Skip to main content
ACS Medicinal Chemistry Letters logoLink to ACS Medicinal Chemistry Letters
. 2022 Apr 14;13(5):833–840. doi: 10.1021/acsmedchemlett.2c00075

Thermal Stability and Utility of Dienes as Protecting Groups for Acrylamides

Annie R Hooper , Alexander S Burns ‡,*
PMCID: PMC9109275  PMID: 35586437

Abstract

graphic file with name ml2c00075_0009.jpg

Acrylamides are privileged electrophiles used in targeted covalent therapies, often installed at the end of a synthetic sequence due to their reactive nature. Herein, we report several diene-acrylamide adducts with a range of thermal stabilities toward retro-Diels–Alder deprotection of the acrylamide, enabling this masked functionality to be introduced early in a synthetic route and deprotected in a specific temperature range. Through kinetic studies, we identify solvent and structural trends that impact the stability of trimethylsilyl cyclopentadiene (TMS-CP) acrylamide adducts. TMS-CP protected acrylamides were installed on several amine-containing drugs, whose acrylamides were thermally unveiled (T = 160 °C, time ≤ 1 h) in moderate to high yields. We also showcase the potential utility of this protection strategy by improving the yield of a base-promoted SNAr reaction when the acrylamide is masked.

Keywords: Acrylamide, protecting group, retro-Diels−Alder, thermal deprotection


Small molecules which covalently bind to biological targets are an important therapeutic class. Within the past 15 years, there has been an increase in the number of drug discovery programs rationally targeting a specific amino acid for covalent modification.1 Many amino acids have been targeted in chemical biology and medicinal chemistry applications, with cysteine being the most common residue due to its high and unique nucleophilicity.2 Various electrophilic warheads have been developed to selectively react with cysteine;3 however, α,β-unsaturated amides are the most commonly employed electrophile to target noncatalytic cysteine residues.4 Acrylamides, such as those found in Zanubrutinib 1, Osimertinib 2, and Dacomitinib 3 (Figure 1), have had great clinical success in recent years as cysteine targeting warheads.5

Figure 1.

Figure 1

Examples of acrylamide-containing FDA-approved drugs.

There is a large body of literature precedent for the protection of electron deficient alkenes, such as maleimides, α,β-unsaturated ketones, and α,β-unsaturated esters, as thermally reversible retro-Diels–Alder (DA) adducts.6,7 There are fewer examples of such a strategy being used to mask α,β-unsaturated amides.810 Though less reactive than their keto or ester counterparts, acrylamides can react via undesired pathways depending on the reaction conditions as well as their electronic and steric environments. Basic and/or nucleophilic conditions such as those in nucleophilic aromatic substitution or palladium mediated cross couplings can be particularly problematic for acrylamides.1113 Most often, this motif is installed at the end of a synthetic sequence to avoid these problems. Notably, there are several examples where a β-chloro or sulfonyl propionamide 4 serves as a masked acrylamide, and the olefin 5 is generated by treatment with strong base (Figure 2A).12,14 To our knowledge, there is a single report disclosing a Diels–Alder adduct as a masked acrylamide in a medicinal chemistry setting.15 As shown in Figure 2B, a patent application from Sandoz Pharmaceuticals describes a synthesis of Ibrutinib 7, in which the acrylamide is protected as the cyclopentadiene (CP) adduct 6. In the final step, the acrylamide is unveiled by heating to 250 °C in diphenyl ether.

Figure 2.

Figure 2

Protecting group strategies for acrylamides.

While the protecting group strategies described above will have settings in which they work, there is a need to develop other protecting groups for acrylamides to address the unique challenges faced by each medicinal chemistry team. As such, it would be advantageous to have a protecting group that would unveil the acrylamide at a lower temperature. We explored the thermal stabilities of several diene-acrylamide DA adducts, enabling this sensitive functionality to be installed early in a synthesis and unmasked within a specific temperature range.

We began by synthesizing model acrylamides protected with various dienes and determining their relative thermal stabilities when heating in Dowtherm A16 (Table 1). Conversion was determined based upon the loss of starting DA-adduct rather than product formation. Conversion to the acrylamide was in some cases accompanied by the formation of unidentified side-products, possibly due to undesired reaction of the reactive acrylamide with the released diene at these elevated temperatures.

Table 1. Thermal Stabilities of Diene-Acrylamide Adductsa in Dowtherm Ab.

graphic file with name ml2c00075_0006.jpg

a

Amide adducts were synthesized through HATU coupling of 4-phenylpiperidine with racemic diene-acrylic acid adducts.

b

Amide adducts (0.03 mmol) and an inert internal standard (0.03 mmol) were diluted in Dowtherm A (0.4 mL) in a vial under a slight positive pressure of nitrogen equipped with a vent needle. The mixtures were heated to the shown temperature for 30 min. Conversion was determined by LC analysis based upon the loss of starting material relative to internal standard.

c

For compounds whose diene was not volatile at the shown temperature, and there was no significant loss of starting material, the reaction was run a second time in the presence of a nucleophilic scavenger and loss of starting material was quantified, in case the forward Diels–Alder was confounding the results (see the SI for details).

d

For compounds whose Diels–Alder adduct eluted at the same retention time as the acrylamide, conversion was quantified by NMR.

As expected, CP adduct 9 was very stable, yielding 21% conversion at 200 °C. While kinetically less stable, anthracene is not volatile and we observed no significant product formation upon heating 10 in Dowtherm A. The retro-DA was inferred by adding a nucleophilic scavenger and looking for the disappearance of starting material. Exo and endo fulvene adducts 11 and 12 were significantly less stable, giving conversions of 15% and 18%, respectively, at 120 °C. When furan was used as a masking group, the adducts showed conversion at 80 °C (14 and 15). Dimethyl substituted furan adduct 16 was the least stable of all substrates, with 28% conversion at 80 °C. There are a range of thermal stabilities that can be achieved for the protected acrylamides through selection of the diene. The desired stability range can be fine-tuned through selection of the endo or exo adduct or by placing substituents at the bridgeheads of the adducts.17

While the thermal stability ranges of both the fulvene and furan adducts would be useful for protecting some acrylamides, we were interested in developing adducts with good stability up to 100 °C. This would allow the adduct to stay intact under many standard chemical transformations involving elevated temperatures. Toward that goal, we investigated trimethylsilyl-cyclopentadiene (TMS-CP), an alkene protecting group first introduced by the Magnus laboratory in their synthesis of aspidosperma-type indole alkaloids.18,19 Relative to unsubstituted CP, adducts with TMS-CP will more readily undergo a retro-DA reaction due to a well oriented σC–Si → σ*C–C donation into the breaking σ-bonds of the DA adduct.20 TMS-CP adduct 13 was more stable than the fulvene substrates, as evidenced by the reduced conversion at 120 and 160 °C. Based on the preliminary ranking of stabilities of adducts in Table 1, we wanted to more closely investigate the deprotection kinetics of TMS-CP adducts.

We synthesized several sterically and electronically differentiated masked acrylamides and determined their stabilities through 1H NMR kinetic experiments (Table 2). There are several points that merit discussion. Most importantly, adducts displayed good to excellent stabilities at 100 °C in d7-DMF (t1/2 = 28–140 h). As expected, the secondary and primary aliphatic amine derived endo adducts (13 and 17, respectively) were more stable than the relatively more electron deficient aniline derived endo adducts (18 and 19).21 Based on our observations from Table 1, we synthesized the exo adducts (20 and 21) of our less stable aniline derived substrates and determined that the half-life can be modestly increased. Secondary amide substrates were more stable than tertiary amide substrates. Finally, though the effects were minor and did not hold true for all substrates, these protected acrylamides were generally most stable in the nonpolar o-xylene.7a,22 Armed with a better understanding of the factors influencing the stability of these adducts, we wanted to verify these deprotections would proceed on more medicinally relevant molecules.

Table 2. Thermal Stabilities of Diene-Acrylamide Adducts in Dowtherm Aa.

graphic file with name ml2c00075_0007.jpg

a

Amide adducts, thiophenol (3 equiv), tributylamine (3.3 equiv), and dibenzylether (1 equiv, internal standard) were dissolved in the indicated deuterated solvent and placed in a preheated NMR spectrometer. 1H spectra were collected at regular time intervals, and half-lives were calculated based upon the first order loss of starting material.

Several TMS-CP-protected acrylamides were installed on marketed drugs bearing aliphatic amines which were then heated to unveil the resulting acrylamides (Scheme 1). We reasoned that since these adducts were generally more stable in o-xylene, the deprotections would be faster in more polar 1,2,4-trichlorobenzene.23 We obtained reasonable yields for many of the initially tested substrates; however, the electron deficient Sitagliptin analogue 29 was produced in only 32% yield. To improve the yield for the deprotection of Sitagliptin adduct 29 and other substrates, we performed a short optimization study using the more electron deficient aniline derived 18 (Table 3).

Scheme 1. Substrate Scope,

Scheme 1

Adducts were synthesized through HATU coupling of amines with racemic TMS-CP Diels–Alder adduct (see the SI for details).

Adducts (0.025 M) were heated to 160 °C in Dowtherm A in a vial under a slight positive pressure of nitrogen equipped with a vent needle. Isolated yields after purification are reported.

Reactions were run in 1,2,4-trichlorobenzene for 12 h.

Table 3. Optimization of Retro-Diels–Alder.

graphic file with name ml2c00075_0008.jpg

solvent time (min) conc (M) relative conversiona yieldb
1,2,4-trichlorobenzene 90 0.05 1 N/D
  30 0.05 1.21 N/D
  30 0.025 1.27 N/D
Dowtherm A 30 0.025 1.41 74%
  30 0.1 N/D 69%
a

18 was diluted in the indicated solvent in a vial under a slight positive pressure of nitrogen equipped with a vent needle. The solutions were heated to 160 °C for the indicated reaction time. An internal standard was added, and the amount of product formed, relative to the internal standard, was determined by LC analysis. Relative conversion is defined as the ratio of product to internal standard, divided by that ratio for the first table entry.

b

Reactions were performed on a 1 mmol scale.

When adduct 18 was heated in trichlorobenzene for 90 min, LCMS analysis of the mixture showed product acrylamide 32 as well as several overlapping peaks with the same mass and approximate retention time of the starting material. We believe that these peaks correspond to DA isomers in which the TMS group is not at its original position and thus would not have the same rate enhancement for the retro-DA reaction that the parent isomer does.24 We speculate that the unproductive forward-DA reaction with more electron deficient acrylamides is competitive with the evaporation of TMS-CP under these reaction conditions.25 Based upon our kinetic studies, we thought the retro-DA would achieve near full conversion sooner than 90 min and prolonged reaction times would not be beneficial. This was borne out experimentally, as the relative formation of 32 increased when the reaction was run for 30 min. There was a small increase in product formation running the reaction under more dilute conditions. Switching the solvent from trichlorobenzene back to Dowtherm A gave an increase in the amount of acrylamide generated. Though we do not fully understand the reasons why this switch improved our relative conversions, decreasing the amount of byproducts isomeric to the starting material, we were happy to exploit these findings for our deprotections. When run on the 1 mmol scale using optimized conditions, we obtained an isolated yield of 74%. Lastly, it is worth noting that this deprotection was run at a higher concentration (0.1 M) on the 1 mmol scale, giving only a minor decrease in yield.

When subjecting previous substrates to our newly optimized conditions, we found yields to improve overall. Most notably, the yield to unveil Sitagliptin acrylamide 29 improved from 32% to 89%. We also examined these deprotections on two aniline containing drugs. While the deprotection of a lenalidomide derivative gave a reasonable yield of acrylamide 30, sulfamethoxazole acrylamide 31 was produced in just 33%.26 Thus, this strategy will not be as effective for very electron deficient anilinic acrylamides.

To demonstrate the utility and advantage of this protecting group strategy, we synthesized both protected and deprotected acrylamides, 33 and 34 respectively, and carried out a base-mediated SNAr (Scheme 2). When we used 2 equiv of amine 35, unprotected acrylamide 34 underwent the desired displacement to give acrylamide 37 in only 4%, in contrast to the 77% yield for protected 36. While the yield could be improved by using 1.1 equiv of amine 35, there will be cases in which the desired reaction will be unable to outcompete side reactivity with the acrylamide, highlighting the need for a protection strategy. Since Boc protecting groups are known to be thermally labile and DA adducts are prone to retro-DA under acidic conditions, we demonstrated that each could be selectively deprotected in the presence of the other giving 38 and 37, respectively.

Scheme 2. Synthetic Utility of TMS-CP Protecting Group.

Scheme 2

Reaction employed 1.1 equiv of N-Boc-piperazine (35).

Reaction employed 2.0 equiv of N-Boc-piperazine (35).

Though an unsubstituted acrylamide is often the targeted warhead, substitutions on the acrylamide are frequently explored. For example, a β-methyleneamine substitution is found in the marketed EGFR, covalent inhibitors Afatinib,27 Dacomitinib 3,5b and Neratinib.28 We wanted to determine how this substitution would affect the thermal stability of our TMS-CP protected acrylamides. As shown in Scheme 3, we synthesized the protected acrylic acid derivative 40 in 3 steps. A Lewis-acid-catalyzed DA reaction between TMS-CP and bromo-crotonate 39, followed by displacement of the bromide with piperidine and then saponification delivered protected acid 40 in 23% overall yield. Acid 40 was coupled to 4-phenylpiperidine to give the TMS-CP protected acrylamide 41. Utilizing our optimized conditions for deprotection, we were able to unveil acrylamide 42 in 75% yield. Pleasingly, substituted adduct 41 shows similar thermal stability to unsubstituted adduct 13 in d7-DMF.

Scheme 3. Thermal Stability and Deprotection of TMS-CP-Masked, Substituted Acrylamide.

Scheme 3

We would like to acknowledge two caveats regarding the use of these masked acrylamides. The first is that the corresponding protected acrylic acid is not commercially available. However, we were able to make endo-TMS-CP protected acrylic acid on the multigram scale in two steps from commercial materials.29 The second is that we have only described the synthesis and use of the racemic carboxylic acids. While the ultimate acrylamides have no stereocenters, when coupled with chiral amines, diastereomeric products are generated. We would point out that resolutions of racemic carboxylic acids30 and enantioselective DA reactions31 are well established in the literature and either strategy would likely be able to produce enantioenriched protected acids.

In summary, we explored the thermal stability of several diene-acrylamide adducts and determined that TMS-CP protected acrylamides are relatively stable up to 100 °C and can be deprotected in 1 h or less at 160 °C. Optimized conditions gave moderate to high yields for the deprotection of acrylamides on several complex drugs, though the yield was reduced for a very electron poor anilinic acrylamide containing substrate. We showed an example where protection of the acrylamide can lead to higher yields for an SNAr reaction and demonstrated protecting group orthogonality to a Boc group. Finally, a TMS-CP protected acrylamide with a β-methyleneamine substitution was shown to have similar thermal stability to the analogous unsubstituted acrylamide. It is our hope that the study described herein will be useful to both the medicinal chemistry and chemical biology communities, shortening drug development timelines and improving access to acrylamide containing probes.

Acknowledgments

We thank Prof. David Sarlah and the University of Illinois at Urbana–Champaign for collaboration to supply laboratory space, chemicals, and equipment. We thank Deszra Shariff and Guoyun Bai for assistance with NMR experiments. We are grateful to Hariharan Venkatesan, Jennifer Venable, Douglas C. Behenna, and Michael Rombola for helpful discussions.

Glossary

Abbreviations

Boc

tert-butoxycarbonyl

CP

cyclopentadiene

DA

Diels–Alder

DCM

dichloromethane

DIPEA

N,N-diisopropylethylamine

DMF

dimethylformamide

EGFR

epidermal growth factor receptor

HATU

1-[bis(dimethylamino)methylene]-1H-1,2,3-triazolo[4,5-b]pyridinium 3-oxide hexafluorophosphate

LCMS

liquid chromatography mass spectrometry

NMR

nuclear magnetic resonance

SNAr

nucleophilic aromatic substitution

TFA

trifluoroacetic acid

TMS-CP

trimethyl-silyl cyclopentadiene

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsmedchemlett.2c00075.

  • Detailed experimental procedures, spectroscopic data, and 1H and 13C spectra (PDF)

Author Contributions

The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.

The authors declare no competing financial interest.

Supplementary Material

ml2c00075_si_001.pdf (10.8MB, pdf)

References

  1. Sutanto F.; Konstantinidou M.; Dömling A. Covalent inhibitors: a rational approach to drug discovery. RSC Med. Chem. 2020, 11, 876–884. 10.1039/D0MD00154F. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. a Gehringer M.; Laufer S. A. Emerging and re-emerging warheads for targeted covalent inhibitors: applications in medicinal chemistry and chemical biology. J. Med. Chem. 2019, 62, 5673–5724. 10.1021/acs.jmedchem.8b01153. [DOI] [PubMed] [Google Scholar]; b Bauer R. A. Covalent inhibitors in drug discovery: from accidental discoveries to avoided liabilities and designed therapies. Drug Discovery Today. 2015, 20, 1061–1073. 10.1016/j.drudis.2015.05.005. [DOI] [PubMed] [Google Scholar]; c Wan X.; Yang T.; Cuesta A.; Pang X.; Balius T. E.; Irwin J. J.; Shoichet B. K.; Taunton J. Discovery of lysine-targeted eIF4E inhibitors through covalent docking. J. Am. Chem. Soc. 2020, 142, 4960–4964. 10.1021/jacs.9b10377. [DOI] [PMC free article] [PubMed] [Google Scholar]; d Poole L. B. The basics of thiols and cysteine in redox biology and chemistry. Free Radic. Biol. Med. 2015, 80, 148–157. 10.1016/j.freeradbiomed.2014.11.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. For some recent examples of nonacrylamide, cysteine targeting electrophiles, see:; a McAulay K.; Hoyt E. A.; Thomas M.; Schimpl M.; Bodnarchuk M. S.; Lewis H. J.; Barratt D.; Bhavsar D.; Robinson D. M.; Deery M. J.; Ogg D. J.; Bernardes G. J. L.; Ward R. A.; Waring M. J.; Kettle J. G. Alkynyl benzoxazines and dihydroquinazolines as cysteine targeting covalent warheads and their application in identification of selective irreversible kinase inhibitors. J. Am. Chem. Soc. 2020, 142, 10358–10372. 10.1021/jacs.9b13391. [DOI] [PubMed] [Google Scholar]; b Zambaldo C.; Vinogradova E. V.; Qi X.; Iaconelli J.; Suciu R. M.; Koh M.; Senkane K.; Chadwick S. R.; Sanchez B. B.; Chen J. S.; Chatterjee A. K.; Liu P.; Schultz P. G.; Cravatt B. F.; Bollong M. J. 2-sulfonylpyridines as tunable, cysteine-reactive electrophiles. J. Am. Chem. Soc. 2020, 142, 8972–8979. 10.1021/jacs.0c02721. [DOI] [PMC free article] [PubMed] [Google Scholar]; c Tokunaga K.; Sato M.; Kuwata K.; Miura C.; Fuchida H.; Matsunaga N.; Koyanagi S.; Ohdo S.; Shindo N.; Ojida A. Bicyclobutane carboxylic amide as a cysteine-directed strained electrophile for selective targeting of proteins. J. Am. Chem. Soc. 2020, 142, 18522–18531. 10.1021/jacs.0c07490. [DOI] [PubMed] [Google Scholar]
  4. a Serafimova I. M.; Pufall M. A.; Krishnan S.; Duda K.; Cohen M. S.; Maglathlin R. L.; McFarland J. M.; Miller R. M.; Frödin M.; Taunton J. Reversible targeting of noncatalytic cysteines with chemically tuned electrophiles. Nat. Chem. Biol. 2012, 8, 471–476. 10.1038/nchembio.925. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Hallenbeck K. K.; Turner D. M.; Renslo A. R.; Arkin M. R. Targeting non-catalytic cysteine residues through structure-guided drug discovery. Curr. Top. Med. Chem. 2016, 17, 4–15. 10.2174/1568026616666160719163839. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. a Guo Y.; Liu Y.; Hu N.; Yu D.; Zhou C.; Shi G.; Zhang B.; Wei M.; Liu J.; Luo L.; Tang Z.; Song H.; Guo Y.; Liu X.; Su D.; Zhang S.; Song X.; Zhou X.; Hong Y.; Chen S.; Cheng Z.; Young S.; Wei Q.; Wang H.; Wang Q.; Lv L.; Wang F.; Xu H.; Sun H.; Xing H.; Li N.; Zhang W.; Wang Z.; Liu G.; Sun Z.; Zhou D.; Li W.; Liu L.; Wang L.; Wang Z. Discovery of Zanubrutinib (BGB3111), a novel, potent, and selective covalent inhibitor of Bruton’s Tyrosine Kinase. J. Med. Chem. 2019, 62, 7923–7940. 10.1021/acs.jmedchem.9b00687. [DOI] [PubMed] [Google Scholar]; b Smaill J. B.; Gonzales A. J.; Spicer J. A.; Lee H.; Reed J. E.; Sexton K.; Althaus I. W.; Zhu T.; Black S. L.; Blaser A.; Denny W. A.; Ellis P. A.; Rivault F.; Schlosser K.; Ellis T.; Thompson A. M.; Trachet E.; Winters R. T.; Tecle H.; Bridges A.; et al. Tyrosine Kinase Inhibitors. 20. Optimization of substituted quinazoline and pyrido[3,4-d]pyrimidine derivatives as orally active, irreversible inhibitors of the epidermal growth factor receptor family. J. Med. Chem. 2016, 59, 8103–8124. 10.1021/acs.jmedchem.6b00883. [DOI] [PubMed] [Google Scholar]; c Cross D. A. E.; Ashton S. E.; Ghiorghiu S.; Eberlein C.; Nebhan C. A.; Spitzler P. J.; Orme J. P.; Finlay R. V.; Ward R. A.; Mel-lor M. J.; Hughes G.; Rahi A.; Jacos V. N.; Brewer M. R.; Ichihara E.; Sun J.; Hailing J.; Ballard P.; Al-Kadhimi K.; Rowlinson R.; Klinowska T.; Richmond G. H. P.; Cantarini M.; Kim D. W.; Ranson M. R.; Pao W. AZD9291, an irreversible EGFR TKI, overcomes T790M-mediated resistance to EGFR inhibitors in lung cancer. Cancer Discovery 2014, 4, 1046–1061. 10.1158/2159-8290.CD-14-0337. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. a Rickborn B. The retro-Diels-Alder reaction: part I. C-C dienophiles. Org. React. 1998, 52, 1–393. 10.1002/0471264180.or052.01. [DOI] [Google Scholar]; b Csende F.; Stájer G.; Fülöp F. 5.12 Retro Diels-Alder reactions. Comprehensive Organic Synthesis 2014, 5, 518–594. 10.1016/B978-0-08-097742-3.00513-9. [DOI] [Google Scholar]
  7. For a comparison of the relative rates of retro-Diels–Alder reactions for various dienophiles, see:; a Chung Y.; Duerr B. F.; McKelvey T. A.; Nanjappan P.; Czarnik A. W. Structural effects controlling the rate of retro-Diels-Alder reaction in anthracene cycloadducts. J. Org. Chem. 1989, 54, 1018–1032. 10.1021/jo00266a006. [DOI] [Google Scholar]; b Nanjappan P.; Czarnik A. W. Reversal of electronic substituent effects in the retro-Diels-Alder reaction. A charge neutral analog of oxyanion-accelerated cycloreversion. J. Org. Chem. 1986, 51, 2851–2853. 10.1021/jo00364a056. [DOI] [Google Scholar]
  8. For an example of a furan protected, substituted α,β-unsaturated amide, see:; Abraham E.; Cooke J. W. B.; Davies S. G.; Naylor A.; Nicholson R. L.; Price P. D.; Smith A. D. Asymmetric synthesis of 4-amino-γ-butyrolactones via lithium amide conjugate addition. Tetrahedron 2007, 63, 5855–5872. 10.1016/j.tet.2007.03.026. [DOI] [Google Scholar]
  9. For an example of both cyclopentadiene and furan protecting an α,β-unsaturated lactam, see:; Arai Y.; Fujii A.; Ohno T.; Koizumi T. Stereocontrol in intermolecular nucleophilic addition of N-acyliminium ion directed by a bicyclo[2.2.1]heptane or 7-oxabicyclo[2.2.1]heptane group. A novel route to 5-substituted 3-pyrrolin-2-ones of high optical purity. Chem. Pharm. Bull. 1992, 40, 1670–1672. 10.1248/cpb.40.1670. [DOI] [Google Scholar]
  10. For examples of retro-Diels–Alder reactions being used to unveil acrylamides as monomers for polymer synthesis, see:; a Yasuhiro K. US Patent, US005177264A, 1993.; b Zhonjun L. Chinese Patent, CN108047075A, 2018.; c Xiao X.; Zhou Z.; Hongli Y.; Zhou X.; Heng W.; Wang Y. Chinese Patent, CN110498776A, 2019.
  11. For an interesting case where undesired chloride and solvent additions to the acrylamide were occurring, see:; Parsons A. T.; Kubryk M.; Hedley S. J.; Thiel O. R.; Bauer D.; Potter-Racine M. S.; Lin Z. An improved process for the preparation of a covalent kinase inhibitor through a C-N bond-forming SNAr reaction. Org. Process Res. Dev. 2018, 22, 898–902. 10.1021/acs.oprd.8b00080. [DOI] [Google Scholar]
  12. For a case where an undesired Michael addition is occurring on an acrylamide, see:; Planken S.; Behenna D. C.; Nair S. K.; Johnson T. O.; Nagata A.; Almaden C.; Bailey S.; Ballard T. E.; Bernier L.; Cheng H.; Cho-Schultz S.; Dalvie D.; Deal J. G.; Dinh D. M.; Edwards M. P.; Ferre R. A.; Gajiwala K. S.; Hemkens M.; Kania R. S.; Kath J. C.; Matthews J.; Murray B. W.; Niessen S.; Orr S. T. M.; Pairish M.; Sach N. W.; Shen H.; Shi M.; Solowiej J.; Tran K.; Tseng E.; Vicini P.; Wang Y.; Weinrich S. L.; Zhou R.; Zientek M.; Liu L.; Luo Y.; Xin S.; Zhang C.; Lafontaine J. Discovery of N-((3R,4R)-4- fluoro-1-(6-((3-methoxy-1-methyl-1H-pyrazol-4-yl)amino)-9-methyl9H-purin-2-yl)pyrrolidine-3-yl)acrylamide (PF-06747775) through structure-based drug design: A high affinity irreversible inhibitor targeting oncogenic EGFR mutants with selectivity over wild-type EGFR. J. Med. Chem. 2017, 60, 3002–3019. 10.1021/acs.jmedchem.6b01894. [DOI] [PubMed] [Google Scholar]
  13. For a case where a Heck reaction is occurring on an acrylamide in the presence of other alkenes, see:; Muddala N. P.; Nammalwar B.; Selvaraju S.; Bourne C. R.; Henry M.; Bunce R. A.; Berlin K. D.; Barrow E. W.; Barrow W. W. Evaluation of new dihydrophthalazine-appended 2,4-diaminopyrimidines against Bacillus anthracis: Improved syntheses using a new pincer omplex. Molecules 2015, 20, 7222–7244. 10.3390/molecules20047222. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. a Chen L.; Chi F.; Wang T.; Wang N.; Li W.; Liu K.; Shu X.; Ma X.; Xu Y. The synthesis of 4-arylamido-2-arylaminoprimidines as potent EGFR T790M/L858R inhibitors for NSCLC. Bioorg. Med. Chem. Lett. 2018, 26, 6087–6095. 10.1016/j.bmc.2018.11.009. [DOI] [PubMed] [Google Scholar]; b Atack T. C.; Raymond D. D.; Blomquist C. A.; Pasaje C. F.; McCarren P. R.; Moroco J.; Befekadu H. B.; Robinson F. P.; Pal D.; Esherick L. Y.; Ianari A.; Niles J. C.; Sellers W. R. Targeted covalent inhibition of plasmodium FK506 binding protein 35. ACS Med. Chem. Lett. 2020, 11, 2131–2138. 10.1021/acsmedchemlett.0c00272. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Felzmann W.; Brunner s.; Lengauer H. International Patent, WO 2016/066726 A2, 2016.
  16. Dowtherm A is a eutectic mixture of diphenyl ether and biphenyl and boils at 257 °C at atmospheric pressure.
  17. a De Pascalis L.; yau M.-K.; Svatunek d.; Tan Z.; Tekkam S.; Houk K. N.; Finn M. G. The influence of substitution on thiol-induced oxanorbornadiene fragmentation. Org. Lett. 2021, 23, 3751–3754. 10.1021/acs.orglett.1c01164. [DOI] [PMC free article] [PubMed] [Google Scholar]; b Boutelle R. C.; Northrop B. H. Substituent effects on the reversibility of furan-maleimide cycloadditions. J. Org. Chem. 2011, 76, 7994–8002. 10.1021/jo201606z. [DOI] [PubMed] [Google Scholar]
  18. Magnus P.; Cairns P. M.; Moursounidis J. Trimethylsilyl accelerated retro-Diels-Alder reaction: A quantitative measure of the β-effect. J. Am. Chem. Soc. 1987, 109, 2469–2471. 10.1021/ja00242a035. [DOI] [Google Scholar]
  19. For examples of TMS-CP being used as an alkene protecting group, see:; a Maehara T.; Motoyama K.; Toma T.; Yokoshima S.; Fukuyama T. Total synthesis of (−)-tetrodotoxin and 11-norTTX-6(R)-ol. Angew. Chem., Int. Ed. 2017, 56, 1549–1552. 10.1002/anie.201611574. [DOI] [PubMed] [Google Scholar]; b Calandra N. A.; King S. M.; Herzon S. B. Development of enantioselective synthetic routes to the hasubanan and acutumine alkaloids. J. Org. Chem. 2013, 78, 10031–10057. 10.1021/jo401889b. [DOI] [PubMed] [Google Scholar]
  20. Lambert J. L.; Zhao Y.; Emblidge R. W.; Salvador L. A.; Liu X.; So J.-H.; Chelius E. C. The β effect of silicon and related manifestations of σ conjugation. Acc. Che. Res. 1999, 32, 183–190. 10.1021/ar970296m. [DOI] [Google Scholar]
  21. The carbonyl groups on aryl-amides are more electron deficient than those on alkyl-amides, which leads to a faster retro-DA rate. A similar argument can be made for the higher electrophilicity of anilinic derived acrylamides.; Flanagan M. E.; Abramite J. A.; Anderson D. P.; Aulabaugh A.; Dahal U. P.; Gilbert A. M.; Li C.; Montgomery J.; Oppenheimer O.; Ryder T.; Schuff B. P.; Uccello D. P.; Walker G. S.; Wu Y.; Brown M. F.; Chen J. M.; Hayward M. M.; Noe M. C.; Obach R. S.; Philippe L.; Shanmugasundaram V.; Shapiro M. J.; Starr J.; Stroh J.; Che Y. Chemical and computational methods for the characterization of covalent reactive groups for the prospective design of irreversible inhibitors. J. Med. Chem. 2014, 57, 10072–10079. 10.1021/jm501412a. [DOI] [PubMed] [Google Scholar]
  22. We hypothesize that small changes in reaction rate with respect to solvent polarity are consistent with a nonionic, concerted transition state for cycloreversion. As the complexes are slightly less stable in more polar solvents, this would imply a higher degree of charge separation in the transition state for cycloreversion than in the starting adduct. N,N-Dimethylformamide (εr = 36.71, μ = 3.8 D), 1,2-dichlorobenzene (εr = 9.93, μ = 2.5 D), ortho-xylene (εr = 2.57, μ = 0.64 D).; a Reichart C.; Welton T.. Solvents and Solvent Effects in Organic Chemistry, 4th ed.; Wiley-VCH: Weinheim, 2010; pp 181–182; appendix A. [Google Scholar]; b Rudolph H. D.; Walzer K.; Krutzik I. Microwave spectrum, barrier for methyl rotation, methyl conformation, and dipole moment of ortho-xylene. J. Mol. Spectrosc. 1973, 47, 314–339. 10.1016/0022-2852(73)90016-7. [DOI] [Google Scholar]; c CRC Handbook of Chemistry and Physics, 56th ed.; Weast R. C., Ed.; CRC Press: Boca Raton, FL, 1975; pp E-57. [Google Scholar]
  23. On small scale, solvent evaporation was problematic. o-Dichlorobenzene (BP = 180 °C, 1 atm) was substituted with 1,2,4-trichlorobenzene (BP = 214.4 °C, 1 atm).; a Wang H.; Liu J.; Han Y. Nano-fibrils formation of pBTTT via adding alkylthiol into solutions: Control of morphology and crystalline structure. Polymer 2013, 54, 948–957. 10.1016/j.polymer.2012.11.073. [DOI] [Google Scholar]; b Sanchez J. M.; Sacks R. D. Development of a multibed sorption trap, comprehensive two-dimensional gas chromatography, and time-of-flight mass spectrometry system for the analysis of volatile organic compounds in human breath. Anal. Chem. 2006, 78, 3046–3054. 10.1021/ac060053k. [DOI] [PubMed] [Google Scholar]
  24. Attempts to individually isolate these peaks chromatographically were not successful.
  25. The boiling point of TMS cyclopentadiene is 138–140 °C.; Palacios F.; de los Santos J. M.. 5-(trimethylsilyl)-1,3-cyclopentadiene. E-EROS 2014.1. 10.1002/047084289X.rn01739 [DOI] [Google Scholar]
  26. For the deprotection reaction generating acrylamide 33, 14% of a mixture of isomeric starting materials was recovered, and we could not account for the rest of the mass balance. When heated to 160 °C for 1 h, the reduced analogue (S16) bearing a pro-pionamide instead of an acrylamide was recovered in quantitative yield. Given the high electron deficiency of the resulting acrylamide, we think it is likely that the product is polymerizing, leading to the reduced yield.
  27. Li D.; Ambrogio L.; Shimamura T.; Kubo S.; Takahashi M.; Chirieac L. R.; Padera R. F.; Shapiro G. I.; Baum A.; Himmelsbach F.; Rettig W. J.; Meyerson M.; Solca F.; Greulich H.; Wong K.-K. BIBW2992, an irreversible EGFR/HER2 inhibitor highly effective in preclinical lung cancer models. Oncogene 2008, 27, 4702–4711. 10.1038/onc.2008.109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Rabindran S. K.; Discafani C. M.; Rosfjord E. C.; Baxter M.; Floyd M. B.; Golas J.; Hallett W. A.; Johnson B. D.; Nilakantan R.; Overbeek E.; Reich M. F.; Shen R.; Shi X.; Tsou H. R.; Wang Y. F.; Wissner A. Antitumor activity of HKI-272, an orally active, irreversible inhibitor of the HER-2 tyrosine kinase. Cancer Res. 2004, 64, 3958–3965. 10.1158/0008-5472.CAN-03-2868. [DOI] [PubMed] [Google Scholar]
  29. See the SI for details.
  30. Jacques J.; Collet A.; Wilen S. H.. Enantiomers, Racemates and Resolutions; Jon Wiley & Sons, Inc.: New York, 1981. [Google Scholar]
  31. Miller J. P.Recent advances in asymmetric Diels-Alder reactions. In Advances in Chemistry Research; Taylor J. C., Ed.; NOVA Science Publishers, Inc.: Hauppauge, New York, 2013; Vol. 18, pp 179−219. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ml2c00075_si_001.pdf (10.8MB, pdf)

Articles from ACS Medicinal Chemistry Letters are provided here courtesy of American Chemical Society

RESOURCES