Improved Preparation of a Key Hydroxylamine Intermediate for

Feb 1, 2018 - ... Johanne Blais, Anthony Casarez, Richard Colvin, Charles R. Dean, Markus Furegati, Luis Gamboa, Ellena Growcott, Cindy Li, Sara Lopez...
0 downloads 0 Views 1MB Size
Article pubs.acs.org/OPRD

Cite This: Org. Process Res. Dev. 2018, 22, 273−277

Improved Preparation of a Key Hydroxylamine Intermediate for Relebactam: Rate Enhancement of Benzyl Ether Hydrogenolysis with DABCO Jianguo Yin, Mark Weisel, Yining Ji, Zhijian Liu, Jinchu Liu, Debra J. Wallace, Feng Xu, Benjamin D. Sherry, and Nobuyoshi Yasuda* Process R&D Department, MRL, Merck & Co., Inc., Rahway, New Jersey 07065, United States S Supporting Information *

ABSTRACT: Previous methods to prepare a bicyclic N-hydroxyl urea intermediate in the synthesis of the potent β-lactamase inhibitor relebactam were effective, but deemed unsuitable for long-term use. Therefore, we developed an in situ protection protocol during hydrogenolysis and a robust deprotection/isolation sequence of this unstable intermediate employing a reactive crystallization. During the hydrogenation studies, we discovered a significant rate enhancement of O-benzyl ether hydrogenolysis in the presence of organic amine bases, especially DABCO. The broader utility of the application of organic bases on the hydrogenolysis of a range of O- and N-benzyl-containing substrates was demonstrated.



INTRODUCTION Infectious disease caused by multidrug resistant bacteria has become one of the most serious life threatening health problems. The increasing resistance of bacteria, especially Gram-negative species, is rooted in their secretion of βlactamases, which selectively hydrolyze β-lactam antibiotics.1 Relebactam 1 (Scheme 1) is currently being evaluated in Phase III clinical trials as a potent β-lactamase inhibitor to restore antibacterial activities of β-lactam antibiotics.2 Large scale preparation of relebactam presented a multitude of synthetic challenges, many stemming from the highly strained diazabicyclooctanone framework. Formation and manipulation of the N-hydroxyl bicyclic urea 2 were particularly complex due to its reactivity, as exemplified in the process development of avibactam which shares a similar substructure.3 We previously reported two methods (Scheme 1) for the preparation of 2.4 The first4a described hydrogenolysis of N-Cbz protected compound 3 in the presence of Pd(OH)2/C and Boc2O in THF, affording 87% of crystalline 2 (Scheme 1a). The use of THF as a reaction solvent was critical to maintain 2 in solution during the hydrogenolysis reaction. The second4b approach did not require a protecting group switch and afforded a 90% yield of 2 (Scheme 1b); however, as we developed additional experience at pilot scale with each of these processes certain liabilities began to surface. Specifically, the solubility of 2 in organic solvents is limited, and even in THF the solubility of 2 was only 31 mg/mL at room temperature. Since a reaction at elevated temperature was not desirable, separation from the heterogeneous catalyst required a relatively large volume of solvent, and consequently an extended time cycle for distillation at manufacturing scale. In addition, if the hydrogenolysis reaction were to stall, degradation of 2 was anticipated to be a significant competing process which could compromise yield and purity. Herein, we report a robust debenzylation process reliant on an in situ protection/deprotection sequence of the N-OH © 2018 American Chemical Society

functional group in 2 which overcomes many of the challenges inherent in manipulating this sensitive compound. During the course of these studies we discovered a significant rate enhancement in the hydrogenolysis reaction when substoichiometric amounts of tertiary amine bases were added, and found that this acceleration was observed with other benzyl ether substrates.



RESULTS AND DISCUSSION At the outset of this work we had the benefit of earlier pilotscale experience which highlighted the challenges associated with handling urea 2 in solution.3,4a To this end, we sought to protect the hydroxylamine in 2 during the debenzylation reaction, only to reveal it at reaction completion in the context of a reactive crystallization. Protection was expected to stabilize 25 and improve the solubility of the intermediate thereby minimizing reaction volume and improving cycle time. A trimethylsilyl (TMS) group was identified as the optimal species due to the mild and rapid protection/deprotection sequence and the expected higher solubility of the silyl protected intermediate in organic solvents. To our delight, in the presence of stoichiometric N,O-bis(trimethylsilyl)acetamide (BSA),6 the TMS protected hydroxyl urea 5 was cleanly formed in nearly quantitative yield by hydrogenolysis of 4 with solid supported Pd catalysts such as Pd/C or Pd/Al2O3. Examining the behavior (with/without catalysts) of 5 indicated improved stability (relative to 2), and the efficient removal of the TMS group upon treatment with water in the presence of acid allowed for an efficient direct isolation of the desired crystalline hydroxyl urea 2. During development of this process significant rate differences in the hydrogenolysis reaction were observed, and ultimately traced back to different sources of BSA. After careful Received: December 7, 2017 Published: February 1, 2018 273

DOI: 10.1021/acs.oprd.7b00381 Org. Process Res. Dev. 2018, 22, 273−277

Organic Process Research & Development

Article

Scheme 1. Preparation of the Key Hydroxylamine Intermediate 2 for Relebactam

In the absence of any additive the reaction proceeded to only 4% conversion (entry 1). Pyridine (entry 2), 2,6-di-tertbutylpyridine (entry 4), and 2,2′-dipyridyl (entry 5) all resulted in lower conversion than the control reaction, with the latter completely suppressing hydrogenolysis. Employing 2,6-lutidine (entry 3) or lithium tert-amoxide (entry 6) afforded greater conversion than in the reaction without base and pointed toward the subtle impact of this additive on conversion. Our early finding on the impact of TEA was confirmed (entry 7), but most interestingly the use of DABCO resulted in 64% conversion after only 4 h, a 16-fold increase in conversion over the control reaction (entry 8). After DABCO was identified as the most efficient additive among this set, a loading study was performed under the standard hydrogenolysis conditions (0.1 M THF solution of 4, 50 psi of H2, 1.8 mol % dry Pd/C (5 wt %), 1.25 equiv BSA, room temperature). Hydrogen uptake was monitored over 3 h8 at four different DABCO loadings (Figure 1). All four reactions proceeded without an induction period, and the reaction in the absence of DABCO reached 28% conversion in 3 h; this difference relative to the data presented in Table 1 is attributed to the presence of BSA (see below). At a charge of only 0.5 mol % DABCO a significant rate acceleration was observed, with hydrogen uptake essentially plateauing after 1 h and a quantitative yield determined by HPLC after 3 h. Only minor differences in rate and conversion were observed from 0.5 to 3 mol % DABCO (Figure 1). Therefore, 1−2 mol % of DABCO was chosen to maximize the rate enhancement in the debenzylation, establishing an approximately 1:1 mole ratio of DABCO to Pd. After establishing the optimal DABCO charge, other solid supported Pd catalysts were evaluated to determine if the rate acceleration was unique to the solid support. The reactions

examination, we discovered certain lots of BSA were contaminated with up to 2 mol % of triethylamine (TEA), while other samples contained none of the amine impurity. Surprisingly, the hydrogenolysis reaction was significantly accelerated when TEA contaminated BSA was employed, a finding in contrast with reported examples of catalyst poisoning in O-benzyl ether hydrogenolysis reactions.7 In an effort to better understand this observation, a selection of basic additives were evaluated in the debenzylation reaction under otherwise standard conditions without BSA (0.1 M substrate in THF, 5 wt % dry Pd/C, 50 psi hydrogen, room temperature). The additive amount employed was 2 mol % for 2,2′-dipyridyl and DABCO and 4 mol % for the other bases examined. All reactions were halted after 4 h, and the conversion was determined by quantitative HPLC analysis (Table 1). Table 1. Impact of Basic Additives on Benzyl Ether Hydrogenolysis

entry

additive

pKa (DMSO)

mol %

conversion (%)

1 2 3 4 5 6 7 8

None pyridine 2,6-lutidine 2,6-di-tert-butylpyridine 2,2′-dipyridyl lithium tert-amoxide TEA DABCO

N/A 3.4 6.6 0.81 4.3 17 9.0 8.9

− 4 4 4 2 4 4 2

4 1 9 1 0 15 40 64 274

DOI: 10.1021/acs.oprd.7b00381 Org. Process Res. Dev. 2018, 22, 273−277

Organic Process Research & Development

Article

Figure 2. Comparison of hydrogen uptake rates for the reactions with/ without DABCO and with/without BSA.

Figure 1. Comparison of hydrogen uptakes at different DABCO loadings.

were conducted using a 0.2 M THF solution of 4 with 1.25 equiv of BSA with and without 2 mol % DABCO (Table 2). Table 2. Screening of Solid Support Pd Catalysts catalyst 5 wt % Pd/C 5 wt % Pd/C 5 wt % Pd/Al2O3 5 wt % Pd/Al2O3 5 wt % Pd/CaCO3 5 wt % Pd/CaCO3 5 wt % Pd/BaSO4 5 wt % Pd/BaSO4 20 wt % Pd(OH)2/C 20 wt % Pd(OH)2/C

loading

DABCO

3 3 5 5 5 5 5 5 1 1

2 mol none 2 mol none 2 mol none 2 mol none 2 mol none

wt wt wt wt wt wt wt wt wt wt

% % % % % % % % % %

% % % % %

time (h)

yield (%)

2 2 2 2 16 16 16 16 8 8

100 8 100 9 38 2 58 9 52 45

Figure 3. Comparison of hydrogen uptake for reactions with different BSA equivalents.

Since DABCO was not added for this study, all the reactions proceeded slowly relative to the optimized conditions. However, the data clearly indicated that without BSA (green line) the reaction is slowest and stalls after 2 h. The hydrogen uptake with 0.6 equiv of BSA (red line) began to plateau at 60% conversion (5 h). Since BSA can transfer one of two TMS groups to the free hydroxylamine 2 under such mild conditions (room temperature),6b the data suggest hydrogenolysis begins to stall once all the BSA was consumed to generate Ntrimethylsilylacetamide. Similar profiles were observed when 1.2 or 2.0 equiv of BSA was employed. The data support that free hydroxylamine intermediate 2 can act as a weak catalyst poison3 and that the role of BSA is to ensure rapid conversion of 2 to TMS protected 5 in situ. On the other hand, the role of DABCO is not clear at the present time; we speculate that DABCO may modify the surface of the solid support, or the Pd distribution between the solid phase and liquid phase. However, no meaningful physical data could be obtained to support these hypotheses. Further studies are underway to better understand the role of DABCO.9 Our ultimate goal was to develop the most efficient and reliable process to manufacture N-hydroxyl urea 2. It is ideal if the process can use a single solvent, as a time-consuming

In all cases, a significant increase in yield was observed when 2 mol % DABCO was present, even though the difference with Pd(OH)2/C was less than with other solid supported catalysts. This study established that either Pd/C or Pd/Al2O3 were equally efficient catalysts for the desired hydrogenolysis reaction. To differentiate the effects of DABCO and BSA, four reactions were conducted with a 0.2 M THF solution of 4, 1.2 mol % Pd/Al2O3 (5 wt %), with/without 1.25 equiv of BSA, and with/without 2 mol % DABCO under 50 psi of hydrogen at room temperature (Figure 2). As depicted in Figure 2, addition of DABCO with or without BSA significantly accelerated the hydrogenolysis reaction, though interestingly, the fastest rate was observed when both DABCO and BSA were present (green line, Figure 2). The data suggest that the free hydroxylamine 2, which is generated without BSA, may act as a weak catalyst poison in the reaction.3 To test this hypothesis, reactions were performed with a 0.2 M THF solution of 4, 1.8 mol % Pd/Al2O3 (5 wt %) without DABCO, and at four different BSA charges (0, 0.6, 1.2, and 2.0 equiv) under 50 psi of hydrogen at room temperature (Figure 3). 275

DOI: 10.1021/acs.oprd.7b00381 Org. Process Res. Dev. 2018, 22, 273−277

Organic Process Research & Development

Article

Scheme 2. Through Process for Preparation of 2

solvent switch could be avoided. A suitable reaction solvent must have good solubility for the TMS protected 5 to ensure easy separation of the product from the solid catalyst. On the other hand, the solubility of 2 in the same solvent should be low enough to enable rapid product precipitation from the solution upon deprotection of 5. We identified isopropyl acetate (IPAc) as an ideal candidate, since the solubility of 2 in IPAc is low (2.1 mg/mL) and hydrogenolysis in IPAc under optimized conditions proceeded with a similar rate and impurity profile to the reaction in THF. Employing 5 wt % Pd/C rather than the alumina supported catalyst was preferred to simplify the recovery of Pd. In the presence of DABCO a catalyst loading of only 0.72 mol % (3 wt %) Pd/C was required to reach complete conversion in a few hours, where in the absence of DABCO more than 12 h were required for reaction completion even with 2.4 mol % of Pd/C (5 wt %). After the hydrogenation was complete, the product solution was separated from the catalyst by filtration. Upon treatment of the filtrate with 2.6 equiv of water and a catalytic amount of acetic acid at room temperature, hydroxyl urea 2 crystallized from the IPAc solution in high purity and was obtained in 88% isolated yield.10 This process was successfully implemented on pilot plant scale, as shown in Scheme 2. The impact of DABCO on the reactivity of Pd/C, and other solid supported Pd catalysts, observed in the relebactam process could have broader implications. Therefore, the generality of this rate acceleration was examined with a series of benzyl ether containing substrates (Table 3). The hydrogenolysis reactions were carried out in the presence of 3 mol % DABCO (condition A) or 0 mol % DABCO (condition B) in 0.1 M THF solution in a glass Parr shaker with Pd/C (5 wt %) under 50 psi hydrogen pressure at room temperature. The reactions were halted after 4 h, and the yields for the corresponding products were determined by HPLC. Phenolic benzyl ethers (entries 1, 2, and 6) and a benzyl ester (entry 3) all showed an increase in conversion in the presence of DABCO. Similarly, an N-Cbz group (entries 4 and 5) was removed more efficiently in the presence of DABCO. In the case of entry 4, the yield was dramatically improved by addition of BSA (condition C), which might be due to the solubility of the product. The debenzylation of an N-benzyl imidazole containing substrate (entry 7) demonstrated a limitation of the current method, showing low conversion with or without DABCO present.

Table 3. Generality of DABCO-Mediated Rate Acceleration in Benzyl Ether Hydrogenolysisa

a

Reaction conditions: (A) DABCO (3 mol %), 0.1 M THF, 5 wt % Pd/C, H2 (50 psi), rt, 4 h; (B) 0.1 M THF, 5 wt % Pd/C, H2 (50 psi), rt, 4 h; (C) DABCO (3 mol %), BSA (2.25 equiv), 0.1 M THF, 5 wt % Pd/C, H2 (50 psi), rt, 4 h.

have allowed for the development of a robust scalable process to secure key bicyclic intermediate 2 in the synthesis of the βlactamase inhibitor relebactam.



EXPERIMENTAL SECTION All catalysts, reagents, and solvents were obtained from commercial supplies and used without further purification prior to use unless otherwise stated. 5 wt % Pd/C was purchased from BASF Co. BSA was purchased from Gelest, Inc. (Contamination of TEA was less than 40 ppm by GC). DABCO and anhydrous THF were purchased from SigmaAldrich. General Procedure for the Preparation of Hydroxylurea 2 from 4. To a 2 L three-necked round-bottom flask containing 5 wt % Pd/C (6.0 g, 50% water) was added IPAc (800 mL). The mixture was distilled at 1 atm, and about 200 mL were collected until the moisture content (determined by KF titration) of the remaining mixture was below 500 ppm.



CONCLUSION A single solvent (IPAc) process for the formation of 2 was developed which relies on in situ TMS ether formation during hydrogenolysis, and reactive crystallization of 2 by acidic deprotection after catalyst removal. A significant rate enhancement of the hydrogenolysis reaction through the addition of organic bases, specifically DABCO, was observed and found to be general for related debenzylation reactions. These findings 276

DOI: 10.1021/acs.oprd.7b00381 Org. Process Res. Dev. 2018, 22, 273−277

Organic Process Research & Development



ACKNOWLEDGMENTS The authors acknowledge helpful discussions with Dr. Kevin R. Campos, Dr. Louis-Charles Campeau, Dr. Paul G. Bulger, and Dr. Michael Kress.

The mixture was cooled to room temperature and additional IPAc (400 mL) was added, followed by compound 4 (100.0 g, 0.218 mol), BSA (68.0 mL, 95 wt %, 0.264 mol), and DABCO (0.48 g, 4.28 mmol). The mixture was then hydrogenated in an autoclave at 50 psi and room temperature for 3 to 5 h until the remaining starting material was less than 0.5% as assessed by HPLC analysis. The mixture was filtered through one inch of Celite in a Buchner funnel under nitrogen, and the cake was rinsed with IPAc (500 mL). The solution was then concentrated at reduced pressure until the volume of the filtrate was 1.4 L. Analysis of the solution by HPLC indicated the solution contained 96.1 g (100%) of 5. A portion (400 mL) of the above filtrate containing 5 was charged into a 2 L three-necked flask fitted with an overhead stirrer. After the solution was cooled to 15 °C, 28% aqueous acetic acid (5.5 g) was added in one portion. The mixture was then agitated at 15 °C for 1 h while adding one-third of the remaining filtrate (333 mL) slowly. Product 2 was precipitated during the addition. To the resulting suspension was added additional 28% aqueous acetic acid (4.17 g) in one portion. The mixture was agitated for another 1 h at 15 °C while another one-third of the filtrate (333 mL) was added slowly. After addition of additional aqueous acetic acid (4.17 g), the remaining filtrate (333 mL) was added over 1 h. The resulting suspension was agitated at 20 °C for 1 h followed by addition of methyl tert-butyl ether (600 mL). The mixture was then cooled to 5 °C over 1 h, and the product 2 was collected by filtration. The cake was washed with dry IPAc (200 mL) and dried in a vacuum oven at 50 °C, to give 2 (70.7 g, 88%; residual Pd less than 5 ppm) as a colorless crystalline solid. The spectroscopic data are consistent with those reported previously.4 Alternatively, analytically pure crystalline 5 could be isolated by concentrating the filtrate and storing the solution at 5 °C overnight. 1H NMR (500 MHz, CDCl3): δ 6.58 (d, J = 7.9 Hz, 1H), 4.10−3.86 (m, 4H), 3.55 (bs, 1H), 3.14 (bd, J = 11.5 Hz, 1H), 2.86 (bt, J = 12.0 Hz, 2H), 2.76 (d, J = 11.5 Hz, 1H), 2.36 (dd, J = 15.1, 7.1 Hz, 1H), 2.12 (m, 1H), 2.00−1.82 (m, 3H), 1.66 (m, 1H), 1.44 (s, 9H), 1.31 (m, 2H), 0.25 (S, 9H). 13C NMR (125 MHz, CDCl3): δ 169.2, 168.3, 154.8, 79.8, 60.7, 60.0, 47.3, 46.9, 42.6 (br, 2C), 32.2, 31.9, 28.5 (3C), 20.5, 17.5, −0.75 (3C). (+)-ESI HRMS: calcd for C20H36N4NaO3Si (M + Na)+, 463.2347; found, 463.2348.



Article



REFERENCES

(1) (a) Karam, G.; Chastre, J.; Wilcox, M. H.; Vincent, J.-L. Critical Care 2016, 20, 136. (b) Zowawi, H. M.; Harris, P. N. A.; Roberts, M. J.; Tambyah, P. A.; Schembri, M. A.; Pezzani, M. D.; Williamson, D. A.; Paterson, D. L. Nat. Rev. Urol. 2015, 12, 570. (2) (a) Haidar, G.; Clancy, C. J.; Chen, L.; Samanta, P.; Shields, R. K.; Kreiswirth, B. N.; Nguyen, M. H. Antimicrob. Agents Chemother. 2017, 61, e00642-17. (b) Lob, S. H.; Hackel, M. A.; Kazmierczak, K. M.; Young, K.; Motyl, M. R.; Karlowsky, J. A.; Sahm, D. F. Antimicrob. Agents Chemother. 2017, 61, e02209-16. (3) During preparation of the penultimate tetrabutylammonium salt of avibactam, an in situ O-sulfonylation was developed during the debenzylation. The hydrogenolysis rate was carefully controlled to avoid build up the unstable N-hydroxyl intermediate. See: (a) Ball, M.; Boyd, A.; Ensor, G. J.; Evans, M.; Golden, M.; Linke, S. R.; Milne, D.; Murphy, R.; Telford, A.; Kalyan, Y.; Lawton, G. R.; Racha, S.; Ronsheim, M.; Zhou, S. H. Org. Process Res. Dev. 2016, 20, 1799. (b) Wang, T.; Du, L.-d.; Wan, D.-j.; Chen, X.; Wu, G. Org. Process Res. Dev., DOI: 10.1021/acs.oprd.7b00290. (4) (a) Mangion, I. K.; Ruck, R. T.; Rivera, N.; Huffman, M. A.; Shevlin, M. Org. Lett. 2011, 13, 5480. (b) Miller, S. P.; Zhong, Y. L.; Liu, Z.; Simeone, M.; Yasuda, N.; Limanto, J.; Chen, Z.; Lynch, J.; Capodanno, V. Org. Lett. 2014, 16, 174. (5) The corresponding O-benzyl compounds are not reported as unstable in refs 3 and 4. (6) (a) Klebe, J. F.; Finkbeiner, H.; White, D. M. J. Am. Chem. Soc. 1966, 88, 3390. (b) El Gihani, M. T.; Heaney, H. Synthesis 1998, 1998, 357. (7) (a) Sajiki, H.; Hirota, K. Tetrahedron 1998, 54, 13981. (b) Sajiki, H.; Hattori, K.; Hirota, K. J. Org. Chem. 1998, 63, 7990. (c) Sajiki, H.; Kuno, H.; Hirota, K. Tetrahedron Lett. 1998, 39, 7127. (d) Researchers at Abbott Laboratories reported rate enhancement in the hydrogenolysis of a benzyl ester in conjunction with the reduction of Ar−NO2 to aniline by addition of TEA. See: Seif, L. S.; Hannick, S. M.; Plata, D. J.; Morton, H. E.; Sharma, P. N. Chemical Industries (Dekker) 2003, 89, 633. (8) The progress of the reaction was also monitored with HPLC analysis, and these data were in good correlation with the hydrogen uptake data. (9) Pd concentrations in solution during the hydrogenation with/ without DABCO were similar, and the highest concentration observed was about 1.8 ppm in both cases. Based on the preliminary studies of this hydrogenolysis, the reaction kinetics were zero-order with respect to the benzyl ether substrate. (10) Water activity was carefully controlled during reactive crystallization to avoid forming a less stable crystalline dihydrate.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.oprd.7b00381. 1

H and 13C NMR spectra of compound 5 (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Zhijian Liu: 0000-0002-5750-9890 Feng Xu: 0000-0003-1949-324X Nobuyoshi Yasuda: 0000-0001-6002-4395 Notes

The authors declare no competing financial interest. 277

DOI: 10.1021/acs.oprd.7b00381 Org. Process Res. Dev. 2018, 22, 273−277