be solved. However, the recent model of Ma- rahiel (2015) gives a possible insight into the structure of the modular organisation of the mega-enzymes.

The model has still not been confirmed or countered, however, by either crystallography or cryo-EM. That means that the model still is debatable.

Also, more recent and larger structures of LgrA (Re- imer et al., 2016), Miller (2016) and Drake (2016), for example, may be taken into account with an evalu- ation of the model. Especially the structure of LgrA might shed a new light on the multi modular compo- sition of NRPSs when superposed on known struc- tures, given the amount of structures available from this complete module and the presence of a formyla- tion domain. Furthermore, Drake et al. show that the interface area between the A- and C-domains varies in the three structures, and also the conformation is not

Figure 5: The helical model of a seven-module NRPS from Marahiel et al., 2015. Each module is turned 120º around the helical axis, with the C-A-domain platforms oriented outwards towards the solvent and the pcp-domains inwards. In the right-above panel a front view is also shown.

Red: A-domain large subunit, brown: A-domain small subunit, grey: C-domain, green: pcp-domain, orange: Te-domain.

Image derived from Marahiel et al., 2015.

(12)

the same. Also they show that the C-domain of the EntF protein is turned some ±25º relative to the A-do- main when compared to SrfA-C. This suggests that the A-C-domain platform, the basis of Marahiels model, is not as rigid and static as previously proposed.

Intrigueing, though, is the fact that in the LgrA structure the pcp-domain makes a translocation of (on average) 61Å when transferring its substrate from one domain to the next. This means that the active serine residues of the pcp-domains of the consecutive modules, 45Å apart in Marahiels model, are in reach of each other.

For further understanding of the organization and dynamics of NRPSs, more structures of condensation domains with bound substrate are needed. Availability of these structures may help in understanding the role of certain residues in binding and catalysis, and also make clear how C-domains are able to cope with the large intermediates usually found at the end of a long NRPS module chain.

This information is vital in the effort of creating en- gineered NRPSs with a changed substrate specificity.

Where large advances are made yearly in engineering A-domains to change the specificity of NRPSs, the turnoves often stays low (Kittilä et al., 2016). Struc- tural data of downstream processing of the intermedi- ates may allow researchers to rationally engineer the downstream NRPS domains and modules to better ac- comodate non-native substrates, which might improve the turnover of these engineered enzymes.

As of now, however, many more advances still need to be made to design a functional custom NRPS. It has become clear that not only interactions within do- mains, but also within modules and between modules play a large role in selectivity and activity of the me- ga-enzymes. More research into this subject is crucial for engineering NRPS machineries, since it seems that knowledge behind the structure and sources of sub- strate specificity is lagging behind the advances in in vivo NRPS reengeneering (Winn, 2016). Fortunately, the ever-growing interest in this subject (shown by the large number of publications released recently) will help us to this end.

(13)

9. REFERENCES

Beld, J., Sonnenschein, E. C., Vickery, C. R., Noel, J. P., & Burkart, M. D.

(2014). The phosphopantetheinyl transferases: catalysis of a post-transla- tional modification crucial for life. Natural product reports, 31(1), 61-108.

Bloudoff, K., Alonzo, D. A., & Schmeing, T. M. (2016). Chemical probes allow structural insight into the condensation reaction of nonribosomal peptide synthetases. Cell chemical biology, 23(3), 331-339.

Bloudoff, K., Rodionov, D., & Schmeing, T. M. (2013). Crystal structures of the first condensation domain of CDA synthetase suggest confor- mational changes during the synthetic cycle of nonribosomal peptide synthetases. Journal of molecular biology, 425(17), 3137-3150.

Bruner, S. D., Weber, T., Kohli, R. M., Schwarzer, D., Marahiel, M. A., Walsh, C. T., & Stubbs, M. T. (2002). Structural basis for the cyclization of the lipopeptide antibiotic surfactin by the thioesterase domain SrfTE.

Structure, 10(3), 301-310.

Drake, E. J., Miller, B. R., Shi, C., Tarrasch, J. T., Sundlov, J. A., Allen, C.

L., ... & Gulick, A. M. (2016). Structures of two distinct conformations of holo-non-ribosomal peptide synthetases. Nature, 529(7585), 235-238.

Goodrich, A. C., Harden, B. J., & Frueh, D. P. (2015). Solution structure of a nonribosomal peptide synthetase carrier protein loaded with its substrate reveals transient, well-defined contacts. Journal of the American Chemical Society, 137(37), 12100-12109.

Gulick, A. M. (2009). Conformational dynamics in the Acyl-CoA syn- thetases, adenylation domains of non-ribosomal peptide synthetases, and firefly luciferase. ACS chemical biology, 4(10), 811-827.

Hur, G. H., Vickery, C. R., & Burkart, M. D. (2012). Explorations of cata- lytic domains in non-ribosomal peptide synthetase enzymology. Natural product reports, 29(10), 1074-1098.

Jaitzig, J., Li, J., Süssmuth, R. D., & Neubauer, P. (2013). Reconstituted biosynthesis of the nonribosomal macrolactone antibiotic valinomycin in Escherichia coli. ACS synthetic biology, 3(7), 432-438.

Kaljunen, H., Schiefelbein, S. H., Stummer, D., Kozak, S., Meijers, R., Christiansen, G., & Rentmeister, A. (2015). Structural Elucidation of the Bispecificity of A Domains as a Basis for Activating Non‐natural Amino Acids. Angewandte Chemie International Edition, 54(30), 8833-8836.

Kittilä, T., & Cryle, M. J. (2016). Capturing the Structure of the Substrate Bound Condensation Domain. Cell chemical biology, 23(3), 315-316.

Kittilä, T., Mollo, A., Charkoudian, L. K., & Cryle, M. J. (2016). New Struc- tural Data Reveal the Motion of Carrier Proteins in Nonribosomal Peptide Synthesis. Angewandte Chemie International Edition, 55(34), 9834-9840.

Lai, J. R., Fischbach, M. A., Liu, D. R., & Walsh, C. T. (2006). A protein interaction surface in nonribosomal peptide synthesis mapped by combi- natorial mutagenesis and selection. Proceedings of the National Academy of Sciences, 103(14), 5314-5319.

Liu, Y., Zheng, T., & Bruner, S. D. (2011). Structural basis for phosphopan- tetheinyl carrier domain interactions in the terminal module of nonribo- somal peptide synthetases. Chemistry & biology, 18(11), 1482-1488.

Marahiel, M. A., Stachelhaus, T., & Mootz, H. D. (1997). Modular peptide synthetases involved in nonribosomal peptide synthesis. Chemical re- views, 97(7), 2651-2674.

Marahiel, M. A. (2015). A structural model for multimodular NRPS assembly lines. Natural product reports, 33(2), 136-140.

Miller, B. R., & Gulick, A. M. (2016). Structural biology of nonribosomal peptide synthetases. Methods of Molecular Biology, 3-29.

Miller, B. R., Drake, E. J., Shi, C., Aldrich, C. C., & Gulick, A. M. (2016).

Structures of a Nonribosomal Peptide Synthetase Module Bound to MbtH-like Proteins Support a Highly Dynamic Domain Architecture.

Journal of Biological Chemistry, 291(43), 22559-22571.

Payne, J. A., Schoppet, M., Hansen, M. H., & Cryle, M. J. (2017). Diversity of nature’s assembly lines–recent discoveries in non-ribosomal peptide synthesis. Molecular BioSystems, 13(1), 9-22.

Reimer, J. M., Aloise, M. N., Harrison, P. M., & Schmeing, T. M. (2016).

Synthetic cycle of the initiation module of a formylating nonribosomal peptide synthetase. Nature, 529(7585), 239-242.

Stachelhaus, T., Mootz, H. D., & Marahiel, M. A. (1999). The specific- ity-conferring code of adenylation domains in nonribosomal peptide synthetases. Chemistry & biology, 6(8), 493-505.

Stachelhaus, T., Mootz, H. D., Bergendahl, V., & Marahiel, M. A. (1998).

Peptide bond formation in nonribosomal peptide biosynthesis catalytic role of the condensation domain. Journal of Biological Chemistry, 273(35), 22773-22781.

Tanovic, A., Samel, S. A., Essen, L. O., & Marahiel, M. A. (2008). Crystal structure of the termination module of a nonribosomal peptide syn- thetase. Science, 321(5889), 659-663.

Uguru, G. C., Milne, C., Borg, M., Flett, F., Smith, C. P., & Micklefield, J.

(2004). Active-site modifications of adenylation domains lead to hydroly- sis of upstream nonribosomal peptidyl thioester intermediates. Journal of the American Chemical Society, 126(16), 5032-5033.

Uhlmann, S., Süssmuth, R. D., & Cryle, M. J. (2013). Cytochrome p450sky interacts directly with the nonribosomal peptide synthetase to generate three amino acid precursors in skyllamycin biosynthesis. ACS chemical biology, 8(11), 2586-2596.

Walsh, C. T., O’Brien, R. V., & Khosla, C. (2013). Nonproteinogenic amino acid building blocks for nonribosomal peptide and hybrid polyketide scaffolds. Angewandte Chemie International Edition, 52(28), 7098-7124.

Winn, M., Fyans, J. K., Zhuo, Y., & Micklefield, J. (2016). Recent advances in engineering nonribosomal peptide assembly lines. Natural Product Reports, 33(2), 317–47.

Zhang, W., Heemstra Jr, J. R., Walsh, C. T., & Imker, H. J. (2010). Activa- tion of the pacidamycin PacL adenylation domain by MbtH-like proteins.

Biochemistry, 49(46), 9946-9947.

Referenties

GERELATEERDE DOCUMENTEN

4 we demonstrated that the proper motions used by Schwan (1991) result in a different convergent point from that derived from the Hipparcos proper motion data, and that the

License: Licence agreement concerning inclusion of doctoral thesis in the Institutional Repository of the University of Leiden. Downloaded

Storm (Technische Universiteit Eindhoven) Overige leden: prof.. Peletier (Technische Universiteit

If we build a membrane out of a single species of lipids which has unsaturated tails, we get a structure which is different from a membrane consisting of lipids with saturated

The energy is in this case given by the (mean) curvature en- ergy with two Lagrange multiplier terms, one for the area (where the multiplier is the surface tension) and one for

(a) Gibbs phase triangle showing the binodal (thick black line), tie lines (thin black lines), spinodal (blue line) and critical points (red dots).. Some optimal quadratic

The aim of the current paper is twofold: 共1兲 to explain previous seemingly contradictory ex- perimental results on the two-dimensionality of shallow flows 3 —in particular, the

Our experimental results suggest that sawtooth crash models should contain two ingredients to be consistent with experimental observations: 共1兲 the 共1,1兲 mode structure should