Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

DOI: 10.1002/cbic.

201402578 Reviews

Using Modern Tools To Probe the Structure–Function


Relationship of Fatty Acid Synthases
Kara Finzel, D. John Lee, and Michael D. Burkart*[a]

ChemBioChem 0000, 00, 0 – 0 1  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim &

These are not the final page numbers! ÞÞ


Reviews

Fatty acid biosynthesis is essential to life and represents one of lack of detailed molecular information behind the ACP–partner
the most conserved pathways in nature, preserving the same protein interactions inherent to the pathway. This review will
handful of chemical reactions across all species. Recent interest focus on recently developed tools for the modification of ACP
in the molecular details of the de novo fatty acid synthase and analysis of protein–protein interactions, such as mecha-
(FAS) has been heightened by demand for renewable fuels and nism-based crosslinking, and the studies exploiting them. Dis-
the emergence of multidrug-resistant bacterial strains. Central cussion specific to each enzymatic domain will focus first on
to FAS is the acyl carrier protein (ACP), a protein chaperone mechanism and known inhibitors, followed by available struc-
that shuttles the growing acyl chain between catalytic en- tures and known interactions with ACP. Although significant
zymes within the FAS. Human efforts to alter fatty acid biosyn- unknowns remain, new understandings of the intricacies of
thesis for oil production, chemical feedstock, or antimicrobial FAS point to future advances in manipulating this complex
purposes has been met with limited success, due in part to a molecular factory.

1. Overview of Fatty Acid Biosynthesis unit by a transacylase. A ketosynthase (KS) elongates the chain
by two carbons through decarboxylative addition with malon-
The de novo fatty acid synthase (FAS) is ubiquitous in nature yl-ACP as a substrate. The resulting ketone is reduced to an
and critical for life. New technologies have emerged to coin- (R)-alcohol by a ketoreductase (KR), which is then eliminated to
cide with renewed interest in FAS because of recent societal a trans-alkene by a dehydratase (DH). A final reduction step by
demands. One such demand is the sustainable production of an enoyl reductase (ER) yields an elongated acyl chain, which
hydrocarbon fuels and feedstocks, due to dwindling supplies can then be subjected to the same cycle again until the de-
of easily accessible fossil fuels. As the central metabolic foun- sired chain length is reached. Termination is achieved through
dry for hydrocarbon production in nature, FAS is seen by many chain release by either a thioesterase (TE) or chain transfer by
as the first target for engineering diverse hydrocarbon produc- an acyl transferase (AT) for incorporation into fats, lipids, and
tion by synthetic biology. Another pressing concern is the other metabolites.
need for new antibiotics. Enzymes within the FAS pathway FASs are classified into two categories, type I and type II (Fig-
offer untapped targets for future inhibitor development, as the ure 1 B). Type I FAS systems use large, multidomain proteins
arms race against antibiotic resistance requires new drugs arranged into a single complex, such that a tethered ACP can
against such linchpin targets. Luckily, tools for interrogating access all required active sites for iteration and synthesis.
and engineering FASs have emerged in recent years to help Type I FAS complexes are common in animals and fungi. For
address bottlenecks in our understanding. Although significant example, the fungal type I FAS is encoded by two genes that
work in the 20th century identified and investigated the pro- assemble as an a6b6-heterododecamer reaching 2.6 MDa (PDB
teins involved in FAS, only recently has it become possible, IDs: 2UVB, 2UVC).[8] The mammalian type I FAS expresses as a
through structural biology combined with mechanistic probes single protein and assembles as a homodimer of 540 kDa (PDB
and advanced kinetics, to interrogate discrete enzymes during ID: 2VZ8).[9]
productive interactions. This review will focus on such new In contrast, type II FASs employ many single domains ex-
tools and progress to date that will inform meaningful future pressed as discrete proteins in the cytosol. Type II FASs are
developments. common in bacteria and eukaryotic organelles, notably chloro-
As a very well-studied system, much of the work discussed plast and mitochondria. Large amounts of FAS proteins are ob-
in this review will focus on the Escherichia coli FAS, with care served in type II systems, with the central and well-character-
taken to denote other organisms as necessary. A thorough un- ized E. coli ACP[10] comprising up to 0.25 % of soluble protein
derstanding of FASs will allow for new engineering and exploi- during logarithmic growth.[11] The ACP shuttles the growing
tation, and many of the tools discussed here should also prove chain through each step of the elongation cycle and must find
useful in the study of FAS-related secondary metabolites from the correct reaction partners in the crowded cytosol. This ne-
polyketide synthases (PKSs). PKSs are beyond the scope of this cessitates multiple, low affinity protein–protein interactions, in-
review, and we refer to recent reviews from Khosla and Kea- volving more than 30 partners once regulatory and other pri-
tinge-Clay.[1–7] mary and secondary metabolic interactions are considered.[12, 13]
Fatty acid biosynthesis (FAB) is an iterative series of enzyme FAB in archaea is largely uncharacterized, and its existence is
reactions (Figure 1 A) in which an acyl chain is extended by controversial due to archaeal membrane phospholipids lacking
a two-carbon unit with each cycle. Initiation occurs when an fatty acids.[14] In 2012, homologues of components of bacterial
acyl carrier protein (ACP) is charged with a two-carbon starting FAS II were found in the archaeal genome, with the exception
of ACP, suggesting the possibility of ACP-independent FAB.[15]
[a] K. Finzel,+ D. J. Lee,+ Prof. M. D. Burkart In 2014, archaeal fatty acid biosynthesis was proposed to be
Department of Chemistry and Biochemistry
performed by bacterial-type enzymes of fatty acid b-oxidation
University of California, San Diego
La Jolla, CA 92093-0358 (USA) and an enzyme from the mevalonate biosynthesis pathway.[16]
E-mail: mburkart@ucsd.edu
[+] These authors contributed equally to this work.

& ChemBioChem 0000, 00, 0 – 0 www.chembiochem.org 2  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ÝÝ These are not the final page numbers!


Reviews

Figure 1. A) Schematic diagram of the FAB cycle. B) Comparison of domain organization in type I and type II FAS systems. C) Nomenclature of E. coli Fab en-
zymes, listed underneath their domain activities. D) A color-coded diagram of the type I mammalian FAS, demonstrating an overall “gingerbread-man” topolo-
gy (PDB ID: 2VZ8). The TE and the ACP are not observed, due to dynamics.

2. The Acyl Carrier Protein: New Tools

Due to ACP’s central position in FAB, consideration of ACP–


partner protein interactions is a critical component of contem-
porary studies of FASs. New tools to examine and modify ACPs
Kara Finzel graduated from the Univer-
have been developed to facilitate this over the past decade.
sity of Richmond in 2011, where she
Classically, ACPs are small (~ 9 kDa) a-helical proteins. When
conducted research under the tutelage
part of a type I megasynthase, the ACP domain is attached by
of Dr. John Gupton and received a BS
flexible N- and C-terminal peptide linkers, allowing it to sample
in chemistry. She has been in the PhD
reaction partners for FAB. Computational work has suggested
program at the University of California,
that the ACP’s motion is stochastic in type I FASs.[17] Discrete
San Diego in the Department of
type II ACPs must travel through the cytosol in bacteria to find
Chemistry and Biochemistry under the
reaction partners. The ACP is expressed as an inactive apo-ACP,
guidance of Prof. Michael Burkart since
which is known to be cytotoxic in E. coli.[18] Activation by post-
2011. Her current work is focused on
the development of new tools for in
vitro and in vivo carrier protein label- Michael Burkart received a BA in
ing to study protein–protein interac- chemistry from Rice University in 1994
tions in fatty acid biosynthesis. and a PhD in organic chemistry from
the Scripps Research Institute in 1999
D. John Lee earned BS degrees in under Prof. Chi-Huey Wong, after
chemistry and biology from the Uni- which he pursued postdoctoral studies
versity of California, Santa Cruz. While at Harvard Medical School with Prof.
there, he studied under Prof. Bakthan Christopher Walsh. Since 2002, he has
Singaram, advancing boronic asym- been a full Professor of Chemistry and
metric reducing agents. He joined Biochemistry and Associate Director of
Prof. Michael Burkart’s lab in 2010 as the California Center for Algae Biotech-
a doctoral student. He currently stud- nology at the University of California,
ies protein–protein interactions of FAS San Diego. His research interests include natural product synthesis
and PKS enzymes by NMR, exploiting and biosynthesis, enzyme mechanism, and metabolic engineering.
chemical tools developed in the Bur- Prof. Burkart has been a fellow of the Alfred P. Sloan and Ellison
kart laboratory. Medical Foundations.

ChemBioChem 0000, 00, 0 – 0 www.chembiochem.org 3  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim &

These are not the final page numbers! ÞÞ


Reviews
translational attachment of a CoA-derived 4’-phosphopante- forming acyl-ACP. Termination of FAS by TE or AT activity re-
theine (PPant) to a conserved serine (Ser36 in E. coli) yields generates holo-ACP.
holo-ACP (Figure 2 B). FAB occurs via thioester-linked inter- Type II ACPs sequester the growing chain within a hydropho-
mediates attached to the terminal thiol of the PPant arm, bic pocket formed in the core of the helical bundle, which
expands based on the acyl chain length.[19] Sequestration has
been proposed to shape ACP for subsequent protein–protein
interactions, define substrate length, and/or protect the labile
thioester bond from hydrolytic cleavage.[20, 21] Recent work
exploiting solvatochromic probes[22] and vibrational probes[23]
appended to the PPant arm of ACPs have allowed direct obser-
vation of both sequestered and nonsequestered microstates.
Type I ACPs are not thought to sequester, as observed by
a lack of structural perturbations by NMR of the excised ACP[24]
and solvatochromic probes.[22] Other studies have revealed that
the structure of the ACP is modulated by its acylation state,
including gel shifts observed by conformationally sensitive
PAGE[25, 26] and resonance shifts observed by NMR spectrosco-
py.[27, 28]
Observation of sequestration poses several intriguing ques-
tions. Are ACP–partner interactions stochastic, or does ACP
communicate its cargo’s status? Some studies have suggested
that cargo-mediated minor changes to the orientation of helix
III might be involved,[29] but the importance of these small
structural changes has not yet been thoroughly validated. How
does the ACP translocate the reactive chain to the partner pro-
tein? This process was originally named the “switchblade
mechanism”,[30] but Cronan recently argued that “chain flip-
ping” (Figure 2 A) might be a more accurate description, as no
evidence of activated release of the acyl chain has been ob-
served.[31] Do structural changes to the ACP, when sequestering
elongated intermediates, signal for TE or AT release of the fatty
acid? Evidence has shown that a fully elongated acyl-ACP
maintains a hairpin-like structure of the acyl chain[32] that the
TE may recognize.
In nature, the ACP is post-translationally modified from apo
to holo form by a phosphopantetheinyl transferase (PPTase). A
timely review discussed this class of enzymes in great detail,[33]
and only a few key points will be mentioned here. PPTases uti-
lize the high-energy CoA as the substrate for transfer. Three
different subfamilies of PPTases have been observed. The AcpS
(holo-ACP synthase)-type PPTases activate type II FAS ACPs and
show very little promiscuity towards non-target ACPs.[34] In
E. coli, AcpS is responsible for activation of apo-ACP (Fig-
ure 2 B). In contrast, Sfp-type PPTases are highly promiscuous
and often used in conjunction with heterologous synthase ex-
pression.[35] Indeed, heterologous expression of homogeneous
holo-ACPs can be achieved by co-expressing recombinant Sfp
with recombinant ACP,[36] whereas without additional Sfp, a
mixture of apo-ACP and holo-ACP is often expressed. The Sfp
Figure 2. A) Demonstration of chain flipping: at left is ACP sequestering a C7 family is named for the archetypical B. subtilis surfactin syn-
chain (PDB ID: 2FAD), and at right is an ACP from the crosslinked complex
thase activator. The third family is comprised of PPTases inte-
with DH (PDB ID: 4KEH). B) Schematic of tools utilized to modify ACP natu-
rally in vivo and in vitro. Often, due to endogenous PPTase activity, heterolo- grated in type I fungal FAS megasynthases.
gous expression of ACPs yields a mixture of apo- and holo-ACP species. R’, Generation of homogeneous apo-, holo-, acyl- and (unnatu-
R’’, and R’’’ are used to denote variability limited by the specificity of the ral cargo bearing) crypto-ACPs has been critical to recent stud-
pathway. R’ can be acyl chains, azides, alkynes, or halogens, limited by the
ies. Many new strategies have been employed to manipulate
specificity of AasS. R’’ is limited by the promiscuity of the specific CoA ligase
used. R’’’ can be many different probes, from acyl chains to fluorescent moi- the cargo attached to the ACP. Certain fatty acids can be di-
eties, as well as replacement of the thioester by an amide. rectly ligated onto CoA by using an acyl-CoA ligase.[37] Once

& ChemBioChem 0000, 00, 0 – 0 www.chembiochem.org 4  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ÝÝ These are not the final page numbers!


Reviews
the acyl-CoA is prepared, addition of Sfp and apo-ACP gener- These mechanism-based crosslinkers covalently attach ACP to
ates crypto-ACP. The primary limitation of this approach is the a partner protein and have been developed for KS,[44, 51]
requirement of an acyl-CoA ligase with activity for the desired DH,[52, 53] and TE[54, 55] activities, allowing for crystallographic and
fatty acid. Many studies of ligases are available,[37–39] with each NMR studies of these complexes. These studies will be dis-
ligase demonstrating substrate specificities. It also requires cussed in the subsequent sections.
apo-ACP, which has historically been challenging to obtain ho- Another ACP cargo-manipulating tool has experienced re-
mogeneously. Finally, this method produces a thioester, which newed interest, with applications both in vitro and in vivo. The
can be beneficial in that it reflects the natural linkage but also acyl ACP synthetase (AasS) uses ATP to ligate fatty acids direct-
suffers from slow hydrolysis in solution. In long experiments, ly onto holo-ACP (Figure 2 B). The Vibrio harveyi AasS has been
such as multidimensional NMR, such hydrolysis might become shown to load a range of unique carboxylic acids, including
detrimental. acids with azide, alkyne, and halogen functionalities, onto
Interest in generating CoA analogues without using acyl- holo-ACP in vitro. AasS allows simple, direct production of
CoA ligase led to the development of a “one-pot” chemoenzy- acyl- and crypto-ACPs with the natural thioester linkage, al-
matic synthesis.[40, 41] Synthetically obtained pantetheine ana- though there are some substrate limitations.[56] In E. coli, no cy-
logues[42, 43] can be converted to the corresponding CoA ana- tosolic AasS activity has been observed, and exogenous fatty
logues by exploiting CoA biosynthesis through the use of a acids can only be degraded or used as substrates for glycero-
pantothenate kinase, phosphopantetheine adenylyl transferase, phospholipid synthesis, depending on chain length. By heterol-
and dephosphocoenzyme A kinase (CoaA, CoaD, and CoaE, re- ogous expression in E. coli with a recombinant AasS, exoge-
spectively, in E. coli).[44] Once prepared (isolated or in situ), the nously supplied unique carboxylic acids of varying structures
CoA analogue can be loaded onto apo-ACP by a PPTase, com- can be imported directly into the FAS pathway and extended.
monly Sfp (Figure 2 B). Due to the inherent promiscuity shown This provides a new avenue to long chain terminally modified
by all of the “one-pot” proteins, probes of varying size and carboxylic acids that can be synthetically inaccessible and
functionality, including fluorescent molecules, intermediate an- a way to incorporate unique moieties into FAS and lipid prod-
alogues, and activity-based warhead moieties, can be loaded ucts.[56]
onto the ACP for visualization, isolation, and functional and
structural studies. Additionally, pantetheine probes with an
3. Structural Techniques for Studying the
amide or oxoester linkage to replace the natural thioester link-
Protein–Protein Interactions of FASs
age can be loaded onto apo-ACP as a means of preventing
slow hydrolysis in solution. The “one-pot” approach greatly Both NMR spectroscopy and X-ray crystallography have been
broadens the scope of loadable substrates, but its application used extensively to study FAS, and the two techniques com-
depends upon access to homogenous apo-ACP. plement each other. Solution-phase NMR is well equipped to
In E. coli, the ratio of apo- to holo-ACP is thought to offer study small, dynamic proteins in a buffered environment,
regulatory control of the FAS.[18] The small protein ACP hydro- though line broadening inhibits the study of large complexes.
lase/phosphodiesterase (AcpH) is responsible for cleaving 4’- Crystallography is well suited for ordered proteins and com-
PPant from holo-ACP to generate apo-ACP.[45] Since its discov- plexes but is not ideal for dynamic proteins and flexible re-
ery in the 1960s, AcpH has been scarcely used in vitro due to gions.
its poor stability characteristics, although it was observed to Computational approaches include molecular dynamics
be very active.[46] In the past, a thiol-Sepharose resin was prin- (MD), docking, and longer timescale accelerated molecular
cipally used to separate holo-ACP from apo-ACP through for- dynamics (AMD) simulations. These approaches have been
mation of a disulfide bond, but this method suffered from the exploited significantly in the last few years, as computational
loss of large quantities of protein and was only possible with power has risen and available structural information has
proteins lacking surface cysteine residues.[47] The use of a par- grown. Microscopic techniques will be discussed briefly.
tially active AcpS mutant in E. coli has also been successfully
leveraged for the overexpression of mostly apo-ACP.[18] Recent-
3.1. X-ray studies and techniques
ly, the Pseudomonas aeruginosa AcpH has been developed as
a much more amenable alternative to generate apo-ACP.[48] X-ray crystallography is a well-established technique for visuali-
Both resin-attached and free protein techniques have been zation of proteins trapped in a rigid crystal lattice. Crystallo-
developed, allowing for the preparation of large quantities of graphic studies of FASs have yielded many structures of both
apo-ACP and the recycling of high-value, isotopically enriched type I and type II FAS proteins. Due to their connected domain
ACPs for NMR use.[49, 50] Reversible tagging has allowed for architecture, only type I FAS complexes have yielded crystal
quantitative “apo-fication” of holo- and crypto-ACPs, yielding structures where most or all of the individual enzymes have
homogenous apo-ACP for further modification (Figure 2 B). been visualized together, whereas type II FAS enzyme struc-
Due to their transient nature, ACP–partner protein interac- tures have mostly been elucidated individually.
tions are difficult to directly observe. In an attempt to trap this Crystal structures of type I megasynthases include the ho-
interaction, we developed crosslinking probes by leveraging modimeric mammalian “gingerbread-man” topology (Fig-
mechanism-based inhibitors incorporated into pantetheine an- ure 1 D; PDB ID: 2VZ8),[9] the Thermomyces lanuginosus fungal
alogues and attaching them to ACP with “one-pot” methods. a6b6-heterododecamer “soccer ball” (PDB IDs: 2UVB, 2UVC),[8]

ChemBioChem 0000, 00, 0 – 0 www.chembiochem.org 5  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim &

These are not the final page numbers! ÞÞ


Reviews
and the differently scaffolded Saccharomyces cerevisiae yeast mentioned above, slow timescale dynamics are observable by
fungal FAS (PDB ID: 2UV8).[30] One critical challenge for crystal- AMD.
lographers studying FASs is the highly dynamic ACP. Indeed, in MD simulations of acylated ACPs have explored the seques-
type I structures, these dynamic regions are not resolved, with tration preference of acyl-ACP and the flexibility in the acyl
the exception of the stalled ACP in the S. cerevisiae structure. conformations.[63] Simulations of the ACP within the type I
Crystal structures of individual type II disparate proteins are fungal FAS have revealed that the ACP’s motion is likely sto-
available and will be discussed in the following sections. Cur- chastic in nature.[17] Optimization by NMR observed residual
rently, only one crosslinked protein crystal structure from dipolar couplings[64, 65] that encode long-timescale motions (ms–
a native FAS pathway has been solved, demonstrating the ACP ms),[66, 67] enabled AMD comparison of long-timescale dynamics
trapped during interaction with a partner (section 8.3).[57] Sig- of acyl-ACP and crosslinked ACP-DH, revealing significant
nificant conformational changes of the ACP are well-visualized losses in flexibility and suggesting the ACP–DH interaction is
in the crosslinked structure with FabA (PDB ID: 4KEH), in which well ordered.[57] Docking studies will be explored in the sec-
helix II and helix III are bound by multiple salt bridges and hy- tions below, as part of the analyses of individual domains.
drophobic interactions to allow chain translocation, and the Electron microscopy (EM) is an important tool that has been
hydrophobic pocket has collapsed. As a tool, X-ray crystallogra- gaining popularity for the visualization of type I FASs. Early
phy has provided much of the key topological information work allowed observation of the mammalian type I FAS and
about FASs, which will be discussed in the subsequent sections the two multidomain monomers, but in limited resolution.[68]
(5–10). Recent breakthroughs include Grininger’s 6  EM map of the
fungal FAS[69] and mycobacterial type I FAS.[70] EM offers a prom-
ising method for study of the FAS in a variety of states and
3.2. NMR studies and techniques conformations that are expected to become more prominent
in the future.
NMR studies enable observation of the dynamics of proteins
and can provide structural data on portions of type I FASs that
cannot be observed by crystallography. Full structural determi-
3.4. Inhibitors
nation by NMR is data- and time-intensive, whereas simpler
techniques such as chemical shift perturbation (CSP) analysis In the partner domain sections (5–10), inhibitors will be dis-
offer easily accessible but limited information. CSP analysis cussed in detail. The search for exploitable drugs has been
allows identification of important residues in protein–protein a primary driving force for inhibitor development in the 20th
or protein–ligand interactions by simple experiments.[58, 59] century. As FAB is essential for life, and humans and bacteria
Structures have been solved for a series of acylated ACPs harbor different machinery, FASs have been a major target of
from the spinach type II FAS, revealing that ACP could comfort- interest. However, given the importance of this target and the
ably sequester ten carbons but no more and that the solvent- sheer number of druggable activities, relatively few FAS-target-
exposed hairpin formed by longer chains might be a recogni- ing drugs have been developed. More numerous are known in-
tion motif for thioesterase activity.[32] Structural calculations hibitor probes that have been discovered, often by accident,
and CSPs observed when comparing different reactive inter- over the course of the last few decades.[71] Mechanism-based
mediates during extension and reduction of C6-ACP to C8-ACP “suicide” inhibitors have recently become additionally impor-
demonstrated very little perturbation of helix II and cargo- tant in the development of useful probes for elucidation of
mediated changes to helix III.[29] The type I mammalian ACP, structure–function relationships in FASs. Noncovalent inhibitors
when truncated from the megasynthase, did not demonstrate have allowed the study of transient interactions important for
CSPs when comparing acylated and holo-ACP, suggesting that enzyme kinetics and structural studies. Protein conformational
type I ACPs do not sequester their cargo.[24] changes that occur upon inhibitor, cofactor, and/or substrate
Dynamics observations are useful for determining which binding can be used together to study important structural, ki-
portions of proteins are undergoing motion on a given time- netic, and thermodynamic properties. We will highlight known
scale. Fast timescale dynamics can be accessed by classical inhibitors of each catalytic function in the following sections.
relaxation experiments.[60] The application of the Lipari-Szabo
Model-free approach[61, 62] allows quantification of flexibility res-
idue-by-residue as a generalized order parameter, revealing 4. FAS Partner Domains
significant flexibility of the loop near the post-translationally
Sections 5–10 focus on six FAS domains (MCAT, KS, KR, DH, ER,
modified serine.[28]
and TE/AT). A brief description of the enzyme will precede in-
formation about its catalytic mechanism, inhibitors with mech-
anisms of action, and what is currently known about interac-
3.3. Computational studies and other techniques
tions with ACP. Tools for modifying ACP and tools for visualiz-
Computational studies have yielded significant information ing protein–protein interactions described earlier will be em-
about protein dynamics and flexibility. Classic MD simulations phasized as applicable for each domain. This review highlights
and docking studies have been used to predict binding and that, whereas the mechanisms and activities of individual do-
develop inhibitors. In addition to the fast-timescale dynamics mains are now well understood, individual domain interactions

& ChemBioChem 0000, 00, 0 – 0 www.chembiochem.org 6  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ÝÝ These are not the final page numbers!


Reviews
with the ACP remain the next important step for studying E. coli, FabD, is specific for malonyl-CoA and unable to use
FASs. acetyl-CoA for loading. In vitro, many reaction partners have
been observed, including CoA, ACP, pantetheine, and N-(N-
acetyl-b-alanyl)cysteamine.[72]
5. Malonyl-CoA:ACP Transacylase (MCAT) Deletion of fabD in E. coli is lethal,[73] whereas overexpression
Malonyl-CoA:ACP transacylase (MCAT) catalyzes the transacyla- of fabD in E. coli leads to an increase in cis-vaccenic acid (cis-
tion of malonate from malonyl CoA to holo-ACP, which is the 11-octadecenoic acid), due to an increase in malonyl-ACP con-
first step towards fatty acyl chain elongation (Figure 3 A). MCAT centration and enhanced ketosynthase activity.[74] However, as
is a serine a/b-hydrolase in which the active site serine (Ser92 MCAT does not catalyze a rate-limiting step in FAB, overexpres-
in E. coli) first becomes malonylated, and the enzyme then sion of fabD in E. coli leads to only an 11 % increase in fatty
transfers the malonyl group to ACP. The MCAT domain in acid content.[75]

5.1. Mechanism
MCAT completes malonyl transfer with the double-displace-
ment or “ping-pong” mechanism common to a/b-hydrolases
by using its His-Ser catalytic domain (Figure 3 B).[76, 77] Double-
displacement is a two-step mechanism: first, the oxyanion hole
of MCAT is occupied by the binding of malonyl-CoA, and trans-
acylation of malonate onto the active site serine forms an acyl-
enzyme intermediate. Second, ACP binds to the surface of the
MCAT/malonyl complex, which triggers structural movement,
allowing transacylation to the PPant sulfhydryl to form malon-
yl-ACP. This transformation has been computationally modeled,
based upon the EcMCAT/malonyl/CoA complex structure (PDB
ID: 2G2Z).[78]
P. falciparum has an apicoplast-localized FAS, and sequence
alignment of the P. falciparum MCAT (PfMCAT) with template
sequences, EcMCAT and HsMCAT, identified conserved motifs
and residues. These important motifs include pentapeptides
GQGXG and GXSXG, along with four key invariant residues:
Gln109, Ser193, Arg218, and His305 in PfMCAT.[79] Glutamine is
responsible for stabilization of the oxyanion hole, and the
nearby arginine is responsible for the recognition and position-
ing of the substrate’s free carboxyl group through electrostatic
interactions.[80] Serine provides the catalytic active site nucleo-
phile, which is stabilized and activated by a histidine.[79] In the
S. coelicolor MCAT (ScMCAT) crystal structure, an acetate mole-
cule is bound to Gln9 and Arg122, mimicking the carboxy end
of the malonyl group corroborating this mechanistic propos-
al.[77]
Although ScMCAT, EcMCAT, and PfMCAT have similar catalytic
mechanisms, the crystal structure of the Mycobacterium tuber-
culosis MCAT (MtMCAT, PDB ID: 2QC3) reveals that the C O
bond of the catalytic Ser91 turns upward and results in a differ-
ent orientation and shape of the active site pocket and, sub-
sequently, a new catalytic diad.[81]

5.2. Inhibition
MCAT has been considered as an antibacterial target.[82] The
Figure 3. A) Schematic of MCAT chemistry, with several known inhibitors.
B) Diagram of MCAT “ping-pong” mechanism. C) Examination of MCAT’s binding of ACP to MCAT was found to be competitively inhibit-
active site. MCAT has been malonylated (cyan), and CoA (green) is present ed by high CoA concentrations.[72] Some MCAT antimicrobial
(PDB ID: 2G2Z). Active-site residues are marked in red. Top left: global topol- drugs have been identified, including trifluoroperazine (Fig-
ogy of MCAT with critical residues marked in red. Top right: surface diagram
ure 3 A).[83]
of the active site, demonstrating depth and accessibility. Bottom left: malo-
nylation of the active-site serine; positioning of CoA. Bottom right: position- Screening against the MCAT from the apicomplexan parasite
ing of catalytic triad. Eimeria tenella (EtMCAT) identified an alkaloid natural product,

ChemBioChem 0000, 00, 0 – 0 www.chembiochem.org 7  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim &

These are not the final page numbers! ÞÞ


Reviews
corytuberine, with moderate inhibitory activity against and ScMCAT NMR titration experiments showed small NMR
EtMCAT.[84] Virtual screening for inhibitors of PfMCAT was per- shifts for negatively charged residues in helix II, suggesting this
formed, and EDTA, p-toluene sulfonyl fluoride, and a thiolac- region interacts with ScMCAT, as shown by other ACP–partner
tone derivative, “TLM-1” (Figure 3 A), were found to position in domain interactions studied.[88] However, all NMR shifts were
the binding pocket and form stable complexes with strong in- inconclusively small, thus suggesting that ACP is in very fast
teractions.[79] Solved crystal structures of small molecules (glyc- exchange between its free and bound forms.[87]
erol and malonate) bound to the active site of EcMCAT re- The protein–protein interactions of MCAT from Xanthomonas
vealed that the length and amphipathic character of the sub- sp. (XoMCAT) and PKS ScACP (PDB ID: 2AF8) were studied by
strate define the specificity of the enzyme, but true targeting using computational docking, revealing that ACP binds to the
of MCAT requires further information about which of these cleft between the two XoMCAT subdomains, as seen in the
structural elements are important to exploit.[80] HpMCAT-ACP model. A positively charged ACP binding site
was observed adjacent to the GQGXQ loop. Gly14 and Ser15
from the GQGXQ loop of XoMCAT hydrogen bond with ACP
5.3. ACP interaction
residues. These residues are near the oxyanion hole, and it is
Four different MCAT-ACP interactions have been studied by hypothesized that this hydrogen bonding could activate the
modeling, revealing two mechanisms of interaction, one large- oxyanion hole to initiate malonyl transfer. Further clarification
ly electrostatic and one largely hydrophobic. of this interaction will require a co-crystal structure.[78]
The interaction between the Helicobacter pylori MCAT
(HpMCAT; PDB ID: 2H1Y) and ACP was studied by using com-
putational docking, glutathione S-transferase (GST) fusion pro-
6. Ketosynthase (KS)
tein pull-down, and surface plasmon resonance (SPR). The
HpMCAT active site is characterized by Ser92 and His198, which Ketosynthase catalyzes carbon–carbon bond formation
are in a deep gorge between two subdomains (Figure 3 C) and through decarboxylative Claisen condensation (Figure 4 A and
part of the GXSXG conserved motif in a/b-hydrolases. Helix II B). Type I FASs contain only one KS, whereas type II FASs have
of HpACP recognizes a conserved hydrophobic pocket be- two or three KSs. In E. coli, KSI, II, and III are encoded by fabB,
tween the two subdomains near the active site entrance of fabF, and fabH, respectively. The product of these reactions is
HpMCAT.[76] Additionally, by computation, the ACP binding site 3-ketoacyl-ACP by condensation of acyl-ACP with malonyl-ACP,
appears to be adjacent to the GQGXQ motif.[76] Similar to other with the exception of FabH (KS III), which condenses malonyl-
FAS proteins, the ACP binding site was found to be a positively ACP and acetyl-CoA (Figure 4 B). FabH is involved in FAS initia-
charged region forming complementary electrostatic interac- tion and FabB and FabF are involved in chain extension. The
tions.[85, 86] acyl group is loaded onto an active site cysteine and malonyl-
Information on the structural basis of the ping-pong mecha- ACP is extended by two carbon units after decarboxylative
nism was obtained computationally by modeling the ScMCAT nucleophilic addition.
interaction with the S. coelicolor actinorhodin polyketide syn- E. coli FabH uses acetyl-CoA as its main substrate, with bu-
thase (act PKS) apo-ACP.[77] The structure of ScMCAT with active tyryl-CoA demonstrating much lower activity and hexanoyl-
site bound acetate, which mimics the carboxyl of the malonyl CoA demonstrating no activity.[89] FabH has been shown to be
(PDB ID: 1NM2), suggests a ping-pong mechanism in which essential to FAB in vivo.[90] Overexpression of fabH leads to a de-
the oxyanion hole is formed upon substrate binding, thereby crease in cis-vaccenate and an increase in myristic acid (C14).[91]
moving the active site from a “closed” to an “open” conforma- FabH is regulated by feedback inhibition by high concentra-
tion. The second half of the mechanism, which involves ACP, tions of long chain acyl-ACPs.[89]
shows a largely hydrophobic interaction between the DSL FabB (KS1) shows activity with C6 to C14 saturated fatty acyl-
motif at the prosthetic group attachment site serine of the PKS ACPs, although it is only weakly active with C14-ACP.[92] FabB is
apo-ScACP and the hydrophobic pocket formed by the helical unsaturation tolerant and catalyzes the condensation of ma-
flap adjacent to the GQGXQ turn of ScMCAT.[77] However, this lonyl-ACP with each of cis-3-decenoyl-ACP, cis-5-dodecenoyl-
study did not use the cognate FAS ACP, and as PKS and FAS ACP, and 7-tetradecenoyl-ACP.[93] fabB deletion leads to auxo-
ACPs are not easily interchangeable, these docking results trophy for unsaturated fatty acids, and overexpression increas-
await experimental validation.[87] es unsaturated fatty acids if fabA is also overexpressed. FabF
NMR CSP studies and further computational work on the in- (KSII) shows activity with C6 to C14 fatty acyl-ACP esters; how-
teraction between the ScMCAT and the FAS ScACP shed some ever, C14 is a weak substrate.[92] FabF also shows low activity for
light on the situation.[87] In silico macromolecular docking of C16 and carries out the last step in the unsaturated pathway by
these proteins returned three different models: the first mimics elongating cis-9-hexadecenoyl-ACP.[92] Overexpression of fabF is
a model from Keatinge-Clay[77] in which the loop region after lethal,[94] although interestingly, expression of increased
helix I of FAS ScACP binds the residues near the ScMCAT active amounts in E. coli did not lead to cell death.[44]
site, the second mimics the binding of HpMCAT and HpACP Reconstitution of the cyanobacterium Synechococcus sp.
with helix II of FAS ScACP packing against ScMCAT, and the PCC 7002 FAS revealed that FabH was the sole rate-limiting
third binds the loop region after helix I of FAS ScACP near the step, in contrast to the E. coli FAS, with rate-limiting FabI (ER)
MCAT active site but in a rotated complex.[87] Holo-ScACP (FAS) and FabZ (DH) steps.[95]

& ChemBioChem 0000, 00, 0 – 0 www.chembiochem.org 8  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ÝÝ These are not the final page numbers!


Reviews
6.1. Mechanism acyl chain. Each has a catalytic triad that lies at the bottom of
the pocket in the structures (Figure 4 D), and the acyl chain
KSs adopt a ping-pong mechanism like MCAT. All three E. coli folds into an extended U-shaped conformation.[96]
KSs have been crystallized and show invariant hydrophobic With all three KS enzymes, the initial transthioesterification
residues that line their acyl pocket to facilitate binding of the occurs by cysteine attack to form a tetrahedral intermediate,
with the oxo group of the bound fatty acid thioester situated
in an oxyanion binding site.[96] FabB and FabF both have a Cys-
His-His active site triad. The entrance of the FabF active site is
blocked by an aromatic “gatekeeping” residue, Phe400, and it
has been proposed that it must shift in order for the acyl sub-
strate to gain access to the active site and the condensation
reaction to occur.[97] A Cys163Gln mutant in EcFabF mimics
acyl-KS. In this structure, the phenylalanine residue blocking
the active site is moved to an open conformation, allowing the
second half of the “ping-pong” mechanism.[98] Of the two
nearby histidine side chains, one is proposed to abstract a
proton from the carboxylic acid leaving group of the malonyl
residue, whereas the other histidine accepts a hydrogen bond
and then makes contact with the thioester oxo group of the
malonyl, inducing a partial negative charge. These two pro-
cesses prompt the decarboxylation reaction, which is followed
by condensation.[96, 97]
FabH, on the other hand, has a different proposed decarbox-
ylating Claisen mechanism. The FabH catalytic triad consists of
Cys-His-Asn. The histidine and asparagine make an oxyanion
binding site, which binds the malonyl thioester oxo group and
promotes decarboxylation.[99]

6.2. Inhibition
The fungal polyketide cerulenin targets KSs of FAS and PKS
pathways (Figure 4 A).[100] Cerulenin is a covalent inhibitor that
selectively targets FabB and FabF through modification of the
active-site cysteine sulfhydryl (Figure 4 C). A crystal structure
with FabF reveals that cerulenin binds in a hydrophobic
pocket at the dimer interface and covalently attaches to the
active site cysteine, mimicking the condensation transition
state. Phe400 and Ile108 completely rotate for this to occur, as
seen with the Cys163Gln EcFabF acyl-KS mimic, providing
access to the active site cysteine and opening a hydrophobic
pocket for the tail of the inhibitor.[101]

Figure 4. A) Schematic of FabB and FabF chemistry, with known inhibitors. R


is an acyl chain, length C1 to C13 (odd values). The two ACP proteins are de-
noted as ACP1 and ACP2 to clarify the transformation. B) Demonstration of
FAS initiation by FabH activity on malonyl-ACP with acetyl-CoA, and a
known inhibitor. C) Inhibition by cerulenin, and mechanism-based chloroa-
crylamide crosslinker based on cerulenin. D) Comparison of FabB, FabF, and
FabH, demonstrating depth of active site (both FabB and FabF exhibit shal-
low active sites, whereas FabH exhibits a deep pocket) and active-site geom-
etry (PDBs: FabB = 2VB8; FabF = 2GFY; FabH = 1HNJ). FabB shows thiolacto-
mycin (green) bound and the catalytic triad (Cys163, His298, His333; red);
FabF shows covalently attached C12 fatty acid (green) and the catalytic triad
(Cys163, His303, His340; red); FabH has malonyl-CoA (green) occupying the
tunnel and the catalytic triad (Cys112, His244, Asn274; red). The active site
geometries of FabB and FabF portray one histidine in position to abstract
a proton from the leaving group and the other in position to provide hydro-
gen-bond stabilization.

ChemBioChem 0000, 00, 0 – 0 www.chembiochem.org 9  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim &

These are not the final page numbers! ÞÞ


Reviews
Thiolactomycin (TLM) is a natural product that inhibits all and ACPs. The warhead portion of the probe is a Michael ac-
FAS condensing enzymes, including FabH.[102] A FabB-TLM crys- ceptor that, when attacked by the KS active site cysteine resi-
tal structure shows that TLM mimics malonyl-ACP in the active due, undergoes tandem 1,4-conjugate addition and b-chloro-
site, and both active site histidines aid the protein–antibiotic elimination to yield a stable covalent linkage between the
interaction. This explains the greater inhibition of FabB than probe and active site (Figure 4 C). The chloroacrylamide com-
FabH with TLM, as FabH harbors different catalytic machi- pound was used to crosslink EcACP with EcFabB and EcFabF.
nery.[103] Additionally, TLM inhibits two ketosynthases from My- As EcFabH did not readily crosslink but does bind ACP subse-
cobacterium tuberculosis, KasA and KasB, which are involved in quent to priming by acetyl-CoA, it is proposed that EcFabH
chain extension like FabB and FabF.[104] Neither cerulenin nor goes through a conformational change upon substrate loading
thiolactomycin are used in the clinic. Cerulenin also inhibits before ACP can bind.[44]
the KS function of the mammalian FAS, and the total synthesis Further studies utilizing the chloroacrylamide probe revealed
of thiolactomycin is problematic, due to its inherent instabili- that select ACPs from polyketide synthases, the frenolicin syn-
ty.[105, 106] thase carrier protein (FrenACP) from Streptomyces roseofulvus,
Platensimycin, a natural product isolated from Streptomyces and the oxytetracycline synthase carrier protein (OtcACP) from
platensis, is a selective FabB/F inhibitor with potent antibiotic Streptomyces rimosus, were able to crosslink with EcFabB.
properties. The formation of the acyl–enzyme complex is es- EcFabB was unable to crosslink with nonribosomal peptide
sential for platensimycin binding. This compound was found synthase carrier proteins.[51] Additionally, Chlamydomonas rein-
to be 200 times more potent than cerulenin and 50 times hardtii chloroplastic cACP crosslinked with EcFabF.[55] Crosslink-
more efficient than thiolactomycin at inhibiting FabF.[98] ing was used as a measure of interaction strength between
ACP and KS, as it is dependent on the ability of KS to recog-
nize the carrier protein through protein–protein interactions
6.3. Interactions with ACP
and react with the substrate analogue. Crosslinking showed
All three KSs from E. coli have been crystallized, but only the that EcFabF-ACP has tighter binding than EcFabB-ACP, as seen
interactions of FabH and FabF with ACP have been shown by by crosslinking efficiency. EcFabF also acts more efficiently
modeling studies.[55, 85] No co-crystal structure of an ACP bound than EcFabB on longer substrate mimics with a fourfold larger
to a type II FAS KS has been published.[13] preference, in accordance with its native substrate specificity.
ACP–protein interactions were investigated by using compu- The two enzymes have equal preference for shorter substrate
tational analysis to dock the NMR structure of EcACP with the mimics.[107]
crystal structure of EcFabH, and the generated model was
experimentally validated by in vitro characterization of FabH
mutants.[85] The EcFabH catalytic triad sits at the bottom of 7. Ketoreductase (KR)
a tunnel, and the entrance to the active site is flanked by four
3-Ketoacyl-ACP is reduced to (R)-3-hydroxyacyl-ACP by an
conserved basic residues (Arg36, Arg151, Lys214, and Arg249)
NAD(P)H-dependent KR, a member of the short-chain dehydro-
and two conserved aromatic residues (Trp32 and Phe213),
genase/reductases (SDR) family (Figure 5 A). FabG, the KR from
thereby giving it hydrophobic and electropositive character.[99]
E. coli, was first purified and characterized in 1966[108] and
EcACP docked to the area adjacent to the active site of
found to be active over a wide range of different 3-ketoacyl-
EcFabH, and a conserved arginine (Arg249) in this patch was
ACPs.[109] FabG can function on acyl-CoAs but with much lower
required for docking, as seen by an ACP-dependent assay and
efficiency.[108]
direct binding studies monitored by SPR. This patch also in-
FabG is ubiquitously expressed in all bacteria, conserved
cluded alanine residues, which were proposed to be important
across all species, and the only known enzyme to catalyze the
for the close approach of the two proteins. Zhang et al. ana-
ketoreduction step in fatty acid biosynthesis.[71] Experimental
lyzed the crystal structures of FabA, FabD, FabI, FabF, and LpxA
evidence for gene essentiality has been reported for E. coli, Sal-
and noted a conserved arginine/lysine residue in an electro-
monella enterica, M. tuberculosis, and Pseudomonas aerugino-
positive/hydrophobic patch adjacent to their active sites
sa.[110–113] Overexpression of fabG in E. coli increases the content
(within 10 to 14 ). ACPs from a variety of species have
of C16 acid twofold and C18 acid threefold.[114]
a highly conserved region from residues 32 to 50, which en-
compasses helix II. Glu41, predicted to interact with Arg249 of
FabH, is conserved in all 49 ACP sequences known at this time.
7.1. Mechanism
Ser36 of ACP, the point of PPant attachment, is oriented cor-
rectly in the docked model, but a 10  movement would be During the catalytic reaction, the b-keto group is reduced to
required for the chain to reach the FabH active site. The a b-hydroxy group by the pro-4S hydride ion from the nicoti-
“chain-flipping” mechanism is proposed, in which dissociation namide cofactor and a proton donated by the hydroxy of an
of the acyl group from ACP, upon binding with FabH, pro- adjacent tyrosine (Figure 5 B).[115–117] FabG is a tetramer and ex-
motes this proposed conformational change in ACP and a sub- hibits negative cooperativity, as seen by the binding of NADPH
sequent 10  movement.[85] to one active site, increasing the affinity at that site for ACP
Wothington et al. designed an E-chloroacrylamide probe as and decreasing the affinity for the cofactor elsewhere.[118] KRs
an active site crosslinking reagent to covalently crosslink KSs exhibit an ordered kinetic mechanism, with the NADPH cofac-

& ChemBioChem 0000, 00, 0 – 0 www.chembiochem.org 10  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ÝÝ These are not the final page numbers!


Reviews
tor binding first and the NADP + being the last product to result from cofactor binding (Figure 5 C). The cofactor binds in
leave.[116] a two-stage mechanism and when bound, the cofactor is in
The structure of E. coli FabG in binary complex with NADPH the syn conformation, and the nicotinamide ribose is stabilized
reveals mechanistically important conformational changes that in the active site by hydrogen bonds from Tyr151 and Lys155.
In the structure of the Y151F EcFabG mutant, the nicotinamide
ribose is disordered and not bound to the active site. Once
NADPH is bound to wild-type FabG, there is a rearrangement
of active site triad residues Ser138, Tyr151, and Lys155 and the
associated b5-a5 loop, as well as a shift of the b4-a4 loop and
dimer interface. These conformational changes are functionally
important because they prime the active site for catalysis and
allow the substrate access to the active site. These changes
also create a hydrogen-bonded network of ribose hydroxyls,
the catalytic triad, and four water molecules to relay a proton
to tyrosine after it donates a proton to the substrate during
catalysis.[116] Finally, the Brassica napus FabG and NADP + cofac-
tor binary complex suggests that the a6/a7 subdomain rotates
to enclose the active site and create a well-defined tunnel as
the final step in ternary complex formation.[119]

7.2. Inhibition
Studies suggest that FabG could be susceptible to inhibition at
its cofactor binding site.[120] Inhibition studies for FabG com-
monly follow the consumption of NADPH by absorbance as
FabG acts on a substrate, acetoacetyl-CoA, for example.[111]
Epigallocatechin gallate, the major component of green tea
extracts, and other related plant polyphenols were found to
inhibit FAS II by targeting FabG and FabI. The galloyl moiety of
the catechins is essential for inhibition, disrupting cofactor
binding by associating with the cofactor free FabG.[121] Isoflavo-
noids and flavonoid derivatives inhibit FabG from H. pylori
(HpFabG) in a noncompetitive manner with respect to aceto-
acetyl-CoA and NADPH.[122] The first inhibitor of microbial
origin, macrolactin S from Bacillus sp. AT28, showed dose-de-
pendent selective inhibition of S. aureus FabG (SaFabG) but did
not inhibit SaFabI.[123]
The crystal structure of FabG from P. falciparum (PfFabG) was
solved, and four structural analogues of triclosan were found
to have 75 % inhibition in vitro against PfFabG: hexachloro-
phene, bithionol, di-resorcinol sulfide, and bromochlorophen.
The inhibitors displayed nonlinear competitive inhibition with
NADPH, but further mechanistic studies performed by varying
substrate concentrations were inconclusive.[122–125] The above
potential inhibitors, however, are largely natural product ex-

Figure 5. A) Schematic of KR chemistry, with several known inhibitors. R is


an acyl chain, length C1 to C15 (odd values). B) Diagram of NAD(P)H-based
reduction mechanism. C) Demonstration of active site restructuring due to
NAD + binding. Top left: KR topology with NADP + highlighted in red;
middle left: surface-filled close-up of NADP + binding and remote Arg129,
Arg172 patch for ACP binding (PDB ID: 1Q7B). For comparison, top right: KR
topology without NADP + ; middle right: surface close-up demonstrating dif-
ferent topology, and highlighted arginine patch Arg129, Arg172 (PDB ID:
1I01). At bottom, demonstration of distance between cofactor binding site
and ACP binding site, requiring chain translocation for activity (PDB ID:
1Q7B).

ChemBioChem 0000, 00, 0 – 0 www.chembiochem.org 11  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim &

These are not the final page numbers! ÞÞ


Reviews
tracts facing drug development challenges and have not 8. Dehydratase (DH)
reached clinical usage.[111]
Recently, an allosteric inhibitor-binding site at the inter-sub- The dehydration of (R)-3-hydroxyacyl-ACP to the enoyl moiety
unit interface in FabG from Pseudomonas aeruginosa (PaFabG) is performed by b-hydroxy-ACP dehydratase through the elimi-
was identified. Two hydrophobic cavities at the dimer–dimer nation of water. The DH catalyzes syn elimination of the pro-2S
interface formed from residues in helices a4 and a5, not seen hydrogen and (3R)-hydroxy group, producing a trans prod-
in apo-PaFabG, were induced upon ligand binding. The size of uct.[127] In E. coli, there are two DHs encoded by fabA and fabZ.
the cavity was dependent on the inhibitor size, demonstrating Besides dehydration, FabA isomerizes trans-2-decenoyl-ACP
induced fit. These conformational changes were shown to into cis-3-decenoyl-ACP, which is the first reaction towards the
propagate to the catalytic triad (Ser141, Tyr154, Lys158), result- synthesis of unsaturated fatty acids (Figure 6 A).[128] Bacteria
ing in an active site conformation that was incompatible with and anaerobes employ a multiple DH system to preserve unsa-
NADPH binding. All ligands tested adopted almost planar con- turation, as seen with FabA in E. coli, whereas other organisms
formations and contained key hydrogen bond donors and utilize desaturases to install unsaturation.[129] The DH domain is
acceptors. Absence of this cavity in the non-inhibited enzyme a homodimer comprised of monomers with double-hotdog
reveals a limitation of most present-day, docking-based virtual topology in which a helixes are wrapped by b sheets.
screening methods for drug discovery.[111] FabA and FabZ exhibit broad overlapping chain length spe-
cificities. The substrate preference for FabA is intermediate
chain length acyl-ACPs centered on ten carbons. FabZ demon-
strates preference for shorter chains, yet also has low activity
for 16-carbon substrates. FabA does not participate in elongat-
ing unsaturated acyl-ACPs, so only FabZ is part of the unsatu-
7.3. Interactions with ACP
rated fatty acid arm of the pathway after FabA performs the
KRs remain poorly investigated, but NMR and in silico studies initial isomerization.[130] Insight into why these two enzymes
of the ACP binding site have been performed. No co-crystal have different substrate specificities has yet to be elucidated.
structure of an ACP with a KR has been solved. E. coli overexpressing both fabA and fabB produces an en-
Two surface residues, Arg129 and Arg172, located in a posi- hanced amount of unsaturated fatty acids.[131] Deletion and
tive hydrophobic patch adjacent to the EcFabG active site (Fig- overproduction of fabA leads to a decrease in unsaturated
ure 5 C), were mutated and showed a decrease in ability to use fatty acid content of phospholipids and an increase in saturat-
ACP thioester substrates but were fully active on a non-ACP ed fatty acids.[130] Deletion of fabZ leads to reduced (3R)-hy-
substrate analogue, b-ketobutyryl-CoA.[86] These two residues droxymyristoyl-ACP activity, thus suggesting that an enhanced
are conserved in PfFabG as well.[124] NMR chemical shift pertur- amount of 3-hydroxymyristoyl-ACP could be found in vivo.[132]
bations (CSPs) revealed that EcFabG-ACP interactions occur Heterologous overexpression of fabZ leads to a twofold in-
along ACP helix II and extend into the adjacent loop 2.[86] The crease in palmitic acid and stearic acid.[114]
acylated prosthetic groups are seen to reside in the hydropho-
bic cavity between helices II and III;[126] therefore, the acyl
8.1. Mechanism
chain of the prosthetic group would interact with residues in
loop 2. NMR titration studies showing a shift in Ile54 from Crystal structures of E. coli FabA (EcFabA), P. aeruginosa FabZ
loop 2 of ACP suggest that FabG interactions alter the acyl (PaFabZ), and P. falciparum hexameric FabZ (PfFabZ) show
chain binding site in ACP to facilitate the release of the pros- active site tunnels formed between two monomers that ac-
thetic group for modification.[86] These studies from 2003 were commodate the growing acyl chain. A catalytic diad, a con-
validated by the X-ray structures of hexanoyl-, heptanoyl-, and served histidine from loop B, and conserved glutamate (FabZ)
decanoyl-ACP solved in 2007 that showed the tethered sub- or aspartate (FabA) from the N-terminal region of the central
strate interacting with loop 2.[19] helix, each from a different subunit, are responsible for dehy-
Based on computational docking of EcACP to EcFabG- drating the acyl substrate.[133–135]
NADPH, the FabG ACP binding site is found closest to the 3-Hydroxydecanoyl-N-acetylcysteamine (3-hydroxydecanoyl-
active site of an adjacent monomer in the tetramer (Figure 5 C). NAC) bound in the pantetheine tunnel in the PaFabA crystal
Substrate delivery via the PPant arm takes place across the structure yields insight into the catalytic mechanism.[136] The
dimer interface. Arg172 of the ACP binding site moves 1.3  NAC molecule hydrogen bonds with loop C, as well as with
towards Ile136, which is at the base of the active site pocket, a water molecule bridging the other subunit. The NAC moiety
when the cofactor binds, showing the conformational change binds closely to catalytic residues Asp84 and His70 (the same
preparing for ACP interaction. The hydrogen bonding of residues found in EcFabA). His70Asn and Asp84Asn mutants
Glu168, located between the two sites, also changes upon co- show a complete loss in PaFabA activity. His70 abstracts the
factor binding and is hypothesized to facilitate ACP interaction proton from C2, whereas Asp84 hydrogen bonds the C3 hy-
and substrate delivery. Additionally in this region of EcFabG is droxy along with a water molecule, and only the R configura-
the b5-a5 loop, which becomes ordered upon cofactor bind- tion puts the hydroxy group in proximity of Asp84. The mecha-
ing and contains the conserved Asn145 residue; this suggests nism for the protonation of the hydroxy has yet to be elucidat-
that it might bind the pantetheine moiety.[116, 119] ed.[136]

& ChemBioChem 0000, 00, 0 – 0 www.chembiochem.org 12  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ÝÝ These are not the final page numbers!


Reviews
FabAs have a conserved a1-b1 loop that is larger than in 8.2. Inhibition
FabZ sequences. PaFabA and EcFabA structures show that this
loop contributes extensively to the alkyl-binding tunnel, Studies of FabZ as a drug target exploit crotonyl-CoA as a sub-
whereas the a3-b3 loop is the major contributor to the alkyl- strate, monitoring the reverse reaction by a decrease in ab-
binding tunnel in FabZ.[136] sorbance at 260 nm. Crotonyl-CoA studies give a lower KM
value than b-hydroxybutyryl-CoA.[137]
H. pylori FabZ (HpFabZ) exists as a hexamer in its native
state, with its active site located in a dimer interface. The sub-
strate binding tunnel is long, narrow, and hydrophobic, with
the catalytic histidines and glutamate residues halfway down
the tunnel.[138] Isoflavonoids, flavonoid derivatives, and synthet-
ic inhibitors have been reported to inhibit HpFabZ as competi-
tive inhibitors of the substrate crotonyl-CoA, blocking the
active site and extending to the groove around the pante-
theine-binding tunnel.[122, 125, 139] The natural product juglone
(Figure 6 A) was found to potently inhibit HpFabZ in a competi-
tive fashion (it also targets HpFabD as an uncompetitive inhibi-
tor). The HpFabZ–juglone crystal structure reveals that juglone
experiences p–p interactions with tyrosine and proline resi-
dues, is stabilized by hydrogen bonds with water, and is insert-
ed deeply into the active-site pocket.[140] The discovery of bind-
ing fragments that occupy various positions in the PaFabA
active site might help future inhibitor design. Two fragments
were found to block the pantetheine-binding tunnel and
partly mimic the flavonoids and synthetic inhibitors reported
for HpFabZ.[136]
The catalytic site of PfFabZ is also located in a deep, narrow
pocket formed at the dimer interface. PfFabZ experiences mul-
timeric state regulation by pH control. An inactive form of
PfFabZ was discovered existing as a dimer rather than as the
active hexamer. As the pH decreases, the dimeric form prevails,
and the two active-site histidines have electrostatic repulsion,
resulting in their expulsion from the active site. This was pro-
posed to help aid in drug design by designing inhibitors to
stabilize the inactive dimeric state.[141]
There are currently no published inhibitors for FabA, al-
though it is an interesting drug target due to its essential iso-
merase activity and absence from type I systems. Enterococcus
faecalis was found to have two FabZ homologue dehydratases,
one with isomerase activity (FabN) and one without (FabZ). Lu
et al. constructed EfFabZ/N chimeric proteins in an attempt to
transform FabZ into an isomerase. Single point mutations of
FabZ were not sufficient, but the addition of the b3–b4 region

Figure 6. A) Schematic of DH chemistry, with known inhibitors. R is an acyl


chain, length C1 to C15 (odd values). FabA isomerization is depicted but ob-
served only when R is C5--C9, with a preference for C7. The anti-periplanar
pro-2R hydrogen is highlighted for elimination. B) Diagram of mechanism-
based crosslinking activity described in text. C) Demonstration of EcFabA
topology and a comparison of conformational changes induced by sub-
strates. The catalytic His70 is shown in gray at the bottom of the active site.
Top left: overall topology of the DH, FabA; top right: wireframe topology of
empty active site (PDB ID: 1MKB). Middle left: wireframe topology of occu-
pied active site, covalently attached to 3-decynoyl-N-acetylcysteamine
(marked in red; (PDB ID: 1MKA). Middle right: wireframe topology of occu-
pied active site when crosslinked to ACP by using the mechanism-based
crosslinker described in (B). Bottom: cartoon of ACP crosslinked to FabA
with overlaid electrostatic surface, demonstrating unfolding of helix III by
interaction with arginine (gray sticks) from the FabA (PDB ID: 4KEH).

ChemBioChem 0000, 00, 0 – 0 www.chembiochem.org 13  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim &

These are not the final page numbers! ÞÞ


Reviews
of EfFabN to the EfFabZ showed a 38 % gain in isomerase conformations, helping to control substrate acceptance and
activity over the original EfFabN. Isomerase activity likely de- product exit.[146]
pends on the shape of the substrate binding tunnel, controlled
by the positioning of the b-strands surrounding the long cen-
9. Enoyl Reductase (ER)
tral helix, and not the catalytic machinery of the active site.
Disruptions to the active site shape eliminate isomerase activi- Enoyl-ACP reductase (ER) is a member of the short-chain dehy-
ty.[142] drogenase (SDR) family and performs the last step in each
fatty acid biosynthetic cycle by reducing 2-enoyl-ACP to fatty
acyl-ACP (Figure 7 A). In E. coli, there is a single enoyl-ACP re-
8.3. Interaction with ACP
ductase, FabI, which is essential.[147] As the KR and DH activities
A mechanism-based crosslinker was used to trap the transient that precede the ER are both reversible, FabI plays a determi-
interaction of AcpP and EcFabA to obtain a crystal structure. nant role in completing each elongation cycle and is thought
The crosslinked AcpP–EcFabA complex shows AcpP in two dif- to be the rate-limiting step.[148] Overexpression of fabI in E. coli
ferent conformations, one presumably a snapshot of the AcpP did not result in growth defects but also did not influence
in transition and one when docking was complete (Fig- fatty acid accumulation.[114] Three other ER isozymes have been
ure 6 C).[57] reported in bacteria: FabL and FabV in the SDR family, and
A dehydratase-specific mechanistic crosslinking probe was FabK, a TIM barrel flavoprotein unrelated to the SDR
designed and synthesized to covalently link FabA and ACP. The family.[149–151]
probe utilizes a 3-alkynyl sulfone warhead in which the sulfone
mimics the natural thioester, yet is safe from hydrolysis. The
9.1. Mechanism
warhead gets deprotonated and forms a reactive allene inter-
mediate, which is attacked by the FabA active site His70, form- In E. coli, NADPH reduction of the double bond proceeds by
ing a covalent linkage (Figure 6 B). Subsequently, when loaded conjugate syn-addition of the pro-4S hydrogen from NAD(P)H
onto EcACP, the 3-alkynyl sulfone probe was demonstrated to to C3 of the trans-2-acyl group, forming an enolate anion inter-
be fully sequestered by ACP, indicating that it mimics the hy- mediate which is stabilized by a lysine residue in the ER active
drophobicity of the natural DH substrate.[52, 53] site. Protonation of C2 to form the final product, acyl-ACP, is
The FabA–AcpP complex crosslinked crystal structure, along performed by an active site tyrosine residue (Figure 7 B).[152, 153]
with chemical shift perturbations from NMR titration experi- In InhA, the ER from M. tuberculosis, the stabilization of the
ments between either holo-ACP or octanoyl-ACP and FabA, enolate anion intermediate is presumed to be by a tyrosine
were used to elucidate a four-step mechanism by which FabA rather than a lysine, because in the X-ray structure of InhA in
prompts the release of the sequestered acyl chain from AcpP complex with NAD + and a C16 fatty acyl substrate (PDB ID:
for dehydration. First, a “positive patch” of EcFabA interacts 1BVR), Tyr158 is seen to make a direct hydrogen bond with
with the PPant attached to Ser36 of AcpP, which is consistent the fatty acyl substrate thioester carbonyl oxygen (Fig-
with previous studies[30, 143–145] that show these charge-comple- ure 7 C).[154]
mentary surfaces dictate protein–protein interactions. Second-
ly, Arg132 and Lys161 of EcFabA form salt bridges with Glu41
9.2. Inhibition
and Glu47 of AcpP to anchor the complex. Thirdly, Arg136 and
Arg137 of EcFabA interact with Ala59 and Glu60 of AcpP to FabI is inhibited by low concentrations of palmitoyl-CoA.[147]
pry away helix III. Finally, Leu138 and Val134 of EcFabA have Additionally, there are currently three well-known small mole-
hydrophobic interactions with Leu37 and Val40 from AcpP he- cule inhibitors of ER: triclosan, isoniazid, and diazoborines (Fig-
lix II to further stabilize the complex.[57] Interestingly, one FabA ure 7 A), which bind to the active site of ER and form a tight
pries apart the AcpP, facilitating acyl chain translocation into complex with the nicotinamide cofactor.[120] Isoniazid, the pri-
the cleft between the FabA monomers, but the other mono- mary anti-tuberculosis drug, is a prodrug that must first under-
mer presents the active-site histidine for crosslinking. go activation to a free radical by KatG prior to acylation of the
These four events together stabilize AcpP in its open confor- NAD + radical, forming a covalent complex with the cofactor
mation so that the acyl substrate can be released into the nicotinamide ring.[71] This species has been found to inhibit
binding pocket of EcFabA. As this occurs, five hydrophobic InhA from M. tuberculosis, as well as ER domains from other
pocket residues between helices II and III of AcpP move mycobacteria. Acyl-ACP substrates can prevent isoniazid inhibi-
inward and collapse the substrate pocket. This provides experi- tion of InhA, suggesting that the activated isoniazid interacts
mental evidence for the previously proposed chain-flipping (or with the substrate binding region.[154]
switchblade) mechanism.[30, 31] Triclosan was added to consumer products, such as antisep-
Additionally, the gatekeeping residues Phe165 and Phe171 tic soaps, toothpaste, and cosmetics, prior to understanding its
at the entrance to the active site were seen to be completely target. In 1999, the EcFabI–NAD + –triclosan complex was re-
rotated or displaced in the crosslinked structure as compared ported.[155] Triclosan was found to non-covalently interact with
to apo-EcFabA. This active site entrance control is also seen in the 2’-hydroxyl of ribose through hydrogen bonds while also
HpFabZ, where the top of the active site is controlled by an ar- hydrogen bonding to the catalytic tyrosine residue and experi-
omatic residue, Tyr100. It switches between open and closed encing p–p stacking interactions with the pyridine ring of the

& ChemBioChem 0000, 00, 0 – 0 www.chembiochem.org 14  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ÝÝ These are not the final page numbers!


Reviews
charged cofactor.[155, 156] In the same year, a conformational Waals interactions with the inhibitor and hydrogen bond the
change of the Ile192–Ser198 loop of EcFabI was shown to bound NAD + . This appears critical for the enhanced binding
close the active site and allow this region to form van der activity of triclosan.[157] In addition, triclosan has been found to
inhibit FabI from the malarial causative apicomplexan P. falcipa-
rum, but the validity of FAS II as a drug target against P. falcipa-
rum is currently under debate after the findings that the micro-
organism can scavenge fatty acids during the blood stage,
leaving no indispensable role for FAS II.[158]
Diazoborines are a class of heterocyclic boron-containing
compounds that inhibit FabI by covalent attachment of the
boron atom to the 2’-hydroxyl of the NAD + ribose.[156] Addi-
tionally, diazoborines p-stack with the nicotinamide ring, and
the boron and hydroxyl group occupy the space of the enolate
intermediate. Diazoborines are not clinically used, due to their
undesirable inhibition of RNA processing.[159] Isoniazid and tri-
closan are used clinically; however, a single-point mutation in
the M. smegmatis and M. bovis FabIs confers resistance to iso-
niazid.[154] As of 2009, 30 % of clinical isolates of M. tuberculosis
were isoniazid resistant.[150] Also, mutants of the EcFabI active
site, G93C for example, acquire resistance to triclosan by inter-
fering with the formation of a stable FabI–NAD + –triclosan ter-
nary complex.[155]
More recently, cyperin, a natural compound of fungal origin,
was found to inhibit ER from Arabidopsis thaliana, mimicking
the binding of triclosan.[160] Cephalochromin, a fungal secon-
dary metabolite, inhibits FabI but the mechanism has not been
elucidated.[161] Pyridomycin, a natural product with potent anti-
tuberculosis activity, inhibits InhA with a new mechanism not
previously seen. Pyridomycin is a competitive inhibitor of
NADH binding in InhA and blocks both the NADH cofactor and
fatty acid substrate binding pockets. This inhibitor’s unique
mechanism can be a starting point for future inhibitor design
(Figure 7 A).[162]

9.3. Interaction with ACP


EcFabI is a tetramer, and a combination of X-ray crystallogra-
phy and molecular dynamics (MD) simulations show that two
ACPs loaded with a 2-dodecenoyl acyl group react with one
EcFabI. This is unusual, as each EcFabI has four active sites. The
FabI–acyl ACP complex of E. coli is stabilized by largely electro-
static interactions between the acidic residues in ACP helix II
(Asp35, Asp38, Glu41, and Glu48) and a patch of basic residues
(Lys201, Arg204, and Lys205) adjacent to the EcFabI substrate-
binding loop, consistent with the structural and modeling
studies of ACP with AcpS, FabA, FabH, and FabG. The ACP

Figure 7. A) Schematic of ER chemistry and several known inhibitors. B) Dia-


gram of ER mechanism with enolate intermediate. C) Comparison of binding
of substrates and inhibitors for InhA, the MtFabI. Top left: overall topology
with NAD + (green) bound, with transparent overlay of electrostatic surface.
Top right: active site occupation by NAD + (green) and the fatty acid trans-2-
hexadecenoyl-(N-acetyl-cysteamine)-thioester (cyan) with the hydrogen
bond positioning tyrosine (Tyr158, red; (PDB ID: 1BVR). Middle left: binding
of triclosan (cyan) and NAD + (green), with positioning tyrosine (Tyr158, red),
forming a tight-binding but noncovalent complex (PDB ID: 2B35). Middle
right: binding of pyridomycin (cyan), with Tyr158 (red) in the absence of
NAD + (PDB ID: 4BII). Transparent surface over cartoon of the EcACP interface
with EcFabI, with residues not resolved (PDB ID: 2FHS).

ChemBioChem 0000, 00, 0 – 0 www.chembiochem.org 15  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim &

These are not the final page numbers! ÞÞ


Reviews
PPant is proposed to deliver the enoyl substrate to the minor
portal between EcFabI helix a8 and the mobile loop of resi-
dues 152–156. However, the absence of electron density for
the ACP pantetheine hindered the ability to predict how the
ACP delivers the substrate into the EcFabI active site (Fig-
ure 7 C).[163]
When the substrate binds, EcFabI does not experience an
overall conformational change, but rather the substrate-bind-
ing loop (residues 191–200) goes through a major conforma-
tional change upon complexation with ACP. In the triclosan–
EcFabI crystal structure, triclosan was shielded from the solvent
by the substrate-binding loop, but in the EcFabI-acyl ACP com-
plex crystal structure, the loop adopts an open lid conforma-
tion, presumably because of the interaction with ACP in this
area. Details of this interaction could not be elucidated from
the crystal structure due to missing side chain density. In
regards to the catalytic triad (Tyr146, Tyr156, and Lys163 in
EcFabI, and Phe149, Tyr158 and Lys165 in M. tuberculosis InhA),
Tyr156 does not play a significant role in the reduction, and
the substrate carbonyl group is pointing away from it in the
crystal structure. Tyr146 is directly involved in catalysis and hy-
drogen bonds with the substrate carbonyl. This differs from
the corresponding Phe149 in InhA, which is proposed to posi-
tion the cofactor for hydride transfer. Finally, the function of
Lys165 is not yet established.[163]
InhA, the enoyl reductase from M. tuberculosis, prefers long
Figure 8. A) Schematic of AT and TE chemistry, with inhibitors of TE. R’ = sub-
chain fatty acyl substrates, C16 and longer. The InhA crystal strate of AT activity, including sugars, proteins, and small molecules. B) Dia-
structure with NAD + and a C16 acyl substrate, trans-2-hexa- gram of TE crosslinker mechanism. C) Demonstration of catalytic triad of a
decenoyl-(N-acetylcysteamine)-thioester, revealed that the sub- Bacteroides thetaiotaomicron TE. Left: overall topology; right: close-up of
triad orientation and nearness to protein surface (PDB ID: 2ESS). The catalyt-
strate binds in a “U-shaped” conformation, with the trans
ic triad (Asn172, His174, Ser209) is denoted in red.
double bond positioned adjacent to the nicotinamide ring of
NAD + . Hydrophobic residues from the substrate binding loop
(residues 196–219) engulf the fatty acyl chain portion of the la] mutant also lacked activity and did not assemble into the
substrate. The substrate binding loop is longer than that of membrane, suggesting that proper folding and membrane in-
EcFabI, creating a deeper binding pocket to accommodate sertion is also dictated by the active site aspartic acid.[166] Addi-
longer chain acyl substrates.[154] The interaction of InhA with tionally, a conserved glycine is involved in catalysis by playing
ACP has not yet been elucidated. a steric role in the active site, and an invariant arginine is criti-
cal for glycerol 3-phosphate binding, as seen by a 13-fold
higher KM value for EcPlsB mutant Arg354Cys.[167] This acyl
10. Thioesterase (TE)/Acyl Transferase (AT)
transferase is one of the proteins that controls chain length in
After the iterative chain elongation cycle, FAB is terminated E. coli.[164] However, in spite of this mechanistic insight, very
either by offloading the fatty acid from the ACP by an acyl- little is known about PlsB protein–protein interactions, and
ACP thioesterase, releasing a free fatty acid or by transesterifi- little has been done with this enzyme in terms of metabolic
cation onto a lipid by an acyltransferase (Figure 8 A). Most engineering. TEs from plants, algae, and bacteria have been
often, prokaryotes utilize ATs, and eukaryotes utilize TEs. used extensively in metabolic engineering of FAB and will be
In E. coli, the acyl group is directly transferred from ACP to the focus of the remainder of this section.
glycerol-3-phosphate by a glycerol-3-phosphate AT, PlsB (a As of 2010, there were 23 families of TEs grouped by their
membrane protein). Free fatty acids are not found as inter- substrate and tertiary structure, and these were almost com-
mediates in bacterial lipid biosynthesis.[164] A plsB mutant in pletely unrelated by primary structure. Of these 23 families,
E. coli leads to the cessation of phospholipid biosynthesis, an only six break bonds between acyl groups and ACP (the other
accumulation of long chain acyl-ACP, and the inhibition of de TEs act on CoA, glutathione, and other proteins). Two of the
novo fatty acid biosynthesis.[165] A conserved domain contain- groups (TE14 and TE15) contain TEs found in plants, algae, and
ing histidine and aspartic acid, separated by four residues bacteria and have a typical hotdog fold, whereas the remain-
(HX4D), was identified for ATs from bacterial, plant, and animal ing groups are found in fungi and mammals and are TEs with
kingdoms. The catalytic histidine is thought to function as an a/b-hydrolase fold (TE16–TE19). The most characterized TEs
a base to deprotonate the hydroxy moiety of the acyl acceptor. that have garnered interest for metabolic engineering fall into
The PlsB[His36Ala] mutant lacked AT activity. The PlsB[Asp311A- the TE14 family.[168] The N-terminal hotdog domain has been

& ChemBioChem 0000, 00, 0 – 0 www.chembiochem.org 16  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ÝÝ These are not the final page numbers!


Reviews
shown by chimeric enzyme construction to control chain 10.3. Interaction with ACP
length specificity.[169]
E. coli has seven hotdog TEs, but none are believed to be There has been considerable interest in the metabolic engi-
dedicated acyl-ACP TEs and are therefore not directly involved neering of plant TEs, but thus far it has been met with limited
in fatty acid or lipid biosynthesis.[170] TesA and TesB do work on success, due to a lack of structural information. In one recent
acyl-ACP and are more active on longer chain fatty acids. How- study, the Jatropha curcas FatA and FatB were docked with 4’-
ever, TesA and TesB activities for palmitoyl-ACP are 0.6 and PPant substrates 4’-PPT-16:0, 4’-PPT-18:0, and 4’-PPT-18:1, and
0.1 % the rate at which they work on palmitoyl-CoA, respec- the top models bound to the active catalytic triad. These
tively.[171] When tesA is overexpressed in E. coli, without target- models were then used for protein–protein docking between
ing the periplasmic sequence, the amount of free fatty acids is spinach ACP and JcFatA or JcFatB. For five JcFatB models and
increased, producing titers as high as 0.32 g L 1.[172, 173] nine JcFatA models, ACP docked near the catalytic triad, and
In plants, TEs can be the sole factor in determining fatty the serine of ACP was oriented towards the TE active site.[184]
acid chain length. There are two distinct TE gene classes in Unlike plants, green algae only have one TE, Fat1. Some
higher plants that are dedicated acyl-ACP TEs: fatA and fatB. plant TEs have been engineered into E. coli and other various
fatA encodes a TE with a preference for 18:1-ACP and a minor plants, and this led to changes in the fatty acid quantity and
preference for 18:0 and 16:0-ACP. fatB encodes a TE with a profile observed.[55] However, engineering plant TEs into algae
preference for short/medium chain saturated fatty acids rang- has been unpredictable, likely due to mismatched protein–pro-
ing from 8 to 16 carbons.[174, 175] When plant TEs are overex- tein interactions. CrcACP and CrTE from Chlamydomonas rein-
pressed in E. coli, the fatty acid profile is altered by the specific- hardtii were docked, identifying a protein–protein recognition
ity of the TE.[172] When the C12:0-ACP-selective TE from Umbel- surface. The fatty acyl thioester attached to Cr-cACP docked
lularia californica (UcFatB1) was overexpressed in E. coli, a 500- closely to the active site cysteine of CrTE.
fold increase in C12:0 fatty acid production was seen.[176, 177] Additionally, an a-bromopalmitic pantetheine activity-based
probe was utilized to investigate ACP-TE interactions. This
probe is very electrophilic at the a-carbon and allows for nu-
10.1. Mechanism
cleophilic attack by the active site TE nucleophile, leading to
The a/b-hydrolase TEs have a conserved catalytic triad: a nucle- covalent attachment of the probe (Figure 8 B). When the probe
ophile-histidine-acid triad, with the nucleophile commonly is chemoenzymatically loaded onto ACP and a partner TE is
being a serine residue. The acid stabilizes the basic histidine, added, a covalent crosslinked complex is achieved. CrTE, UcTE,
which accepts a proton from the nucleophile. The nucleophilic and ChTE oriented in a similar conformation to CrcACP in com-
serine forms a tetrahedral intermediate with the substrate putational docking studies. However, only Cr-cACP and CrTE
before it is attacked by water.[178] The active site residues in TEs successfully formed a crosslinked complex, showing that vas-
are on the surface of the protein, rather than buried in an acyl cular plant TEs do not interact with CrACP in vitro. Far less
binding pocket, as seen with other FAS domains (Figure 8 C). complex formation of CrcACP–CrTE was seen with a a-bromo-
The hotdog-fold TEs lack defined nonsolvated binding pockets hexanoic pantetheine probe than with the C16 analogue, show-
and conserved catalytic residues, and therefore, a variety of ing that protein–protein interactions, along with chain length,
catalytic residues and mechanisms exist.[170] Plant TEs likely govern hydrolysis by TE.[54]
have a papain-like catalytic triad, as they are inhibited by thiol
inhibitors and carry a conserved histidine required for cataly-
11. Conclusion and Outlook
sis.[179] A bioinformatics-guided site-directed mutagenesis study
proposed the catalytic triad for the A. thaliana TE (FatB) to be This review has established that the understanding of mecha-
Cys264, His229, and Asn227, but there is currently no structural nistic, structural, and inhibition characteristics of type II FASs is
verification.[180, 181] substantial. However, drug resistance, a lack of clinical viability,
and unachieved FAS engineering prospects highlight the need
for a new generation of studies focusing on protein–protein
10.2. Inhibition
interactions with ACP. Future investigations will delve deeper
Plant TEs have a cysteine catalytic residue and therefore are into the fundamentals of the FAS to guide future inhibitor
generally inhibited by compounds that are known to react design and metabolic engineering efforts.
with thiols. Animal and bacterial TEs have a serine catalytic res- It is evident that protein–protein interactions are necessary
idue and are inhibited by serine reactive reagents.[179] for modification of substrates attached to ACP in the FAS.[13]
When the Umbellularia californica plant TE (UcTE) was pre-in- Each interaction between ACP and its partners must be specif-
cubated with diethylpyrocarbonate, a 97 % loss in activity was ic enough to allow precise chain flipping and substrate presen-
seen (Figure 8 A).[182] Diethylpyrocarbonate acts by covalently tation, but weak enough to allow rapid turnover.[85] In 2000,
modifying histidine at the N-W-2 nitrogen of the imidazole the ACP–AcpS crystal structure allowed the first visualization of
ring, showing evidence for a catalytic histidine residue. UcTE domain–ACP interactions and served as a starting point for
was also completely inhibited by N-ethylmaleimide, suggesting future work. However, AcpS acts only on serine, not the PPant
an active cysteine mechanism, as it acts as a Michael acceptor arm of ACP, raising questions about the generality of the ob-
for soft nucleophiles.[183] served interactions. In 2001, Zhang et al. anticipated that ACP

ChemBioChem 0000, 00, 0 – 0 www.chembiochem.org 17  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim &

These are not the final page numbers! ÞÞ


Reviews
conformational changes facilitate the presentation of cargo [7] J. Piel, Nat. Prod. Rep. 2010, 27, 996 – 1047.
and the release post-modification, but could not demonstrate [8] S. Jenni, M. Leibundgut, D. Boehringer, C. Frick, B. Mikolsek, N. Ban,
Science 2007, 316, 254 – 261.
how.[85] It was not until 2006 that a crystal structure of a partner [9] T. Maier, M. Leibundgut, N. Ban, Science 2008, 321, 1315 – 1322.
domain (FabI) and ACP was solved, but the lack of resolution [10] C. O. Rock, J. Cronan, Methods Enzymol. 1981, 71, 341 – 351.
of the PPant and nearby side chain density precluded insight [11] P. Lu, C. Vogel, R. Wang, X. Yao, E. M. Marcotte, Nat. Biotechnol. 2007,
into substrate delivery (section 9.3).[163] It was not until 2014 25, 117 – 124.
[12] G. Butland, J. M. Peregrn-Alvarez, J. Li, W. Yang, X. Yang, V. Canadien,
that the crosslinked crystal structure of ACP and FabA was elu- A. Starostine, D. Richards, B. Beattie, N. Krogan, M. Davey, J. Parkinson,
cidated, allowing a snapshot of the AcpP in transition and J. Greenblatt, A. Emili, Nature 2005, 433, 531 – 537.
when fully interacting with a native partner (section 8.3). By [13] J. Crosby, M. P. Crump, Nat. Prod. Rep. 2012, 29, 1111 – 1137.
using new technologies, including a functional AcpH resin, [14] M. Falb, K. Mller, L. Kçnigsmaier, T. Oberwinkler, P. Horn, S. von Gro-
nau, O. Gonzalez, F. Pfeiffer, E. Bornberg-Bauer, D. Oesterhelt, Extremo-
a “one-pot” chemoenzymatic reaction for crypto-ACP prepara- philes 2008, 12, 177 – 196.
tion and mechanistic crosslinking probes (section 2), along [15] J. Lombard, P. Lpez-Garca, D. Moreira, Mol. Biol. Evol. 2012, 29, 3261 –
with protein crystallography and NMR experiments (section 3), 3265.
among others, we have been able to directly observe substrate [16] D. V. Dibrova, M. Y. Galperin, A. Y. Mulkidjanian, Environ. Microbiol.
2014, 16, 907 – 918.
delivery. [17] C. Anselmi, M. Grininger, P. Gipson, J. D. Faraldo-Gmez, J. Am. Chem.
Directions for future work include the design, synthesis, and Soc. 2010, 132, 12357 – 12364.
application of mechanistic crosslinking probes for cofactor de- [18] D. H. Keating, M. R. Carey, J. E. Cronan, J. Biol. Chem. 1995, 270,
pendent ER and KR domains, as these domains are essential, 22229 – 22235.
[19] A. Roujeinikova, W. J. Simon, J. Gilroy, D. W. Rice, J. B. Rafferty, A. R.
and their protein–protein interactions could prove to be ideal Slabas, J. Mol. Biol. 2007, 365, 135 – 145.
targets for inhibitor design. Additionally, cryo-electron micros- [20] S. E. Evans, C. Williams, C. J. Arthur, S. G. Burston, T. J. Simpson, J.
copy has been used recently to study large type I FAS com- Crosby, M. P. Crump, ChemBioChem 2008, 9, 2424 – 2432.
plexes in motion.[70, 185] The application of mechanistic crosslink- [21] G. A. Zornetzer, J. Tanem, B. G. Fox, J. L. Markley, Biochemistry 2010, 49,
470 – 477.
ing probes to stabilize type I FAS complexes for better resolu- [22] J. Beld, H. Cang, M. D. Burkart, Angew. Chem. Int. Ed. 2014, 53, 14456 –
tion can enhance future cryo-EM studies. To build upon current 14461, Angew. Chem. 2014, 126, 14684 – 14689.
in vitro studies, AasS can be utilized as a means to study the [23] M. N. R. Johnson, C. H. Londergan, L. K. Charkoudian, J. Am. Chem. Soc.
FAS and subsequent products in vivo by loading unique visual- 2014, 136, 11240 – 11243.
[24] E. Płoskoń, C. J. Arthur, S. E. Evans, C. Williams, J. Crosby, T. J. Simpson,
ization- and activity-based acids onto ACP in a fast and effi- M. P. Crump, J. Biol. Chem. 2008, 283, 518 – 528.
cient manner.[56] [25] C. O. Rock, J. E. Cronan, I. M. Armitage, J. Biol. Chem. 1981, 256, 2669 –
The outlook for FAS research is bright. New tools are avail- 2674.
able, allowing visualization of previously unobservable states. [26] D. Post-Beittenmiller, J. G. Jaworski, J. B. Ohlrogge, J. Biol. Chem. 1991,
266, 1858 – 1865.
Our understanding of the ACP interaction with partner pro- [27] P. J. Jones, T. A. Holak, J. H. Prestegard, Biochemistry 1987, 26, 3493 –
teins allows for future studies of processivity, sequestration, 3500.
and new cargo-dependent conformational changes. Specifical- [28] Y. Kim, E. L. Kovrigin, Z. Eletr, Biochem. Biophys. Res. Commun. 2006,
ly, studies on the ACP interaction with the TE will yield crucial 341, 776 – 783.
[29] E. Płoskoń, C. J. Arthur, A. L. P. Kanari, P. Wattana-amorn, C. Williams, J.
information for engineering attempts to modify fatty acid pro- Crosby, T. J. Simpson, C. L. Willis, M. P. Crump, Chem. Biol. 2010, 17,
files, in the same manner that unsaturation may be controlled 776 – 785.
by careful engineering of dehydratases. Finally, although the [30] M. Leibundgut, S. Jenni, C. Frick, N. Ban, Science 2007, 316, 288 – 290.
chemistry of FAB is conserved between type I and type II FAS [31] J. E. Cronan, Biochem. J. 2014, 460, 157 – 163.
[32] G. A. Zornetzer, B. G. Fox, J. L. Markley, Biochemistry 2006, 45, 5217 –
systems, sequestration and ACP–partner interactions are differ- 5227.
ent, providing an opportunity for directed development of [33] J. Beld, E. C. Sonnenschein, C. R. Vickery, J. P. Noel, M. D. Burkart, Nat.
antibiotics. Prod. Rep. 2014, 31, 61 – 108.
[34] A. M. Gehring, R. H. Lambalot, K. W. Vogel, D. G. Drueckhammer, C. T.
Walsh, Chem. Biol. 1997, 4, 17 – 24.
[35] S. C. Wenzel, R. Mller, Curr. Opin. Biotechnol. 2005, 16, 594 – 606.
Acknowledgements [36] K. J. Weissman, H. Hong, M. Oliynyk, A. P. Siskos, P. F. Leadlay, ChemBio-
Chem 2004, 5, 116 – 125.
We thank Joris Beld for critical advice. This work was supported [37] M. Fernndez-Valverde, A. Reglero, H. Martinez-Blanco, J. M. Luengo,
by NIH R01M094924 and R01M086225. Appl. Environ. Microbiol. 1993, 59, 1149 – 1154.
[38] M. K. Go, J. Y. Chow, V. W. N. Cheung, Y. P. Lim, W. S. Yew, Biochemistry
2012, 51, 4568 – 4579.
Keywords: acyl carrier protein · biosynthesis · crosslinkers · [39] G. L. Holliday, S. A. Rahman, N. Furnham, J. M. Thornton, J. Mol. Biol.
2014, 426, 2098 – 2111.
fatty acids · inhibitors
[40] K. M. Clarke, A. C. Mercer, J. J. La Clair, M. D. Burkart, J. Am. Chem. Soc.
2005, 127, 11234 – 11235.
[1] B. J. Dunn, C. Khosla, J. R. Soc. Interface 2013, 10, 20130297. [41] A. S. Worthington, M. D. Burkart, Org. Biomol. Chem. 2006, 4, 44 – 46.
[2] F. T. Wong, C. Khosla, Curr. Opin. Chem. Biol. 2012, 16, 117 – 123. [42] A. L. Mandel, J. J. La Clair, M. D. Burkart, Org. Lett. 2004, 6, 4801 – 4803.
[3] A. Das, C. Khosla, Acc. Chem. Res. 2009, 42, 631 – 639. [43] J. L. Meier, A. C. Mercer, H. Rivera, M. D. Burkart, J. Am. Chem. Soc.
[4] C. Khosla, J. Org. Chem. 2009, 74, 6416 – 6420. 2006, 128, 12174 – 12184.
[5] C. Khosla, S. Kapur, D. E. Cane, Curr. Opin. Chem. Biol. 2009, 13, 135 – [44] A. S. Worthington, H. Rivera, J. W. Torpey, M. D. Alexander, M. D. Bur-
143. kart, ACS Chem. Biol. 2006, 1, 687 – 691.
[6] A. T. Keatinge-Clay, Nat. Prod. Rep. 2012, 29, 1050 – 1073. [45] P. Vagelos, A. Larrabee, J. Biol. Chem. 1967, 242, 1776 – 1781.

& ChemBioChem 0000, 00, 0 – 0 www.chembiochem.org 18  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ÝÝ These are not the final page numbers!


Reviews
[46] J. Thomas, J. E. Cronan, J. Biol. Chem. 2005, 280, 34675 – 34683. [85] Y.-M. Zhang, M. S. Rao, R. J. Heath, A. C. Price, A. J. Olson, C. O. Rock,
[47] R. S. Flugel, Y. Hwangbo, R. H. Lambalot, J. E. Cronan, C. T. Walsh, J. S. W. White, J. Biol. Chem. 2001, 276, 8231 – 8238.
Biol. Chem. 2000, 275, 959 – 968. [86] Y.-M. Zhang, B. Wu, J. Zheng, C. O. Rock, J. Biol. Chem. 2003, 278,
[48] E. Murugan, R. Kong, H. Sun, F. Rao, Z.-X. Liang, Protein Expression 52935 – 52943.
Purif. 2010, 71, 132 – 138. [87] C. J. Arthur, C. Williams, K. Pottage, E. Płoskoń, S. C. Findlow, S. G. Bur-
[49] N. M. Kosa, R. W. Haushalter, A. R. Smith, M. D. Burkart, Nat. Methods ston, T. J. Simpson, M. P. Crump, J. Crosby, ACS Chem. Biol. 2009, 4,
2012, 9, 981 – 984. 625 – 636.
[50] M. Rothmann, N. M. Kosa, M. D. Burkart, RSC Adv. 2014, 4, 9092 – 9097. [88] K. D. Parris, L. Lin, A. Tam, R. Mathew, J. Hixon, M. Stahl, C. C. Fritz, J.
[51] A. S. Worthington, D. F. Porter, M. D. Burkart, Org. Biomol. Chem. 2010, Seehra, W. S. Somers, Structure 2000, 8, 883 – 895.
8, 1769 – 1772. [89] R. J. Heath, C. O. Rock, J. Biol. Chem. 1996, 271, 10996 – 11000.
[52] F. Ishikawa, R. W. Haushalter, M. D. Burkart, J. Am. Chem. Soc. 2012, [90] C.-Y. Lai, J. E. Cronan, J. Biol. Chem. 2003, 278, 51494 – 51503.
134, 769 – 772. [91] J. T. Tsay, W. Oh, T. J. Larson, S. Jackowski, C. O. Rock, J. Biol. Chem.
[53] F. Ishikawa, R. W. Haushalter, D. J. Lee, K. Finzel, M. D. Burkart, J. Am. 1992, 267, 6807 – 6814.
Chem. Soc. 2013, 135, 8846 – 8849. [92] P. Edwards, J. Sabo Nelsen, J. G. Metz, K. Dehesh, FEBS Lett. 1997, 402,
[54] J. L. Blatti, J. Beld, C. A. Behnke, M. Mendez, S. P. Mayfield, M. D. Bur- 62 – 66.
kart, PLoS One 2012, 7, e42949. [93] Y. Feng, J. E. Cronan, J. Biol. Chem. 2009, 284, 29526 – 29535.
[55] J. Beld, J. Blatti, C. Behnke, M. Mendez, M. Burkart, J. Appl. Phycol. [94] S. Subrahmanyam, J. E. Cronan, J. Bacteriol. 1998, 180, 4596 – 4602.
2014, 26, 1619 – 1629. [95] J. Kuo, C. Khosla, Metab. Eng. 2014, 22, 53 – 59.
[56] J. Beld, K. Finzel, M. D. Burkart, Chem. Biol. 2014, 21, 1293 – 1299. [96] J. G. Olsen, A. Kadziola, P. von Wettstein-Knowles, M. Siggaard-Ander-
[57] C. Nguyen, R. W. Haushalter, D. J. Lee, P. R. L. Markwick, J. Bruegger, G. sen, S. Larsen, Structure 2001, 9, 233 – 243.
Caldara-Festin, K. Finzel, D. R. Jackson, F. Ishikawa, B. O’Dowd, J. A. [97] W. Huang, J. Jia, P. Edwards, K. Dehesh, G. Schneider, Y. Lindqvist,
McCammon, S. J. Opella, S.-C. Tsai, M. D. Burkart, Nature 2014, 505, EMBO J. 1998, 17, 1183 – 1191.
427 – 431. [98] J. Wang, S. M. Soisson, K. Young, W. Shoop, S. Kodali, A. Galgoci, R.
[58] M. P. Williamson, Prog. Nucl. Magn. Reson. Spectrosc. 2013, 73, 1 – 16. Painter, G. Parthasarathy, Y. S. Tang, R. Cummings, S. Ha, K. Dorso, M.
[59] A. M. J. J. Bonvin, R. Boelens, R. Kaptein, Curr. Opin. Chem. Biol. 2005, 9, Motyl, H. Jayasuriya, J. Ondeyka, K. Herath, C. Zhang, L. Hernandez, J.
501 – 508. Allocco, . Basilio, J. R. Tormo, O. Genilloud, F. Vicente, F. Pelaez, L. Col-
[60] V. A. Jarymowycz, M. J. Stone, Chem. Rev. 2006, 106, 1624 – 1671. well, S. H. Lee, B. Michael, T. Felcetto, C. Gill, L. L. Silver, J. D. Hermes, K.
[61] G. Lipari, A. Szabo, J. Am. Chem. Soc. 1982, 104, 4559 – 4570. Bartizal, J. Barrett, D. Schmatz, J. W. Becker, D. Cully, S. B. Singh, Nature
[62] G. Lipari, A. Szabo, J. Am. Chem. Soc. 1982, 104, 4546 – 4559. 2006, 441, 358 – 361.
[63] D. I. Chan, T. Stockner, D. P. Tieleman, H. J. Vogel, J. Biol. Chem. 2008, [99] C. Davies, R. J. Heath, S. W. White, C. O. Rock, Structure 2000, 8, 185 –
283, 33620 – 33629. 195.
[64] J. R. Tolman, J. M. Flanagan, M. A. Kennedy, J. H. Prestegard, Proc. Natl. [100] J. P. Torella, T. J. Ford, S. N. Kim, A. M. Chen, J. C. Way, P. A. Silver, Proc.
Acad. Sci. USA 1995, 92, 9279 – 9283. Natl. Acad. Sci. USA 2013, 110, 11290 – 11295.
[65] P. M. Gasper, B. Fuglestad, E. A. Komives, P. R. L. Markwick, J. A. McCam- [101] M. Moche, G. Schneider, P. Edwards, K. Dehesh, Y. Lindqvist, J. Biol.
mon, Proc. Natl. Acad. Sci. USA 2012, 109, 21216 – 21222. Chem. 1999, 274, 6031 – 6034.
[66] P. R. L. Markwick, J. A. McCammon, Phys. Chem. Chem. Phys. 2011, 13, [102] S. Jackowski, C. M. Murphy, J. E. Cronan, C. O. Rock, J. Biol. Chem. 1989,
20053 – 20065. 264, 7624 – 7629.
[67] M. Blackledge, Prog. Nucl. Magn. Reson. Spectrosc. 2005, 46, 23 – 61. [103] A. C. Price, K.-H. Choi, R. J. Heath, Z. Li, S. W. White, C. O. Rock, J. Biol.
[68] F. J. Asturias, J. Z. Chadick, I. K. Cheung, H. Stark, A. Witkowski, A. K. Chem. 2001, 276, 6551 – 6559.
Joshi, S. Smith, Nat. Struct. Mol. Biol. 2005, 12, 225 – 232. [104] L. Kremer, J. D. Douglas, A. R. Baulard, C. Morehouse, M. R. Guy, D.
[69] P. Gipson, D. J. Mills, R. Wouts, M. Grininger, J. Vonck, W. Khlbrandt, Alland, L. G. Dover, J. H. Lakey, W. R. Jacobs, P. J. Brennan, D. E. Minni-
Proc. Natl. Acad. Sci. USA 2010, 107, 9164 – 9169. kin, G. S. Besra, J. Biol. Chem. 2000, 275, 16857 – 16864.
[70] D. Boehringer, N. Ban, M. Leibundgut, J. Mol. Biol. 2013, 425, 841 – 849. [105] K. Young, H. Jayasuriya, J. G. Ondeyka, K. Herath, C. Zhang, S. Kodali, A.
[71] R. J. Heath, S. W. White, C. O. Rock, Appl. Microbiol. Biotechnol. 2002, Galgoci, R. Painter, V. Brown-Driver, R. Yamamoto, L. L. Silver, Y. Zheng,
58, 695 – 703. J. I. Ventura, J. Sigmund, S. Ha, A. Basilio, F. Vicente, J. R. Tormo, F.
[72] V. C. Joshi, S. J. Wakil, Arch. Biochem. Biophys. 1971, 143, 493 – 505. Pelaez, P. Youngman, D. Cully, J. F. Barratt, D. Schmatz, S. B. Singh, J.
[73] I. I. G. S. Verwoert, E. F. Verhagen, K. H. van der Linden, E. C. Verbree, Wang, Antimicrob. Agents Chemother. 2006, 50, 519 – 526.
H. J. J. Nijkamp, A. R. Stuitje, FEBS Lett. 1994, 348, 311 – 316. [106] R. J. Heath, S. W. White, C. O. Rock, Prog. Lipid Res. 2001, 40, 467 – 497.
[74] J. L. Garwin, A. L. Klages, J. E. Cronan, J. Biol. Chem. 1980, 255, 3263 – [107] A. S. Worthington, G. H. Hur, J. L. Meier, Q. Cheng, B. S. Moore, M. D.
3265. Burkart, ChemBioChem 2008, 9, 2096 – 2103.
[75] E. Jeon, S. Lee, J.-I. Won, S. O. Han, J. Kim, J. Lee, Enzyme Microb. Tech- [108] R. E. Toomey, S. J. Wakil, Biochim. Biophys. Acta Lipids Lipid Metab.
nol. 2011, 49, 44 – 51. 1966, 116, 189 – 197.
[76] L. Zhang, W. Liu, J. Xiao, T. Hu, J. Chen, K. Chen, H. Jiang, X. Shen, Pro- [109] C. T. Nomura, K. Taguchi, Z. Gan, K. Kuwabara, T. Tanaka, K. Takase, Y.
tein Sci. 2007, 16, 1184 – 1192. Doi, Appl. Environ. Microbiol. 2005, 71, 4297 – 4306.
[77] A. T. Keatinge-Clay, A. A. Shelat, D. F. Savage, S.-C. Tsai, L. J. W. Miercke, [110] Y. Zhang, J. E. Cronan, J. Bacteriol. 1998, 180, 3295 – 3303.
J. D. O’Connell, III, C. Khosla, R. M. Stroud, Structure 2003, 11, 147 – [111] C. D. Cukier, A. G. Hope, A. A. Elamin, L. Moynie, R. Schnell, S. Schach,
154. H. Kneuper, M. Singh, J. H. Naismith, Y. Lindqvist, D. W. Gray, G.
[78] S. Natarajan, J.-K. Kim, T.-K. Jung, T. Doan, H.-P.-T. Ngo, M.-K. Hong, S. Schneider, ACS Chem. Biol. 2013, 8, 2518 – 2527.
Kim, V. Tan, S. Ahn, S. Lee, Y. Han, Y.-J. Ahn, L.-W. Kang, Mol. Cells 2012, [112] C.-Y. Lai, J. E. Cronan, J. Bacteriol. 2004, 186, 1869 – 1878.
33, 19 – 25. [113] T. Parish, G. Roberts, F. Laval, M. Schaeffer, M. Daff , K. Duncan, J. Bac-
[79] M. A. L. Sreshty, A. Surolia, G. N. Sastry, U. S. Murty, Mol. Inform. 2012, teriol. 2007, 189, 3721 – 3728.
31, 281 – 299. [114] E. Jeon, S. Lee, S. Lee, S. O. Han, Y. J. Yoon, J. Lee, J. Microbiol. Biotech-
[80] C. Oefner, H. Schulz, A. D’Arcy, G. E. Dale, Acta Crystallogr. Sect. D Biol. nol. 2012, 22, 990 – 999.
Crystallogr. 2006, 62, 613 – 618. [115] U. Oppermann, C. Filling, M. Hult, N. Shafqat, X. Wu, M. Lindh, J. Shaf-
[81] Z. Li, Y. Huang, J. Ge, H. Fan, X. Zhou, S. Li, M. Bartlam, H. Wang, Z. qat, E. Nordling, Y. Kallberg, B. Persson, H. Jçrnvall, Chem.-Biol. Interact.
Rao, J. Mol. Biol. 2007, 371, 1075 – 1083. 2003, 143 – 144, 247 – 253.
[82] J. W. Campbell, J. E. Cronan, Annu. Rev. Microbiol. 2001, 55, 305 – 332. [116] A. C. Price, Y.-M. Zhang, C. O. Rock, S. W. White, Structure 2004, 12,
[83] I. Sinha, T. Dick, J. Antimicrob. Chemother. 2004, 53, 1072 – 1075. 417 – 428.
[84] M. Sun, G. Zhu, Z. Qin, C. Wu, M. Lv, S. Liao, N. Qi, M. Xie, J. Cai, Mol. [117] K. P. Nambiar, D. M. Stauffer, P. A. Kolodziej, S. A. Benner, J. Am. Chem.
Biochem. Parasitol. 2012, 184, 20 – 28. Soc. 1983, 105, 5886 – 5890.

ChemBioChem 0000, 00, 0 – 0 www.chembiochem.org 19  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim &

These are not the final page numbers! ÞÞ


Reviews
[118] A. C. Price, Y.-M. Zhang, C. O. Rock, S. W. White, Biochemistry 2001, 40, [153] K. Saito, A. Kawaguchi, Y. Seyama, T. Yamakawa, S. Okuda, Eur. J. Bio-
12772 – 12781. chem. 1981, 116, 581 – 586.
[119] M. Fisher, J. T. M. Kroon, W. Martindale, A. R. Stuitje, A. R. Slabas, J. B. [154] D. A. Rozwarski, C. Vilcheze, M. Sugantino, R. Bittman, J. C. Sacchettini,
Rafferty, Structure 2000, 8, 339 – 347. J. Biol. Chem. 1999, 274, 15582 – 15589.
[120] H. T. Wright, Structure 2004, 12, 358 – 359. [155] R. J. Heath, J. R. Rubin, D. R. Holland, E. Zhang, M. E. Snow, C. O. Rock,
[121] Y.-M. Zhang, C. O. Rock, J. Biol. Chem. 2004, 279, 30994 – 31001. J. Biol. Chem. 1999, 274, 11110 – 11114.
[122] D. Tasdemir, G. Lack, R. Brun, P. Redi, L. Scapozza, R. Perozzo, J. Med. [156] C. W. Levy, A. Roujeinikova, S. Sedelnikova, P. J. Baker, A. R. Stuitje, A. R.
Chem. 2006, 49, 3345 – 3353. Slabas, D. W. Rice, J. B. Rafferty, Nature 1999, 398, 383 – 384.
[123] M.-J. Sohn, C.-J. Zheng, W.-G. Kim, J. Antibiot. 2008, 61, 687 – 691. [157] X. Qiu, S. S. Abdel-Meguid, C. A. Janson, R. I. Court, M. G. Smyth, D. J.
[124] S. R. Wickramasinghe, K. A. Inglis, J. E. Urch, S. Mller, D. M. F. van Aal- Payne, Protein Sci. 1999, 8, 2529 – 2532.
ten, A. H. Fairlamb, Biochem. J. 2006, 393, 447 – 457. [158] C. Y. Bott , F. Dubar, G. I. McFadden, E. Mar chal, C. Biot, Chem. Rev.
[125] L. Zhang, Y. Kong, D. Wu, H. Zhang, J. Wu, J. Chen, J. Ding, L. Hu, H. 2012, 112, 1269 – 1283.
Jiang, X. Shen, Protein Sci. 2008, 17, 1971 – 1978. [159] B. Pertschy, G. Zisser, H. Schein, R. Kçffel, G. Rauch, K. Grillitsch, C. Mor-
[126] B.-N. Wu, Y.-M. Zhang, C. O. Rock, J. J. Zheng, Protein Sci. 2009, 18, genstern, M. Durchschlag, G. Hçgenauer, H. Bergler, Mol. Cell. Biol.
240 – 246. 2004, 24, 6476 – 6487.
[127] B. Sedgwick, C. Morris, S. J. French, J. Chem. Soc. Chem. Commun. [160] F. E. Dayan, D. Ferreira, Y.-H. Wang, I. a Khan, J. a McInroy, Z. Pan, Plant
1978, 193 – 194. Physiol. 2008, 147, 1062 – 1071.
[128] L. R. Kass, K. Bloch, Proc. Natl. Acad. Sci. USA 1967, 58, 1168 – 1173. [161] C. J. Zheng, M.-J. Sohn, S. Lee, Y.-S. Hong, J.-H. Kwak, W.-G. Kim, Bio-
[129] H. Marrakchi, Y.-M. Zhang, C. O. Rock, Biochem. Soc. Trans. 2002, 30, chem. Biophys. Res. Commun. 2007, 362, 1107 – 1112.
1050 – 1055. [162] R. C. Hartkoorn, F. Pojer, J. A. Read, H. Gingell, J. Neres, O. P. Horlacher,
[130] R. J. Heath, C. O. Rock, J. Biol. Chem. 1996, 271, 27795 – 27801. K.-H. Altmann, S. T. Cole, Nat. Chem. Biol. 2014, 10, 96 – 98.
[131] Y. Cao, J. Yang, M. Xian, X. Xu, W. Liu, Appl. Microbiol. Biotechnol. 2010, [163] S. Rafi, P. Novichenok, S. Kolappan, X. Zhang, C. F. Stratton, R. Rawat, C.
87, 271 – 280. Kisker, C. Simmerling, P. J. Tonge, J. Biol. Chem. 2006, 281, 39285 –
[132] S. Mohan, T. M. Kelly, S. S. Eveland, C. R. Raetz, M. S. Anderson, J. Biol. 39293.
Chem. 1994, 269, 32896 – 32903. [164] K. Magnuson, S. Jackowski, C. O. Rock, J. E. Cronan, Microbiol. Rev.
[133] M. Leesong, B. S. Henderson, J. R. Gillig, J. M. Schwab, J. L. Smith, Struc- 1993, 57, 522 – 542.
ture 1996, 4, 253 – 264. [165] R. J. Heath, C. O. Rock, J. Biol. Chem. 1995, 270, 15531 – 15538.
[134] M. S. Kimber, F. Martin, Y. Lu, S. Houston, M. Vedadi, A. Dharamsi, K. M. [166] R. J. Heath, C. O. Rock, J. Bacteriol. 1998, 180, 1425 – 1430.
Fiebig, M. Schmid, C. O. Rock, J. Biol. Chem. 2004, 279, 52593 – 52602. [167] T. M. Lewin, P. Wang, R. A. Coleman, Biochemistry 1999, 38, 5764 –
[135] D. Kostrewa, F. K. Winkler, G. Folkers, L. Scapozza, R. Perozzo, Protein 5771.
Sci. 2005, 14, 1570 – 1580. [168] D. C. Cantu, Y. Chen, P. J. Reilly, Protein Sci. 2010, 19, 1281 – 1295.
[136] L. Moyni , S. M. Leckie, S. A. McMahon, F. G. Duthie, A. Koehnke, J. W. [169] L. Yuan, T. A. Voelker, D. J. Hawkins, Proc. Natl. Acad. Sci. USA 1995, 92,
Taylor, M. S. Alphey, R. Brenk, A. D. Smith, J. H. Naismith, J. Mol. Biol. 10639 – 10643.
2013, 425, 365 – 377. [170] Z. Zhuang, F. Song, H. Zhao, L. Li, J. Cao, E. Eisenstein, O. Herzberg, D.
[137] S. K. Sharma, M. Kapoor, T. N. C. Ramya, S. Kumar, G. Kumar, R. Modak, Dunaway-Mariano, Biochemistry 2008, 47, 2789 – 2796.
S. Sharma, N. Surolia, A. Surolia, J. Biol. Chem. 2003, 278, 45661 – [171] A. K. Spencer, A. D. Greenspan, J. E. Cronan, J. Biol. Chem. 1978, 253,
45671. 5922 – 5926.
[138] W. Liu, C. Luo, C. Han, S. Peng, Y. Yang, J. Yue, X. Shen, H. Jiang, Bio- [172] X. Lu, H. Vora, C. Khosla, Metab. Eng. 2008, 10, 333 – 339.
chem. Biophys. Res. Commun. 2005, 333, 1078 – 1086. [173] E. J. Steen, Y. Kang, G. Bokinsky, Z. Hu, A. Schirmer, A. McClure, S. B. del
[139] L. He, L. Zhang, X. Liu, X. Li, M. Zheng, H. Li, K. Yu, K. Chen, X. Shen, H. Cardayre, J. D. Keasling, Nature 2010, 463, 559 – 562.
Jiang, H. Liu, J. Med. Chem. 2009, 52, 2465 – 2481. [174] A. Jones, H. M. Davies, T. A. Voelker, Plant Cell 1995, 7, 359 – 371.
[140] Y. Kong, L. Zhang, Z. Yang, C. Han, L. Hu, H. Jiang, X. Shen, Acta Phar- [175] T. Voelker in Genetic Engineering: Principles and Methods, Vol. 18 (Ed.: J.
macol. Sin. 2008, 29, 870 – 876. Setlow), Springer, New York, 1996, pp. 111 – 133.
[141] P. L. Swarnamukhi, S. K. Sharma, P. Bajaj, N. Surolia, A. Surolia, K. [176] T. A. Voelker, A. C. Worrell, L. Anderson, J. Bleibaum, C. Fan, D. J. Haw-
Suguna, FEBS Lett. 2006, 580, 2653 – 2660. kins, S. E. Radke, H. M. Davies, Science 1992, 257, 72 – 74.
[142] Y.-J. Lu, S. W. White, C. O. Rock, J. Biol. Chem. 2005, 280, 30342 – 30348. [177] T. A. Voelker, H. M. Davies, J. Bacteriol. 1994, 176, 7320 – 7327.
[143] V. Agarwal, S. Lin, T. Lukk, S. K. Nair, J. E. Cronan, Proc. Natl. Acad. Sci. [178] D. L. Ollis, E. Cheah, M. Cygler, B. Dijkstra, F. Frolow, S. M. Franken, M.
USA 2012, 109, 17406 – 17411. Harel, S. J. Remington, I. Silman, J. Schrag, J. L. Sussman, K. H. G. Ver-
[144] M. J. Cryle, I. Schlichting, Proc. Natl. Acad. Sci. USA 2008, 105, 15696 – schueren, A. Goldman, Protein Eng. 1992, 5, 197 – 211.
15701. [179] L. Yuan, B. A. Nelson, G. Caryl, J. Biol. Chem. 1996, 271, 3417 – 3419.
[145] M. Babu, J. F. Greenblatt, A. Emili, N. C. J. Strynadka, R. A. F. Reithmeier, [180] K. M. Mayer, J. Shanklin, J. Biol. Chem. 2005, 280, 3621 – 3627.
T. F. Moraes, Structure 2010, 18, 1450 – 1462. [181] K. M. Mayer, J. Shanklin, BMC Plant Biol. 2007, 7, 1.
[146] L. Zhang, W. Liu, T. Hu, L. Du, C. Luo, K. Chen, X. Shen, H. Jiang, J. Biol. [182] H. M. Davies, L. Anderson, C. Fan, D. J. Hawkins, Arch. Biochem. Biophys.
Chem. 2007, 283, 5370 – 5379. 1991, 290, 37 – 45.
[147] H. Bergler, S. Fuchsbichler, G. Hçgenauer, F. Turnowsky, Eur. J. Biochem. [183] M. R. Pollard, L. Anderson, C. Fan, D. J. Hawkins, H. M. Davies, Arch. Bio-
1996, 242, 689 – 694. chem. Biophys. 1991, 284, 306 – 312.
[148] R. J. Heath, C. O. Rock, J. Biol. Chem. 1995, 270, 26538 – 26542. [184] K. G. Srikanta Dani, K. S. Hatti, P. Ravikumar, A. Kush, Plant Biol. 2011,
[149] R. J. Heath, N. Su, C. K. Murphy, C. O. Rock, J. Biol. Chem. 2000, 275, 13, 453 – 461.
40128 – 40133. [185] P. Kumari, J. Vonck, R. R. Schrçder, W. Khlbrandt, Microsc. Microanal.
[150] R. P. Massengo-Tiass , J. E. Cronan, Cell. Mol. Life Sci. 2009, 66, 1507 – 2007, 13, 158 – 159.
1517.
[151] H. Marrakchi, W. E. Dewolf, C. Quinn, J. West, B. J. Polizzi, C. Y. So, D. J.
Holmes, S. L. Reed, R. J. Heath, D. J. Payne, C. O. Rock, N. G. Wallis, Bio-
chem. J. 2003, 370, 1055 – 1062.
[152] A. Quemard, J. C. Sacchettini, A. Dessen, C. Vilcheze, R. Bittman, W. R. Received: October 6, 2014
Jacobs, J. S. Blanchard, Biochemistry 1995, 34, 8235 – 8241. Published online on && &&, 0000

& ChemBioChem 0000, 00, 0 – 0 www.chembiochem.org 20  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

ÝÝ These are not the final page numbers!


REVIEWS
K. Finzel, D. J. Lee, M. D. Burkart*
&& – &&
Using Modern Tools To Probe the
Structure–Function Relationship of
Fatty Acid Synthases
A complex molecular factory: Review- linking probes. These can be used to
ing the mechanism of fatty acid syn- study acyl carrier protein–partner pro-
thase (FAS) partner domains and current tein interactions and provide critical in-
inhibitors provides insight for the de- formation for engineering FASs for fuels,
sign and synthesis of mechanistic cross- feedstocks, and drug development.

ChemBioChem 0000, 00, 0 – 0 www.chembiochem.org 21  0000 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim &

These are not the final page numbers! ÞÞ

You might also like