Abstract
Together with protein tyrosine kinases, protein tyrosine phosphatases (PTPs) control protein tyrosine phosphorylation and regulate numerous cellular functions. Dysregulated PTP activity is associated with the onset of multiple human diseases. Nevertheless, understanding of the physiological function and disease biology of most PTPs remains limited, largely due to the lack of PTP-specific chemical probes. In this study, starting from a well-known nonhydrolyzable phosphotyrosine (pTyr) mimetic, phosphonodifluoromethyl phenylalanine (F2Pmp), we synthesized 7 novel phosphonodifluoromethyl-containing bicyclic/tricyclic aryl derivatives with improved cell permeability and potency toward various PTPs. Furthermore, with fragment- and structure-based design strategies, we advanced compound 9 to compound 15, a first-in-class, potent, selective, and bioavailable inhibitor of human CDC14A and B phosphatases. This study demonstrates the applicability of the fragment-based design strategy in creating potent, selective, and bioavailable PTP inhibitors and provides a valuable probe for interrogating the biological roles of hCDC14 phosphatases and assessing their potential for therapeutic interventions.
Introduction
Protein tyrosine phosphorylation is a critical component of cellular signaling that is tightly controlled by the opposing actions of protein tyrosine kinases (PTKs) and protein tyrosine phosphatases (PTPs) and is essential for the regulation of a variety of cellular functions1. The roles of several PTPs, such as PTP1B2–4, TC-PTP3–7, SHP28,9, and LYP10–12, have been clearly established in the etiology of several human diseases, highlighting the therapeutic potential of these PTPs. However, there are over 100 PTPs in the human genome13 and the roles of the vast majority of PTPs in the development of human diseases are still unexplored or incompletely understood. Although small molecule inhibitors are valuable chemical tools for interrogating the role of specific PTPs in the pathogenesis and validating their therapeutic potential under a more clinically relevant setting, the development of potent, selective, and bioavailable PTP inhibitors is a difficult endeavor. Traditionally, targeting the phosphotyrosine (pTyr) binding pocket (i.e. the PTP active site) with nonhydrolyzable pTyr mimetics is the most straight forward strategy to realize PTP inhibition. However, due to the positively charged and highly conserved nature of PTP active site14–16, most nonhydrolyzable pTyr mimetics bear at least one negative charge and can indiscriminately bind to multiple PTPs, which hinders their development into potent, selective, and bioavailable inhibitors for the PTP of interest. Although allosteric regulatory mechanisms can be harnessed to bypass the aforementioned challenges for some PTPs17–20, the uniqueness of each allosteric mechanism also limits broad application of allosteric inhibition to all PTPs. Moreover, many PTPs do not possess known allosteric pockets, and full-length structures are not available to facilitate detection of new allosteric sites. Therefore, a more general method is still warranted for the development of potent, selective and bioavailable inhibitors for in-depth interrogation of PTP biology and therapeutic translation.
Among the “untapped” PTPs, the cell division cycle 14 (CDC14) phosphatases are highly conserved members of the dual-specificity phosphatases (DUSPs) within the PTP superfamily that exist in most eukaryotes. Although DUSPs generally dephosphorylate both phosphotyrosine and phosphoserine (pSer)/phosphothreonine (pThr) with the same active site and employ a catalytic mechanism identical to that of pTyr-specific PTPs21, the CDC14 phosphatases evolved a unique substrate specificity. An ancient duplication of the DUSP fold, a signature feature of CDC14 enzymes22 created a novel peptide binding groove at the domain interface adjacent to the catalytic pocket23,24. This interface renders the catalytic pocket highly selective for protein substrates containing pSer followed by the docking motif Pro-X-Lys/Arg25. For example, the catalytic efficiency of CDC14 enzymes is typically >1,000-fold higher towards pSer-containing peptide substrates compared to identical pThr-containing sequences and >100-fold higher than identical pTyr-containing sequences22,25. Consistent with this, most characterized biological targets of CDC14 are substrates of cyclin-dependent kinases, which share recognition of the Ser-Pro-X-Lys/Arg motif. We recently suggested that the unique active site selectivity and nearby binding pockets for the +1 Pro and +3 Lys/Arg may allow highly selective inhibitor development towards the CDC14 phosphatases22.
The human genome encodes two widely expressed CDC14 homologs (hCDC14A and B)26 and a hominid-specific hCDC14C that is nearly identical to hCDC14B and expressed only in testis and brain27. The biological function of hCDC14A/B and their connections with the etiology of diseases are an active area of investigation. While early studies with gene overexpression or RNAi knockdown suggested that both enzymes regulate essential processes in mitosis such as spindle formation28–30, centrosome duplication31 and separation29, chromosome segregation23, nuclear organization32, mitotic exit33, and cytokinesis23, more recent work with double knockout mice and human cell lines indicate that hCDC14A and B are completely dispensable for cell division34. Instead, hCDC14A and B are essential for differentiation of neuronal stem cells during mouse embryogenesis35. In addition, hCDC14A, in particular, appears to play important roles in ciliogenesis and cilia maintenance, and human families with hCDC14A mutations suffer from hearing loss and male infertility associated with loss of ciliated cells in the inner ear and reproductive tract36. The connection between hCDC14A/B and cancer is complicated. Existing studies suggest that both hCDC14A and hCDC14B are potential oncoproteins, for example promoting tumorigenesis via dephosphorylation of p53 at Ser315, which disrupts p53-Pin1 and p53-E2F1 interactions, reducing p53 stability and transactivation activity.37–42 On the other hand, hCDC14A was reported to restrain cell migration associated with malignant tumor growth43 and reduced hCDC14A and B expression is correlated with poor prognosis in certain cancers44,45. Additionally, genetic knockout of hCDC14A or B also impaired the DNA damage response in certain tumor cells and reduced their viability after ionizing radiation46–48, which indicated that both enzymes can be utilized as therapeutic targets that sensitize tumors to radiotherapy and overcome the radiation resistance49,50. Potent and selective hCDC14A/B inhibitors would be invaluable chemical probes to help clarify contributions of hCDC14A/B to cancer and other physiological processes and test various therapeutic hypotheses.
To offer a general approach that can overcome the challenges associated with orthosteric PTP inhibitor development, our group previously advanced a fragment-based PTP inhibitor design strategy1,51. It is well known that pTyr plays an essential role in PTP-substrate recognition and the amino acids flanking the pTyr residue increase the binding affinity and selectivity of pTyr-containing peptides52–56. Accordingly, we hypothesized that molecules engaging both the conserved PTP active site and nearby distinct peripheral pockets, which can be generated by tethering a nonhydrolyzable pTyr mimetic with appropriate diversity fragments, will possess enhanced potency and selectivity due to the additivity of free energy of binding and the unique interaction involved with the less conserved peripheral sites1,51. Additionally, owing to the uncharged nature of most peripheral pockets, many of the corresponding binding fragments are neutral and better at balancing inhibitor polarity and lipophilicity. With this strategy, a number of potent, selective, and bioavailable orthosteric PTP inhibitors have been developed10,57–63 and recently, a PTP1B/TC-PTP dual targeting active site-directed competitive inhibitor ABBV-CLS-484 has successfully proceeded to clinical trials64. Although peripheral site ligands undoubtedly play a crucial role in augmenting the potency, selectivity, and cell permeability of PTP inhibitors, the pTyr mimetic remains the cornerstone in the fragment-based design of competitive PTP inhibitors. Therefore, a pTyr mimetic possessing superior PTP binding affinity and cell permeability can greatly facilitate the development of potent, selective, and drug-like inhibitors for a variety of therapeutically relevant PTPs.
Among the reported nonhydrolyzable pTyr mimetics, phosphonodifluoromethyl phenylalanine (F2Pmp)65 is widely utilized in PTP inhibitor development1. By virtue of its structural resemblance to pTyr, F2Pmp retains the ability to bind the PTP active sites66. The amino acid-like structure of F2Pmp also facilitates its further structural expansion and decoration. However, this structural characteristic also imposes restrictions on feasible approaches for its subsequent elaboration (i.e. limited to peptide incorporation). Moreover, the difluorophosphonic acid within F2Pmp is di-ionized at physiological pH, which generally causes poor cell membrane penetration and bioavailability. Lastly, the structural simplicity of the difluorobenzyl moiety of F2Pmp also limits its interactions with the PTP active site, leading to relatively low PTP binding affinity56. Due to these limitations, most of the existing F2Pmp-based potent, selective, and cell-active PTP inhibitors are peptide-like molecules with high molecular weight.59,60,66–68. As a result, the ligand efficiency and metabolic stability of these inhibitors were usually compromised. To create high affinity and cell permeable pTyr mimetics for further PTP inhibitor development, we designed and synthesized 7 novel phosphonodifluoromethyl-containing bicyclic/tricyclic aryl derivatives that were more cell permeable and exhibited significantly improved binding affinity toward one or multiple screened PTPs. Using fragment- and structure-based design, we developed the first potent and selective hCDC14A/B inhibitor 15, which interacts with both the active site and the nearby P+1 Pro binding pocket. In addition, compound 15 blocked hCDC14A/B mediated p53 phosphorylation and stabilized p53 protein in U2OS cells. Importantly, compound 15 exhibited good pharmacokinetic properties and was also orally bioavailable. Taken together, our study demonstrated that potent, selective, and bioavailable active site-directed PTP inhibitors can be obtained using a fragment-based design approach. The yielded hCDC14A/B inhibitor can serve as a chemical probe for further interrogation of the biological function and assessment of the therapeutic potential of these two enzymes. Furthermore, the reported novel pTyr mimetics can also be applied to the design of inhibitors for other therapeutically relevant PTPs.
Results and Discussion
Design, Synthesis, and Characterization of Novel Phosphonodifluoromethyl-Containing pTyr Mimetics
As depicted in Figure 1A, our work started by simplifying the F2Pmp or Ac-F2Pmp-OMe (compound 1, which better mimics the F2Pmp moiety in most reported PTP inhibitors59,60,67–69) structure to P-(Difluorophenylmethyl) phosphonic acid (PhFP, compound 2), the segment crucial for the interaction between PTP active site and F2Pmp. This simplification was driven by the recognition that the amino acid-like structure present in F2Pmp constrains further elaboration and offers only a limited contribution to binding affinity. To further enhance PTP binding while reducing the polarity of PhFP, we retained the essential difluorophosphonic acid moiety within PhFP while replacing the benzene ring with bicyclic or tricyclic aromatic rings. This strategy is supported by earlier studies indicating that PTPs have a greater affinity for bicyclic/tricyclic aryl phosphate substrates and nonhydrolyzable pTyr mimetics compared to their single-ring counterparts70–72. The basis for this approach is also grounded in the idea that the more lipophilic fused aromatic ring system can potentially better offset the pronounced polarity of the difluorophosphonic acid moiety and thereby enhance cell permeability.
Figure 1.

Design and synthesis of the novel phosphonodifluoromethyl-containing pTyr mimetics. (A) Design strategy for the novel pTyr Mimetics. (B) Structure of the 7 novel pTyr mimetics.
As shown in Figure 1B, 7 novel pTyr mimetics were synthesized with the methods described in the Chemistry section. Upon obtaining the 7 new pTyr mimetics, we set out to determine their PTP inhibitory activity. Since self-aggregation of small organic molecules can lead to the formation of colloidal aggregates and cause 85–95% of the false positives in enzymatic assays73, we began by screening all of the pTyr mimetics with a previously described dynamic light scattering (DLS)-based small molecule aggregation detection protocol74 to exclude small molecule aggregators, if any. All new pTyr mimetics were dissolved in 3,3-Dimethylglutaric acid (DMG) buffer (pH 7.0) (same as our enzymatic inhibition assay) at 1 mM concentration. As shown in Figure S1, no DLS signal decay was observed in any solution, indicating that no particle with >1.0 nm diameter was detected in these solutions, thus none of the compounds can form self-aggregates. In contrast, a known aggregator (Ketoconazole) gave strong DLS signal decay at concentrations as low as 15 μM, agreeing with previous reports using various detection methods75. We next profiled the inhibitory activity of PhFP and all 7 new pTyr mimetics against a panel of 10 major mammalian PTPs with 6,8-Difluoro-4-Methylumbelliferyl Phosphate (DiFMUP) as substrate. As shown in Table 1, PhFP exhibited low millimolar half-inhibitory concentration values (IC50s) toward most tested PTPs (except CD45, for which PhFP showed no inhibition at 10 mM concentration). Moreover, no significant inhibition selectivity was observed with this compound. In contrast, all of the new bicyclic or tricyclic pTyr mimetics (compounds 3-9) exhibited low- to mid- micromolar IC50s for one or more PTPs with initial selectivity over other tested PTPs. These findings agreed with previous studies showing PTPs prefer bicyclic/tricyclic aryl phosphate substrates and pTyr mimetics compared to their single-ring counterparts70–72. Interestingly, several of these novel pTyr mimetics demonstrated distinct patterns of inhibition selectivity across specific PTPs. For instance, compounds 3, 5, and 6 displayed modest inhibition selectivity towards SHP1, while compound 4 exhibited a predilection for binding with CD45. Additionally, compound 7 showed a preference for inhibiting both PTP1B and Laforin, and compound 9 demonstrated remarkable selectivity for hCDC14A.
Table 1.
IC50s of PhFP (compound 2) and 7 new pTyr mimetics on 10 major mammalian PTPs.a
| Cmpd | IC50,μM | |||||||||
|---|---|---|---|---|---|---|---|---|---|---|
|
| ||||||||||
| CD45 | Laforin | MKP5 | SHP2 | SHP1 | hCDC14A | TC-PTP | PTP1B | LYP | STEP | |
|
| ||||||||||
| 2 | >>10,000 | 8384 ± 759 | 8226 ± 626 | 7985 ± 672 | 5730 ± 477 | 3269 ± 303 | 4853 ± 196 | 2217 ± 126 | 3205 ± 169 | 2860 ± 119 |
| 3 | 234.8 ± 16.0 | >500 | >500 | 165.9 ± 12.4 | 57.0 ± 6.2 | 220.6 ± 26.3 | >500 | >500 | >500 | >500 |
| 4 | 21.2 ± 1.4 | 61.3 ± 24.1 | 267.9 ± 32.9 | 167.2 ± 2.3 | 64.7 ± 5.2 | 98.7 ± 2.9 | 402.3 ± 17.4 | >500 | 41.9 ± 4.9 | 206.9 ± 20.6 |
| 5 | 42.9 ± 3.7 | 122.4 ± 9.0 | 214.7 ± 12.0 | 71.8 ± 6.5 | 22.7 ± 2.4 | 119.6 ± 14.6 | 397.2 ± 27.4 | 447.2 ± 23.9 | 37.5 ± 2.0 | 334.7 ± 28.5 |
| 6 | 107.1 ± 8.1 | 197.8 ± 18.4 | 431.3 ± 47.2 | 75.6 ± 16.4 | 36.6 ± 2.3 | 273.2 ± 18.8 | >500 | >500 | 139.4 ± 24.8 | >500 |
| 7 | 339.0 ± 25.8 | 59.7 ± 3.2 | 332.2 ± 28.6 | 181.0 ± 13.2 | 138.7 ± 5.7 | 339 ± 19.9 | 89.1 ± 6.7 | 40.2 ± 1.4 | 98.4 ± 6.0 | 287.2 ± 21.0 |
| 8 | 477.65 ± 41.9 | 116.85 ± 13.7 | 101.7 ± 8.9 | 164.1 ± 17.9 | 167.3 ± 15.6 | 360 ± 28.8 | 427.45 ± 26.4 | 189.25 ± 12.7 | 173.7 ± 14.1 | 245.4 ± 29 |
| 9 | >500 | 205.1 ± 23.5 | >500 | >500 | >500 | 10.4 ± 0.6 | >500 | >500 | >500 | >500 |
Details of the enzyme assays are provided under Experimental Section. IC50 values are mean ± standard deviation (SD) of the results from 3 independent experiments.
After validating the PTP inhibition potency of the novel pTyr mimetics, their polarity and cell permeability were also characterized. As suggested by the calculated physicochemical properties including topological polar surface area (tPSA) and cLogP (Table 2), all newly obtained pTyr mimetics are less polar than Ac-F2Pmp-OMe (compound 1) and PhFP (compound 2), which confirmed our hypothesis that fused aromatic ring systems can better balance the high polarity of the difluorophosphonic acid. Expectedly, the decreased polarity of these pTyr mimetics led to a significantly improved cell permeability in our PAMPA assay. As shown in Table 2, Ac-F2Pmp-OMe and PhFP have permeability coefficient (Pe) values of 0.6 × 10−6 cm/s and 0.5 × 10−6 cm/s, respectively, indicating both are poorly cell permeable (compounds with Pe > 1.5 × 10−6 cm/s are defined as cell permeable76,77). On the other hand, the cell permeability of the new pTyr mimetics was significantly improved with Pe values between 1.5 × 10−6 cm/s and 1.9 × 10−6 cm/s. Collectively, we showed that the new phosphonodifluoromethyl-containing bicyclic/tricyclic aryl derivatives are genuine PTP inhibitors with increased binding affinity and selectivity than their single-ring counterparts. They also possess improved cell permeability.
Table 2.
Calculated and determined physicochemical properties of the selected compounds
| Compound | M.Wa | tPSAb | cLogPc | Pe(×10−6 cm/s)d |
|---|---|---|---|---|
|
| ||||
| 1 | 351.1 | 112.9 | 1.1 | 0.6 |
| 2 | 208.0 | 57.5 | 1.9 | 0.6 |
| 3 | 325.9 | 86.2 | 2.6 | 1.7 |
| 4 | 248.0 | 70.7 | 2.7 | 1.5 |
| 5 | 264.1 | 57.5 | 3.1 | 1.6 |
| 6 | 293.0 | 70.4 | 3.1 | 1.9 |
| 7 | 391.9 | 57.5 | 5.0 | 1.7 |
| 8 | 409.9 | 57.5 | 4.8 | 1.6 |
| 9 | 298.0 | 70.7 | 3.8 | 1.7 |
| Verapamil | 454.3 | 64.0 | 5.1 | 12.9 |
| Furosemide | 330.7 | 122.6 | 1.8 | 0.3 |
Molecular weight.
Topological PSA (Å2).
Molinspiration calculated Log P.
Permeability coefficient determined with PAMPA assay in pH 7.4 PBS buffer. The calcium channel blocker Verapamil and loop diuretic Furosemide were used as positive control and negative control, respectively.
Compound 9 Competitively and Selectively Inhibited Both hCDC14A and B
Among the newly obtained pTyr mimetics, compound 9 inhibited hCDC14A with an IC50 value of 10.4 ± 0.6 μM. Further characterization revealed that compound 9 also inhibited hCDC14B with comparable potency (IC50 = 11.2 ± 0.6 μM). Subsequently, we determined the mode of hCDC14A/B inhibition by compound 9 through steady-state kinetic analyses. As shown in Figures 2A and B, the Lineweaver–Burk plots showed that compound 9 is a classic competitive inhibitor that only affects the apparent Km value, while the Vmax remains unchanged. This observation is consistent with the expectation that compound 9 binds to the active site of hCDC14A/B due to its pTyr mimetic properties. Consistent with competitive inhibition, compound 9 displays inhibition constant (Ki) values of 5.8 ± 0.9 and 7.3 ± 0.9 μM for hCDC14A and B, respectively.
Figure 2.

Compound 9 competitively inhibited both hCDC14A and B with selectivity over the other 10 major mammalian PTPs. (A-B) Effect of compound 9 on hCDC14A (A) and hCDC14B (B) catalyzed DiFMUP hydrolysis. Compound 9 concentrations were 0 (●), 3 (■), 6 (▲), 9 (▼), and 12 (◆) μM, respectively. The Lineweaver-Burk plot displayed the characteristic intersecting line pattern, consistent with competitive inhibition. (C) Selectivity of compound 9 over a panel of 10 mammalian PTPs. IC50 values of compound 9 tested against indicated PTPs. IC50 values are mean ± standard deviation (SD) of the result from 3 independent experiments. (D) Putative binding pose of compound 9 (green stick) bound to hCDC14B (PDB 1OHC) colored by element with the nearby P+1 pocket shown in the transparent surface. Hydrogen bonds and ionic bonds are represented by yellow dashes. Cation-π interactions are shown with purple dashes. The 4-position of the dibenzofuran centroid points towards the P+1 pocket. (E) P-loop sequence alignment of hCDC14A, hCDC14B, and additional DUSPs (MKP5, Laforin, and VHZ). UniProt accession codes used for alignment are as follows: CDC14A – Q9UNH5; CDC14B – O60729; MKP5 – Q9Y6W6; Laforin – O95278; VHZ - Q9BVJ7. Key: red box with white text = identical residues; white box with black text = different residues; white box with red text = similar residues; text outlined with blue box = similar residues across the group. Figure was prepared using ESPript 3.0.
In addition to its notable hCDC14A/B inhibition potency, compound 9 also displays at least 20-fold selectivity over the other tested PTPs (Figure 2C), including both closely related dual-specificity phosphatases and more distantly related classic pTyr-specific PTPs. To gain insights into the molecular basis of selective hCDC14A/B inhibition by compound 9, and guide further modification, molecular docking studies were performed using Glide78. The hCDC14A and B active sites are strongly conserved, and no hCDC14A crystal structure exists. In light of this, our molecular docking studies were performed on an existing hCDC14B crystal structure (PDB code 1OHC). As shown in Figure 2D, compound 9 makes extensive interactions with hCDC14B. Expectedly, the phosphonate group of compound 9 forms an ionic bond with the R320 side chain and several hydrogen bonds with the amide -NH backbones of K315, A316, G319, and R320. Further elevating compound 9’s binding affinity, the dibenzofuran core forms two cation-π interactions with the K315 side chain. Lastly, the dibenzofuran core also makes hydrophobic contact with the alkane side chains of A316 and I353. As shown in the sequence alignment (Figure 2E), the K315 residue in human CDC14A and B, which is involved in two cation-π interactions with compound 9, is absent in the other closely related DUSPs, which could largely contribute to the binding selectivity of compound 9 for hCDC14A and B. Moreover, despite the PTPs and DUSPs sharing the conserved signature catalytic motif HCXXGXXRS(T) and exhibiting similar mechanisms of catalysis21, the active site topology of these two classes of phosphatases is distinctive. Extensive structural biology studies have previously revealed that the active site in DUSPs is wide and shallow for the hydrolysis of phosphorylated serine, threonine, or tyrosine protein residues, whereas the pTyr-specific PTPs possess deeper active sites that restrict substrate specificity to only phosphotyrosine79–81. As shown in Figure S2, we docked compound 9 into a classical nonreceptor PTP (PTP1B) and a receptor-like PTP (CD45). In both cases, compound 9 sterically clashes with multiple active-site residues as a result of poor steric complementarity between the ligand and these pTyr-specific active sites. Moreover, the planarity of the dibenzofuran core is distorted to allow binding in the obtained docking poses, further highlighting unfavorable binding to PTP1B and CD45. To gain additional insights into the selectivity for hCDC14 among the various tested DUSPs, we performed docking studies on three dual-specificity phosphatases including Laforin, VHZ, and MKP5. The docking studies revealed that, while the Laforin active site is expectedly broader than that of the pTyr-specific PTP1B and CD45 and better able to accommodate compound 9, it lacks the key lysine residue (Figure 2E), and the ligand clashes with a nearby active site tyrosine residue. Notably, docking studies for VHZ and MKP-5 failed to generate a putative pose due to the inability of compound 9 to fit into their active sites, which are significantly smaller than the large peptide-binding groove of hCDC14B. Computational docking studies show that this poor steric complementarity and clash is a result of the perpendicular 1-phosphonodifluoromethyl substitution of compound 9. In contrast, due to the more linear alignment of the phosphonodifluoromethyl head and the aromatic ring system, compounds 7 and 8 exhibited moderate inhibition potency against most of the tested PTPs. These results could help explain the selectivity of compound 9 over both the tested pTyr-specific PTPs and DUSPs.
Fragment-Based Elaboration of Compound 9 Led to Improved hCDC14A/B Inhibitors
To further improve the hCDC14A and B inhibition potency and amplify inhibition selectivity over other PTPs, fragment- and structure-based elaboration of compound 9 was deployed. According to the docking result (Figure 2D), the 4-position of compound 9 pointed toward the aromatic/hydrophobic P+1 pocket of hCDC14B (also exists in hCDC14A). This pocket is composed of aromatic residues F85, Y86, and Y170, all of which are located in the A-domain that is unique to the CDC14s23. L318 and I353 also contribute to this hydrophobic pocket. Previous studies have shown that this P+1 pocket contributes to the substrate selectivity of CDC14s23, therefore, we envisioned that introducing hydrophobic, aromatic functional groups at the 4-position of compound 9 could establish extra hydrophobic/π-π interactions with the P+1 pocket, thus improving hCDC14A/B binding potency and selectivity.
As shown in Table 3, with compound 10 as the starting material, a variety of aromatic rings were introduced to the 4-position of compound 9 and rapidly yielded compounds 11-17. Compared with the parent compound 9 and the 4-bromo substituted derivative compound 10, directly attaching a benzene ring (compound 11) increases the hCDC14A/B inhibition potency by about 5-fold. Substituting this benzene ring with either an electron donating group (compound 12) or electron withdrawing group (compound 13), as well as replacing it with a substituted heteroaromatic ring, 2-chlorothiophene (compound 14) has a very marginal impact on the inhibition potency. Considering the docking pose of compound 9 and the nearby hydrophobic P+1 pocket, these results implied that the directly attached aromatic rings only participated in hydrophobic contacts with the pocket but were not ideally situated for the desired π-π stacking interaction with the deeper posited aromatic residues in the P+1 pocket. In contrast, compound 15 has a 4-styrene substitution and inhibits hCDC14A and B with IC50 values of 0.09 ± 0.01 μM and 0.22 ± 0.02 μM, respectively. While the 4-styrene moiety enjoys a 108-fold increase in binding affinity, attaching phenyl rings via an alkyne or alkene linker (compounds 16 and 17) likely precludes the phenyl functionality from engaging the P+1 pocket and forming π-π stacking interactions. As a result, compared to compound 15, these compounds exhibited decreased hCDC14A/B inhibition potency.
Table 3.
Structures and IC50s of compound 9 and its derivatives on hCDC14A and B.a
| |||||||
|---|---|---|---|---|---|---|---|
|
| |||||||
| Compound | R Structure | IC50, μM |
Compound | R Structure | IC50, μM |
||
| hCDC14A | hCDC14B | hCDC14A | hCDC14B | ||||
|
| |||||||
| 9 | -H | 9.85 ± 0.30 | 12.78 ± 0.41 | 14 |
|
3.21 ± 0.25 | 5.85 ± 0.33 |
| 10 | -Br | 9.41 ± 0.41 | 11.88 ± 0.44 | 15 |
|
0.09 ± 0.01 | 0.22 ± 0.02 |
| 11 |
|
1.92 ± 0.16 | 2.17 ± 0.24 | 16 |
|
2.00 ± 0.42 | 4.19 ± 0.17 |
| 12 |
|
2.17 ± 0.24 | 5.21 ± 0.46 | 17 |
|
2.74 ± 0.27 | 3.77 ± 0.39 |
| 13 |
|
2.78 ± 0.44 | 3.76 ± 0.25 | ||||
Details of the enzyme assays are provided under Experimental Section. IC50 values are mean ± standard deviation (SD) of the results from 3 independent experiments.
Compound 15 is the Most Potent and Selective hCDC14A/B Inhibitor
After obtaining the most potent compound 15, we next sought to confirm the mode of hCDC14A/B inhibition by this molecule. Expectedly, compound 15 behaved as a reversible and competitive inhibitor for both hCDC14A and B with Ki values of 57.4 ± 7.6 and 90.0 ± 8.8 nM, respectively (Figure 3A and B). To be qualified as a chemical probe that can be applied to interrogate the function and therapeutic potential of hCDC14A/B, the inhibition specificity of compound 15 is of utmost importance. To demonstrate this, the inhibition activity of compound 15 against a panel of 16 representative PTPs including the closely related DUSPs, more distantly related non-receptor PTPs, receptor-like PTPs, and low molecular weight PTP, was profiled (Figure 3C). To our delight, compound 15 showed at least 50-fold selectivity over all of the PTPs tested. Although the IC50 values of compound 15 on all enzymes were measured with 0.01% Triton X-100 in the assay buffer, which is known to be capable of eliminating most of the false positive caused by the self-aggregation of organic molecules82, we still confirmed that compound 15 was not an aggregator with the DLS-based small molecule aggregation detection method. As shown in Figure 3D, no DLS signal decay was observed with 100 μM of compound 15 solution. This DLS testing concentration of compound 15 is higher than its IC50s on all the tested PTPs, suggesting that compound 15 is a bona fide inhibitor. We further validated the target engagement of compound 15 with the Differential Scanning Fluorimetry (DSF) assay. As shown in Figure 3E, the melting temperature (Tm) for hCDC14A was 44.7 ± 0.1 °C. Compound 15 treatment increased the thermal stability of hCDC14A and resulted in a positive thermal shift with ΔTm of 3.4 ± 0.3 °C, hence validating direct binding between compound 15 and the hCDC14A protein. Due to the limited hCDC14B protein solubility and the relatively high protein concentration required for the DSF assay, we did not run this assay with hCDC14B.
Figure 3.

Compound 15 is a potent and selective competitive inhibitor for human CDC14A and B. (A-B) Effect of compound 15 on hCDC14A (A) and hCDC14B (B) catalyzed DiFMUP hydrolysis. For CDC14A Ki measurement, compound 15 concentrations were 0 (●), 25 (■), 50 (▲), 75 (▼), and 100 (◆) nM, respectively. For CDC14B Ki measurement, compound 15 concentrations were 0 (●), 100 (■), 200 (▲), 300 (▼), and 400 (◆) nM, respectively. The Lineweaver-Burk plot displayed the characteristic intersecting line pattern, consistent with competitive inhibition. (C) Selectivity of compound 15 over a panel of 16 mammalian PTPs. IC50 values are mean ± standard deviation (SD) of the result from 3 independent experiments. (D) DLS signal of 100 μM compound 15 dissolved in DMG assay buffer. (E) Thermal shift first-derivative curves of 5 μM hCDC14A (green) and 5 μM hCDC14A + 50 μM compound 15 (red).
To investigate the significant improvement in potency and selectivity conferred by the added styrene ring, we performed additional molecular docking studies on compound 15, and the result is shown in Figure 4A. Analogous to compound 9, the phosphonate group forms an ionic bond with the R320 side chain and several hydrogen bonds with the amide -NH backbones of K315, A316, G319, and R320. The dibenzofuran core forms two cation-π interactions with K315 while also making hydrophobic contacts with the alkane side chains of A316 and I353. According to our docking model, the added styrene moiety engages the hydrophobic P+1 pocket and is suitably oriented to form an edge-to-face π-π interaction with F85. Additional van der Waals interactions between the alkene and A168 and Y170 side chains further strengthen the binding affinity of compound 15. Notably, this P+1 pocket is only present in the CDC14s, thus explaining the over 50-fold selectivity observed for compound 15 (compared to over 20-fold for compound 9 which lacks the styrene ring). To assess the stability of our docking pose, we performed a 50 ns molecular dynamic (MD) simulation on the docked pose of compound 15 to hCDC14B. As shown in Figure S3, the ligand root mean square deviation (RMSD) (as aligned on the protein) reveals that the docked pose is stable and does not diffuse away during the simulation. Notably, the protein root mean square fluctuation (RMSF) during the simulation correlates well with the experimental B factor, a measure of protein flexibility, in hCDC14B. Alignment of our docking result and the cocrystal structure of hCDC14B complexed with a phosphoserine peptide that is known to engage the P+1 pocket (Figure 4B) show that the styrene ring binds deeper into the P+1 pocket compared to the prolyl ring of the peptide, thus making greater hydrophobic contacts. Unlike the prolyl ring side chain, styrene is also able to participate in the aforementioned π-π interaction, which contributes to the high potency and selectivity of the scaffold. Notably, our docking model does not predict this π-π stacking interaction to be formed by the scaffolds with a directly attached aromatic ring (compounds 11-14, Table 3) or those with an alkene (16) or alkyne (17) linker. These findings support the biochemical potencies reported in Table 3 and suggest that these other derivatives only gain additional hydrophobic contacts with the hCDC14 binding site.
Figure 4.

The docking pose of compound 15 bound to hCDC14B suggests engagement with the P+1 pocket. (A) Docking pose of compound 15 (cyan stick) with hCDC14B (PDB code 1OHC) colored by element. Hydrogen bonds and ionic bonds are depicted by yellow dashes, cation-π interactions are depicted by purple dashes, and the edge-to-face π-π stacking is shown with a red dash. (B) Docking pose of compound 15 (cyan stick) aligned to the cocrystal structure of hCDC14B (surface colored by element) and a phosphoserine peptide ligand (green stick) (PDB code 1OHE).
Mutagenesis of Key Binding Residues in the P+1 Pocket in hCDC14A and B Supports the Binding Mode of Compound 15
To ascertain the importance of the unique P+1 pocket for the binding of compound 15 and further corroborate our docking model, we carried out site-directed mutagenesis experiments. In particular, we individually substituted the Phe48 (hCDC14A)/Phe85 (hCDC14B) residues, which were anticipated to capture π-π stacking interactions with the benzene ring of the styrene group in compound 15, to alanine. We also replaced the Phe134 (hCDC14A)/Tyr170 (hCDC14B) residues, which were predicted to establish hydrophobic contacts with the ethylene of the styrene group in compound 15, with alanine residues. Additionally, as shown in Figure S4, residues of the P+1 pocket in hCDC14A and B are relatively conserved. Among the P+1 pocket residues that directly interact with compound 15, Phe134 (hCDC14A)/Tyr170 (hCDC14B) is the only pair of amino acids that is distinct between these two enzymes, which can potentially contribute to the slight (2.4-fold) inhibition preference of compound 15 for hCDC14A. Notably, this sequence difference results in a more lipophilic P+1 pocket for hCDC14A compared to hCDC14B due to the absence of the polar hydroxyl group. To determine the impact of these residues on the selectivity of compound 15, we also mutated the Phe134 residue in hCDC14A to tyrosine to mimic the structure of hCDC14B.
The wild-type and mutants of hCDC14A and B were expressed in E. coli and purified to homogeneity. As shown in Table S1, the kinetic parameters of hCDC14A F134A and F134Y mutants and hCDC14B Y170A mutant are comparable to those of the wild type hCDC14A and B, indicating that these mutations do not perturb the overall enzyme structure and function. On the other hand, the Km values of hCDC14A F48A and hCDC14B F85A mutants are similar to the corresponding wild-type enzyme albeit the kcat values for these mutants are lower than those of the wild-type enzymes. A similar observation was also documented in an earlier publication, wherein the hCDC14B F85A mutant displayed decreased kcat when assessed with a phosphopeptide as substrate.23 After characterizing these mutants, we determined the impact of altering the above-mentioned P+1 site residues on the IC50 values of compounds 15. As shown in Table 4, compared to the wild-type enzymes, hCDC14A F48A and hCDC14B F85A mutants displayed reduced sensitivity towards compound 15, resulting in IC50 values that were 27.3- and 15.6-fold higher, respectively. To ascertain that this diminished effectiveness of compound 15 could be attributed to the absence of the π-π interaction, we also assessed the IC50 of vanadate (a known general phosphate mimicking PTP competitive inhibitor83,84), PhFP, and compound 9 (both PhFP and compound 9 are not supposed to engage the P+1 pocket) for the hCDC14A F48A and hCDC14B F85A mutants. As anticipated, the IC50 values of these compounds for the hCDC14A F48A and hCDC14B F85A mutants were comparable to those observed with the corresponding wild-type enzymes (data shown in Table S2). Additionally, the hCDC14A F134A and hCDC14B Y170A mutants, which disrupt the aforementioned hydrophobic interaction with the styrene alkene, experienced a 2 to 3-fold reduction in the inhibitory potency of compound 15 (Table 4). These data supported the proposed binding mode for compound 15, namely it interacts with both the active site and the nearby P+1 Pro binding pocket. Finally, the introduction of the F134Y mutation in hCDC14A resulted in an IC50 value of 0.19 ± 0.02 μM, which is comparable to the IC50 value of compound 15 observed for the wild-type hCDC14B (0.22 ± 0.02 μM). These results indicated that the Phe134 (hCDC14A)/Tyr170 (hCDC14B) residues indeed contribute to the slight (2.4-fold) hCDC14A binding preference of compound 15 for hCDC14A.
Table 4.
Comparison of IC50 data of compound 15 against wild-type and mutant hCDC14A and B.a
| Enzyme | Mutation | IC50,μM |
|---|---|---|
|
| ||
| CDC14A | WT | 0.09 ± 0.01 |
| F48A | 2.46 ± 0.21 | |
| F134A | 0.26 ± 0.03 | |
| F134Y | 0.19 ± 0.02 | |
|
| ||
| CDC14B | WT | 0.22 ± 0.02 |
| F85A | 3.43 ± 0.28 | |
| Y170A | 0.51 ± 0.03 | |
IC50 values are mean ± standard deviation (SD) of the result from 3 independent experiments.
Compound 15 is Cell Active and Bioavailable
Previous studies have revealed that hCDC14 dephosphorylates Ser315 of p5337. Genomic stresses such as DNA damage induce p53 Ser315 phosphorylation, which promotes the binding of Pin1, leading to a conformational change that facilitates additional modifications and influences its interaction with Mdm2, a regulatory partner of p53, resulting in full stabilization of p53.39,40 To determine the cellular efficacy of our hCDC14A/B inhibitor compound 15, we treated the U2OS cells with the DNA damage reagent Mitomycin C along with various doses of compound 15 for 24 hours. As expected, Mitomycin C treatment causes p53 stabilization along with increased Ser315 phosphorylation (Figure 5A). compound 15 dose-dependently increased both p53 Ser315 phosphorylation and p53 protein levels with maximum 20- and 13-fold enhancement, respectively (Figure 5A). To increase confidence that the engagement of the hCDC14A/B is responsible for the observed p53 Ser315 phosphorylation enhancement and p53 stabilization, we designed and prepared a negative control compound 18 (structure shown in Figure 5A), which is closely related in structure to compound 15 but showed no hCDC14A/B inhibition potency up to 100 μM due to the lack of the PTP active site binding-essential difluorophosphonic acid moiety. Satisfactorily, compound 18 showed negligible impact on total p53 and p53 Ser315 phosphorylation levels at 20 μM concentration. Taken together, these results agreed with previous studies and indicated that hCDC14s mediated dephosphorylation of p53 on Ser315 was abolished upon compound 15 treatment.
Figure 5.

Compound 15 is a cell active and bioavailable hCDC14A and B inhibitor. (A) Immunoblots of whole cell lysates from U2OS cells with indicated treatment for 24 hours. Compound 15 was able to further amplify Mitomycin C-induced p53 ser315 phosphorylation and stabilize p53 protein, whereas no effect on the negative control (compound 18) treated cells were observed. (B) In vivo pharmacokinetics data based on mass spectrometry quantification at 0, 0.5, 1, 3, 6, and 24 h from the time of intraperitoneal injection (I.P.) or oral administration (P.O.) of compound 15 with 50 mg/kg dose. Values are mean ± standard deviation (SD), n=3.
Having obtained a first-in-class, potent, selective, and cell-active hCDC14A and B inhibitor, we next preliminarily assessed the safety and pharmacokinetics (PK) of compound 15 to enable and support future in-depth mechanistic studies. For safety assessment, the impact of compound 15 treatment on the proliferation of 4 human and murine cell lines (HEK293, U2OS, Jurkat, and Raw 264.7) was tested. As shown in Figure S5, following a 24-hour treatment period, compound 15 showed no considerable cytotoxic effects across all tested cells at a maximum concentration of 50 μM. To characterize the pharmacokinetics, compound 15 was dosed by intraperitoneal injection (I.P.) or oral administration (P.O.) in C57BL/6 mice. As shown in Figure 5B, a single I.P. injection of compound 15 at 50 mg/kg achieved a peak plasma concentration (Cmax) of 21.9 ± 3.7 μM with a half-life of 2.2 ± 0.4 hours. Moreover, compound 15 was also orally bioavailable. A single oral administration of compound 15 at 50 mg/kg gave a Cmax of 4.1 ± 0.6 μM with a half-life of 3.0 ± 1.2 hours (Figure 5B). Overall, compound 15 showed a good PK profile with acceptable exposure and elimination to enable its use as a chemical probe for future in vivo studies.
Chemistry
As depicted in Scheme 1, the P-(Difluorophenylmethyl) phosphonic acid (PhFP, compound 2) and compound 2–10 were synthesized via a CuI-promoted coupling reaction85 followed by the hydrolysis of the phosphonates. Briefly, the cadmium-containing compound 19 was first prepared by stirring a DMA solution of diethyl bromodifluoromethylphosphonate with cadmium shots under argon. Subsequently, compound 19 was coupled to commercially available iodo-substituted aromatic rings in the presence of copper (I) iodide to yield intermediates 2a-10a. After 19 hours of reaction at 50 °C, the yields of indicated compounds were between 63% and 74% (determined by LC-MS). Next, compounds 2a-10a were dissolved in DCM and treated with trimethylsilyl iodide to furnish the final pTyr mimetics 2-10.
Scheme 1.

Synthesis of intermediates 2a-10a and pTyr mimetics 2–10a
aReagents and conditions: (a) cadmium shot, DMA, argon, rt, 3 h. (b) compound 19, CuBr, argon, DMA, rt to 50 °C, 15 h. 63–74%. (c) TMSI, DCM, 0 °C, 4 h. 41–64%.
The synthesis of Ac-F2Pmp-OMe (compound 1) was depicted in Scheme 2. Starting with the commercially available starting material N-[(1,1-dimethylethoxy)carbonyl]-4-iodo-L-phenylalanine methyl ester, the phosphonate intermediate 20 was obtained with the above-described CuI-promoted coupling reaction. Ac-F2Pmp-OMe (compound 1) was obtained by the sequence of Boc removal, acetylation of the N-terminal amine, and hydrolysis of the diethyl phosphonate.
Scheme 2.

Synthesis of Ac-F2Pmp-OMe (1) a
aReagents and conditions: (a) compound 19, CuBr, argon, DMA, rt to 50 °C, 19 h. 77%. (b) 20% TFA/DCM, rt, 6h. (c) acetyl chloride, triethylamine, DCM 0 °C to rt, 4 h. 64% (b and c 2 steps). (c) TMSI, DCM, 0 °C, 4 h. 59%.
The synthesis of compound 11-18 was depicted in Scheme 3. Started with compound 10, the substitutions on the 4-position of the dibenzofuran rings were introduced through palladium-catalyzed Suzuki (compound 11–16 and compound 18) or Sonogashira cross-coupling (compound 17) reactions with the corresponding boronic acids or alkyne. Similarly, the negative control compound 18 was synthesized from 4-bromodibenzo[b,d]furan by the Suzuki coupling reaction.
Scheme 3.

Synthesis of compound 11–18 a
aReagents and conditions: (a) Aryl boronic acid, K2CO3, Pd(PPh3)4, argon, DMF, H2O, 90°C, 16h, 39–49%. (b) ethynylbenzene, CuI, Pd(PPh3)4, triethylamine, argon, DMF, 90°C, 16h, 56%. (c) K2CO3, Pd(PPh3)4, argon, DMF, H2O, 90°C, 16h, 71%.
Conclusions
PTPs are pivotal regulators of protein tyrosine phosphorylation, which plays indispensable roles in modulating cellular signaling. Abnormal perturbation of the level of protein tyrosine phosphorylation is intricately associated with the pathogenesis of various human diseases. Gaining an improved understanding of the involvement of PTPs in dysregulated signaling events provides valuable insights into the etiology of diseases and holds the potential to unveil novel therapeutic targets. In contrast to the extensively studied PTKs, research on PTPs is relatively underdeveloped, which is partially due to the limited availability of high-quality small molecule inhibitors that can serve as invaluable chemical probes to interrogate PTP function and evaluate therapeutic hypothesis. The challenge of identifying active site-directed, potent, selective and pharmacologically efficacious inhibitors for PTPs is well known in drug discovery. Although success in harnessing PTP allosteric regulatory mechanisms for PTP inhibitor development has been documented in a few cases17–20, the discovery of allosteric inhibitors has been mostly serendipitous. Previous efforts aimed at the highly conserved and positively charged PTP active sites have often led to predominantly negatively charged inhibitors with limited cell permeability, which makes them unsuitable as chemical probes and therapeutics. Thus, an improved general approach targeting the PTP active site is highly desirable for the development of orthosteric PTP inhibitors as chemical probes to advance our understanding of PTPs in disease biology and target validation.
We have previously shown that despite the significant structural conservation among the PTP active sites, potent, selective and bioavailable inhibitors for PTPs can be generated by employing fragment-based approaches to capitalize on subtle sequence differences in the outer region of the active site.1 To that end, we have developed a number of highly selective and efficacious orthosteric PTP inhibitors by tethering a negatively charged pTyr mimetics, such as F2Pmp, salicylic acids and α-sulfophenylacetic amide with appropriate diversity fragments to engage both the active site and nearby unique peripheral pockets.1,10,57–63 Excitingly, an orally bioavailable acidic thiadiazolidinone dioxide moiety-containing PTP1B/TC-PTP active site inhibitor, ABBV-CLS-484, has entered clinical evaluation for cancer immunotherapy in human64, which significantly boosts the druggability of targeting the PTP active site. To further enhance our ability to target the PTP active site, we sought to develop novel nonhydrolyzable pTyr mimetics with improved binding affinity and cell permeability.
Here, starting with a classic nonhydrolyzable pTyr mimetic, F2Pmp, we synthesized 7 novel phosphonodifluoromethyl-containing bicyclic/tricyclic aryl derivatives in order to better balance the lipophilicity and binding affinity. As expected, these new bicyclic/tricyclic pTyr mimetics display increased cell permeability and greater intrinsic PTP binding potency. Among these derivatives, compound 9 with a dibenzofuran scaffold exhibited promising hCDC14A and B inhibition potency and exceptional selectivity over 10 other PTPs. Subsequent molecular docking studies elucidated the binding mode between compound 9 and hCDC14B, explained the hCDC14A/B biased inhibition potency, and guided the following fragment-based optimization. Among the compound 9 derivatives, compound 15 showed a Ki of 57.4 ± 7.6 and 90.0 ± 8.8 nM for hCDC14A and B, respectively, with >50-fold selectivity over a panel of 16 major mammalian PTPs. Through kinetic and biophysical analyses, molecular docking, and site-directed mutagenesis, compound 15 was confirmed to be a genuine active site-directed competitive inhibitor of both hCDC14A and hCDC14B. Consistent with the design strategy, the extraordinarily high selectivity of compound 15 for the CDC14 phosphatases derives from its unique interactions with the P+1 pocket in addition to the active site. Moreover, despite its negative charges, compound 15 can engage its target and block CDC14-mediated p53 dephosphorylation in a cell-based assay with low μM efficacy. Lastly, compound 15 exhibits good pharmacokinetic properties and oral bioavailability. Collectively, the results qualify compound 15 as a chemical probe for the CDC14 phosphatases because it meets the specific criteria recommended for a small molecule probe86,87. Given its potency, selectivity, and in vivo bioavailability, compound 15 holds great potential as a molecular probe for further investigation of the roles of hCDC14A and hCDC14B in normal physiological processes and diseases. Additionally, it could be utilized to validate the therapeutic potential of targeting these enzymes. Finally, we demonstrate that the newly developed phosphonodifluoromethyl-containing bi/tri-cyclic pTyr mimetics are sufficiently polar to bind the PTP active site yet remain capable of efficiently crossing cell membranes. Importantly, these novel nonhydrolyzable pTyr mimetics offer new starting points for the development of active site-directed inhibitors for other PTPs of therapeutic interest. Our work also furnishes another example of the acquisition of highly potent, selective, and efficacious active site-directed PTP inhibitors by exploiting the specific structural features of different PTP active sites and less conserved peripheral pockets.
Experimental Section
General Information
Unless otherwise noted, all reagents and starting materials were purchased from commercial suppliers and used without further purification. Thin-layer chromatography was performed using glass precoated Merck silica gel 60 F254 plates. Flash column chromatography was performed on Biotage prepacked columns using the automated flash chromatography system Biotage Isolera One. Organic solvents were evaporated using rotary evaporation at 50–55 °C. The 1H- and 13C NMR spectra were recorded on a Bruker AVANCE 500 MHz spectrometer using CDCl3 or dimethyl sulfoxide (DMSO-d6) as the solvent. Chemical shifts are expressed in ppm (δ scale) and referenced to the residual protonated solvent. Peak multiplicities are reported using the following abbreviations: s (singlet), d (doublet), t (triplet), q (quartet), m (multiplet), or br (broad singlet). Low-resolution mass spectra and purity data were obtained using an Agilent Technologies 6470 series, triple quadrupole LC–MS. The purity of all final tested compounds was determined to be >95% (UV, λ = 254 nm). Anti-p53 (catalog# sc-126), antiphospho-p53 (catalog# sc-135772), and anti-GAPDH (catalog# SC-59541) antibodies were purchased from Santa Cruz. 6,8-Difluoro-4-Methylumbelliferyl Phosphate (DiFMUP) was purchased from Invitrogen (catalog# D6567). bpV(Phen) was purchased from Sigma-Aldrich (catalog# SML0889). Mitomycin C was purchased from Thermo Scientific (catalog# J63193). Phosphatase inhibitor cocktail was purchased from Selleck Chemicals (catalog# B150001). Protease inhibitor mixture was purchased from Roche Applied Science (catalog# 4693116001)
Chemical Synthesis
Synthesis of cadmium-containing intermediate (19)
Cadmium metal (6.00 g, 53.4 mmol) was washed by stirring under argon with 1 N HCl (5 mL, 15 min), H2O (5 mL, 3 × 1 min) and acetone (5 mL, 3 × 1 min). The metal was dried overnight under high vacuum while stirring until a metallic shine was observed. While warming the reaction vessel by hand, anhydrous DMF (10 mL) and diethyl bromodifluoromethylphosphonate (12.5 mL, 31.2 mmol) were added dropwise over 15 min to the metal under argon. The slightly exothermic reaction proceeded and was allowed to stir at r.t. for 3 h. The supernant of this reaction was directly used in the following step.
Synthesis of Methyl (S)-2-((tert-butoxycarbonyl)amino)-3-(4-((diethoxyphosphoryl)difluoromethyl)phenyl)propanoate (20)
N-[(1,1-Dimethylethoxy)carbonyl]-4-iodo-L-phenylalanine methyl ester (4.10 g, 10.0 mmol, 1.0 eq) and CuBr (2.87 g, 20.0 mmol, 2 equiv.) were suspended in anhydrous DMF (100 mL) under argon atmosphere, then the supernatant solution of the Cd reagent 19 (ca. 1.7 equiv.) was added in a dropwise manner. After 3 hours, further CuBr (1.43 g, 10.0 mmol, 1 equiv.) and the Cd reagent solution (ca. 1.7 equiv.) were added. The reaction was allowed to stir at r.t. for 19 hours in total. The progress of the reaction was monitored by LC-MS. Upon completion, the reaction mixture was diluted with EtOAc (500 mL), filtered through Celite, and extracted with aq NH4Cl (2 × 500 mL) and brine (500 mL). The organic layer was dried over anhydrous Na2SO4 and evaporated in vacuo. The product was then purified by flash chromatography (EtOAc/Hexanes, 0%→40%). Yield 3.58 g (77%). 1H NMR (500 MHz, CDCl3) δ 7.52 (d, J = 8.0 Hz, 2H), 7.20 (d, J = 7.9 Hz, 2H), 5.01 (d, J = 8.3 Hz, 1H), 4.57 (q, J = 6.8 Hz, 1H), 4.20 – 4.07 (m, 4H), 3.68 (s, 3H), 3.20 – 2.99 (m, 2H), 1.39 (s, 9H), 1.27 (tt, J = 7.2, 1.6 Hz, 6H). MS (ESI): calcd for C20H30F2NO7P [M + H]+ 466.17; found, 466.16.
Synthesis of Methyl (S)-2-acetamido-3-(4-((diethoxyphosphoryl)difluoromethyl)phenyl)propanoate (21)
To a solution of Intermediate 20 (3.58 g, 7.7 mmol) in DCM (40 mL), trifluoroacetic acid (10 mL) was added and stirred for 6 hours at rt. Then the excess reagent and solvent were evaporated under reduced pressure to give the crude deprotected amine, which was used in the next step without further purification. Acetyl chloride (824 μL, 11.55 mmol,1.5 equiv.) was added at 0 °C to a stirred solution of the deprotected amine and triethylamine (1.6 mL, 11.55 mmol, 1.5 equiv.) in dry DCM (50 mL). The reaction mixture was then allowed to warm to rt and stirred for 4 hours. Upon completion, the reaction mixture was diluted with EtOAc and washed with brine. The organic layer was dried over anhydrous Na2SO4 and evaporated in vacuo. The product was then purified by flash chromatography (EtOAc/Hexanes, 0%→40%). Two steps yield 2.01 g (64%). 1H NMR (500 MHz, CDCl3) δ 7.51 (d, J = 7.4 Hz, 2H), 7.18 (d, J = 8.2 Hz, 2H), 6.24 (d, J = 7.8 Hz, 1H), 4.85 (dt, J = 7.7, 5.9 Hz, 1H), 4.21 – 4.05 (m, 4H), 3.68 (s, 3H), 3.20 – 3.04 (m, 2H), 1.95 (s, 3H), 1.28 (td, J = 7.1, 4.2 Hz, 6H). MS (ESI): calcd for C17H24F2NO6P [M + H]+ 408.13; found, 408.12.
Synthesis of (S)-((4-(2-acetamido-3-methoxy-3-oxopropyl)phenyl)difluoromethyl)phosphonic acid (Ac-F2Pmp-OMe, 1)
The solution of Intermediate 21 (2.01 g, 4.9 mmol) in anhydrous DCM (25 ml) was cooled to 0 °C and stirred vigorously. Then trimethylsilyl iodide (4.9 mL, 34.3 mmol, 7.0 eq) was added to the solution dropwise. The reaction was kept at 0 °C and monitored with LC-MS. Upon completion, the reaction mixture was added to a 50% MeCN/H2O mixture dropwise and stirred at r.t. for 30 minutes. Then the water and organic solvent was evaporated in vacuo to give crude product which was further purified by prep HPLC. Yield 1.01 g (59%). 1H NMR (500 MHz, DMSO-d6) δ 8.37 (d, J = 7.7 Hz, 1H), 7.44 (d, J = 7.9 Hz, 2H), 7.31 (d, J = 8.0 Hz, 2H), 4.53 – 4.41 (m, 1H), 3.59 (s, 3H), 3.09 – 2.88 (m, 2H), 1.79 (s, 3H). MS (ESI): calcd for C13H16F2NO6P [M - H]+ 350.07; found, 350.08.
General procedure for the synthesis of intermediate 2a-10a
Iodo-substituted aromatic rings (0.2 mmol, 1.0 eq) and CuBr (57.4 mg, 0.4 mmol, 2.0 equiv.) were suspended in anhydrous DMA (2 mL) under argon atmosphere, then the supernatant solution of the Cd reagent 19 (ca. 1.7 equiv.) was added in a dropwise manner. After 3 hours, further CuBr (28.7 mg, 0.2 mmol, 1.0 equiv.) and the Cd reagent solution (ca. 1.7 equiv.) were added. The reaction was then allowed to stir at 50 °C for 12 hours. The progress of the reaction was monitored by LC-MS. Upon completion, the reaction mixture was diluted with EtOAc (50 mL), filtered through Celite, and extracted with aq NH4Cl (2 × 50 mL) and brine (50 mL). The organic layer was dried over anhydrous Na2SO4 and evaporated in vacuo. The product was then purified by flash chromatography (EtOAc/Hexanes, 0%→40%).
Compound 2a. Yield 37.5 mg (71%). 1H NMR (500 MHz, CDCl3) δ 7.62 (m, 2 H), 7.46 (m, 3 H), 4.25–4.10 (m, 4 H), 1.29 (td, J = 7.1, 4.2 Hz, 6H). MS (ESI): calcd for C11H15F2O3P [M + H]+ 265.07; found, 265.08.
Compound 3a. Yield 52.9 mg (69%). 1H NMR (500 MHz, CDCl3) δ 10.21 (s, 1H), 8.25 (s, 1H), 7.83 (s, 1H), 7.53 (s, 1H), 4.34 – 4.20 (m, 4H), 1.36 (t, J = 7.1 Hz, 6H). (ESI): calcd for C12H14BrF2O3P [M + H]+ 382.99; found, 383.00.
Compound 4a. Yield 38.3 mg (63%). 1H NMR (500 MHz, CDCl3) δ 7.90 (s, 1H), 7.68 (d, J = 2.2 Hz, 1H), 7.56 (d, J = 1.2 Hz, 2H), 6.82 (d, J = 2.2 Hz, 1H), 4.25 – 4.09 (m, 4H), 1.33 – 1.28 (m, 6H). MS (ESI): calcd for C13H15F2O4P [M + H]+ 305.07; found, 305.09.
Compound 5a. Yield 44.8 mg (70%). 1H NMR (500 MHz, CDCl3) δ 8.09 (s, 1H), 7.95 (d, J = 8.5 Hz, 1H), 7.58 (d, J = 8.5 Hz, 1H), 7.52 (d, J = 5.5 Hz, 1H), 7.40 (d, J = 5.4 Hz, 1H), 4.26 – 4.10 (m, 4H), 1.31 (t, J = 7.1 Hz, 6H). MS (ESI): calcd for C13H15F2O3PS [M + H]+ 321.04; found, 321.06.
Compound 6a. Yield 45.5 mg (65%). 1H NMR (500 MHz, CDCl3) δ 9.13 (s, 1H), 8.29 (d, J = 8.3 Hz, 1H), 8.12 – 8.01 (m, 2H), 7.57 (dd, J = 8.4, 4.1 Hz, 1H), 4.30 – 4.18 (m, 4H), 1.34 (td, J = 7.1, 0.7 Hz, 6H). MS (ESI): calcd for C14H15ClF2NO3P [M + H]+ 350.04; found, 350.04.
Compound 7a. Yield 66.5 mg (74%). 1H NMR (500 MHz, CDCl3) δ 8.40 – 8.31 (m, 2H), 7.93 (d, J = 8.4 Hz, 1H), 7.77 – 7.70 (m, 2H), 7.59 (dd, J = 8.4, 1.9 Hz, 1H), 4.30 – 4.14 (m, 4H), 1.33 (td, J = 7.0, 0.6 Hz, 6H). MS (ESI): calcd for C17H16BrF2O3PS [M + H]+ 450.97; found, 450.99.
Compound 8a. Yield 61.7 mg (66%). 1H NMR (500 MHz, CDCl3) δ 7.84 (s, 1H), 7.80 – 7.75 (m, 2H), 7.66 – 7.62 (m, 2H), 7.49 (d, J = 8.1 Hz, 1H), 4.30 – 4.16 (m, 4H), 1.34 (t, J = 7.1 Hz, 6H). MS (ESI): calcd for C18H16BrF4O3P [M + H]+ 467.00; found, 467.03.
Compound 9a. Yield 51.0 mg (72%). 1H NMR (500 MHz, CDCl3) δ 8.31 (d, J = 7.2 Hz, 1H), 7.69 (d, J = 8.2 Hz, 1H), 7.63 – 7.44 (m, 4H), 7.36 (ddd, J = 8.2, 7.2, 1.2 Hz, 1H), 4.24 – 4.06 (m, 4H), 1.22 (td, J = 7.0, 0.7 Hz, 6H). MS (ESI): calcd for C17H17F2O4P [M + H]+ 355.08; found, 355.09.
Compound 10a. Yield 62.9 mg (71%). 1H NMR (500 MHz, CDCl3) δ 8.25 (d, J = 8.2, 1H), 7.65 (d, J = 8.3 Hz, 1H), 7.60 (d, J = 8.3 Hz, 1H), 7.50 – 7.40 (m, 2H), 7.34 (ddd, J = 8.2, 7.3, 1.1 Hz, 1H), 4.23 – 4.02 (m, 4H), 1.19 (t, J = 7.1 Hz, 6H). MS (ESI): calcd for C17H16BrF2O4P [M + H]+ 432.99; found, 432.98.
General procedure for the synthesis of compound 2–10
The solution of Intermediate 2a-10a (0.1 mmol) in anhydrous DCM (1 ml) was cooled to 0 °C and stirred vigorously. Then trimethylsilyl iodide (100 μL, 0.7 mmol, 7.0 eq) was added to the solution dropwise. The reaction was kept at 0 °C and monitored with LC-MS. Upon completion, the reaction mixture was added to a 50% MeCN/H2O mixture dropwise and stirred at r.t. for 30 minutes. Then the water and organic solvent was evaporated in vacuo to give crude product which was further purified by prep HPLC.
Compound 2. Yield 13.3 mg (64%)1H NMR (500 MHz, DMSO-d6): δ = 7.47 (m, 5H); 13C NMR (126 MHz, DMSO-d6): δ = 130.5, 129.2,126.7, 126.9, 126.5. MS (ESI): calcd for C7H7F2O3P [M - H]- 207.01; found, 207.00.
Compound 3. Yield 17.9 mg (55%). 1H NMR (500 MHz, DMSO-d6) δ 8.11 (s, 1H), 7.71 (s, 1H), 7.41 (s, 1H). 13C NMR (126 MHz, DMSO-d6) δ 153.72, 140.22, 133.57, 124.72, 121.11, 113.33, 108.67. MS (ESI): calcd for C8H6BrF2N2O3P [M - H]- 324.93; found, 324.91.
Compound 4 Yield 13.1 mg (53%). 1H NMR (500 MHz, DMSO-d6) δ 8.06 (d, J = 2.2 Hz, 1H), 7.85 (s, 1H), 7.69 (d, J = 8.6 Hz, 1H), 7.48 (d, J = 8.6 Hz, 1H), 7.05 (d, J = 1.3 Hz, 1H). 13C NMR (126 MHz, DMSO-d6) δ 141.20, 139.42, 129.27, 124.77, 122.95, 122.30, 121.88. MS (ESI): calcd for C9H7F2O4P [M - H]- 247.01; found, 247.02.
Compound 5. Yield 16.1 mg (61%). 1H NMR (500 MHz, DMSO-d6) δ 8.09 (d, J = 8.4 Hz, 1H), 8.04 (s, 1H), 7.85 (d, J = 5.4 Hz, 1H), 7.57 (d, J = 5.4 Hz, 1H), 7.49 (d, J = 8.5 Hz, 1H). 13C NMR (126 MHz, DMSO-d6) δ 155.46, 147.65, 129.22, 127.46, 122.93, 120.04, 119.71, 111.55, 107.48. MS (ESI): calcd for C9H7F2O3PS [M - H]- 262.98; found, 262.99.
Compound 6. Yield 12.0 mg (41%). 1H NMR (500 MHz, DMSO-d6) δ 9.09 (d, J = 3.2 Hz, 1H), 8.62 (d, J = 8.0 Hz, 1H), 8.19 (s, 1H), 7.96 (s, 1H), 7.72 (dd, J = 8.2, 4.1 Hz, 1H). 13C NMR (126 MHz, DMSO-d6) δ 153.20, 144.29, 138.12, 133.09, 132.83, 128.91, 127.05, 126.19, 123.72. MS (ESI): calcd for C10H7ClF2NO3P [M - H]- 291.98; found, 291.97.
Compound 7. Yield 20.1 mg (51%). 1H NMR (500 MHz, DMSO-d6) δ 8.74 (d, J = 1.9 Hz, 1H), 8.56 (s, 1H), 8.16 (d, J = 8.4 Hz, 1H), 8.04 (d, J = 8.5 Hz, 1H), 7.76 – 7.62 (m, 2H). 13C NMR (126 MHz, DMSO-d6) δ 141.63, 138.48, 137.07, 134.07, 131.60, 130.56, 125.97, 125.60, 125.55, 123.52, 120.62, 120.35, 118.84. MS (ESI): calcd for C13H8BrF2O3PS [M - H]- 390.91; found, 390.90.
Compound 8. Yield 23.4 mg (57%). 1H NMR (500 MHz, DMSO-d6) δ 8.03 – 7.99 (m, 2H), 7.89 (d, J = 8.1 Hz, 1H), 7.84 (dd, J = 8.1, 1.8 Hz, 1H), 7.77 – 7.72 (m, 2H). 13C NMR (126 MHz, DMSO-d6) δ 140.21, 139.32, 137.58, 136.72, 136.46, 135.82, 131.45, 127.61, 124.53, 123.33, 122.59, 122.11, 121.87, 120.20. MS (ESI): calcd for C14H8BrF4O3P [M - H]- 408.93; found, 408.92.
Compound 9. Yield 18.2 mg (61%). 1H NMR (500 MHz, DMSO-d6) δ 8.28 (d, J = 8.2 Hz, 1H), 7.86 (d, J = 8.2 Hz, 1H), 7.70 (d, J = 8.2 Hz, 1H), 7.63 (t, J = 8.0 Hz, 1H), 7.57 – 7.47 (m, 2H), 7.40–7.34 (m, 1H). 13C NMR (126 MHz, DMSO-d6) δ 156.25, 156.20, 128.95, 128.51, 127.40, 126.13, 123.28, 122.57, 122.33, 121.85, 120.32, 114.35, 111.78. MS (ESI): calcd for C13H9F2O4P [M - H]- 297.02; found, 297.01.
Compound 10. Yield 20.3 mg (54%). 1H NMR (500 MHz, DMSO-d6) δ 8.27 (d, J = 7.0 Hz, 1H), 7.89 (d, J = 8.3 Hz, 1H), 7.81 (d, J = 8.3 Hz, 1H), 7.63 – 7.57 (m, 1H), 7.48 – 7.39 (m, 2H). 13C NMR (126 MHz, DMSO-d6) δ 156.11, 153.08, 130.10, 129.28, 128.48, 126.56, 123.93 (2C), 123.35, 122.44, 120.03, 112.13, 106.90. MS (ESI): calcd for C13H8BrF2O4P [M - H]- 374.93; found, 374.92.
General procedure for the synthesis of compound 11–16
Compound 10 (50mg, 0.13 mmol, 1.0 equiv.), boronic acid (0.20 mmol, 1.5 equiv.), K2CO3 (71.9 mg, 0.52 mmol, 4.0 equiv.) and Pd(PPh3)4 (11.6 mg, 0.01 mmol, 0.1 equiv.) were suspended in 2 mL of solvent (DMF: 1.8 mL, H2O: 0.2 mL) under argon atmosphere. The reaction mixture was heated at 90°C and the reaction progress was monitored with LC-MS. Upon completion, 1 ml of 1M HCl aqueous solution was added into the reaction mixture. The resulted mixture was purified by prep HPLC to afford the products.
Compound 11. Yield 23.8 mg (49%). 1H NMR (500 MHz, DMSO-d6) δ 8.34 (d, J = 8.0 Hz, 1H), 7.91 – 7.86 (m, 2H), 7.76 (d, J = 8.0 Hz, 1H), 7.72 (d, J = 8.2 Hz, 1H), 7.63 (d, J = 8.1 Hz, 1H), 7.61 – 7.52 (m, 3H), 7.49 (m, 1H), 7.39 (ddd, J = 8.2, 7.2, 1.1 Hz, 1H). 13C NMR (126 MHz, DMSO-d6) δ 162.79, 156.28, 153.21, 135.67, 129.33 (3C), 128.87, 128.61, 128.03, 127.56, 126.88, 126.29, 123.38, 123.14, 122.66, 122.36, 120.42, 111.89. MS (ESI): calcd for C19H13F2O4P [M - H]- 373.05; found, 373.03.
Compound 12. Yield 20.8 mg (41%). 1H NMR (500 MHz, DMSO-d6) δ 9.78 (brd, 1H), 8.32 (d, J = 8.0 Hz, 1H), 7.78 – 7.65 (m, 4H), 7.63 – 7.49 (m, 2H), 7.42 – 7.33 (m, 1H), 7.01 – 6.92 (m, 2H). 13C NMR (126 MHz, DMSO-d6) δ 162.79, 158.30, 156.22, 153.05, 130.55 (2C), 128.45, 127.68, 127.04, 126.25, 126.18, 126.13, 123.26, 123.11, 122.54, 122.46, 119,65, 116.16, 111.83. MS (ESI): calcd for C19H13F2O5P [M - H]- 389.05; found, 389.04.
Compound 13. Yield 23.3 mg (39%). 1H NMR (500 MHz, DMSO-d6) δ 8.33 (d, J = 7.8 Hz, 1H), 8.00 – 7.90 (m, 2H), 7.79 (d, J = 8.0 Hz, 1H), 7.73 (d, J = 7.7 Hz, 1H), 7.70 – 7.62 (m, 2H), 7.54 (t, J = 7.7 Hz, 1H), 7.40 (t, J = 7.6 Hz, 1H). 13C NMR (126 MHz, DMSO-d6) δ 160.62, 158.63, 156.28, 153.37, 133.63, 131.55, 129.84, 128.87, 128.22, 127.84, 127.72, 126.36, 124.88, 123.56, 122.82, 122.45, 122.19, 120.59, 114.15, 111.88. MS (ESI): calcd for C20H11F6O4P [M - H]- 459.03; found, 459.06.
Compound 14. Yield 24.2 mg (45%). 1H NMR (500 MHz, DMSO-d6) 1H NMR (500 MHz, DMSO) δ 8.37 (d, J = 8.0 Hz, 1H), 7.86 (d, J = 8.1 Hz, 1H), 7.82 (d, J = 4.0 Hz, 1H), 7.78 (d, J = 8.2 Hz, 1H), 7.58 – 7.52 (m, 2H), 7.37 (t, J = 7.6 Hz, 1H), 7.29 (d, J = 4.0 Hz, 1H). 13C NMR (126 MHz, DMSO-d6) δ 162.78, 156.15, 151.44, 136.01, 129.97, 128.69, 128.32, 127.21, 126.71, 123.90, 123.62, 123.29, 122.89, 122.33, 119.24, 111.82. MS (ESI): calcd for C17H10ClF2O4PS [M - H]- 412.97; found, 412.98.
Compound 15. Yield 21.3 mg (41%). 1H NMR (500 MHz, DMSO-d6) δ 8.33 (d, J = 8.0 Hz, 1H), 7.58 – 7.51 (m, 2H), 7.47 (t, J = 7.7 Hz, 1H), 7.42 (d, J = 7.7 Hz, 1H), 7.38 – 7.27 (m, 6H), 6.01 (s, 1H), 5.70 (s, 1H). 13C NMR (126 MHz, DMSO-d6) δ 156.11, 153.70, 143.80, 139.85, 129.01 (3C), 128.59, 128.56, 128.34, 127.93, 127.23 (3C), 126.20, 123.33, 122.70, 122.28, 120.34, 118.74, 111.79. MS (ESI): calcd for C21H15F2O4P [M - H]- 399.07; found, 399.07.
Compound 16 Yield 25.5 mg (49%). 1H NMR (500 MHz, DMSO-d6) δ 8.34 (d, J = 7.1 Hz, 1H), 7.87 (d, J = 7.4 Hz, 1H), 7.80 (d, J = 8.2 Hz, 1H), 7.76 – 7.70 (m, 3H), 7.63 (d, J = 16.6 Hz, 1H), 7.56 (s, 2H), 7.45 – 7.41 (m, 2H), 7.39 – 7.31 (m, 2H). 13C NMR (126 MHz, DMSO-d6) δ 156.14, 153.44, 137.21, 133.20, 129.35 (3C), 128.84, 128.50, 127.35 (3C), 126.39, 124.62, 123.94, 123.41, 122.99, 122.45, 121.84, 111.86. MS (ESI): calcd for C21H15F2O4P [M - H]- 399.07; found, 399.06.
Synthesis of compound 17.
Intermediate 10 (50 mg, 0.13 mmol, 1.0 equiv.), ethynylbenzene (16.3 mg, 0.16 mmol, 1.2 equiv.), CuI (1.9 mg, 0.01 mmol, 0.1 equiv.), Pd(PPh3)4 (11.6 mg, 0.01 mmol, 0.1 equiv.) and triethylamine (109 μL, 0.78 mmol, 6.0 equiv.) were suspended in 2 mL of DMF under argon atmosphere. The reaction mixture was heated at 90°C and the reaction progress was monitored with LC-MS. Upon completion, 1 ml of 1M HCl aqueous solution was added into the reaction mixture. The resulted mixture was purified by prep HPLC to afford the products. Yield 29.0 mg (56%). 1H NMR (500 MHz, DMSO-d6) δ 8.35 (d, J = 8.0 Hz, 1H), 7.85 – 7.77 (m, 2H), 7.70 – 7.64 (m, 2H), 7.59 (d, J = 7.1 Hz, 1H), 7.55 (d, J = 7.9 Hz, 1H), 7.51 – 7.44 (m, 3H), 7.40 (t, J = 7.6 Hz, 1H).13C NMR (126 MHz, DMSO-d6). δ 162.79, 156.09, 155.75, 133.06, 132.05 (3C), 129.79, 129.38 (3C), 128.55, 127.25, 123.36, 122.78, 122.32, 111.61, 108.02, 95.21, 84.02. MS (ESI): calcd for C21H13F2O4P [M - H]- 397.05; found, 397.04.
Synthesis of compound 18.
4-bromodibenzo[b,d]furan (50 mg, 0.20 mmol, 1.0 equiv.), (1-phenylvinyl)boronic acid (44.4 mg, 0.30 mmol, 1.5 equiv.), K2CO3 (41.5 mg, 0.30 mmol, 1.5 equiv.) and Pd(PPh3)4 (23.2 mg, 0.02 mmol, 0.1 equiv.) were suspended in 2 mL of solvent (DMF: 1.8 mL, H2O: 0.2 mL) under argon atmosphere. The reaction mixture was heated at 90°C and the reaction progress was monitored with LC-MS. Upon completion, 1 ml of 1M HCl aqueous solution was added into the reaction mixture. The resulted mixture was purified by prep HPLC to afford the products. Yield 38.4 mg (71%). 1H NMR (500 MHz, DMSO-d6) δ 7.57 (d, J = 8.3 Hz, 1H), 7.49 – 7.44 (m, 1H), 7.42 – 7.34 (m, 4H), 7.32 (d, J = 11.3 Hz, 5H), 7.22 (dt, J = 6.3, 2.2 Hz, 1H), 5.92 (s, 1H), 5.73 (s, 1H). 13C NMR (126 MHz, DMSO-d6) δ 155.79, 153.49, 144.35, 140.33, 133.80, 129.46, 129.27, 128.94, 128.63, 128.46, 128.14, 127.44, 125.86, 124.50, 123.93, 123.68, 121.67, 121.20, 118.17, 112.17. MS (ESI): calcd for C20H14O [M + H]+ 271.10; found, 271.11.
Recombinant Protein Production
The coding sequences of human CDC14A isoform 2 (codons 1–413) and CDC14B isoform 3 (codons 54–448) were codon-optimized for E. coli, synthesized, and cloned into an entry vector for Gateway cloning (Invitrogen) by Twist Biosciences. The coding sequences were sub-cloned into pDEST15 for expression as N-terminal glutathione-S-transferase (GST) fusion proteins using Gateway cloning. Mutations were made in pDEST15 vectors by inverse PCR using the In-Fusion cloning system (Takara Bio) as recommended by the supplier and confirmed by whole plasmid sequencing. GST-Cdc14 enzymes were expressed in 1 L 2xYT cultures (10 g/L yeast extract, 16 g/L tryptone, 5 g/L NaCl) of BL21-AI cells by induction with 0.1% L-arabinose for 18 hours at 25°C. Cells were lysed with 1 mg/mL lysozyme for 30 min on ice in 30 mL 25 mM HEPES pH 7.5, 500 mM NaCl, 2 mM EDTA, 0.1% Triton X-100, 10% glycerol, 1 mM PMSF, 10 μM leupeptin, 1 μM pepstatin, 5 μM bestatin, 1 mM benzamidine and 1,000 units Universal Nuclease (Thermo Fisher Scientific). Extracts were clarified by centrifugation at 35,000xg for 30 min at 4°C and the soluble fraction was loaded onto 0.5 mL bed volume of Glutathione Superflow Agarose (Thermo Fisher Scientific) equilibrated with 10 mL 25 mM HEPES pH 7.5, 500 mM NaCl, and 10% glycerol. The column was washed with 10 mL 25 mM HEPES pH 7.5, 500 mM NaCl, and 10% glycerol. Bound protein was subsequently eluted with 25 mM HEPES pH 7.5, 500 mM NaCl, 10% glycerol, and 20 mM reduced glutathione. Peak fractions determined by Bradford assay, were pooled, and dialyzed overnight into 25 mM HEPES pH 7.5, 300 mM NaCl, 2 mM EDTA, 0.1% 2-mercaptoethanol, and 40% glycerol prior to storage at −80°C in small aliquots.
The proteins for the selectivity panel (TC-PTP, Aa 1–387; PTP1B, Aa 1–321; SHP-1, Aa 245–543; SHP-2, Aa 224–528; LYP, Aa 1–294; STEP, Aa 258–539; HePTP, Aa 22–360; PTP-PEST, Aa 5–304; FAP-1, Aa 2124–2485; PTPα, Aa 173–793; PTPε, Aa 107–697; CD45, Aa 620–1236; mouse MKP5, Aa 320–647; VHZ, Aa 1–150; Larforin, Aa 1–331; LMPTP, Aa 1–158.) were cloned into pET-21a(+) vector. Bacterial BL21(DE3) (Novagen) was used as an expression host, and the induction of protein expression was carried out in LB media with 1 mM IPTG at 18 °C overnight. Cell pellets were stored at −80 °C for subsequent protein purification. Protein purification was conducted at 4 °C. Frozen cell pellets were lysed by sonication in 40 ml cold lysis buffer (50 mM Tris-HCl, pH 8.0, 150 mM NaCl, 5 mM imidazole, and 1 mM PMSF) per liter cell pellet. Cell lysates were clarified by centrifugation using a Bechman JA-18 rotor for 15 min at 6,000 rpm. The supernatant was incubated with HisPur Ni-NTA resin (Thermo Scientific) for 2 h, and then packed onto a column and washed with 50 resin volume of buffer A (50 mM Tris-HCl, pH 8.0, 500 mM NaCl, 5 mM imidazole). The HIS-tagged proteins were eluted with Buffer B (50 mM Tris-HCl, pH=8.0, 500 mM NaCl, 300 mM imidazole,). Pooled HIS-protein-containing fractions were concentrated, loaded onto a HiLoad 26/600 Superdex 75 column (GE Healthcare Biosciences), and eluted with storage buffer (50 mM Tris-HCl, pH 8.0, 150 mM NaCl, 1 mM DTT, 10% glycerol). Proteins used for inhibition assays were purified using Ni-NTA resin (Qiagen) followed by size exclusion column chromatography (ÄKTA pure, cytiva) and the purity was determined to be >95% by SDS-PAGE and Coomassie staining. The protein was aliquoted and stored at −80 °C.
Dynamic Light Scattering (DLS) Assay
In a 384-well, low-volume, non-treated, black with clear bottom assay plate (Corning 3540), compounds were dissolved in DMG buffer (50 mM DMG, pH 7.0, 1 mM EDTA, 18 mM NaCl, 0.01% Triton X-100) to reach a final volume of 30 μL and a concentration of 1 mM (15 μM for ketoconazole). The DMSO concentration was maintained at 1% (v/v) for all solutions. The bottom of the plate was wiped off using Kimwipe to eliminate interferences from dust and other particles. The plate was centrifuged using a ThermoFisher Scientific Heraeus Multifuge X3 centrifuge and incubated in the instrument for 30 min to ensure temperature equilibrium. The data was collected using DyanaPro Plate Reader II with the following parameter:
Number of times to measure each well: 1, Number of times to scan the selected regions: 1, Enable Auto-attenuation: Yes, Acquire an image of each well: Yes, Number of DLS acquisitions: 10, Temperature: 25°C, DLS Acquisition Time: 5 s.
The data cutoffs were set between 10 μs and 1 s with a filter to mask acquisitions with an autocorrelation intensity outside of the baseline limit of 1 ± 0.02. The data was analyzed and plotted using GraphPad Prism 9.2.0
IC50 and Ki Determination
PTP activity was assayed using 6,8-Difluoro-4-Methylumbelliferyl Phosphate (DiFMUP) as a substrate in 3,3-Dimethylglutaric acid (DMG) buffer (50 mM DMG, pH 7.0, 1 mM EDTA, 18 mM NaCl, 0.01% Triton X-100) at 25 °C. To determine the IC50 values, the assays were performed in 96-well plates (Corning Costar 3915). The reaction was initiated by the addition of enzyme (the final concentration was 2.5 nM) to a reaction mixture containing DiFMUP (final concentration close to Km values of each enzyme) and inhibitors (serially diluted by 2-fold/step for a total of 11 concentrations). The final reaction volume is 200 μL. The reaction was allowed to proceed for 10 minute and then quenched by the addition of 40 μL of a 160 μM solution of bpV(Phen). The fluorescence signal was measured using a CLARIOstar Plus Microplate Spectrophotometer (BMG Labtech) using excitation and emission wavelength of 340 nm and 450 nm, respectively. Data were fitted using Prism GraphPad 9.2.0 with the build in 4-parameter IC50 equation.
To determine the mode of inhibition, the reactions were initiated by the addition of enzyme (2.5 nM final concenrtation) to the reaction mixtures (0.2 mL) containing various concentrations of DiFMUP (serially diluted by 2-fold/step for a total of 8 concentrations) with inhibitor at the specified concentration.. The reaction was allowed to proceed for 10 minute and then quenched by the addition of 40 μL of a 160 μM solution of bpV(Phen). The fluorescence signal was measured using a CLARIOstar Plus Microplate Spectrophotometer (BMG Labtech) using excitation and emission wavelength of 340 nm and 450 nm, respectively. Data were fitted using Prism GraphPad 9.2.0 with the build in competitive inhibition equation.
Molecular Docking Studies
Molecular docking studies were performed on a previously reported hCDC14B crystal structure (PDB 1OHC23) using the Maestro suite version 2021–3 (Schrödinger, LLC, New York, NY). The protein was prepared using the Protein Preparation Workflow at pH 7.4 ± 1.0. Missing side chains and loops were filled using Prime. Hydrogen bond assignments were optimized using PROPKA, and the catalytic C314 was manually ionized via the 3D Builder module. Lastly, a restrained minimization was performed to 0.30Å RMSD using the OPLS4 force field, and all waters further than 5 Å from heteroatoms were removed. The receptor grid was generated using the Receptor Grid Generation module. The prepared 1OHC structure was aligned to PDB 1OHE23, and the coordinates of the 1OHE peptide ligand was used to define the grid site in 1OHC (using default settings). Ligand preparation was conducted using the LigPrep module with the OPSL4 force field. Ionization states were generated using Epik at pH 7.0 ± 1.0 Docking was performed using Glide in the standard precision (SP) mode using default settings. The top 10 poses were retained for visual evaluation.
Docking studies for compound 9 with other PTPs was conducted in a similar manner. Docking studies were performed on PTP1B (PDB 1PXH88), MKP-5 (1ZZW89), VHZ (2IMG90), CD45 (1YGU91), and Laforin (4RKK92). Prior to preparation, organic solvents and ions present from the crystallography experiments were removed. For 4RKK, the maltohexose chains and phosphates were removed from the crystal structure, and S266 was mutated to cysteine to mimic the enzymatically active protein. All proteins were subsequently prepared using the above protocol, and the catalytic cysteines for each PTP were manually ionized using the 3D Builder similar as above. With the exception of MKP-5 (1ZZW) and Laforin (4RKK), the receptor grid was generated under default settings using the co-crystallized ligand to define the grid. For MKP-5 (1ZZW), the grid was centered on residues A410 and N379. For Laforin (4RKK), the grid was centered on the XYZ coordinates 52.47, 71.22, and 122.78, and a hydrogen bond constraint was placed on the R272 side chain, which is known to interact with substrate phosphate groups. compound 9 was docked into each PTP using the standard precision (SP) setting of Glide, and up to 10 top poses were retained for visual evaluation.
Molecular Dynamics
Molecular dynamics were carried out using Desmond (Desmond Molecular Dynamics System, D. E. Shaw Research, New York, NY, 2021. Maestro-Desmond Interoperability Tools, Schrödinger, New York, NY, 2021.). The docking model of compound 15 bound to hCDC14B was set up for MD studies with the System Builder. The TIP3P solvent model was utilized, and an orthorhombic box shape was used when constructing the boundary conditions. The box size was calculated using buffer, the box volume was minimized, and the OPSL4 force field was utilized. Counterions were added to neutralize the system and were excluded from the region within 20Å of the ligand. The salt concentration used was 0.15M. A 50ns simulation time was used, with a trajectory of 100ps and approximately 500 frames. The simulation was run at 300K, and the interactions analysis was selected prior to the run.
Cell Permeability Studies
The cell permeability of tested compounds was determined with Corning® BioCoat® Pre-coated PAMPA Plate System as per the manufacturer’s guidance. In specific, the compound solutions were prepared by diluting 10 mM DMSO stock solutions in PBS (in most cases we used a final concentration of 200 μM). The compound solutions were added to the wells (300 μL/well) of the receiver plate and PBS was added to the wells (200 μL/well) of the pre-coated filter plate. The filter plate was then coupled with the receiver plate, and the plate assembly was incubated at room temperature without agitation for five hours. At the end of the incubation, the plates were separated, and 150 μL solution from each well of both the filter plate and the receiver plate was sampled. The final concentrations of compounds in both donor wells and acceptor wells were analyzed by LC-MS analysis. Permeability of the compounds was calculated using the following formula:
Permeability (cm/s):
whereas: A = filter area (0.3 cm2), = donor well volume (0.3 mL), = acceptor well volume (0.2 mL), t = incubation time (seconds), = compound concentration in acceptor well at time t, = compound concentration in donor well at time t.
Differential Scanning Fluorimetry (DSF) Assay
DSF experiments were performed in a Roche LC480 Light Cycler II qPCR machine. Solutions of protein (assay buffer: 50 mM DMG, pH 7.0, 1 mM EDTA, 18 mM NaCl), Sypro Orange (Thermo Fisher Scientific, S6650) and tested compounds or DMSO were added to the wells of a 96-well PCR plate (USA Scientific, 1402–9590). The final concentration for protein, compound, and Sypro Orange are 10 μM, 100 μM, and 15x, respectively. The plates were sealed with transparent films (EXCEL Scientific, TS-RT2–100) and heated in the qPCR machine from 25 to 95 °C in increments of 1.0 °C. The 96 well-plate assay data was analyzed using Roche LC480 Light Cycler II software.
Cell Culture
The U2OS and HEK293 cells were grown in DMEM cell culture media supplemented with 10% fetal bovine serum (no fetal bovine serum supplement for cytotoxicity assay), penicillin (50 units/mL), and streptomycin (50 μg/mL) in a 37°C incubator containing 5% CO2. The Jurkat cells were grown in RPMI 1640 cell culture media supplemented with 10% fetal bovine serum (no fetal bovine serum supplement for cytotoxicity assay), penicillin (50 units/mL), and streptomycin (50 μg/mL) in a 37°C incubator containing 5% CO2. 5 μM Mitomycin C was used as needed to induce DNA damage.
Immunoblotting
Cultured cells were lysed with ice cold lysis buffer (50 mM Tris (pH 8.0), 150 mM NaCl, 10% Glycerol, 1% Triton-X-100) supplied with phosphatase inhibitor and protease inhibitor mixture. Equal amounts of protein were resolved by SDS-PAGE, transferred to nitrocellulose membrane and subjected to immunoblotting.
Cytotoxicity Evaluation
For cytotoxicity test, 5,000 cells were plated onto 96-well plates in 100 μL medium and treated with compound 15 concentration varying from 0 to 50 μM. After 24 hours of incubation, the compound was washed off followed by addition of 100 μL fresh media and 10 μL cell counting kit 8 (CCK-8) reagent. OD470 was measured after 3 hours of incubation and used to calculate cytotoxicity.
PK Study
All the in vivo studies were performed under an animal protocol (1511001324) approved by the Institutional Animal Care & Use Committee of the Purdue University, in accordance with the recommendations in the Guide for the Care and Use of Laboratory Animals of the National Institutes of Health. For PK study, C57BL6 female mice (25–30g body weight) were injected intraperitoneally with 25 or 50 mg/kg DU-14 dissolved in 0.4 ml saline. Blood samples were collected through tail vein at indicated time points after injection. Isoflurane was used as an anesthetic. All blood samples were centrifuged at 1,500 g for 5 minutes, and plasma was separated and stored at –80°C until analysis by a validated method based on reversed-phase liquid chromatography coupled to mass-spectrometric detection (LC/MS) using a previously published procedure93.
Supplementary Material
Acknowledgements
This work was supported in part by NIH RO1CA069202 and the Robert C. and Charlotte Anderson Chair Endowment. M.C.H is supported by NIH 1R01AI168050. The authors gratefully acknowledge the support of the Chemical Genomics Facility at the Purdue Institute for Drug Discovery.
Abbreviations
- DiFMUP
6,8-Difluoro-4-Methylumbelliferyl Phosphate
- DLS
dynamic light scattering
- F2Pmp
phosphonodifluoromethyl phenylalanine
- hCDC14A
human Cell Division Cycle 14A
- hCDC14B
human Cell Division Cycle 14B
- IP
intraperitoneal
- MD
molecular dynamic
- PAMPA
parallel artificial membrane permeability assay
- PhFP
p-(difluorophenylmethyl) phosphonic acid
- PK
pharmacokinetics
- PO
oral administration
- PTKs
protein tyrosine kinases
- PTPs
protein tyrosine phosphatases
- pSer
phosphoserine
- pTyr
phosphotyrosine
- RMSD
root mean square deviation
- RMSF
root mean square fluctuation
- SD
standard deviation
- tPSA
topological polar surface area
Footnotes
The authors declare no competing financial interest.
Associated Content
- Docking of compound 9 with hCDC14B (from PDB entry 1OHC), CD45 (from PDB entry 1YGU), Laforin (from PDB entry 4RKK), and PTP1B (from PDB entry1PXH). Docking of compound 15 with hCDC14B (from PDB entry 1OHC).
- Table of molecular formula strings with biochemical data (CSV).
- Supplementary figures and tables, NMR descriptions for all reported compounds. 1H NMR and 13C NMR spectrums for all final compounds, LC-MS spectra of compound 9 and compound 15 (PDF).
Complete contact information is available at: https://pubs.acs.org/doi/10.1021/acs.jmedchem.xxx.
References
- (1).Zhang Z-Y Drugging the Undruggable: Therapeutic Potential of Targeting Protein Tyrosine Phosphatases. Acc. Chem. Res. 2017, 50 (1), 122–129. 10.1021/acs.accounts.6b00537. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (2).Stuible M; Doody KM; Tremblay ML PTP1B and TC-PTP: Regulators of Transformation and Tumorigenesis. Cancer Metastasis Rev 2008, 27 (2), 215–230. 10.1007/s10555-008-9115-1. [DOI] [PubMed] [Google Scholar]
- (3).Dodd GT; Xirouchaki CE; Eramo M; Mitchell CA; Andrews ZB; Henry BA; Cowley MA; Tiganis T Intranasal Targeting of Hypothalamic PTP1B and TCPTP Reinstates Leptin and Insulin Sensitivity and Promotes Weight Loss in Obesity. Cell Reports 2019, 28 (11), 2905–2922.e5. 10.1016/j.celrep.2019.08.019. [DOI] [PubMed] [Google Scholar]
- (4).Galic S; Hauser C; Kahn BB; Haj FG; Neel BG; Tonks NK; Tiganis T Coordinated Regulation of Insulin Signaling by the Protein Tyrosine Phosphatases PTP1B and TCPTP. MCB 2005, 25 (2), 819–829. 10.1128/MCB.25.2.819-829.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (5).Manguso RT; Pope HW; Zimmer MD; Brown FD; Yates KB; Miller BC; Collins NB; Bi K; LaFleur MW; Juneja VR; Weiss SA; Lo J; Fisher DE; Miao D; Van Allen E; Root DE; Sharpe AH; Doench JG; Haining WN In Vivo CRISPR Screening Identifies Ptpn2 as a Cancer Immunotherapy Target. Nature 2017, 547 (7664), 413–418. 10.1038/nature23270. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (6).LaFleur MW; Nguyen TH; Coxe MA; Miller BC; Yates KB; Gillis JE; Sen DR; Gaudiano EF; Al Abosy R; Freeman GJ; Haining WN; Sharpe AH PTPN2 Regulates the Generation of Exhausted CD8+ T Cell Subpopulations and Restrains Tumor Immunity. Nat Immunol 2019, 20 (10), 1335–1347. 10.1038/s41590-019-0480-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (7).Wiede F; Shields BJ; Chew SH; Kyparissoudis K; van Vliet C; Galic S; Tremblay ML; Russell SM; Godfrey DI; Tiganis T T Cell Protein Tyrosine Phosphatase Attenuates T Cell Signaling to Maintain Tolerance in Mice. J. Clin. Invest. 2011, 121 (12), 4758–4774. 10.1172/JCI59492. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (8).Chan G; Kalaitzidis D; Neel BG The Tyrosine Phosphatase Shp2 (PTPN11) in Cancer. Cancer Metastasis Rev 2008, 27 (2), 179–192. 10.1007/s10555-008-9126-y. [DOI] [PubMed] [Google Scholar]
- (9).Asmamaw MD; Shi X-J; Zhang L-R; Liu H-M A Comprehensive Review of SHP2 and Its Role in Cancer. Cell Oncol. 2022, 45 (5), 729–753. 10.1007/s13402-022-00698-1. [DOI] [PubMed] [Google Scholar]
- (10).Yu X; Sun J-P; He Y; Guo X; Liu S; Zhou B; Hudmon A; Zhang Z-Y Structure, Inhibitor, and Regulatory Mechanism of Lyp, a Lymphoid-Specific Tyrosine Phosphatase Implicated in Autoimmune Diseases. Proc. Natl. Acad. Sci. U.S.A. 2007, 104 (50), 19767–19772. 10.1073/pnas.0706233104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (11).Smyth D; Cooper JD; Collins JE; Heward JM; Franklyn JA; Howson JMM; Vella A; Nutland S; Rance HE; Maier L; Barratt BJ; Guja C; Ionescu-Tîrgovişte C; Savage DA; Dunger DB; Widmer B; Strachan DP; Ring SM; Walker N; Clayton DG; Twells RCJ; Gough SCL; Todd JA Replication of an Association Between the Lymphoid Tyrosine Phosphatase Locus ( LYP/PTPN22 ) With Type 1 Diabetes, and Evidence for Its Role as a General Autoimmunity Locus. Diabetes 2004, 53 (11), 3020–3023. 10.2337/diabetes.53.11.3020. [DOI] [PubMed] [Google Scholar]
- (12).Jassim BA; Lin J; Zhang Z-Y PTPN22: Structure, Function, and Developments in Inhibitor Discovery with Applications for Immunotherapy. Expert Opinion on Drug Discovery 2022, 17 (8), 825–837. 10.1080/17460441.2022.2084607. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (13).He R; Yu Z; Zhang R; Zhang Z Protein Tyrosine Phosphatases as Potential Therapeutic Targets. Acta Pharmacol Sin 2014, 35 (10), 1227–1246. 10.1038/aps.2014.80. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (14).Zhang Z-Y Protein Tyrosine Phosphatases: Structure and Function, Substrate Specificity, and Inhibitor Development. Annu. Rev. Pharmacol. Toxicol. 2002, 42 (1), 209–234. 10.1146/annurev.pharmtox.42.083001.144616. [DOI] [PubMed] [Google Scholar]
- (15).Ghattas M; Raslan N; Sadeq A; Al Sorkhy M; Atatreh N Druggability Analysis and Classification of Protein Tyrosine Phosphatase Active Sites. DDDT 2016, Volume 10, 3197–3209. 10.2147/DDDT.S111443. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (16).Tautz L; Critton DA; Grotegut S Protein Tyrosine Phosphatases: Structure, Function, and Implication in Human Disease. In Phosphatase Modulators; Millán JL, Ed.; Methods in Molecular Biology; Humana Press: Totowa, NJ, 2013; Vol. 1053, pp 179–221. 10.1007/978-1-62703-562-0_13. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (17).Wiesmann C; Barr KJ; Kung J; Zhu J; Erlanson DA; Shen W; Fahr BJ; Zhong M; Taylor L; Randal M; McDowell RS; Hansen SK Allosteric Inhibition of Protein Tyrosine Phosphatase 1B. Nat Struct Mol Biol 2004, 11 (8), 730–737. 10.1038/nsmb803. [DOI] [PubMed] [Google Scholar]
- (18).Krishnan N; Koveal D; Miller DH; Xue B; Akshinthala SD; Kragelj J; Jensen MR; Gauss C-M; Page R; Blackledge M; Muthuswamy SK; Peti W; Tonks NK Targeting the Disordered C Terminus of PTP1B with an Allosteric Inhibitor. Nat Chem Biol 2014, 10 (7), 558–566. 10.1038/nchembio.1528. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (19).Singh JP; Lin M-J; Hsu S-F; Peti W; Lee C-C; Meng T-C Crystal Structure of TCPTP Unravels an Allosteric Regulatory Role of Helix Α7 in Phosphatase Activity. Biochemistry 2021, 60 (51), 3856–3867. 10.1021/acs.biochem.1c00519. [DOI] [PubMed] [Google Scholar]
- (20).Chen Y-NP; LaMarche MJ; Chan HM; Fekkes P; Garcia-Fortanet J; Acker MG; Antonakos B; Chen CH-T; Chen Z; Cooke VG; Dobson JR; Deng Z; Fei F; Firestone B; Fodor M; Fridrich C; Gao H; Grunenfelder D; Hao H-X; Jacob J; Ho S; Hsiao K; Kang ZB; Karki R; Kato M; Larrow J; La Bonte LR; Lenoir F; Liu G; Liu S; Majumdar D; Meyer MJ; Palermo M; Perez L; Pu M; Price E; Quinn C; Shakya S; Shultz MD; Slisz J; Venkatesan K; Wang P; Warmuth M; Williams S; Yang G; Yuan J; Zhang J-H; Zhu P; Ramsey T; Keen NJ; Sellers WR; Stams T; Fortin PD Allosteric Inhibition of SHP2 Phosphatase Inhibits Cancers Driven by Receptor Tyrosine Kinases. Nature 2016, 535 (7610), 148–152. 10.1038/nature18621. [DOI] [PubMed] [Google Scholar]
- (21).Zhang Z-Y Mechanistic Studies on Protein Tyrosine Phosphatases. In Progress in Nucleic Acid Research and Molecular Biology; Elsevier, 2003; Vol. 73, pp 171–220. 10.1016/S0079-6603(03)01006-7. [DOI] [PubMed] [Google Scholar]
- (22).DeMarco AG; Milholland KL; Pendleton AL; Whitney JJ; Zhu P; Wesenberg DT; Nambiar M; Pepe A; Paula S; Chmielewski J; Wisecaver JH; Tao WA; Hall MC Conservation of Cdc14 Phosphatase Specificity in Plant Fungal Pathogens: Implications for Antifungal Development. Sci Rep 2020, 10 (1), 12073. 10.1038/s41598-020-68921-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (23).Gray CH The Structure of the Cell Cycle Protein Cdc14 Reveals a Proline-Directed Protein Phosphatase. The EMBO Journal 2003, 22 (14), 3524–3535. 10.1093/emboj/cdg348. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (24).Kobayashi J; Matsuura Y Structure and Dimerization of the Catalytic Domain of the Protein Phosphatase Cdc14p, a Key Regulator of Mitotic Exit in Saccharomyces Cerevisiae. Protein Science 2017, 26 (10), 2105–2112. 10.1002/pro.3244. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (25).Bremmer SC; Hall H; Martinez JS; Eissler CL; Hinrichsen TH; Rossie S; Parker LL; Hall MC; Charbonneau H Cdc14 Phosphatases Preferentially Dephosphorylate a Subset of Cyclin-Dependent Kinase (Cdk) Sites Containing Phosphoserine. Journal of Biological Chemistry 2012, 287 (3), 1662–1669. 10.1074/jbc.M111.281105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (26).Li L; Ernsting BR; Wishart MJ; Lohse DL; Dixon JE A Family of Putative Tumor Suppressors Is Structurally and Functionally Conserved in Humans and Yeast. Journal of Biological Chemistry 1997, 272 (47), 29403–29406. 10.1074/jbc.272.47.29403. [DOI] [PubMed] [Google Scholar]
- (27).Rosso L; Marques AC; Weier M; Lambert N; Lambot M-A; Vanderhaeghen P; Kaessmann H Birth and Rapid Subcellular Adaptation of a Hominoid-Specific CDC14 Protein. PLoS Biol 2008, 6 (6), e140. 10.1371/journal.pbio.0060140. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (28).Kaiser BK; Zimmerman ZA; Charbonneau H; Jackson PK Disruption of Centrosome Structure, Chromosome Segregation, and Cytokinesis by Misexpression of Human Cdc14A Phosphatase. MBoC 2002, 13 (7), 2289–2300. 10.1091/mbc.01-11-0535. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (29).Mailand N; Lukas C; Kaiser BK; Jackson PK; Bartek J; Lukas J Deregulated Human Cdc14A Phosphatase Disrupts Centrosome Separation and Chromosome Segregation. Nat Cell Biol 2002, 4 (4), 318–322. 10.1038/ncb777. [DOI] [PubMed] [Google Scholar]
- (30).Cho HP; Liu Y; Gomez M; Dunlap J; Tyers M; Wang Y The Dual-Specificity Phosphatase CDC14B Bundles and Stabilizes Microtubules. Molecular and Cellular Biology 2005, 25 (11), 4541–4551. 10.1128/MCB.25.11.4541-4551.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (31).Wu J; Cho HP; Rhee DB; Johnson DK; Dunlap J; Liu Y; Wang Y Cdc14B Depletion Leads to Centriole Amplification, and Its Overexpression Prevents Unscheduled Centriole Duplication. The Journal of Cell Biology 2008, 181 (3), 475–483. 10.1083/jcb.200710127. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (32).Nalepa G; Harper JW Visualization of a Highly Organized Intranuclear Network of Filaments in Living Mammalian Cells. Cell Motil. Cytoskeleton 2004, 59 (2), 94–108. 10.1002/cm.20023. [DOI] [PubMed] [Google Scholar]
- (33).Dryden SC; Nahhas FA; Nowak JE; Goustin A-S; Tainsky MA Role for Human SIRT2 NAD-Dependent Deacetylase Activity in Control of Mitotic Exit in the Cell Cycle. Molecular and Cellular Biology 2003, 23 (9), 3173–3185. 10.1128/MCB.23.9.3173-3185.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (34).Partscht P; Uddin B; Schiebel E Human Cells Lacking CDC14A and CDC14B Show Differences in Ciliogenesis but Not in Mitotic Progression. Journal of Cell Science 2021, 134 (2), jcs255950. 10.1242/jcs.255950. [DOI] [PubMed] [Google Scholar]
- (35).Villarroya-Beltri C; Martins AFB; García A; Giménez D; Zarzuela E; Novo M; Del Álamo C; González-Martínez J; Bonel-Pérez GC; Díaz I; Guillamot M; Chiesa M; Losada A; Graña-Castro O; Rovira M; Muñoz J; Salazar-Roa M; Malumbres M Mammalian CDC14 Phosphatases Control Exit from Stemness in Pluripotent Cells. The EMBO Journal 2023, 42 (1), e111251. 10.15252/embj.2022111251. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (36).Imtiaz A; Belyantseva IA; Beirl AJ; Fenollar-Ferrer C; Bashir R; Bukhari I; Bouzid A; Shaukat U; Azaiez H; Booth KT; Kahrizi K; Najmabadi H; Maqsood A; Wilson EA; Fitzgerald TS; Tlili A; Olszewski R; Lund M; Chaudhry T; Rehman AU; Starost MF; Waryah AM; Hoa M; Dong L; Morell RJ; Smith RJH; Riazuddin S; Masmoudi S; Kindt KS; Naz S; Friedman TB CDC14A Phosphatase Is Essential for Hearing and Male Fertility in Mouse and Human. Human Molecular Genetics 2018, 27 (5), 780–798. 10.1093/hmg/ddx440. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (37).Li L; Ljungman M; Dixon JE The Human Cdc14 Phosphatases Interact with and Dephosphorylate the Tumor Suppressor Protein P53. Journal of Biological Chemistry 2000, 275 (4), 2410–2414. 10.1074/jbc.275.4.2410. [DOI] [PubMed] [Google Scholar]
- (38).Fogal V; Hsieh J-K; Royer C; Zhong S; Lu X Cell Cycle-Dependent Nuclear Retention of P53 by E2F1 Requires Phosphorylation of P53 at Ser315. EMBO J 2005, 24 (15), 2768–2782. 10.1038/sj.emboj.7600735. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (39).Zheng H; You H; Zhou XZ; Murray SA; Uchida T; Wulf G; Gu L; Tang X; Lu KP; Xiao Z-XJ The Prolyl Isomerase Pin1 Is a Regulator of P53 in Genotoxic Response. Nature 2002, 419 (6909), 849–853. 10.1038/nature01116. [DOI] [PubMed] [Google Scholar]
- (40).Zacchi P; Gostissa M; Uchida T; Salvagno C; Avolio F; Volinia S; Ronai Z; Blandino G; Schneider C; Sal GD The Prolyl Isomerase Pin1 Reveals a Mechanism to Control P53 Functions after Genotoxic Insults. Nature 2002, 419 (6909), 853–857. 10.1038/nature01120. [DOI] [PubMed] [Google Scholar]
- (41).Blaydes JP; Luciani MG; Pospisilova S; Ball HM-L; Vojtesek B; Hupp TR Stoichiometric Phosphorylation of Human P53 at Ser315Stimulates P53-Dependent Transcription. Journal of Biological Chemistry 2001, 276 (7), 4699–4708. 10.1074/jbc.M003485200. [DOI] [PubMed] [Google Scholar]
- (42).Slee EA; Benassi B; Goldin R; Zhong S; Ratnayaka I; Blandino G; Lu X Phosphorylation of Ser312 Contributes to Tumor Suppression by P53 in Vivo. Proc. Natl. Acad. Sci. U.S.A. 2010, 107 (45), 19479–19484. 10.1073/pnas.1005165107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (43).Chen N-P; Uddin B; Voit R; Schiebel E Human Phosphatase CDC14A Is Recruited to the Cell Leading Edge to Regulate Cell Migration and Adhesion. Proc. Natl. Acad. Sci. U.S.A. 2016, 113 (4), 990–995. 10.1073/pnas.1515605113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (44).Chen N-P; Uddin B; Hardt R; Ding W; Panic M; Lucibello I; Kammerer P; Ruppert T; Schiebel E Human Phosphatase CDC14A Regulates Actin Organization through Dephosphorylation of Epithelial Protein Lost in Neoplasm. Proc. Natl. Acad. Sci. U.S.A. 2017, 114 (20), 5201–5206. 10.1073/pnas.1619356114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (45).Zhu C; Zhao Y; Zheng W CDC14B Is a Favorable Biomarker for Recurrence and Prognosis of GBM. Clinical Neurology and Neurosurgery 2023, 227, 107665. 10.1016/j.clineuro.2023.107665. [DOI] [PubMed] [Google Scholar]
- (46).Mocciaro A; Schiebel E Cdc14: A Highly Conserved Family of Phosphatases with Non-Conserved Functions? Journal of Cell Science 2010, 123 (17), 2867–2876. 10.1242/jcs.074815. [DOI] [PubMed] [Google Scholar]
- (47).Ramos F; Villoria MT; Alonso-Rodríguez E; Clemente-Blanco A Role of Protein Phosphatases PP1, PP2A, PP4 and Cdc14 in the DNA Damage Response. CST 2019, 3 (3), 70–85. 10.15698/cst2019.03.178. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (48).Lin H; Ha K; Lu G; Fang X; Cheng R; Zuo Q; Zhang P Cdc14A and Cdc14B Redundantly Regulate DNA Double-Strand Break Repair. Molecular and Cellular Biology 2015, 35 (21), 3657–3668. 10.1128/MCB.00233-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (49).Liu Y; Zheng C; Huang Y; He M; Xu WW; Li B Molecular Mechanisms of Chemo- and Radiotherapy Resistance and the Potential Implications for Cancer Treatment. MedComm 2021, 2 (3), 315–340. 10.1002/mco2.55. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (50).Mladenov E; Magin S; Soni A; Iliakis G DNA Double-Strand Break Repair as Determinant of Cellular Radiosensitivity to Killing and Target in Radiation Therapy. Front. Oncol. 2013, 3. 10.3389/fonc.2013.00113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (51).Puius YA; Zhao Y; Sullivan M; Lawrence DS; Almo SC; Zhang Z-Y Identification of a Second Aryl Phosphate-Binding Site in Protein-Tyrosine Phosphatase 1B: A Paradigm for Inhibitor Design. Proc. Natl. Acad. Sci. U.S.A. 1997, 94 (25), 13420–13425. 10.1073/pnas.94.25.13420. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (52).Zhang ZY; Thieme-Sefler AM; Maclean D; McNamara DJ; Dobrusin EM; Sawyer TK; Dixon JE Substrate Specificity of the Protein Tyrosine Phosphatases. Proc. Natl. Acad. Sci. U.S.A. 1993, 90 (10), 4446–4450. 10.1073/pnas.90.10.4446. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (53).Zhang ZY; Maclean D; McNamara DJ; Sawyer TK; Dixon JE Protein Tyrosine Phosphatase Substrate Specificity: Size and Phosphotyrosine Positioning Requirements in Peptide Substrates. Biochemistry 1994, 33 (8), 2285–2290. 10.1021/bi00174a040. [DOI] [PubMed] [Google Scholar]
- (54).Jia Z; Barford D; Flint AJ; Tonks NK Structural Basis for Phosphotyrosine Peptide Recognition by Protein Tyrosine Phosphatase 1B. Science 1995, 268 (5218), 1754–1758. 10.1126/science.7540771. [DOI] [PubMed] [Google Scholar]
- (55).Wu L; Buist A; Den Hertog J; Zhang Z-Y Comparative Kinetic Analysis and Substrate Specificity of the Tandem Catalytic Domains of the Receptor-like Protein-Tyrosine Phosphatase α. Journal of Biological Chemistry 1997, 272 (11), 6994–7002. 10.1074/jbc.272.11.6994. [DOI] [PubMed] [Google Scholar]
- (56).Sarmiento M; Zhao Y; Gordon SJ; Zhang Z-Y Molecular Basis for Substrate Specificity of Protein-Tyrosine Phosphatase 1B. Journal of Biological Chemistry 1998, 273 (41), 26368–26374. 10.1074/jbc.273.41.26368. [DOI] [PubMed] [Google Scholar]
- (57).Zhang X; He Y; Liu S; Yu Z; Jiang Z-X; Yang Z; Dong Y; Nabinger SC; Wu L; Gunawan AM; Wang L; Chan RJ; Zhang Z-Y Salicylic Acid Based Small Molecule Inhibitor for the Oncogenic Src Homology-2 Domain Containing Protein Tyrosine Phosphatase-2 (SHP2). J. Med. Chem. 2010, 53 (6), 2482–2493. 10.1021/jm901645u. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (58).Black E; Breed J; Breeze AL; Embrey K; Garcia R; Gero TW; Godfrey L; Kenny PW; Morley AD; Minshull CA; Pannifer AD; Read J; Rees A; Russell DJ; Toader D; Tucker J Structure-Based Design of Protein Tyrosine Phosphatase-1B Inhibitors. Bioorganic & Medicinal Chemistry Letters 2005, 15 (10), 2503–2507. 10.1016/j.bmcl.2005.03.068. [DOI] [PubMed] [Google Scholar]
- (59).Zhang S; Chen L; Luo Y; Gunawan A; Lawrence DS; Zhang Z-Y Acquisition of a Potent and Selective TC-PTP Inhibitor via a Stepwise Fluorophore-Tagged Combinatorial Synthesis and Screening Strategy. J. Am. Chem. Soc. 2009, 131 (36), 13072–13079. 10.1021/ja903733z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (60).Zhang S; Liu S; Tao R; Wei D; Chen L; Shen W; Yu Z-H; Wang L; Jones DR; Dong XC; Zhang Z-Y A Highly Selective and Potent PTP-MEG2 Inhibitor with Therapeutic Potential for Type 2 Diabetes. J. Am. Chem. Soc. 2012, 134 (43), 18116–18124. 10.1021/ja308212y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (61).Zeng L-F; Zhang R-Y; Yu Z-H; Li S; Wu L; Gunawan AM; Lane BS; Mali RS; Li X; Chan RJ; Kapur R; Wells CD; Zhang Z-Y Therapeutic Potential of Targeting the Oncogenic SHP2 Phosphatase. J. Med. Chem. 2014, 57 (15), 6594–6609. 10.1021/jm5006176. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (62).He R; Wang J; Yu Z-H; Moyers JS; Michael MD; Durham TB; Cramer JW; Qian Y; Lin A; Wu L; Noinaj N; Barrett DG; Zhang Z-Y Structure-Based Design of Active-Site-Directed, Highly Potent, Selective, and Orally Bioavailable Low-Molecular-Weight Protein Tyrosine Phosphatase Inhibitors. J. Med. Chem. 2022, 65 (20), 13892–13909. 10.1021/acs.jmedchem.2c01143. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (63).Kym PR Protein Tyrosine Phosphatase Inhibitors and Methods of Use Thereof. WO2022056281A1.
- (64).Baumgartner CK; Ebrahimi-Nik H; Iracheta-Vellve A; Hamel KM; Olander KE; Davis TGR; McGuire KA; Halvorsen GT; Avila OI; Patel CH; Kim SY; Kammula AV; Muscato AJ; Halliwill K; Geda P; Klinge KL; Xiong Z; Duggan R; Mu L; Yeary MD; Patti JC; Balon TM; Mathew R; Backus C; Kennedy DE; Chen A; Longenecker K; Klahn JT; Hrusch CL; Krishnan N; Hutchins CW; Dunning JP; Bulic M; Tiwari P; Colvin KJ; Chuong CL; Kohnle IC; Rees MG; Boghossian A; Ronan M; Roth JA; Wu M-J; Suermondt JSMT; Knudsen NH; Cheruiyot CK; Sen DR; Griffin GK; Golub TR; El-Bardeesy N; Decker JH; Yang Y; Guffroy M; Fossey S; Trusk P; Sun I-M; Liu Y; Qiu W; Sun Q; Paddock MN; Farney EP; Matulenko MA; Beauregard C; Frost JM; Yates KB; Kym PR; Manguso RT The PTPN2/PTPN1 Inhibitor ABBV-CLS-484 Unleashes Potent Anti-Tumour Immunity. Nature 2023, 622 (7984), 850–862. 10.1038/s41586-023-06575-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (65).Burke T Design and Synthesis of Phosphonodifluoromethyl Phenylalanine (F2Pmp):A Useful Phosphotyrosyl Mimetic. CTMC 2006, 6 (14), 1465–1471. 10.2174/156802606777951091. [DOI] [PubMed] [Google Scholar]
- (66).Chen L; Wu L; Otaka A; Smyth MS; Roller PP; Burke TR; Denhertog J; Zhang ZY Why Is Phosphonodifluoromethyl Phenylalanine a More Potent Inhibitory Moiety Than Phosphonomethyl Phenylalanine Toward Protein-Tyrosine Phosphatases. Biochemical and Biophysical Research Communications 1995, 216 (3), 976–984. 10.1006/bbrc.1995.2716. [DOI] [PubMed] [Google Scholar]
- (67).Meyer C; Hoeger B; Temmerman K; Tatarek-Nossol M; Pogenberg V; Bernhagen J; Wilmanns M; Kapurniotu A; Köhn M Development of Accessible Peptidic Tool Compounds To Study the Phosphatase PTP1B in Intact Cells. ACS Chem. Biol. 2014, 9 (3), 769–776. 10.1021/cb400903u. [DOI] [PubMed] [Google Scholar]
- (68).Liao H; Pei D Cell-Permeable Bicyclic Peptidyl Inhibitors against T-Cell Protein Tyrosine Phosphatase from a Combinatorial Library. Org. Biomol. Chem. 2017, 15 (45), 9595–9598. 10.1039/C7OB02562A. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (69).Shen K; Keng Y-F; Wu L; Guo X-L; Lawrence DS; Zhang Z-Y Acquisition of a Specific and Potent PTP1B Inhibitor from a Novel Combinatorial Library and Screening Procedure. Journal of Biological Chemistry 2001, 276 (50), 47311–47319. 10.1074/jbc.M106568200. [DOI] [PubMed] [Google Scholar]
- (70).Wang W-Q; Bembenek J; Gee KR; Yu H; Charbonneau H; Zhang Z-Y Kinetic and Mechanistic Studies of a Cell Cycle Protein Phosphatase Cdc14. Journal of Biological Chemistry 2004, 279 (29), 30459–30468. 10.1074/jbc.M402217200. [DOI] [PubMed] [Google Scholar]
- (71).Sarmiento M; Wu L; Keng Y-F; Song L; Luo Z; Huang Z; Wu G-Z; Yuan AK; Zhang Z-Y Structure-Based Discovery of Small Molecule Inhibitors Targeted to Protein Tyrosine Phosphatase 1B. J. Med. Chem. 2000, 43 (2), 146–155. 10.1021/jm990329z. [DOI] [PubMed] [Google Scholar]
- (72).Liang F; Huang Z; Lee S-Y; Liang J; Ivanov MI; Alonso A; Bliska JB; Lawrence DS; Mustelin T; Zhang Z-Y Aurintricarboxylic Acid Blocks in Vitro and in Vivo Activity of YopH, an Essential Virulent Factor of Yersinia Pestis, the Agent of Plague. Journal of Biological Chemistry 2003, 278 (43), 41734–41741. 10.1074/jbc.M307152200. [DOI] [PubMed] [Google Scholar]
- (73).Ghattas MA; Al Rawashdeh S; Atatreh N; Bryce RA How Do Small Molecule Aggregates Inhibit Enzyme Activity? A Molecular Dynamics Study. J. Chem. Inf. Model. 2020, 60 (8), 3901–3909. 10.1021/acs.jcim.0c00540. [DOI] [PubMed] [Google Scholar]
- (74).Allen SJ; Dower CM; Liu AX; Lumb KJ Detection of Small-Molecule Aggregation with High-Throughput Microplate Biophysical Methods. Current Protocols in Chemical Biology 2020, 12 (1). 10.1002/cpch.78. [DOI] [PubMed] [Google Scholar]
- (75).Allen SJ; Dower CM; Liu AX; Lumb KJ Detection of Small-Molecule Aggregation with High-Throughput Microplate Biophysical Methods. Current Protocols in Chemical Biology 2020, 12 (1), e78. 10.1002/cpch.78. [DOI] [PubMed] [Google Scholar]
- (76).Chen X; Murawski A; Patel K; Crespi CL; Balimane PV A Novel Design of Artificial Membrane for Improving the PAMPA Model. Pharm Res 2008, 25 (7), 1511–1520. 10.1007/s11095-007-9517-8. [DOI] [PubMed] [Google Scholar]
- (77).Senaweera S; Edwards TC; Kankanala J; Wang Y; Sahani RL; Xie J; Geraghty RJ; Wang Z Discovery of N-Benzyl Hydroxypyridone Carboxamides as a Novel and Potent Antiviral Chemotype against Human Cytomegalovirus (HCMV). Acta Pharmaceutica Sinica B 2022, 12 (4), 1671–1684. 10.1016/j.apsb.2021.08.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (78).Friesner RA; Banks JL; Murphy RB; Halgren TA; Klicic JJ; Mainz DT; Repasky MP; Knoll EH; Shelley M; Perry JK; Shaw DE; Francis P; Shenkin PS Glide: A New Approach for Rapid, Accurate Docking and Scoring. 1. Method and Assessment of Docking Accuracy. J. Med. Chem. 2004, 47 (7), 1739–1749. 10.1021/jm0306430. [DOI] [PubMed] [Google Scholar]
- (79).Yuvaniyama J; Denu JM; Dixon JE; Saper MA Crystal Structure of the Dual Specificity Protein Phosphatase VHR. Science 1996, 272 (5266), 1328–1331. 10.1126/science.272.5266.1328. [DOI] [PubMed] [Google Scholar]
- (80).Camps M; Nichols A; Arkinstall S Dual Specificity Phosphatases: A Gene Family for Control of MAP Kinase Function. FASEB j. 2000, 14 (1), 6–16. 10.1096/fasebj.14.1.6. [DOI] [PubMed] [Google Scholar]
- (81).Jeong DG; Wei CH; Ku B; Jeon TJ; Chien PN; Kim JK; Park SY; Hwang HS; Ryu SY; Park H; Kim D-S; Kim SJ; Ryu SE The Family-Wide Structure and Function of Human Dual-Specificity Protein Phosphatases. Acta Crystallogr D Biol Crystallogr 2014, 70 (2), 421–435. 10.1107/S1399004713029866. [DOI] [PubMed] [Google Scholar]
- (82).Feng BY; Shoichet BK A Detergent-Based Assay for the Detection of Promiscuous Inhibitors. Nat Protoc 2006, 1 (2), 550–553. 10.1038/nprot.2006.77. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (83).Huyer G; Liu S; Kelly J; Moffat J; Payette P; Kennedy B; Tsaprailis G; Gresser MJ; Ramachandran C Mechanism of Inhibition of Protein-Tyrosine Phosphatases by Vanadate and Pervanadate. Journal of Biological Chemistry 1997, 272 (2), 843–851. 10.1074/jbc.272.2.843. [DOI] [PubMed] [Google Scholar]
- (84).Yang X-G; Wang K Chemical, Biochemical, and Biological Behaviors of Vanadate and Its Oligomers. In Biomedical Inorganic Polymers; Müller WEG, Wang X, Schröder HC, Eds.; Progress in Molecular and Subcellular Biology; Springer Berlin Heidelberg: Berlin, Heidelberg, 2013; Vol. 54, pp 1–18. 10.1007/978-3-642-41004-8_1. [DOI] [PubMed] [Google Scholar]
- (85).Meyer C; Köhn M Efficient Scaled-Up Synthesis of N-a-Fmoc-4-Phosphono(Difluoromethyl)-L- Phenylalanine and Its Incorporation into Peptides. New York 2011, No. 20, 6. [Google Scholar]
- (86).Workman P; Collins I Probing the Probes: Fitness Factors For Small Molecule Tools. Chemistry & Biology 2010, 17 (6), 561–577. 10.1016/j.chembiol.2010.05.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (87).Arrowsmith CH; Audia JE; Austin C; Baell J; Bennett J; Blagg J; Bountra C; Brennan PE; Brown PJ; Bunnage ME; Buser-Doepner C; Campbell RM; Carter AJ; Cohen P; Copeland RA; Cravatt B; Dahlin JL; Dhanak D; Edwards AM; Frederiksen M; Frye SV; Gray N; Grimshaw CE; Hepworth D; Howe T; Huber KVM; Jin J; Knapp S; Kotz JD; Kruger RG; Lowe D; Mader MM; Marsden B; Mueller-Fahrnow A; Müller S; O’Hagan RC; Overington JP; Owen DR; Rosenberg SH; Ross R; Roth B; Schapira M; Schreiber SL; Shoichet B; Sundström M; Superti-Furga G; Taunton J; Toledo-Sherman L; Walpole C; Walters MA; Willson TM; Workman P; Young RN; Zuercher WJ The Promise and Peril of Chemical Probes. Nat Chem Biol 2015, 11 (8), 536–541. 10.1038/nchembio.1867. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (88).Sun J-P; Fedorov AA; Lee S-Y; Guo X-L; Shen K; Lawrence DS; Almo SC; Zhang Z-Y Crystal Structure of PTP1B Complexed with a Potent and Selective Bidentate Inhibitor. Journal of Biological Chemistry 2003, 278 (14), 12406–12414. 10.1074/jbc.M212491200. [DOI] [PubMed] [Google Scholar]
- (89).Jeong DG; Yoon T-S; Kim JH; Shim MY; Jung S-K; Son JH; Ryu SE; Kim SJ Crystal Structure of the Catalytic Domain of Human MAP Kinase Phosphatase 5: Structural Insight into Constitutively Active Phosphatase. Journal of Molecular Biology 2006, 360 (5), 946–955. 10.1016/j.jmb.2006.05.059. [DOI] [PubMed] [Google Scholar]
- (90).Agarwal R; Burley SK; Swaminathan S Structure of Human Dual Specificity Protein Phosphatase 23, VHZ, Enzyme-Substrate/Product Complex*. Journal of Biological Chemistry 2008, 283 (14), 8946–8953. 10.1074/jbc.M708945200. [DOI] [PubMed] [Google Scholar]
- (91).Nam H-J; Poy F; Saito H; Frederick CA Structural Basis for the Function and Regulation of the Receptor Protein Tyrosine Phosphatase CD45. Journal of Experimental Medicine 2005, 201 (3), 441–452. 10.1084/jem.20041890. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (92).Raththagala M; Brewer MK; Parker MW; Sherwood AR; Wong BK; Hsu S; Bridges TM; Paasch BC; Hellman LM; Husodo S; Meekins DA; Taylor AO; Turner BD; Auger KD; Dukhande VV; Chakravarthy S; Sanz P; Woods VL; Li S; Vander Kooi CW; Gentry MS Structural Mechanism of Laforin Function in Glycogen Dephosphorylation and Lafora Disease. Molecular Cell 2015, 57 (2), 261–272. 10.1016/j.molcel.2014.11.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- (93).Leblanc AF; Sprowl JA; Alberti P; Chiorazzi A; Arnold WD; Gibson AA; Hong KW; Pioso MS; Chen M; Huang KM; Chodisetty V; Costa O; Florea T; de Bruijn P; Mathijssen RH; Reinbolt RE; Lustberg MB; Sucheston-Campbell LE; Cavaletti G; Sparreboom A; Hu S OATP1B2 Deficiency Protects against Paclitaxel-Induced Neurotoxicity. Journal of Clinical Investigation 2018, 128 (2), 816–825. 10.1172/JCI96160. [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.
