Abstract
Tyrosyl-DNA phosphodiesterase 1 (TDP1) is a molecular target for the sensitization of cancer cells to the FDA-approved topoisomerase inhibitors topotecan and irinotecan. High-throughput screening of natural product extract and fraction libraries for inhibitors of TDP1 activity resulted in the discovery of a new class of knotted cyclic peptides from the marine sponge Axinella sp. Bioassay-guided fractionation of the source extract resulted in the isolation of the active component which was determined to be an unprecedented 42-residue cysteine-rich peptide named recifin A. The native NMR structure revealed a novel fold comprising a four strand antiparallel β-sheet and two helical turns stabilized by a complex disulfide bond network that creates an embedded ring around one of the strands. The resulting structure, which we have termed the Tyr-lock peptide family, is stabilized by a tyrosine residue locked into three-dimensional space. Recifin A inhibited the cleavage of phosphodiester bonds by TDP1 in a FRET assay with an IC50 of 190 nM. Enzyme kinetics studies revealed that recifin A can specifically modulate the enzymatic activity of full-length TDP1 while not affecting the activity of a truncated catalytic domain of TDP1 lacking the N-terminal regulatory domain (Δ1–147), suggesting an allosteric binding site for recifin A on the regulatory domain of TDP1. Recifin A represents both the first of a unique structural class of knotted disulfide-rich peptides and defines a previously unseen mechanism of TDP1 inhibition that could be productively exploited for potential anticancer applications.
Graphical Abstract
INTRODUCTION
Relaxation of supercoiled DNA by topoisomerases is necessary for the normal cell functions of DNA transcription, replication, recombination, and repair.1 Topoisomerase I (TOP1) mediates both DNA strand break and religation by forming a transient, covalent 3′-;-phosphotyrosyl bond with the DNA substrate (Figure 1).2 This TOP1-DNA cleavage complex is the target of chemotherapeutic TOP1 inhibitors such as the natural product camptothecin. Irinotecan, an analogue of camptothecin, is a widely used anticancer agent that stabilizes the TOP1-DNA cleavage complex, causing irreversible double-strand DNA breaks, eventually leading to the death of replicating cancer cells.3,4 Tyrosyl-DNA phosphodiesterase 1 (TDP1) is an enzyme that, upon recognizing stalled TOP1-DNA cleavage complexes, catalyzes the cleavage of the 3′-phosphotyrosyl bond between DNA and TOP1 (Figure 1).5 TDP1 is composed of an as-yet nonstructurally characterized N-terminal regulatory domain whose function has been reported to be modulated by both phosphorylation and SUMOylation and a C-terminal catalytic domain that utilizes two histidine residues to effect phosphodiester cleavage at Tyr723 of TOP1 (Figure 1).6 After removal of the 3′ adduct, polynucleotide kinase phosphatase prepares the degraded DNA strands for further repair by DNA polymerase β and DNA ligase III.7 The clearance of TOP1-DNA complexes (Figure 1) results in escape from TOP1 inhibitor-induced cell death. This activity has led to the consideration of TDP1 as a molecular target for the sensitization of replicating cancer cells to camptothecin and related chemotherapeutic agents.6,8,9
Marine organisms are known to be a diverse source of bioactive peptides and proteins including lectins, defensins, conotoxins, cyclic peptides, and depsipeptides.10,11 Previously reported cysteine-rich peptides (CRPs) isolated from marine sponges include asteropine A,12 the asteropsins,13–15 the neopetrosiamides,16,17 and the barrettides.18 Within the Porifera, the genus Axinella itself has yielded many bioactive molecules, including cyclic peptides, lipopolysaccharides, lectins, alkaloids, terpenes, lipids, and polyether macrolides.19–27 We recently conducted a high-throughput screen to identify natural product inhibitors of TDP1 activity.28 During this effort, inhibitory activity was identified in the aqueous extract of the marine sponge Axinella sp. Here we report the first CRP isolated from the genus Axinella, which we have named recifin A after the Cape Recife Nature Reserve in South Africa from which the source sponge was harvested. The 42-residue peptide was sequenced, the disulfide connectivity was elucidated, and the unique three-dimensional structure of the peptide was solved using solution state homonuclear NMR spectroscopy. The isolated peptide recifin A is shown to specifically modulate the enzymatic activity of full-length TDP1 but not an enzymatically active N-terminal truncated variant (Δ147TDP1) lacking the regulatory domain, suggesting an allosteric recifin A binding site within the regulatory domain of TDP1.
RESULTS AND DISCUSSION
Isolation and Primary Structure Determination.
The aqueous extract of the marine sponge Axinella sp. was identified as active in a previously published high throughput screen for TDP1 inhibitory activity.28 The aqueous extract was first subjected to vacuum-assisted, wide-pore C4 chromatography and eluted using a stepwise methanol gradient. Fractions were tested for TDP1 inhibitory activity and subjected to liquid chromatography-mass spectrometry (LC-MS) analysis. Peptidic molecular charge envelopes were observed in LC-MS analysis of the active fractions, and the proteinaceous composition of active components was confirmed by SDS-PAGE. A family of four main Axinella peptides was isolated from the initial chromatographic step with observed average masses of 4683.87, 4785.89, 4915.95, and 5674.47 Da (Figure S1).
A combination of reversed-phase-high performance liquid chromatography (RP-HPLC) and bioassay-guided fractionation28 was used to isolate the most abundant and most active peptide, recifin A (MW 4915.95 Da), to homogeneity. The yield of purified recifin A from the crude aqueous extract was approximately 0.1% w/w. Recifin A was found to inhibit full-length recombinant human TDP1 enzymatic activity in a concentration-dependent manner with an IC50 of 2.4 μM in a biochemical assay for cleavage of a 5′-radiolabeled oligonucleotide DNA substrate containing a 3′-phosphotyrosyl residue (Figure 2).29 Importantly, recifin A retained the ability to inhibit TDP1 processing of the radiolabeled oligonucleotide within a whole-cell extract assay context, indicating the specificity and stability of the molecule. This is significant as it shows that recifin A could exert its inhibitory activity against native TDP1 in the presence of other cellular macromolecules and against an enzyme whose regulatory domain had potentially been post-translationally modified. The other main Axinella-derived peptides showed weaker TDP1 inhibitory activity, indicating they are likely additional members of the same structural class of peptides and will be characterized more fully in subsequent research studies (Figure S2).
The primary amino acid sequence of recifin A was determined using a combination of tandem mass spectrometry (MS/MS) and automated Edman degradation. The monoisotopic mass of the intact recifin A peptide was observed to be 4912.9661 Da. Upon disulfide reduction and alkylation with 4-vinylpyridine (4-VP, 105.06 Da), a mass increase of 636.38 Da was observed, which indicated the reduction of three disulfide bonds and conversion of six cysteine residues to S-pyridylethyl cysteine (Figure S3). Neither the native peptide nor the 4-VP alkylated peptide was amenable to N-terminal amino acid sequencing by Edman degradation, which suggested a blocked N-terminus. A trypsin digest of the 4-VP alkylated peptide was performed which generated three fragments A, B, and C with molecular weights of 1205.48, 2136.94, and 2242.95 Da, respectively (Figure S4).
Sequencing of the tryptic fragments by LC/MS/MS and confirmation by Edman degradation (Figure S4A–C)30 indicated the presence of a pyroglutamic acid residue (pGlu) on the N-terminus of fragment A explaining the lack of success with N-terminal Edman degradation of recifin A. This was confirmed by selective cleavage of the pGlu with Pfu pyroglutamate aminopeptidase. Upon successful enzymatic removal of the N-terminal pGlu from the reduced, alkylated recifin A, 35 contiguous amino acids of the N-terminally truncated peptide were able to be sequenced by Edman degradation. In addition to trypsin digestion, the alkylated peptide was subjected to digestion with chymotrypsin, glutamic acid C-terminal (Glu-C), and proline endopeptidases. The resultant fragments were sequenced by CID MS/MS only (Figure S5) and confirmed the full sequence of recifin A. The theoretical mass of the proposed amino acid sequence of recifin A was 4918.9994 Da, which differed from the observed mass by 6.0333 Da, confirming the presence of three disulfide bonds (2.1 ppm mass error).
Disulfide Bond Mapping.
The disulfide bonding pattern of recifin A was determined using a partial reduction and alkylation approach as previously reported.31 A combination of 80 μM recifin A and 50 mM tris(2-carboxyethyl)phosphine (TCEP) yielded one (two intact cystines, 2-SS), and two disulfide bond (one intact cystine, 1-SS) reduction events and the fully reduced species (zero intact cystines, 0-SS) as shown in Figure 3A.
The main 2-SS, N-ethylmaleimide (NEM) alkylated peptide isoform (Figure 3A) was fully reduced, alkylated, and digested with trypsin. The resultant trypsin fragments were sequenced to map the positions of the alkylation events (Figure 3B and Figure S6). NEM-alkylated cysteine residues were found to be at positions 22 and 42, which mapped a projected cystine linkage at Cys IV–VI. The 2-SS, NEM-alkylated peptide was digested with chymotrypsin to map the remaining, intact disulfide linkages by LC-MS. MassHunter (Agilent Technologies, Inc.) software was used to construct a database of the three possible disulfide-linked sequence permutations (Cys I–II, Cys III–V, Cys IV–VI; Cys I–III, Cys II–V, Cys IV–VI; and Cys I–V, Cys II–III, Cys IV–VI) and to match the observed chymotrypsin fragment masses to a set of theoretical digest fragment masses. A limitation of 5 ppm mass error was applied to the fragment matching process. Only fragments that linked Cys I–III and Cys II–V were observed (Supporting Table 1). Taken together, the data indicated the disulfide bond connectivity of recifin A to be Cys I–III, Cys II–V, and Cys IV–VI (Figure 3C).
The molecular weight, number of cysteine residues, and the stability of recifin A are similar to those reported for members of the inhibitory cystine knot (ICK) family, comprising protease inhibitors, toxins, and antimicrobial peptides. However, the ICK family is characterized by the intertwined or “knotted” Cys I–IV, Cys II–V, Cys III–VI disulfide bond arrangement.32 The recifin A disulfide bond framework is Cys I–III, Cys II–V, and Cys IV–VI, so while recifin A is a CRP, the peptide is not a member of the ICK family. Indeed, the primary amino acid sequence of recifin A is not homologous to any sequence within the nonredundant GenBank translated protein database (BLASTp search). While recifin A does display some component similarities to CRPs isolated from the marine sponge genus Asteropus,12–15 there were no identified amino acid sequence alignments of recifin A to Asteropus-derived CRPs (or ICK peptides) within the KNOTTIN database (Figure S7).33
Solution NMR Spectroscopy Structure Determination.
Given the lack of sequence homology to known proteins and unexpected disulfide array when compared to other CRPs, we subjected recifin A to solution NMR spectroscopy in order to characterize its three-dimensional structure. The one-dimensional 1H NMR spectrum showed excellent signal dispersion across the entire spectral region indicating a well-structured peptide (Figure S8A). Homonuclear 1H TOCSY and NOESY data were used for sequential assignments as described previously (Figure S9).34 This process proved a significant challenge due to a number of unusual chemical shifts and features in the NMR data for recifin A. Chemical shift anomalies included the Gly16 HN proton at 5.51 ppm, upfield of several Hα protons. The Hβ resonances of Tyr40 and Pro35 were observed at 1.15 and −0.33 ppm, respectively, the latter being the most upfield resonance in the spectrum. Finally, the Hα of Cys11 was essentially overlapping one of the Hβ resonance at 2.68 ppm. Resonances observed at 4.94 and 5.58 were, after identification of TOCSY peaks to their respective Hβ protons, assigned as the hydroxyl protons of Ser27 and Ser29, and a resonance at 7.96 was assigned as the phenolic proton of Tyr6 because of a lack of TOCSY peaks but strong NOESY connections to Tyr6 Hε. These protons are all not expected to be visible in the spectra due to fast exchange with the solvent but in the recifin A structure must clearly be involved in strong hydrogen bonds and protected from the solvent. Finally, four individual aromatic 1H signals were identified for Tyr6 (Hδ1, Hδ2, Hε1, Hε2) revealing that it is positioned in a tightly packed environment where ring-flips are sufficiently slowed down to prevent averaging into the typically observed single Hδ* and Hε* resonances. Line broadening, suggesting dynamics, was also observed around residues 21–25, with the HN proton of Arg25 broadened beyond detection. In addition to the homonuclear data, a 1H–13C HSQC data set was recorded at natural abundance, which was essential for confirming all proton assignments and provided 13C chemical shift information for dihedral restraints.
Initial analysis of secondary Hα chemical shifts suggested secondary structural features in the form of short β-strands and α-helices/turns, as indicated by significant positive and negative shifts, respectively (Figure S8B).35 The three-dimensional solution structure of recifin A was calculated using torsion angle dynamics in Cyana followed by refinement in a watershell using crystallography and NMR system (CNS).36 A total of 425 distance restraints, including 403 distance restraints derived from NOEs, 22 hydrogen bond restraints, and 75 dihedral angle restraints (ϕ, ψ, χ) were included in the calculations (Table 1). A family of 20 structures were chosen to represent the solution structure of recifin A based on energies, stereochemical quality, and consistency with the experimental data (Table 1). As seen from the superposition of the ensemble, the structure is well-defined except a loop region comprising residues 21–25, consistent with the observed line broadening (Figure 4). The structure is dominated by a central, antiparallel β-sheet comprising four strands involving residues 4–6, 14–16, 27–29, and 40–41 and two short 310 helical turns involving residues 21–23 and 36–38. The elements of secondary structure are stabilized by the three disulfide bonds, with the Cys5-Cys21 and Cys22-Cys42 disulfides bracing the 21–23 turn to strands 1 and 4, respectively, and the Cys11-Cys39 cross-bracing two loops. Intriguingly, the disulfides form an embedded ring together with their backbone segments, through which the third strand (27–29) is threaded. This arrangement gives rise to a previously not observed fold and represents a new type of cysteine-rich peptide knot. Although this is somewhat reminiscent of the inhibitory cystine knot, where two of the disulfide bonds form a ring structure through, which the third disulfide bond is threaded forming the knot,37,38 it bears perhaps even more resemblance to the lasso peptides, in which the peptide backbone is threaded through a ring formed by an N-terminus to side-chain carboxyl lactam bond (Figure 5) and locked in place by two aromatic residues.39
Table 1.
distance restraints | |
intraresidue (|i – j| = 0) | 126 |
sequential (|i – j| = 1) | 115 |
medium range (|i – j| ≤ 5) | 50 |
long range (|i – j| > 5) | 112 |
hydrogen bonds | 22 |
total | 425 |
dihedral angle restraints | |
Φ | 25 |
φ | 25 |
χ | 25 |
total | 75 |
structure statistics | |
energies (kcal/mol, mean ± SD) | |
overall | −1422.3 ± 36.5 |
bonds | 16.9 ± 1.2 |
angles | 49.8 ± 4.4 |
improper | 19.3 ± 2.4 |
dihedral | 181.4 ± 1.6 |
van de Waals | −221.2 ± 3.9 |
electrostatic | −1469.1 ± 35.2 |
NOE (experimental) | 0.03 ± 0.01 |
constrained dihedrals (experimental) | 0.6 ± 0.3 |
atomic RMSD (Å) | |
mean global backbone (1–42)a | 0.93 ± 0.31 |
mean global heavy (1–42)a | 1.57 ± 0.26 |
mean global backbone (3–17, 27–42) | 0.44 ± 0.08 |
mean global heavy (3–17, 27–42) | 1.17 ± 0.15 |
MolProbity statistics | |
Clash score, all atomsb | 10.02 ± 1.9 |
poor rotamers | 0 ± 0 |
favored rotamers | 95.8 ± 1.6 |
Ramachandran outliers (%) | 0 ± 0 |
Ramachandran favored (%) | 94.9 ± 3.3 |
MolProbityc score | 1.8 ± 0.2 |
MolProbity percentile | 82.9 ± 8.3 |
violations | |
distance constraints (>0.5 Å) | 0 |
dihedral-angle constraints (>5°) | 0 |
Pairwise RMSD from 20 refined structures over amino acids 1–42.
Number of steric overlaps (>0.4 Å)/1000 atoms.
100% is the best among structures of comparable resolution. 0% is the worst.
However, the embedded ring in recifin A (Figure 5A) is bigger than both lasso-peptides (e.g., microcin J25)40 and prototypic ICK peptides (e.g., kalata B1) (Figure 5B and Figure 5C, respectively).41 The unusual fold of recifin A is further stabilized by Tyr6, which is deeply buried in the middle of the peptide (Figure 6) and locked in place by a number of residues, most notably Cys11, Tyr14, Ser29, and Leu32. It is because of this tight packing that Tyr6 does not undergo the usual fast “ring flips” typically observed for aromatic residues, which results in single resonance lines and sets of NOEs for the geminal Hδ* and Hε* protons. Instead, recifin A has extensive individual NOEs from surrounding residues to each of the four aromatic (Hδ1/2 and Hε1/2) protons locking Tyr6 in a specific conformation. In addition, a series of NOEs from the phenolic proton of Tyr6 to other surrounding residues can be observed, further highlighting the structurally stabilizing role of Tyr6 as these types of NOEs are rarely seen in a NOESY spectrum. The buried Tyr6 phenol group serves both as hydrogen bond donor, to the backbone carbonyl of Glu31, and as hydrogen bond acceptor for the HN proton of Gln33, while the hydroxyl groups of Ser27 and Ser29 serve as hydrogen bond donors to the carbonyls of Asp8 and Glu31, respectively. Ring current effects from aromatic residues are responsible for the unusual chemical shifts with Tyr6 packing against the Hα of Cys11, while the positioning of the side chains of Tyr14, Tyr28, and Trp37 is consistent with ring current effects on the HN of Gly16 and the Hβ resonances of Tyr40 and Pro35, respectively. This unprecedented structural arrangement and integral locked position of Tyr6 in the recifin A structure led us to the term structural class Tyr-lock peptides of which recifin A is the first.
Although there is currently no information available as to which region of recifin A interacts with the regulatory domain of TDP1, there are potential points on recifin A where residues commonly involved in protein-protein interactions (i.e., Arg9, Phe10, Arg25, and Trp37) are clustered.42,43 The identification of key binding residues will require the capability to produce this peptide either synthetically or recombinantly, facilitating thorough structure-activity relationship analysis.
Biological Activity and Mechanism of Inhibition.
To characterize the effect of recifin A on TDP1 catalytic activity, a steady-state enzyme kinetics study was undertaken. The initial TDP1 screening assay was conducted at a single time-point involving product capture and optimized for use with crude natural product extracts which can result in inflated effective concentrations; it was therefore necessary to employ a previously described fluorescence resonance energy transfer (FRET) assay28 to more accurately measure the kinetic parameters of the recifin A-TDP1 interaction. Recifin A inhibitory activity was confirmed in the FRET assay format as shown in Figure 7. Recifin A inhibited full-length TDP1 enzymatic activity in a concentration-dependent manner with an apparent IC50 of 190 nM. We also evaluated the ability of recifin A to inhibit the enzymatic activity of a N-terminal truncated form of TDP1 (Δ147TDP1) in which the regulatory domain had been removed44 and found only a minimal effect (approximately 20% maximal inhibition) at the highest concentration (1500 nM) assessed. Initial kinetic evaluation of the effect of recifin A on full-length TDP1 activity revealed that submicromolar concentrations of recifin A increased the Km for the substrate, broadly defined as an inhibitory characteristic. In addition, a modest increase of the observed Vmax value was also detected. This second observation is most often associated with allosteric enzymatic activators (Figure 8A).45,46
Further analysis of the data revealed that the recifin A-dependent increase in the Km for the substrate is significantly more pronounced (approximately 6-fold higher) than the modest effect on the observed Vmax (approximately 1.6-fold higher, Figure S10), consistent with our initial discovery of this peptide as a TDP1 inhibitor. To further characterize the inhibitory effects of recifin A on TDP1, we used the FRET based assay to determine whether recifin A had any effect on the enzymatic activity of Δ147TDP1, lacking the regulatory domain of TDP1. While smaller, Δ147TDP1 retains the substrate binding cleft and dual histidine-lysine-aspartic acid (HKD) motifs responsible for phosphodiesterase catalysis.47,48 As shown in Figure 8B, recifin A had overlapping 95% confidence intervals for both Km and Vmax with the untreated controls, indicating that it does not affect the enzymatic activity of truncated TDP1. This suggests that the binding site for recifin A on TDP1 is outside the active site region common to both the truncated and full-length forms of TDP1, consistent with our results suggesting that recifin A acts as an allosteric modulator of TDP1 enzymatic activity and is binding to the N-terminal TDP1 regulatory domain. Additionally, evaluation of extract of the marine sponge Axinella sp. that yielded recifin A, in an assay to identify inhibitors of the related enzyme tyrosyl-DNA phosphodiesterase II (TDP2),6 indicated lack of inhibition of TDP2 (data not shown). The lack of activity against this related phosphodiesterase suggests another level of specificity for recifin A against TDP1.
Mechanistically, the recifin A-TDP1 interaction is interesting in that modulators that increase the Km of an enzyme for the substrate are most often characterized as competitive inhibitors. However, the fact that an enzymatically active but truncated form of the protein, with an identical active site, was unaffected by recifin A indicates that the peptide was not directly competing for substrate binding at the active site. Additionally, the observation that recifin A treatment increased the Vmax of the enzyme is a general characteristic of an enzymatic activator, further highlighting the novelty of the recifin A-TDP1 interaction and reinforcing the evidence that recifin A does not compete with the phosphotyrosyl-DNA TDP1 substrate in contrast to recently discovered TDP1 inhibitors.49 These attributes together in a single interaction are unusual but not without precedent when considering that recifin A is not a small molecule but a complex peptide. There are several classes of enzymes for which a protein-protein interaction is known to change substrate specificity, catalytic efficiency, or both.50–53 That recifin A may bind TDP1 allosterically suggests that there may be more to understand about the allosteric regulation of cellular TDP1 activity and that more of the TDP1 protein may be both pharmacologically accessible and therapeutically relevant. It is worth noting that the importance and major topological features present in the first 147 amino acids (deleted from the truncated variant) have not been resolved in a published crystal structure. The few existing publications about this region suggest that it has several known and potential post-translational modification sites (12 predicted according to prosite.expasy.org), including phosphorylation of serine 81 and SUMOylation of lysine 111, which are important for the regulation of TDP1 intracellular activity.54–57 As our data demonstrate, there are substantial enzymatic differences with regard to both Km and Vmax of the truncated and full-length TDP1 enzymes. These catalytic differences as well as the specific interaction between the full-length TDP1 and recifin A suggest that a more complete enzymatic and structural characterization of the effect of modifications and protein-protein interactions in the regulatory region of TDP1 is warranted.
In conclusion, we have isolated and characterized recifin A, the first member of a new family of cysteine-rich peptides which we have named Tyr-lock peptides, from the genus Axinella, and which exhibit the ability to selectively inhibit the enzyme TDP1. Recifin A was determined to have a novel, non-ICK peptide disulfide-bonding pattern. The unusual disulfide linkage gives recifin A an unprecedented three-dimensional structure, with a central four-stranded β-sheet motif sandwiched by two helical turns. The disulfides create an embedded ring that encloses one of the strands, defining a new type of cystine knot peptide that also locks Tyr6 in three-dimensional space. Furthermore, recifin A was shown to modulate the activity of full-length human TDP1 enzyme but not an N-terminal truncated form of TDP1 (Δ147TDP1) lacking the regulatory domain. Further kinetic analysis confirmed recifin A to be an allosteric modulator of TDP1 suggesting that the N-terminal regulatory domain of TDP1 could be of possible pharmacological relevance for therapeutic inhibition of TDP1 to enhance sensitivity to anticancer topoisomerase I inhibitors.
EXPERIMENTAL SECTION
General Procedures.
All purification solvents were of HPLC and spectrophotometry grade. Mass spectrometry solvents were LC-MS grade and purchased from either Thermo Fisher or Burdick & Jackson. Mass spectrometry measurements were performed using an Accurate-Mass Q-TOF Dual-ESI 6530B instrument with an online 1260 Infinity binary HPLC system (Agilent Technologies, Inc., Santa Clara, CA), calibrated daily and operated with continual, internal calibration using reference mass ions at 121 and 1221 m/z. For MS, chromatographic separations were performed using linear gradients from 0 to 60% acetonitrile (0.1% v/v formic acid modified) at 1.00 mL/min on a Poroshell 300SB-C18, 5 μm, 2.1 mm × 75 mm column (Agilent Technologies, Inc., Santa Clara, CA) maintained at 40 °C. Source parameters for dual-ESI(+) were the following: capillary 4000 V, fragmentor 150–175 V, skimmer 65 V. Nitrogen flow was 12 L/min at 350 °C. High-resolution measurements (minimum of 20 000 resolution at 1521 m/z) were acquired in the range 100–3200 m/z at a scan rate of 1 spectra/s, and for MS/MS it was 50–3200 m/z at a scan rate of 3 spectra/s for both MS and MS/MS. Collision induced dissociation was accomplished using nitrogen gas and ramped collision energies (CE) calculated using the equation
Extraction and Isolation.
The sponge Axinella sp., (voucher ID 0CDN7410, NSC C020686) was harvested at a depth of 40 m at the Thunderbolt reef, south-southwest of Cape Recife Nature Reserve, Port Elizabeth, South Africa. A voucher specimen for this collection is maintained at the Smithsonian Institution (Suitland, MD). Aqueous extracts of Axinella sp. were provided by the Natural Products Branch of the National Cancer Institute and were prepared as previously reported.58 The dried extract was reconstituted in water at a concentration of 10 mg/mL and then subjected to vacuum-assisted chromatography using Bakerbond C4 wide-pore media (Mallinckrodt Baker, Inc., Phillipsburg, NJ). Compounds were eluted using a stepwise methanol gradient of five column volumes (CV) each of 100% water, 40% methanol, 60% methanol, and 100% methanol, and the resulting fractions were evaporated under vacuum and then lyophilized to dryness. A high-throughput biochemical assay for inhibition of TDP1 enzymatic activity was utilized to track fraction activity.28 Active fractions were subjected to RP-HPLC at room temperature, first using a Dynamax 300 Å, 5 μm, C4 column (Rainin, Woburn, MA), eluted with a 0–60% methanol gradient over 20 CV and then purified to homogeneity using a Vydac Protein&Peptide, 300 Å, 5 μm, C18 column (Grace Davison Discovery Science, Deerfield, IL), eluted with either a 0–60% methanol, 20 CV gradient or a 5–40%, 20 CV acetonitrile gradient. Purified peptides were lyophilized and stored at −20 °C.
Amino Acid Sequencing and Disulfide Assignments.
Purified recifin A was dissolved in 0.25 M Tris HCl, 1 mM EDTA, 6 M guanidine HCl, reduced at room temperature with 2-mercaptoethanol, and alkylated with 4-vinylpyridine according to standard techniques.59 The peptide was purified by RP-HPLC using a Vydac C18 column and eluted using an acetonitrile gradient as described above. Reduced and alkylated peptide was subjected to digestion with various proteases per manufacturer’s protocol. Roche Diagnostics, Indianapolis, IN: trypsin, chymotrypsin. Thermo Scientific, Rockford, IL: glu-c. Sigma-Aldrich, St. Louis, MO: proline specific endopeptidase. Clontech Takara Bio USA, Inc., Mountain View, CA: pfu-pyroglutamate aminopeptidase. Peptide fragments were sequenced by MS/MS CID or purified by RP-HPLC and sequenced by automated N-terminal Edman degradation on an Applied Biosystems 494 protein sequencer (Applied Biosystems, Foster City, CA) according to the manufacturer’s protocols. PEAKS software version 7.5 was used for de novo peptide sequencing (Bioinformatics Solutions, Inc., Waterloo, ON, Canada). Precursor mass error tolerances were set to 5 ppm, and fragment ion error tolerance was set to 0.1 Da.
Disulfide bonds were mapped using a partial reduction and sequential alkylation technique.31 A quantity of 1 nmol of recifin A (81 μM final concentration) was incubated in 0.1 M glycine HCl, pH 2.5, with 5, 10, 20, or 50 mM TCEP at 37 °C for 30 min. N-Ethylmaleimide, freshly prepared in acetonitrile, was added to the reaction to a final concentration of 250 mM and incubated at 37 °C for 15 min. Partially alkylated species were desalted and separated by RP-HPLC using a Vydac Protein & Peptide, 300 Å, 5 μm, C18 column at 40 °C, using a linear gradient of water with 0.05% (v/v) TFA to 50% acetonitrile with 0.05% (v/v) TFA. The partially reduced/alkylated species were either combined with 0.1 M Tris HCl, pH 8.0, 1 M urea and digested with chymotrypsin for 18 h at room temperature or fully reduced with 5 mM DTT, alkylated with 14 mM iodoacetamide, and digested with trypsin for 18 h at 37 °C. Fragments were sequenced by LC-MS/MS collision induced dissociation and PEAKS de novo sequencing software as described above. Intact disulfide-bridged peptides were analyzed by LC-MS and assigned using MassHunter qualitative analysis software with BioConfirm, version B.07.00 (Agilent Technologies, Inc., Santa Clara, CA). Input amino acid sequences of the disulfide isoforms were constructed with a fixed N-terminal pyroglutamic acid residue, and amino acid numbers 22 and 42 were fixed as N-ethylmaleimide alkylated cysteine residues.
NMR Spectroscopy and Structure Determination.
All spectra were acquired on a 600 MHz Bruker Avance III equipped with a cryogenically cooled probe (Bruker Biospin, Billerica, MA). An approximately 1 mg sample of recifin A was dissolved in 90% H2O/10% D2O at pH 4.85, and 1D 1H and 2D 1H–1H TOCSY (mixing time 80 ms) and 1H–1H NOESY (mixing time 200 ms) experiments were acquired at 298 K. In addition, a series of 1H–1H TOCSY experiments were acquired over 24 h directly after adding lyophilized recifin A to 100% D2O to investigate slow exchange of HN protons. This was followed by acquisition of 1H–13C HSQC and 1H–1H NOESY (200 ms mixing time) experiments in 100% D2O. TopSpin 3.5 (Bruker) was used to process the spectra, and the data were referenced to water at δH 4.76 ppm. Sequential assignments were completed using CCPNMR analysis 2.4.1 (CCPN, University of Cambridge, Cambridge, U.K.) and Xeasy.60 Distance restraints were derived from 1H–1H NOESY experiments acquired in 90% H2O/10% D2O and 100% D2O, and ϕ and ψ dihedral angle restraints were derived from chemical shifts from 1H–1H NOESY and 1H–13C HSQC experiments analyzed by the online version of TALOS-N61 to derive ϕ and ψ dihedral angle restraints. χ1 and χ2 dihedral restraints for Cys residues were derived from DISH,62 and additional χ1 dihedral restraints were derived from a combination of TALOS-N, patterns of NOE intensities, and preliminary structure calculations. Hydrogen bonds were introduced based on D2O exchange data or in the case of hydroxyl groups based on exchange behavior in the H2O sample and preliminary structure calculations. An initial 20 structures were calculated using the using automated assignments in Cyana.63,64 After manual assessment of the output all remaining NOEs could be unambiguously assigned. Structural refinement was carried out in a watershell using CNS65 where 50 structures were calculated and 20 representative structures selected based on MolProbity scores66 and energies. Root mean square deviations (RMSDs) were calculated using MOLMOL,67 and structural visualization was carried out using MOLMOL and PyMOL (the PyMOL Molecular Graphics System, version 1.7.4, Schrödinger, LLC). Recifin A structure has been deposited into the PDB68 (rcsb.org, code 6XN9), and NMR data have been deposited into the Biological Magnetic Resonance Bank69 (bmrb.wisc.edu, code 30767).
Biological Activity and Kinetics Assays.
TDP1 enzymatic activity inhibition assays using a radiolabeled oligonucleotide DNA substrate have been previously described.29 Briefly, serial dilutions of recifin A were incubated with 1 nM 5′–32P-labeled DNA oligos (P14Y:5′-[32P]-GATCTAAAAGACTT(3′-pTyr)-3′), 30 pM recombinant human TDP1 or 2 μg/mL of hTDP1 WCE which were collected from TDP1 knockout (TDP1−/−) DT40 cells complemented with human TDP1. The reactions were carried out in a final volume of 10 μL in 1× LMP 1 reaction buffer (50 mM Tris HCl, pH 7.5, 80 mM KCl, 2 mM EDTA, 1 mM DTT, 40 μg/mL BSA, 0.01% Tween 20) at room temperature for 15 min and terminated by adding 10 μL of 2× stop buffer (99.5% formamide, 10 mM EDTA, 0.01% methylene blue, 0.01% bromophenol blue). A 20% DNA sequencing gel was used to load the samples and exposed to a PhosphorImager screen for further analysis by Typhoon FLA 9500 (GE Healthcare).
FRET-based TDP1 enzymatic activity inhibition assays have been described in detail previously.28 Briefly, for Michaelis-Menten analysis, an eight-point FRET substrate concentration response was used (from 0.01–3 μM substrate) in the presence of 0, 0.2, 0.5, 1, and 2 μM recifin A. Quadruplicate reactions were set up in which a 1.25× concentration of either full-length TDP1 or Δ1–147TDP1 was diluted to 1× by the addition of a 6× solution of substrate and recifin A to reach a final concentration of 0.5 nM TDP1 (full length or truncated) and the indicated substrate and recifin A concentration in 1× phosphate buffered saline (PBS), pH 7.4, 80 mM potassium chloride, 1 mM TCEP, referred to as “1× TDP1 buffer”. After dilution these reactions were transferred to a black small volume 384-well plate (Greiner Bio-One, Monroe, NC). Fluorescence measurements (excitation 520 nM, emission 550 nm) were taken at 30 s intervals for 1 h using a i3x SpectraMax plate reader (Molecular Devices, Sunnyvale, CA). Reaction progression curves for each condition were examined for linearity over the time course, and the reaction rate for each condition was determined by linear regression using GraphPad Prism software (version 8.3.1, San Diego, CA). Reaction rates were replotted in terms of substrate concentration, and kinetic parameters for each recifin A treatment concentration were calculated by nonlinear regression (GraphPad Prism) according to the following equation:
For IC50 determinations, a 12-point concentration-response curve was prepared over a recifin A concentration range of 0–15 μM. This was accomplished by diluting a 5× stock solution of recifin A and TDP1 FRET substrate into a stock solution of 1.25× TDP1 buffer containing 0.625 nM full-length TDP1 or Δ147TDP1, bringing the final concentration to 1× TDP1 buffer, 0.5 nM enzyme (or a no enzyme control), 1 μM FRET substrate, and 0–15 μM recifin A. Reactions were set up in triplicate using the same plates and plate reader described above for the kinetic measurements. Reaction wells were read at 0 (T0) and 15 (T15) min after initiation. The T15 data were background corrected by subtracting T0 fluorescence measurements. Corrected data were normalized to a control with no enzyme present (0% activity) and a vehicle control (100% activity). Recifin A concentrations were converted to log10 values, normalized data were fitted to the following equation by nonlinear regression (least-squares fit with a variable slope), and an IC50 value was calculated using GraphPad Prism software:
Supplementary Material
ACKNOWLEDGMENTS
We thank Dr. Wendy Henry, Molecular Targets Program, for initial evaluation of the Axinella sp. extract and the NCI Natural Products Branch for supply and resupply of the Axinella sp. extract. This project has been funded in whole or in part with federal funds from the Frederick National Laboratory for Cancer Research and the National Cancer Institute, National Institutes of Health, under Contract HHSN26120080001E. The content of this publication does not necessarily reflect the views or policies of the Department of Health and Human Services, nor does mention of trade names, commercial products, or organizations imply endorsement by the U.S. Government. This research was supported [in part] by the Intramural Research Program of the NIH, Frederick National Lab, and the National Cancer Institute, Center for Cancer Research (Grants Z01-BC 006150 and BC 006161). C.I.S. was supported by an Australian Research Council Future Fellowship (Grant FT160100055).
Footnotes
Notes
The authors declare the following competing financial interest(s): A patent on the protein recifin has been filed by the United States Government. All rights are owned by the U.S. Government.
ASSOCIATED CONTENT
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.0c10418.
LC-MS and biological activity data for recifin A and other family members, reduced/alkylated recifin A endoprotease digest peptides tandem mass spectra and de novo sequencing results, sequence alignment of recifin A with CRPs and ICK peptides, NMR spectra and chemical shifts, Km and Vmax ratios for recifin A-TDP1 interaction, partially reduced and NEM alkylated recifin A chymotrypsin digest observed peptides (table) (PDF)
Contributor Information
Lauren R. H. Krumpe, Basic Science Program, Leidos Biomedical Research, Inc., Frederick National Laboratory for Cancer Research, Frederick, Maryland 21702, United States; Molecular Targets Program, Center for Cancer Research, NCI-Frederick, NIH, Frederick, Maryland 21702, United States
Brice A. P. Wilson, Molecular Targets Program, Center for Cancer Research, NCI-Frederick, NIH, Frederick, Maryland 21702, United States
Christophe Marchand, Developmental Therapeutics Branch, Laboratory of Molecular Pharmacology, NCI, NIH, Bethesda, Maryland 20892, United States.
Suthananda N. Sunassee, Molecular Targets Program, Center for Cancer Research, NCI-Frederick, NIH, Frederick, Maryland 21702, United States
Alun Bermingham, Molecular Targets Program, Center for Cancer Research, NCI-Frederick, NIH, Frederick, Maryland 21702, United States.
Wenjie Wang, Developmental Therapeutics Branch, Laboratory of Molecular Pharmacology, NCI, NIH, Bethesda, Maryland 20892, United States.
Edmund Price, Molecular Targets Program, Center for Cancer Research, NCI-Frederick, NIH, Frederick, Maryland 21702, United States.
Tad Guszczynski, Molecular Targets Program, Center for Cancer Research, NCI-Frederick, NIH, Frederick, Maryland 21702, United States.
James A. Kelley, Chemical Biology Laboratory, Center for Cancer Research, NCI-Frederick, NIH, Frederick, Maryland 21702, United States
Kirk R. Gustafson, Molecular Targets Program, Center for Cancer Research, NCI-Frederick, NIH, Frederick, Maryland 21702, United States
Yves Pommier, Developmental Therapeutics Branch, Laboratory of Molecular Pharmacology, NCI, NIH, Bethesda, Maryland 20892, United States.
K. Johan Rosengren, School of Biomedical Sciences, The University of Queensland, Brisbane, QLD 4072, Australia.
Christina I. Schroeder, Institute for Molecular Bioscience, The University of Queensland, Brisbane, QLD 4072, Australia; Chemical Biology Laboratory, Center for Cancer Research, NCI-Frederick, NIH, Frederick, Maryland 21702, United States
Barry R. O’Keefe, Molecular Targets Program, Center for Cancer Research, NCI-Frederick, NIH, Frederick, Maryland 21702, United States; Natural Products Branch, Developmental Therapeutics Program, Division of Cancer Treatment and Diagnosis, National Cancer Institute, Frederick, Maryland 21702, United States
REFERENCES
- (1).Capranico G; Marinello J; Chillemi G Type I DNA Topoisomerases. J. Med. Chem. 2017, 60 (6), 2169–2192. [DOI] [PubMed] [Google Scholar]
- (2).Pommier Y; Sun Y; Huang SN; Nitiss JL Roles of eukaryotic topoisomerases in transcription, replication and genomic stability. Nat. Rev. Mol. Cell Biol. 2016, 17 (11), 703–721. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (3).Wall ME; Wani MC; Cook CE; Palmer KH; Mcphail AT; Sim GA Plant Antitumor Agents. I. Isolation and Structure of Camptothecin a Novel Alkaloidal Leukemia and Tumor Inhibitor from Camptotheca Acuminata. J. Am. Chem. Soc. 1966, 88 (16), 3888–3890. [Google Scholar]
- (4).Hsiang YH; Hertzberg R; Hecht S; Liu LF Camptothecin induces protein-linked DNA breaks via mammalian DNA topoisomerase I. J. Biol. Chem. 1985, 260 (27), 14873–8. [PubMed] [Google Scholar]
- (5).Flett FJ; Ruksenaite E; Armstrong LA; Bharati S; Carloni R; Morris ER; Mackay CL; Interthal H; Richardson JM Structural basis for DNA 3′-end processing by human tyrosyl-DNA phosphodiesterase 1. Nat. Commun. 2018, 9 (1), 24. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (6).Pommier Y; Huang SY; Gao R; Das BB; Murai J; Marchand C Tyrosyl-DNA-phosphodiesterases (TDP1 and TDP2). DNA Repair 2014, 19, 114–29. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (7).Dexheimer TS; Antony S; Marchand C; Pommier Y Tyrosyl-DNA phosphodiesterase as a target for anticancer therapy. Anti-Cancer Agents Med. Chem. 2008, 8 (4), 381–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (8).Yang SW; Burgin AB Jr.; Huizenga BN; Robertson CA; Yao KC; Nash HA A eukaryotic enzyme that can disjoin dead-end covalent complexes between DNA and type I topoisomerases. Proc. Natl. Acad. Sci. U. S. A. 1996, 93 (21), 11534–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (9).Zakharenko A; Dyrkheeva N; Lavrik O Dual DNA topoisomerase 1 and tyrosyl-DNA phosphodiesterase 1 inhibition for improved anticancer activity. Med. Res. Rev. 2019, 39 (4), 1427–1441. [DOI] [PubMed] [Google Scholar]
- (10).Sable R; Parajuli P; Jois S Peptides, Peptidomimetics, and Polypeptides from Marine Sources: A Wealth of Natural Sources for Pharmaceutical Applications. Mar. Drugs 2017, 15 (4), 124. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (11).Cheung RC; Wong JH; Pan W; Chan YS; Yin C; Dan X; Ng TB Marine lectins and their medicinal applications. Appl. Microbiol. Biotechnol. 2015, 99 (9), 3755–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (12).Takada K; Hamada T; Hirota H; Nakao Y; Matsunaga S; van Soest RW; Fusetani N Asteropine A, a sialidase-inhibiting conotoxin-like peptide from the marine sponge Asteropus simplex. Chem. Biol. 2006, 13 (6), 569–74. [DOI] [PubMed] [Google Scholar]
- (13).Li H; Bowling JJ; Su M; Hong J; Lee BJ; Hamann MT; Jung JH Asteropsins B-D, sponge-derived knottins with potential utility as a novel scaffold for oral peptide drugs. Biochim. Biophys. Acta, Gen. Subj 2014, 1840 (3), 977–84. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (14).Li H; Bowling JJ; Fronczek FR; Hong J; Jabba SV; Murray TF; Ha NC; Hamann MT; Jung JH Asteropsin A: an unusual cystine-crosslinked peptide from porifera enhances neuronal Ca2+ influx. Biochim. Biophys. Acta, Gen. Subj 2013, 1830 (3), 2591–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (15).Su M; Li H; Wang H; Kim E; Kim HS; Kim EH; Lee J; Jung JH Stable and biocompatible cystine knot peptides from the marine sponge Asteropus sp. Bioorg. Med. Chem. 2016, 24 (13), 2979–2987. [DOI] [PubMed] [Google Scholar]
- (16).Williams DE; Austin P; Diaz-Marrero AR; Soest RV; Matainaho T; Roskelley CD; Roberge M; Andersen RJ Neopetrosiamides, peptides from the marine sponge Neopetrosia sp. that inhibit amoeboid invasion by human tumor cells. Org. Lett. 2005, 7 (19), 4173–6. [DOI] [PubMed] [Google Scholar]
- (17).Liu H; Boudreau MA; Zheng J; Whittal RM; Austin P; Roskelley CD; Roberge M; Andersen RJ; Vederas JC Chemical synthesis and biological activity of the neopetrosiamides and their analogues: revision of disulfide bond connectivity. J. Am. Chem. Soc. 2010, 132 (5), 1486–7. [DOI] [PubMed] [Google Scholar]
- (18).Carstens BB; Rosengren KJ; Gunasekera S; Schempp S; Bohlin L; Dahlstrom M; Clark RJ; Goransson U Isolation, Characterization, and Synthesis of the Barrettides: Disulfide-Containing Peptides from the Marine Sponge Geodia barretti. J. Nat. Prod. 2015, 78 (8), 1886–93. [DOI] [PubMed] [Google Scholar]
- (19).Milanowski DJ; Rashid MA; Gustafson KR; O’Keefe BR; Nawrocki JP; Pannell LK; Boyd MR Cyclonellin, a new cyclic octapeptide from the marine sponge Axinella carteri. J. Nat. Prod. 2004, 67 (3), 441–4. [DOI] [PubMed] [Google Scholar]
- (20).Urban S; de Almeida Leone P; Carroll AR; Fechner GA; Smith J; Hooper JN; Quinn RJ Axinellamines A-D, Novel Imidazo-Azolo-Imidazole Alkaloids from the Australian Marine Sponge Axinella sp. J. Org. Chem. 1999, 64 (3), 731–735. [DOI] [PubMed] [Google Scholar]
- (21).Supriyono A; Schwarz B; Wray V; Witte L; Muller WE; van Soest R; Sumaryono W; Proksch P Bioactive alkaloids from the tropical marine sponge Axinella carteri. Z. Naturforsch., C: J. Biosci 1995, 50 (9–10), 669–74. [DOI] [PubMed] [Google Scholar]
- (22).Pettit GR; Herald CL; Boyd MR; Leet JE; Dufresne C; Doubek DL; Schmidt JM; Cerny RL; Hooper JN; Rutzler KC Isolation and structure of the cell growth inhibitory constituents from the western Pacific marine sponge Axinella sp. J. Med. Chem. 1991, 34 (11), 3339–40. [DOI] [PubMed] [Google Scholar]
- (23).Gold ER; Phelps CF; Khalap S; Balding P Observations on Axinella sp. hemagglutinin. Ann. N. Y. Acad. Sci. 1974, 234 (0), 122–8. [DOI] [PubMed] [Google Scholar]
- (24).Bretting H Serological and immunoelectrophoretical investigations of the haemagglutinins of the two sponges Aaptos papillata (Keller) and Axinella polypoides (Schmidt) (author’s transl). Z. Immunitatsforsch. Exp. Klin. Immunol. 1973, 146 (3), 239–59. [PubMed] [Google Scholar]
- (25).Rudi A; Goldberg I; Stein Z; Kashman Y; Benayahu Y; Schleyer M; Garcia Gravalos MD Sodwanones G, H, and I, new cytotoxic triterpenes from a marine sponge. J. Nat. Prod. 1995, 58 (11), 1702–12. [DOI] [PubMed] [Google Scholar]
- (26).Warabi K; Hamada T; Nakao Y; Matsunaga S; Hirota H; van Soest RW; Fusetani N Axinelloside A, an unprecedented highly sulfated lipopolysaccharide inhibiting telomerase, from the marine sponge, Axinella infundibula. J. Am. Chem. Soc. 2005, 127 (38), 13262–70. [DOI] [PubMed] [Google Scholar]
- (27).Costantino V; D’Esposito M; Fattorusso E; Mangoni A; Basilico N; Parapini S; Taramelli D Damicoside from Axinella damicornis: the influence of a glycosylated galactose 4-OH group on the immunostimulatory activity of alpha-galactoglycosphingolipids. J. Med. Chem. 2005, 48 (23), 7411–7. [DOI] [PubMed] [Google Scholar]
- (28).Bermingham A; Price E; Marchand C; Chergui A; Naumova A; Whitson EL; Krumpe LRH; Goncharova EI; Evans JR; McKee TC; Henrich CJ; Pommier Y; O’Keefe BR Identification of Natural Products That Inhibit the Catalytic Function of Human Tyrosyl-DNA Phosphodiesterase (TDP1). SLAS Discov 2017, 22, 1093–1105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (29).Marchand C; Huang SY; Dexheimer TS; Lea WA; Mott BT; Chergui A; Naumova A; Stephen AG; Rosenthal AS; Rai G; Murai J; Gao R; Maloney DJ; Jadhav A; Jorgensen WL; Simeonov A; Pommier Y Biochemical assays for the discovery of TDP1 inhibitors. Mol. Cancer Ther. 2014, 13 (8), 2116–26. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (30).Zhang J; Xin L; Shan B; Chen W; Xie M; Yuen D; Zhang W; Zhang Z; Lajoie GA; Ma B PEAKS DB: de novo sequencing assisted database search for sensitive and accurate peptide identification. Mol. Cell. Proteomics 2012, 11 (4), M111.010587. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (31).Gray WR Disulfide structures of highly bridged peptides: a new strategy for analysis. Protein Sci. 1993, 2 (10), 1732–48. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (32).Pallaghy PK; Nielsen KJ; Craik DJ; Norton RS A common structural motif incorporating a cystine knot and a triple-stranded beta-sheet in toxic and inhibitory polypeptides. Protein Sci. 1994, 3 (10), 1833–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (33).Postic G; Gracy J; Perin C; Chiche L; Gelly JC KNOTTIN: the database of inhibitor cystine knot scaffold after 10 years, toward a systematic structure modeling. Nucleic Acids Res. 2018, 46 (D1), D454–D458. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (34).Schroeder CI; Rosengren KJ Three-Dimensional Structure Determination of Peptides Using Solution Nuclear Magnetic Resonance Spectroscopy. Methods Mol. Biol. 2020, 2068, 129–162. [DOI] [PubMed] [Google Scholar]
- (35).Wishart DS; Bigam CG; Holm A; Hodges RS; Sykes BD 1H, 13C and 15N random coil NMR chemical shifts of the common amino acids. I. Investigations of nearest-neighbor effects. J. Biomol. NMR 1995, 5 (1), 67–81. [DOI] [PubMed] [Google Scholar]
- (36).Brunger AT; Adams PD; Clore GM; DeLano WL; Gros P; Grosse-Kunstleve RW; Jiang JS; Kuszewski J; Nilges M; Pannu NS; Read RJ; Rice LM; Simonson T; Warren GL Crystallography & NMR system: A new software suite for macromolecular structure determination. Acta Crystallogr., Sect. D: Biol. Crystallogr. 1998, 54 (Part 5), 905–21. [DOI] [PubMed] [Google Scholar]
- (37).Daly NL; Craik DJ Bioactive cystine knot proteins. Curr. Opin. Chem. Biol. 2011, 15 (3), 362–8. [DOI] [PubMed] [Google Scholar]
- (38).Craik DJ; Du J Cyclotides as drug design scaffolds. Curr. Opin. Chem. Biol. 2017, 38, 8–16. [DOI] [PubMed] [Google Scholar]
- (39).Maksimov MO; Pan SJ; James Link A Lasso peptides: structure, function, biosynthesis, and engineering. Nat. Prod. Rep. 2012, 29 (9), 996–1006. [DOI] [PubMed] [Google Scholar]
- (40).Rosengren KJ; Clark RJ; Daly NL; Goransson U; Jones A; Craik DJ Microcin J25 has a threaded sidechain-to-backbone ring structure and not a head-to-tail cyclized backbone. J. Am. Chem. Soc. 2003, 125 (41), 12464–74. [DOI] [PubMed] [Google Scholar]
- (41).Rosengren KJ; Daly NL; Plan MR; Waine C; Craik DJ Twists, knots, and rings in proteins. Structural definition of the cyclotide framework. J. Biol. Chem. 2003, 278 (10), 8606–16. [DOI] [PubMed] [Google Scholar]
- (42).Talavera D; Robertson DL; Lovell SC Characterization of protein-protein interaction interfaces from a single species. PLoS One 2011, 6 (6), e21053. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (43).Jayashree S; Murugavel P; Sowdhamini R; Srinivasan N Interface residues of transient protein-protein complexes have extensive intra-protein interactions apart from inter-protein interactions. Biol. Direct 2019, 14 (1), 1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (44).Huang SN; Pommier Y; Marchand C Tyrosyl-DNA Phosphodiesterase 1 (Tdp1) inhibitors. Expert Opin. Ther. Pat. 2011, 21 (9), 1285–92. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (45).Segel IH Enzyme Kinetics: Behavior and Analysis of Rapid Equilibrium and Steady State Enzyme Systems; Wiley: New York, 1975; p xxii, 957 pp. [Google Scholar]
- (46).Henage LG; Exton JH; Brown HA Kinetic analysis of a mammalian phospholipase D: allosteric modulation by monomeric GTPases, protein kinase C, and polyphosphoinositides. J. Biol. Chem. 2006, 281 (6), 3408–17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (47).Davies DR; Interthal H; Champoux JJ; Hol WG The crystal structure of human tyrosyl-DNA phosphodiesterase, Tdp1. Structure 2002, 10 (2), 237–48. [DOI] [PubMed] [Google Scholar]
- (48).Interthal H; Pouliot JJ; Champoux JJ The tyrosyl-DNA phosphodiesterase Tdp1 is a member of the phospholipase D superfamily. Proc. Natl. Acad. Sci. U. S. A. 2001, 98 (21), 12009–14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (49).Lountos GT; Zhao XZ; Kiselev E; Tropea JE; Needle D; Pommier Y; Burke TR; Waugh DS Identification of a ligand binding hot spot and structural motifs replicating aspects of tyrosyl-DNA phosphodiesterase I (TDP1) phosphoryl recognition by crystallographic fragment cocktail screening. Nucleic Acids Res. 2019, 47 (19), 10134–10150. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (50).Pawson T; Nash P Protein-protein interactions define specificity in signal transduction. Genes Dev. 2000, 14 (9), 1027–47. [PubMed] [Google Scholar]
- (51).Haendeler J; Hoffmann J; Rahman S; Zeiher AM; Dimmeler S Regulation of telomerase activity and anti-apoptotic function by protein-protein interaction and phosphorylation. FEBS Lett. 2003, 536 (1–3), 180–6. [DOI] [PubMed] [Google Scholar]
- (52).Moscat J; Diaz-Meco MT; Wooten MW Signal integration and diversification through the p62 scaffold protein. Trends Biochem. Sci. 2007, 32 (2), 95–100. [DOI] [PubMed] [Google Scholar]
- (53).Grimsby J; Berthel SJ; Sarabu R Glucokinase activators for the potential treatment of type 2 diabetes. Curr. Top. Med. Chem. 2008, 8 (17), 1524–32. [DOI] [PubMed] [Google Scholar]
- (54).Das BB; Antony S; Gupta S; Dexheimer TS; Redon CE; Garfield S; Shiloh Y; Pommier Y Optimal function of the DNA repair enzyme TDP1 requires its phosphorylation by ATM and/or DNA-PK. EMBO J. 2009, 28 (23), 3667–80. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (55).Chiang SC; Carroll J; El-Khamisy SF TDP1 serine 81 promotes interaction with DNA ligase IIIalpha and facilitates cell survival following DNA damage. Cell Cycle 2010, 9 (3), 588–595. [DOI] [PubMed] [Google Scholar]
- (56).Das BB; Huang SY; Murai J; Rehman I; Ame JC; Sengupta S; Das SK; Majumdar P; Zhang H; Biard D; Majumder HK; Schreiber V; Pommier Y PARP1-TDP1 coupling for the repair of topoisomerase I-induced DNA damage. Nucleic Acids Res. 2014, 42 (7), 4435–49. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (57).Hudson JJ; Chiang SC; Wells OS; Rookyard C; El-Khamisy SF SUMO modification of the neuroprotective protein TDP1 facilitates chromosomal single-strand break repair. Nat. Commun. 2012, 3, 733. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (58).McCloud TG High throughput extraction of plant, marine and fungal specimens for preservation of biologically active molecules. Molecules 2010, 15 (7), 4526–63. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (59).Crimmins DL; Mische SM; Denslow ND Chemical cleavage of proteins in solution. Current Protocols in Protein Science; Wiley, 2005; Chapter 11, Unit 11.4. [DOI] [PubMed] [Google Scholar]
- (60).Bartels C; Xia TH; Billeter M; Guntert P; Wuthrich K The program XEASY for computer-supported NMR spectral analysis of biological macromolecules. J. Biomol. NMR 1995, 6 (1), 1–10. [DOI] [PubMed] [Google Scholar]
- (61).Shen Y; Bax A Protein backbone and sidechain torsion angles predicted from NMR chemical shifts using artificial neural networks. J. Biomol. NMR 2013, 56 (3), 227–41. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (62).Armstrong DA; Kaas Q; Rosengren KJ Prediction of disulfide dihedral angles using chemical shifts. Chem. Sci. 2018, 9 (31), 6548–6556. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (63).Guntert P Automated NMR structure calculation with CYANA. Methods Mol. Biol. 2004, 278, 353–78. [DOI] [PubMed] [Google Scholar]
- (64).Guntert P; Mumenthaler C; Wuthrich K Torsion angle dynamics for NMR structure calculation with the new program DYANA. J. Mol. Biol. 1997, 273 (1), 283–98. [DOI] [PubMed] [Google Scholar]
- (65).Linge JP; Williams MA; Spronk CA; Bonvin AM; Nilges M Refinement of protein structures in explicit solvent. Proteins: Struct., Funct., Genet 2003, 50 (3), 496–506. [DOI] [PubMed] [Google Scholar]
- (66).Chen VB; Arendall WB 3rd; Headd JJ; Keedy DA; Immormino RM; Kapral GJ; Murray LW; Richardson JS; Richardson DC MolProbity: all-atom structure validation for macromolecular crystallography. Acta Crystallogr., Sect. D: Biol. Crystallogr. 2010, 66 (Part 1), 12–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (67).Koradi R; Billeter M; Wuthrich K MOLMOL: a program for display and analysis of macromolecular structures. J. Mol. Graphics 1996, 14 (1), 51–5. 29–32 [DOI] [PubMed] [Google Scholar]
- (68).Berman HM; Westbrook J; Feng Z; Gilliland G; Bhat TN; Weissig H; Shindyalov IN; Bourne PE The Protein Data Bank. Nucleic Acids Res. 2000, 28 (1), 235–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (69).Ulrich EL; Akutsu H; Doreleijers JF; Harano Y; Ioannidis YE; Lin J; Livny M; Mading S; Maziuk D; Miller Z; Nakatani E; Schulte CF; Tolmie DE; Kent Wenger R; Yao H; Markley JL BioMagResBank. Nucleic Acids Res. 2008, 36 (Database), D402–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.