Structure-based design of peptide inhibitors of ...

2 downloads 0 Views 245KB Size Report
peptide inhibitors; protease inhibitors; tetrapeptide inhibitors; botulinum neurotoxin inhibitors; arginine hydroxamate. Reference to this paper should be made as ...
The Botulinum J., Vol. 1, No. 3, 2009

297

Structure-based design of peptide inhibitors of botulinum neurotoxin serotype A proteolytic activity Matthew L. Ludivico Department of Cell Biology and Biochemistry, Integrated Toxicology Division, US Army Medical Research Institute of Infectious Diseases, 1425 Porter Street, Fort Detrick, MD 21702, USA E-mail: [email protected]

Leonard A. Smith Medical Countermeasures Technology, US Army Medical Research and Materiel Command, Fort Detrick, MD 21702-5011, USA Fax No. 301-619-2348 E-mail: [email protected]

S. Ashraf Ahmed* Department of Cell Biology and Biochemistry, Integrated Toxicology Division, US Army Medical Research Institute of Infectious Diseases, 1425 Porter Street, Fort Detrick, MD 21702, USA Fax: (301)-619-2348 E-mail: [email protected] *Corresponding author Abstract: P1’arginine of the substrate SNAP-25 is an absolute requirement for catalysis, and the active site of botulinum neurotoxin serotype A (BoNT/A) is populated by acidic residues. We reasoned that arginine derivatives, and basic peptides might bind to the proteolytic domain, and act as inhibitors. A systematic investigation of amino acids, derivatives and basic peptides provided proofs of our concept. D-arginine, arginine hydroxamate and basic peptides effectively inhibited BoNT/A activity. Optimisation of the peptide inhibitors resulted in a RRGC tetrapeptide displaying 90% inhibition at 20 microM. Our studies provide an inhibitor scaffold for potential development as a BoNT/A drug candidate. Keywords: botulinum neurotoxin catalytic domain; basic tetrapeptides; peptide inhibitors; protease inhibitors; tetrapeptide inhibitors; botulinum neurotoxin inhibitors; arginine hydroxamate. Reference to this paper should be made as follows: Ludivico, M.L., Smith, L.A. and Ahmed, S.A. (2009) ‘Structure-based design of peptide inhibitors of botulinum neurotoxin serotype A proteolytic activity’, The Botulinum J., Vol. 1, No. 3, pp.297–308.

Copyright © 2009 Inderscience Enterprises Ltd.

298

M.L. Ludivico et al. Biographical notes: Matthew L. Ludivico received a BSc in Molecular Biology from Lehigh University, PA. He later joined the US Army, received training as a Medical Laboratory Technician, and was stationed at the US Army Medical Research Institute of Infectious Diseases (USAMRIID). For six years, he worked on structural and functional aspects of botulinum neurotoxin catalytic domain, and on development of inhibitors against it. Leonard A. Smith received his PhD in Biochemistry from Georgetown University School of Medicine. He is the Senior Research Scientist (ST) for Medical Countermeasures Technology at the US Army Research Institute of Infectious Diseases. His research efforts focus on the development of toxin therapeutics and toxin vaccine products. He currently is the Chairman of the NATO HFM-ET91 working group. S. Ashraf Ahmed is a Research Chemist at USAMRIID. He had his MSc from Dhaka University, Bangladesh, and a PhD in microbial biochemistry from Kyoto University, Japan. He extensively worked on structure and mechanism of enzymes of amino acid and oxygen metabolism and of botulinum neurotoxin. A Research Chemist, his current research is focused on therapeutic development against botulism. He has published over 70 peer reviewed scientific papers.

1

Introduction

Of all the natural and synthetic toxins, botulinum neurotoxins are the most lethal. There are seven immunologically distinct botulinum neurotoxins (BoNT), designated BoNT/A-G, produced by strains of Clostridium botulinum (see Montecucco and Schiavo, 1995; Schiavo et al., 1995 for a review). While they are potential military threats and bioterrorism agents (Arnon et al., 2001), these neurotoxins are also widely used as laboratory research tools (Steinhardt et al., 1994), as clinical drugs in a variety of neuromuscular disorders and in cosmetic applications (Verheyden et al., 2001; Brin et al., 1999). BoNTs are expressed by bacteria as 150-kDa single polypeptides. Post-translational cleavage by a trypsin-like protease generates a 50-kDa N-terminal Light Chain (LC) and a 100-kDa C-terminal Heavy Chain (HC) that are still connected by a disulfide bond. The 100-kDa HC can further be proteolysed into a 50-kDa N-terminal membrane spanning domain (Hn) and a 50-kDa C-terminal receptor-binding domain (Hc). The LC possesses the toxic, zinc endopeptidase catalytic domain, but in the absence of HC, it is non-toxic. Because of its potential bio-weapon and bio-terrorism use, these neurotoxins are targets of inhibitor and drug design. Since the full-length BoNT is highly toxic but its separated catalytic LC moiety is not, we expressed BoNT/A LC in Eschericia coli and purified it to homogeneity from inclusion bodies (Ahmed and Smith, 2000) or soluble extracts (Ahmed et al., 2001; Jensen et al., 2003) as stable and highly active in vitro and in vivo (Alderton et al., 2000), and then characterised them further (Ahmed et al., 2003). In characterising a recombinant BoNT/A LC, we demonstrated earlier that organic and inorganic mercury compounds such as p-chloro-mercuribenzoate and mercuric chloride completely inhibited (Ki 90% by (Quality Controlled Biochemicals, Hampton, MS).

3

Methods

3.1 Protein concentration To determine protein concentration and to assess purity, UV-visible absorption spectra were recorded at 22°C with a Hewlett-Packard 8452 diode array spectrophotometer. LC concentration was determined using A0.1% (1 cm light path) value of 1.0 at 278 nm or by BCA assay (Pierce) with Bovine Serum Albumin (BSA) as standard. Both methods give the same result (Ahmed et al., 2001).

3.2 Enzymatic activity assays The enzymatic assay was based on HPLC separation and measurement of the cleaved products from a 17-residue C-terminal peptide corresponding to residue #187-203 of SNAP-25 (Ahmed and Smith, 2000; Schmidt and Bostian, 1995). A master reaction mixture lacking the LC was made and aliquots were stored at –20°C. At the time of assay, an aliquot of the master mix was thawed and 25 µl was added to 5 µl of the LC (see above) to initiate the enzymatic reaction. Components and final concentration in this 30 µl reaction mixture were 0.9 mM substrate peptide, 0.25 mM ZnCl2, 5 mM dithiothreitol, and 50 mM Na-HEPES, pH 7.4. Amino acids, their derivatives and peptides were added to the reaction mixtures at the indicated concentrations before starting the reaction by adding the substrate.

4

Results

4.1 Amino acids We measured the BoNT/A LC activity using a 17-residue synthetic peptide substrate (Schmidt and Bostian, 1997) encompassing the glutamine-arginine cleavage site of SNAP-25 in the presence of all natural L-amino acids and their D-enantiomers. Neither L-glutamine nor L-arginine, the two residues N- and C-terminal, respectively, to the BoNT/A cleavage site, was an inhibitor (Table 1). Although L-arginine was not an inhibitor, D-arginine inhibited the reaction by 50%. We have no explanation why D-arginine is an inhibitor and not L-arginine, but find an analogy with a peptide having D-cysteine as a better inhibitor than the one having L-cysteine in the sequence CRATKML (Schmidt and Stafford, 2002). Except for D- and L-cysteine, L-serine, and D-glutamate, inhibition by other L-amino acids was insignificant. Combining the results with L- and D-amino acids, the following generalisation of amino acids as inhibitors of BoNT/A protease activity was made: D-arginine > D-glutamate = D-cysteine > D-tryptophan > L-serine = D-aspartate = L-cysteine. Inhibitory action of cysteine is most likely due to its interaction with the active-site zinc (Ahmed and Smith, 2000; Schmidt and Stafford, 2002).

Structure-based design of peptide inhibitors Table 1

301

Effects of L- and D-amino acids on the catalytic activity of BoNT/A LC Percent activity

Amino acid

L-

Control

100

100

Arginine

90

50

D-

Lysine

105

110

Histidine

115

120

Proline

105

100

Hydroxyproline

125

n.d.

Alanine

120

115

Valine

118

n.d.

Leucine

120

110

Isoleucine

120

135

Norleucine

115

n.d.

Methionine1

105

115

Glycine

120

Cysteine

80

70

Cystine2

n.d.

90

Serine

80

115

100

100

1003

1304

90

145

Tryptophan

125

75

Aspartate

120

80

Glutamate

120

70

Asparagine

95

105

Glutamine

95

130

Threonine Tyrosine Phenylalanine

1

Each amino acid (1 mM) was incubated at 37°C for 5 min in a 30 µl reaction mixture containing 1 mM substrate peptide, 0.375 µg of LC and 50 mM HEPES pH 7.4. The control 100% activity represents a specific activity of 3.8 µmol/min/mg. Results represent average of three assays that was rounded to the nearest 5. 1 0.5 mM; 20.2 mM; 30.1 mM; 40.3 mM.

4.2 Arginine derivatives and hydroxamates Because arginine was shown to be an essential residue in the BoNT/A substrate (Schmidt and Bostian, 1997), and D-arginine displayed significant inhibition, we investigated several commercially available arginine and guanido derivatives in order to identify more potent inhibitors (Table 2). Among all the compounds tested, hydroxamates of both enantiomers of arginine were equally effective and most potent inhibitors. Although a derivative of dipeptide hydroxamate was reported to be ineffective as a BoNT inhibitor (Deshpande et al., 1995), hydroxamates are generally known as

302

M.L. Ludivico et al.

metalloprotease inhibitors (Butler et al., 2004). We measured the activity of BoNT/A LC as a function of substrate concentration at several fixed concentrations of L-arginine hydroxamate (Figure 1(a)). With the limitation of sub-Km concentration of substrate employed in this experiment, the reciprocal data points in the Linewever-Burke plot appeared to intersect on the y-axis suggesting a competitive inhibition. Replot of the apparent Km calculated from Figure 1(a) as a function of L-hydroxamate concentration (Figure 1(b)) was linear yielding an average a Ki of 0.16 mM. Lack of significant effects of hydroxamates of lysine, aspartate, glutamate or nicotinic acid (Table 2) suggest that inhibition by arginine hydroxamate is specific for BoNT/A LC. Figure 1

(a) Double reciprocal plots of reaction rate vs. substrate concentration at fixed L-arginine hydroxamate concentrations of 0. mM (open circle), 0.083 mM (closed circle), 0.167 mM (open triangle), and 0.33 mM (closed triangle) and (b) the apparent Km values calculated from (a) are plotted against L-arginine hydroxamate concentration. An average Ki of 0.16 mM was calculated from the relation Ki = (Km*[inhibitor]/(apparent Km – Km)

(a) Table 2

(b)

Effects of arginine and its derivatives on BoNT/A LC activity

Arginine derivative

% Activity

Control

100

L-Arginine

105

D-Arginine

50

L-Argininamide

110

L-Arginine ethyl ester

80

L-arginine p-nitroanilide

95

L-arginino succinate

80

L-arginine 7-amido-4 methyl

100

L-Arginine N-phosphate

100

L-arginine hydroxamate N-α-benzoyl DL-arginine

20 1

40 100

N-α-acetyl DL-arginine 1

80 Nω−Nitro D-arginine Me-ester Final concentration of each compound was 1 mM in the assay mixture. The control 100% activity represents a specific activity of 3.8 µmol/min/mg. Results represent average of three assays that was rounded to nearest 5. 1 0.1 mM.

Structure-based design of peptide inhibitors

303

Because inhibition by D-arginine and hydroxamates of D- or L-arginine but not by D-lysine or other hydroxamates (Tables 1 and 3) may suggest a specific inhibition of BoNT/A LC activity, we measured the activity of BoNT/B LC in the presence of these compounds. Our results showed that neither enantiomer of arginine and lysine nor arginine hydroxamate was an inhibitor of BoNT/B activity. This result also suggested that although the overall structures are very similar (Swaminathan and Eswaramoorthy, 2000), the active sites of BoNT/A and BoNT/B LCs are different enough to show this discrepancy. Arphamenine A and arphamenine B are extremely potent inhibitors of arginine aminopeptidase having very low (2.5–0.84 nM) Ki (Harbeson and Rich, 1988). Because both arginine aminopeptidase and BoNT/A LC share arginine (in the P1 and P1′ positions, respectively) in the cleavable bond, we tested the effects of arphamenine A & B on the activity of BoNT/A LC: at 4.5 mM concentration, they decreased the activity by 25–35% only. This result may serve as an indirect proof that inhibition of BoNT/A LC activity by D-arginine and hydroxamates of L- and D-arginine is specific. Moreover, arginine aminopeptidase is not known to be inhibited by L- or D-arginine. Table 3

Effects of hydroxamates on BoNT/A LC activity

Hydroxamate Control L-arginine hydroxamate* D-arginine hydroxamate* L-lysine hydroxamate

% Activity 100 20 20 110

D-lysine hydroxamate

140

DL-aspartate hydroxamate

100

D-aspartate β-hydroxamate

90

L-glutamate γ-hydroxamate

100

L-Tryptophan hydroxamate

90

DL-phenylalanine hydroxamate DL-tryptophan hydroxamate L-tyrosine hydroxamate

105 90 65

DL-methionine hydroxamate

110

Nicotinic acid hydroxamate

75

Final concentration of each compound was 1.0 mM in the assay mixture. The control 100% activity represents a specific activity of 3.8 µmol/min/mg. Results represent an average of three assays that was rounded to nearest 5. *0.1 mM.

We also investigated many commercially available guanidine derivatives and sulfur containing compounds (Table 4) as probable inhibitors of BoNT/A LC activity, and compared their effects with that of L-arginine hydroxamate and cysteine, respectively. None of the compounds tested were better than L-arginine hydroxamate as inhibitors. Cysteine is an inhibitor (Table 1) that acts as a chelator of zinc at the active site of BoNT/A LC (Ahmed and Smith, 2000; Schmidt and Stafford, 2002). The small protein (∼60 amino acids), metallothionein, containing 20 cysteine residues is well characterised physiological metal chelators including zinc (Chen and Song, 2009).

304

M.L. Ludivico et al.

Metal free form of this protein was also ineffective as an inhibitor (Table 4). All these results thus proved our prediction that arginine or its derivative(s) should act as inhibitor(s) of BoNT/A LC activity. Indeed, recent X-ray crystallographic studies provided structural proof of the inhibition showing that arginine hydroxamate binds at the active site of BoNT/A LC with its hydroxamate chelating the active site zinc (Fu et al., 2006; Silvaggi et al., 2007). Table 4

Guanidino derivatives and sulphur containing compounds

Compound

% Activity

Control

100

L-Arginine hydroxamate1 Nγ-aminoflavinate L-arginine

20 2

100

α-Guanidino glutarate

105

α-Guanidino butyrate

45

4-Guanidino benzoate

2

Guanidino succinate

60 75

L-Citruline

185

D-Citruline

185

Guanidine thiocyanate2

80

D-Cysteine

70

Thioglycollate

70

Thiazolidine COOH2

140

Metallothionein1

75

δ-Amino levulinic acid

50

Final concentration of each compound was 1 mM in the assay mixture. The control 100% activity represents a specific activity of 3.4 µmol/min/mg. Results represent average of three assays that was rounded to nearest 5. 1 0.2 mM; 20.1 mM.

4.3 Protamine sulphate and basic peptides Protamine sulphate is a small protein of 5100 Da that is often used in precipitating nucleic acids from cell extracts. When used at 20 µM concentration with standard BoNT/A LC assay mixtures, it inhibited 85% of the activity. We next tried two basic peptides, one GRKKRRQRRRPPQC, is derived from HIV-tat (Albini et al., 1998), and the other a polymer of arginine. As shown in Table 5, both inhibited the activity remarkably. These basic peptides might bind at the acidic access route to the active site restricting substrate access. Because these peptides contain arginine residues, they might also inhibit by interacting with the active site residues normally occupied by P1’ arginine of the substrate. Figure 2 shows that inhibition of BoNT/A LC activity by the GRKKRRQRRRPPQC peptide is biphasic probably suggesting a high affinity and a low affinity binding of the inhibitor to the protein.

Structure-based design of peptide inhibitors Table 5

305

Basic protein and peptides

Peptide

% Activity

Control

100

Protamine sulfate

15

GRKKRRQRRRPPQC-amide

10

RRRRRRRRGC-amide

35

RKRKRKRKGC-amide

60

dRKdRKdRKdRKGC-amide

45

dRdKdRdKdRdKdRdKGC-amide

35

RRGC-amide

10

Final concentration of each peptide was 20 µM in the assay mixture. Results represent an average of five assays that was rounded to nearest 5. Figure 2

Inhibition of BoNT/A LC activity by the peptide GRKKRRQRRRPPQC at a substrate concentration of 0.9 mM

Because the GRKKRRQRRRPPQC peptide containing both arginine and lysine residues was a better inhibitor than the peptide with eight arginine residues, we modified the latter peptide with alternating arginine and lysine residues both as D- and L-enatiomers. However, none of these modifications yielded any inhibitor better than the HIV-tat peptide. We also varied the length of the eight arginine containing peptide (not shown), and found that a basic peptide containing two arginine residues was as effective as the GRKKRRQRRRPPQC peptide. All of the peptides listed in Table 5 are highly soluble in aqueous buffers. The RRGC tetrapeptide competitively inhibits BoNT/A LC with a Ki of 157 nM (Kumaran et al., 2008b). The sequence of this peptide suggested that it could mimic the corresponding sequence of the substrate, QRAT, and might inhibit the BoNT/A LC reaction by competing with the substrate. Indeed, our recent high resolution 3-dimensional structure showed that RRGC was bound at the active site of BoNT/A LC that would normally be expected to be occupied by the substrate sequence (Kumaran et al., 2008). Moreover, the peptide was bound with extensive polar interactions with the protein residues.

5

Discussions

The enzymatic assay in this study used a very dependable quantitative method on a 17-mer SNAP-25 derived substrate (Schmidt and Bostian, 1995) instead of the

306

M.L. Ludivico et al.

full length SNAP-25. Therefore, some of the inhibitors may or may not be behave the same way on these two substrates. As potential drugs, peptides often exhibit highly desirable physical and pharmacological characteristics, such as good aqueous solubility, low toxicity, and high specificity for the targeted disease agent (Ayoub and Scheidegger, 2006; Marx, 2005). They are easily synthesised by conventional techniques, readily soluble in aqueous buffers, and are target-specific. Our results provide a proof of the concept that arginine derivatives and basic peptides should act as inhibitors of BoNT/A LC activity. In addition, low Ki of the tetrapeptide RRGC make it one of the highest affinity inhibitor thus far demonstrated for BoNT/A. Thus, this tetrapeptide or any of its variants (Kumaran et al., 2008b) will be a good candidate for drug development. Poor oral bioavailability of peptides may be a concern in their use as drugs (Hamman et al., 2005). In such a situation, the structural coordinates of the optimised tetrapeptide inhibitor should provide a starting point to design high affinity small molecule inhibitors.

Acknowledgement and disclaimer The research described herein was supported by JSTO-CBD Project number 3.10012_06_RD_B. Opinions, interpretations, conclusions, and recommendations are those of the author and are not necessarily endorsed by the US Army.

References Adler, M., Nicholson, J.D., Cornille, F. and Hackley Jr., B.E. (1998) ‘Efficacy of a novel metalloprotease inhibitor on botulinum neurotoxin B activity’, FEBS Lett., Vol. 429, pp.234–238. Ahmed, S.A. and Smith, L.A. (2000) ‘Light chain of botulinum A neurotoxin expressed as an inclusion body from a synthetic gene is catalytically and functionally active’, J. Protein Chem., Vol. 19, pp.475–487. Ahmed, S.A., Byrne, M.P., Jensen, M., Hines, H.B., Brueggemann, E. and Smith, L.A. (2001) ‘Enzymatic autocatalysis of botulinum A neurotoxin light chain’, J. Protein Chem., Vol. 20, pp.221–231. Ahmed, S.A., Mcphie, P. and Smith, L.A. (2003) ‘Autocatalytically fragmented light chain of botulinum A neurotoxin is enzymatically active’, Biochemistry, Vol. 42, pp.12539–12549. Ahmed, S.A., Olson, M.A., Ludivico, M.L., Gilsdorf, J. and Smith, L.A. (2008) ‘Identification of residues surrounding the active site of type a botulinum neurotoxin important for substrate recognition and catalytic activity’, Protein J., Vol. 27, pp.151–162. Albini, A., Benelli, R., Giunciuglio, D., Cai, T., Mariani, G., Ferrini, S. and Noonan, D.M. (1998) ‘Identification of a novel domain of HIV tat involved in monocyte chemotaxis’, J. Biol. Chem., Vol. 273, pp.15895–15900. Alderton, J.M., Ahmed, S.A., Smith, L.A. and Steinhardt, R.A. (2000) ‘Evidence for a vesicle-mediated maintenance of store-operated calcium channels in a human embryonic kidney cell line’, Cell Calcium, Vol. 28, pp.161–169. Arnon, S.S., Schechter, R., Inglesby, T.V., Henderson, D.A., Bartlett, J.G., Ascher, M.S., Eitzen, E., Fine, A.D., Hauer, J., Layton, M., Lillibridge, S., Osterholm, M.T., O’toole, T., Parker, G., Perl, T.M., Russell, P.K., Swerdlow, D.L. and Tonat, K. (2001) ‘Botulinum toxin as a biological weapon: medical and public health management’, Jama., Vol. 285, pp.1059–1070.

Structure-based design of peptide inhibitors

307

Ayoub, M. and Scheidegger, D. (2006) ‘Peptide drugs, overcoming the challenges, a growing business’, Chemistry Today, Vol. 24, pp.46–48. Boldt, G.E., Kennedy, J.P., Hixon, M.S., Mcallister, L.A., Barbieri, J.T., Tzipori, S. and Janda, K.D. (2006) ‘Synthesis, characterization and development of a high-throughput methodology for the discovery of botulinum neurotoxin a inhibitors’, J. Comb. Chem., Vol. 8, pp.513–521. Brin, M.F., Lew, M.F., Adler, C.H., Comella, C.L., Factor, S.A., Jankovic, J., O’brien, C., Murray, J.J., Wallace, J.D., Willmer-Hulme, A. and Koller, M. (1999) ‘Safety and efficacy of NeuroBloc (botulinum toxin type B) in type A-resistant cervical dystonia’, Neurology, Vol. 53, pp.1431–1438. Burnett, J.C., Schmidt, J.J., Stafford, R.G., Panchal, R.G., Nguyen, T.L., Hermone, A.R., Vennerstrom, J.L., Mcgrath, C.F., Lane, D.J., Sausville, E.A., Zaharevitz, D.W., Gussio, R. and Bavari, S. (2003) ‘Novel small molecule inhibitors of botulinum neurotoxin A metalloprotease activity’, Biochem. Biophys. Res. Commun., Vol. 310, pp.84–93. Butler, G.S., Tam, E.M. and Overall, C.M. (2004) ‘The canonical methionine 392 of matrix metalloproteinase 2 (gelatinase A) is not required for catalytic efficiency or structural integrity: probing the role of the methionine-turn in the metzincin metalloprotease superfamily’, J. Biol. Chem., Vol. 279, pp.15615–15620. Chen, M.D. and Song, Y.M. (2009) ‘Tissue metallothionein concentrations in mice and humans with hyperglycemia’, Biol. Trace Elem. Res., Vol. 127, pp.251–256. Deshpande, S.S., Sheridan, R.E. and Adler, M. (1995) ‘A study of zinc-dependent metalloendopeptidase inhibitors as pharmacological antagonists in botulinum neurotoxin poisoning’, Toxicon., Vol. 33, pp.551–557. Fu, Z., Chen, S., Baldwin, M.R., Boldt, G.E., Crawford, A., Janda, K.D., Barbieri, J.T. and Kim, J.J. (2006) ‘Light chain of botulinum neurotoxin serotype A: structural resolution of a catalytic intermediate’, Biochemistry, Vol. 45, pp.8903–8911. Hamman, J.H., Enslin, G.M. and Kotze, A.F. (2005) ‘Oral delivery of peptide drugs: barriers and developments’, BioDrugs, Vol. 19, pp.165–177. Harbeson, S.L. and Rich, D.H. (1988) ‘Inhibition of arginine aminopeptidase by bestatin and arphamenine analogues. Evidence for a new mode of binding to aminopeptidases’, Biochemistry, Vol. 27, pp.7301–7310. Jensen, M.J., Smith, T.J., Ahmed, S.A. and Smith, L.A. (2003) ‘Expression, purification, and efficacy of the type A botulinum neurotoxin catalytic domain fused to two translocation domain variants’, Toxicon, Vol. 41, pp.691–701. Kumaran, D., Rawat, R., Ahmed, S.A. and Swaminathan, S. (2008a) ‘Substrate binding mode and its implication on drug design for botulinum neurotoxin A’, PLoS Pathog, Vol. 4, pp.e1000165. Kumaran, D., Rawat, R., Ludivico, M.L., Ahmed, S.A. and Swaminathan, S. (2008b) ‘Structure and substrate based inhibitor design for clostridium botulinum neurotoxin serotype A’, J. Biol. Chem., Vol. 283, pp.18883–18891. Lacy, D.B., Tepp, W., Cohen, A.C., Dasgupta, B.R. and Stevens, R.C. (1998) ‘Crystal structure of botulinum neurotoxin type A and implications for toxicity’, Nat. Struct. Biol., Vol. 5, pp.898–902. Marx, V. (2005) ‘Watching peptide drugs grow up: Peptide therapeutics market grows in fits and starts for drug firms and contract manufacturers’, Chemical and Engineering News, Vol. 83, pp.17–24. Montecucco, C. and Schiavo, G. (1995) ‘Structure and function of tetanus and botulinum neurotoxins’, Q. Rev. Biophys., Vol. 28, pp.423–472. Schiavo, G., Shone, C.C., Bennett, M.K., Scheller, R.H. and Montecucco, C. (1995) ‘Botulinum neurotoxin type C cleaves a single Lys-Ala bond within the carboxyl-terminal region of syntaxins’, J. Biol. Chem., Vol. 270, pp.10566–10570.

308

M.L. Ludivico et al.

Schmidt, J.J. and Bostian, K.A. (1995) ‘Proteolysis of synthetic peptides by type A botulinum neurotoxin’, J. Protein Chem., Vol. 14, pp.703–708. Schmidt, J.J. and Bostian, K.A. (1997) ‘Endoproteinase activity of type A botulinum neurotoxin: substrate requirements and activation by serum albumin’, J. Protein Chem., Vol. 16, pp.19–26. Schmidt, J.J. and Stafford, R.G. (2002) ‘A high affinity competitive inhibitor of type A botulinum neurotoxin protease activity’, FEBS Lett., Vol. 532, pp.423–426. Schmidt, J.J. and Stafford, R.G. (2005) ‘Botulinum neurotoxin serotype F: identification of substrate recognition requirements and development of inhibitors with low nanomolar affinity’, Biochemistry, Vol. 44, pp.4067–4073. Silvaggi, N.R., Boldt, G.E., Hixon, M.S., Kennedy, J.P., Tzipori, S., Janda, K.D. and Allen, K.N. (2007) ‘Structures of clostridium botulinum neurotoxin serotype a light chain complexed with small-molecule inhibitors highlight active-site flexibility’, Chem. Biol., Vol. 14, pp.533–542. Steinhardt, R.A., Bi, G. and Alderton, J.M. (1994) ‘Cell membrane resealing by a vesicular mechanism similar to neurotransmitter release’, Science, Vol. 263, pp.390–393. Swaminathan, S. and Eswaramoorthy, S. (2000) ‘Structural analysis of the catalytic and binding sites of Clostridium botulinum neurotoxin B’, Nat. Struct. Biol., Vol. 7, pp.693–699. Verheyden, J., Blitzer, A. and Brin, M.F. (2001) ‘Other noncosmetic uses of BOTOX’, Semin. Cutan. Med. Surg., Vol. 20, pp.121–126. Zalups, R.K. (2000) ‘Molecular interactions with mercury in the kidney’, Pharmacol Rev., Vol. 52, pp.113–143. Zuniga, J.E., Schmidt, J.J., Fenn, T., Burnett, J.C., Arac, D., Gussio, R., Stafford, R.G., Badie, S.S., Bavari, S. and Brunger, A.T. (2008) ‘A potent peptidomimetic inhibitor of botulinum neurotoxin serotype a has a very different conformation than snap-25 substrate’, Structure, Vol. 16, pp.1588–1597.