首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 375 毫秒
1.
3-Aryl-2-phosphinoimidazo[1,2-a]pyridine ligands were synthesized from 2-aminopyridine via two complementary routes. The first synthetic route involves the copper-catalyzed iodine-mediated cyclizations of 2-aminopyridine with arylacetylenes followed by palladium-catalyzed cross-coupling reactions with phosphines. The second synthetic route requires the preparation of 2,3-diiodoimidazo[1,2-a]pyridine or 2-iodo-3-bromoimidazo[1,2-a]pyridine from 2-aminopyridine followed by palladium-catalyzed Suzuki/phosphination or a phosphination/Suzuki cross-coupling reactions sequence, respectively. Preliminary model studies on the Suzuki synthesis of sterically-hindered biaryl and Buchwald–Hartwig amination compounds are presented with these ligands.

3-Aryl-2-phosphoimidazo[1,2-a]pyridine ligands were prepared via two complimentary synthetic routes and were evaluated in the Suzuki–Miyaura and Buchwald–Hartwig amination cross-coupling reactions.

Palladium-catalyzed cross-coupling reactions have revolutionized the formation of C–C and C–X bond formation in the academic and industrial synthetic organic chemistry sectors.1,2 Applications such as synthesis of natural products,3 active pharmaceutical ingredients (API),4 agrochemicals,5 and materials for electronic applications6 are showcased. Snieckus described in his 2010 Nobel Prize review that privileged ligand scaffolds represented the “third wave” in the cross-coupling reactions where the “first wave” was the investigation of the metal catalyst-the rise of palladium and the “second wave” was the exploration of the organometallic coupling partner.1 In the last twenty years, it was recognized that the choice of ligand facilitated the oxidative addition and reductive-elimination steps of the catalytic cycle of transition metal-catalyzed cross-coupling reactions, increasing the overall rate of the reaction. For example, bulky trialkylphosphines facilitated the oxidative addition processes of electron-rich, unactivated substrates such as aryl chlorides.7,8 Sterically demanding ligands also provided enhanced rates of reductive elimination from [(L)nPd(aryl)(R), R = aryl, amido, phenoxo, etc.] species by alleviation of steric congestion.9 Privileged ligands such as Buchwald''s biarylphosphines,10,11 Fu''s trialkylphosphines,7,8,12 Nolan–Hermann''s N-heterocyclic carbenes (NHC),13–15 Hartwig''s ferrocenes,16,17 Beller''s bis(adamantyl)phosphines18,19 and N-aryl(benz)imidazolyl or N-pyrrolylphosphines,20,21 Zhang''s ClickPhos ligands,22,23 and Stradiotto''s biaryl P–N phosphines,24,25 to mention a few, have found wide-spread use in Suzuki–Miyaura, Corriu–Kumada, Heck, Negishi, Sonogashira, C–X (X = S, O, P) cross-coupling and Buchwald–Hartwig amination reactions (Fig. 1). Preformed catalysts with these ligands attached to the palladium metal center are also recognized as well-defined entities in cross-coupling reactions.26Open in a separate windowFig. 1Privileged ligands for palladium-catalyzed cross-coupling reactions.The term privileged structure was first coined by Evans et al. in 1988 and was defined as “a single molecular framework able to provide ligands for diverse receptors”.27 In the last three decades, it is clear that privileged structures are exploited as opportunities in drug discovery programs.28–31 For example, imidazo[1,2-a]pyridines are privileged structures in medicinal chemistry programs (Fig. 2).32 Imidazo[1,2-a]pyridines are a represented motif in several drugs on the market such as zolpidem, marketed as Ambien™ for the treatment of insomnia,33 minodronic acid, marketed as Bonoteo™ for oral treatment of osteoporosis,34 and olprinone, sold as Coretec™ as a cardiotonic agent.35Open in a separate windowFig. 2Imidazo[1,2-a]pyridines as privileged structures in medicinal chemistry and in our cross-coupling reactions approach.Our group is interested in a long-term research program directed at the use of key privileged structures that are employed in drug discovery programs as potential phosphorus ligands for cross-coupling reactions. In our entry into the use of privileged structures from the medicinal chemistry literature for our investigation into new phosphorus ligands, we have developed two complementary synthetic routes for the preparation of 3-aryl-2-phosphinoimidazo[1,2-a]pyridine ligands from 2-aminopyridine as our initial substrate.Our first synthetic route for the preparation of 3-aryl-2-phosphinoimidazo[1,2-a]pyridine ligands 3a–3l required the copper(ii) acetate iodine-mediated double oxidative C–H amination of 2-aminopyridine (1) with arylacetylenes under an oxygen atmosphere to give 3-aryl-2-iodoimidazo[1,2-a]pyridines 2a–2d (Scheme 1).36,37Open in a separate windowScheme 1Preparation of 3-aryl-2-phosphinoimidazo[1,2-a]pyridine ligands 3a–3l from 2-aminopyridine via copper-catalyzed arylacetylene cyclizations/palladium-catalyzed phosphination reactions sequences.Phenylacetylene and 2-/3-/4-methoxyphenylacetylenes were commercially available reagents. With intermediates 2a–d in hand, we explored several cross-coupling phosphination reactions and we found that palladium-catalyzed phosphination with DIPPF ligand in the presence of cesium carbonate as the base in 1,4-dioxane under reflux provided twelve new ligands 3a–3l as shown in 38 Moderate to good yields were obtained under these cross-coupling conditions. There are few commercially available dimethoxyphenylacetylenes, and most are prohibitively expensive, and so an alternative synthetic strategy was explored.Palladium-catalyzed phosphination of 3-aryl-2-iodoimidazo[1,2-a]pyridines 2a–2da
EntryArR3 (% yield)
1Ph (2a) t-Bu3a (41)
2Ph (2a)Cy3b (50)
3Ph (2a)Ph3c (61)
42-OMeC6H4 (2b) t-Bu3d (53)
52-OMeC6H4 (2b)Cy3e (83)
62-OMeC6H4 (2b)Ph3f (69)
73-OMeC6H4 (2c) t-Bu3g (62)
83-OMeC6H4 (2c)Cy3h (72)
93-OMeC6H4 (2c)Ph3i (79)
104-OMeC6H4 (2d) t-Bu3j (73)
114-OMeC6H4 (2d)Cy3k (55)
124-OMeC6H4 (2d)Ph3l (59)
Open in a separate windowaReaction conditions: 2a–2d (1 equiv.), HPR2 (1 equiv.), Pd(OAc)2 (2 mol%), Cs2CO3 (1.2 equiv.), DIPPF (2.5 mol%), 1,4-dioxane, 80 °C.2-Iodoimidazo[1,2-a]pyridine (4) was conveniently prepared in three steps from 2-aminopyridine (1) following literature procedures, which was then converted into either iodo 5 or bromo 6 with NIS or NBS, respectively (Scheme 2).39,40Open in a separate windowScheme 2Preparation of 2,3-diiodoimidazo[1,2-a]pyridine (5) and 3-bromo-2-iodoimidazo[1,2-a]pyridine (6).When the phosphorus ligands 3 contained tert-butyl or cyclohexyl groups, method 1 was followed where 2,3-diiodoimidazo[1,2-a]pyridine (5) underwent Suzuki cross-coupling reactions with arylboronic acids to yield aryl intermediates 7a–7f, which was followed by palladium-catalyzed cross-coupling phosphination reactions with di-tert-butylphosphine or dicyclohexylphosphine to give C-2 substituted phosphorus ligands 3m–3u in low to moderate yields (Scheme 3, 38 The phosphorus ligands 3v–3ab were prepared from 3-bromo-2-iodoimidazo[1,2-a]pyridine (6) via a palladium-catalyzed phospination with diphenylphosphine (method 2) to give intermediate 8 (X = Br, I becomes PPh2) followed by Suzuki palladium-catalyzed cross-coupling reactions with arylboronic acids. Note that the change in reactivity of the core when switching between bromo and iodo at C3 results in a change in the order of cross-coupling steps.Open in a separate windowScheme 3Preparation of 3-aryl-2-phosphinoimidazo[1,2-a]pyridine ligands 3m–3ab from 2-iodo-3-iodo(or bromo)imidazo[1,2-a]pyridines 5 or 6via palladium-catalyzed Suzuki/phosphination or a phosphination/Suzuki cross-coupling reactions sequences.Palladium-catalyzed Suzuki/phosphination or phosphination/Suzuki reactions sequences of 2,3-diiodoimidazo[1,2-a]pyridine (5) or 3-bromo-2-iodoimidazo[1,2-a]pyridine (6)a
EntryRArMethod/substrateStep 1 (% yield)Step 2 (% yield)
1 t-Bu2,3-diOMeC6H31, 57a (59)3m (64)
2 t-Bu3,4-diOMeC6H31, 57b (54)3n (31)
3 t-Bu2,5-diOMeC6H31, 57c (58)3o (61)
4 t-Bu3,4,5-triOMeC6H21, 57d (50)3p (62)
5Cy2,3-diOMeC6H31, 57a (59)3q (46)
6Cy2,6-diOMeC6H31, 57e (40)3r (52)
7Cy3,4-diOMeC6H31, 57b (54)3s (52)
8Cy2,3,4-triOMeC6H21, 57f (58)3t (21)
9Cy3,4,5-triOMeC6H21, 57d (50)3u (55)
10Ph2,3-diOMeC6H32, 68 (70)3v (52)
11Ph2,5-diOMeC6H32, 68 (70)3w (68)
12Ph3,4-diOMeC6H32, 68 (70)3x (67)
13Ph2,3,4-triOMeC6H22, 68 (70)3y (52)
14Ph3,4,5-triOMeC6H22, 68 (70)3z (64)
15Ph4-FC6H42, 68 (70)3aa (40)
16Ph3-F,5-OMeC6H32, 68 (70)3ab (39)
Open in a separate windowaReaction conditions: 5, ArB(OH)2, Pd(PPh3)4 (5 mol%), Na2CO3 (2 equiv.), 1,4-dioxane/H2O (2 : 1) and HPR2 (1 equiv.), Pd(OAc)2 (2.5–5 mol%), Cs2CO3 (1.2 equiv.), DIPPF (2.5–10 mol%), 1,4-dioxane, 80 °C or 6, reverse sequence of reactions.With our library of functionalized imidazo[1,2-a]pyridine phosphorus ligands 3a–3ab in hand, we began to screen these ligands in Suzuki–Miyaura cross-coupling reactions to prepare sterically-hindered biaryl compounds. We chose the Suzuki–Miyaura cross-coupling reactions of m-bromo-xylene (9) and 2-methoxyphenylboronic acid (10) to give 2,6-dimethyl-(2-methoxy)biphenyl (11) as our model reaction as outlined in ii) acetate with 2.5 equivalents of base in 1,4-dioxane at 80 °C for 12–24 h. As expected, SPhos and XPhos were employed as our initial ligands to confirm our GC analyses of >99% conversion in our chosen model reaction (Entries 14–15). With the GC conditions validated, we screened selected ligands from 3a–3ab. It was clearly evident that the di-tert-butyl phosphorus ligands represented by 3a, 3m, and 3p were ineffective ligands in our model reactions (Entries 1–3). Furthermore, the diphenyl phosphorus ligands such as 3w, 3y, 3z, and 3ab showed low to moderate conversions in the model cross-coupling reactions (Entries 6–9). However, the dicyclohexyl phosphorus ligands shown by 3r and 3t showed greater than 99% conversions by GC analyses (Entries 4–5). Further exploration of ligand 3r with K3PO4 as the base, stirring the reaction overnight at room temperature or for 3 h at 80 °C showed inferior conversions (Entries 10–12). There was no conversion when a ligand was not used in the model reaction (Entry 13).Optimization of conditions for the Suzuki–Miyaura cross-coupling model reaction
EntryLigandConditionsConversiona (%)
13a12
23m20
33p14
4 3r >99 b
53t>99
63w21
73y55
83z46
93ab11
103rK3PO4 was used as base reaction was performed at 25 °C reaction was stirred for 3 h no ligand91
113r4
123r39
130
14SPhos>99
15XPhos>99
Open in a separate windowaBased on GC analyses of consumed 9.bIsolated yield of 96% was obtaisned.Furthermore, a Buchwald–Hartwig amination model study was investigated with our new imidazo[1,2-a]pyridine phosphorus ligands 3a–3ab. The Buchwald–Hartwig amination reaction of 4-chlorotoluene (12) with aniline (13) to give 4-methyl-N-phenylaniline (14) was screened with our ligands (
EntryLigandConditionsConversiona (%)
13a38
23d26
3 3e >99 b
43g29
53h54
63k71
73n0
83p0
93q>99
103r92
113s>99
123sK3PO4 was used as base83
133sK2CO3 was used as base0
143sKOt-Bu was used as base>99
153sNaOt-Bu was used as base>99
Open in a separate windowaBased on GC analyses of consumed 13.bIsolated yield of 76% was obtained.In summary, we have disclosed two complementary synthetic routes to 3-aryl-2-phosphinoimidazo[1,2-a]pyridine ligands 3a–3ab from 2-aminopyridine (1). In one method, 2-aminopyridine (1) underwent a copper-catalyzed iodine-mediated cyclization with arylacetylenes followed by palladium-catalyzed cross-coupling reactions with phosphines. In the second protocol, 2,3-diiodoimidazo[1,2-a]pyridine (5) or 3-bromo-2-iodoimidazo[1,2-a]pyridine (6) were prepared from 2-aminopyridine (1) followed by palladium-catalyzed phosphination/Suzuki or Suzuki/phosphination reactions sequences, respectively. We are currently exploring the scope and limitations of the 3-aryl-2-phosphinoimidazo[1,2-a]pyridine ligand 3r and 3e in our Suzuki–Miyaura and Buchwald–Hartwig amination cross-coupling reactions, respectively.  相似文献   

2.
Oxidative cyanation of N-aryltetrahydroisoquinoline induced by visible light for the synthesis of α-aminonitrile using potassium thiocyanate as a “CN” agent     
Bing Yi  Ning Yan  Niannian Yi  Yanjun Xie  Xiaoyong Wen  Chak-Tong Au  Donghui Lan 《RSC advances》2019,9(51):29721
A novel method for the synthesis of α-aminonitrile through visible-light-induced oxidative cyanation of N-aryltetrahydroisoquinoline with potassium thiocyanate has been developed. The process does not require the use of a photocatalyst, transition metal reagent, strong oxidizing agent, or toxic cyano-containing compound, which makes the reaction simple and green.

A novel method for the synthesis of α-aminonitrile, through visible-light-induced oxidative cyanation of N-aryltetrahydroisoquinoline with potassium thiocyanate, has been developed.

α-Aminonitriles are versatile intermediates which can be readily converted into a range of multifunctional organics,1 such as α-aminocarbonyl compounds, α-amino acids, 1,2-diamines and many others. Moreover, many nitrile-containing compounds have been widely used as drugs for treating various diseases.2 For example, amphetaminil3 is a stimulant that can be employed to treat obesity and narcolepsy; others like Saxagliptin,4 NVP-DPP728 (ref. 5) and Vildagliptin6 are potent reversible inhibitors of dipeptidyl peptidase-4 (DPP-4), and can be applied as novel antidiabetic drugs. Therefore, the development of novel, efficient and clean methods for the synthesis of α-aminonitriles has attracted wide attention.Traditionally, α-aminonitriles are synthesized through the Strecker reaction.7 Another momentous strategy for synthesis of α-aminonitriles is oxidative cyanation of tertiary amines with the involvement of transition metal reagents (such as Ru,8 Au,9 Cu,10 Fe,11 Co,12 V,13 Mo,14etc.) and non-metallic reagents (such as PhI(OAc)2,15 tropylium salt,16 AIBN,17 TBAI,18 AcOH,19 PIFA,20 TBHP,21 thiourea,22 DDQ,23 TBPB,24aetc.). Among the reported cyanidation of tertiary amines, the cyano sources mainly includes (i) metal-cyanides such as NaCN,8a–8c,10c KCN,16 K3[Fe(CN)6],25 KSCN,21etc., (ii) organic cyanides such as trimethylcyanosilane,9a,10e,11a–11c,11f malononitrile,10a,15,23 ethyl cyanoformate,11d,8e,11d,12 benzoyl cyanide,11e phenylacetonitrile,18a cyanoacetic acid,18b tetrabutylammonium cyanide,19 cyanobenziodoxolones,24 AIBN,26etc., and (iii) combined cyano source such as the combination of 1,2-dichloroethane and trimethylsilyl azide.10b (Scheme 1). However, most of these methods require transition metal reagents, strong oxidizing agents, and cyano sources that are highly toxic and difficult to obtain. Therefore, it is necessary to find a cyanide source that is easily available and of low toxicity for the cyanidation of tertiary amines.Open in a separate windowScheme 1Cyano sources for cyanidation reactions of tertiary amine.In organic synthesis, thiocyanate has received extensive attention due to its low toxicity, low cost and easy availability.27 Using the thiocyano group of thiocyanate, thiocyanation of organic molecules can be achieved by nucleophilic substitution or through a free radical pathway.28 More importantly, thiocyanate can be utilized as a clean and safe source of cyano group.21,29 However, these methods have the following disadvantages: using a transition metal catalyst, or a strong oxidizing agent, or heating. The utilization of light energy in organic synthesis is highly commendable because it is sustainable, clean, environmentally benign, as well as widely and easily available. In recent years, great progresses have been made in the formation of carbon–carbon bond that is induced by visible light,30 among which the oxidative cyanation of tertiary amine is a good example.31 However, to the best of our awareness, the cyanidation of tertiary amine with thiocyanate induced by visible light has not been documented yet. Herein, we report the visible-light-induced oxidative cyanation reaction of tertiary amines with potassium thiocyanate.We began our investigation with the use of N-phenyltetrahydroisoquinoline (1a) and potassium thiocyanate (2a) as the model reaction ( EntryOxidantLightSolventYieldb (%)1Air30 W CFLC2H5OH502O230 W CFLC2H5OH553TBHP30 W CFLC2H5OH254H2O230 W CFLC2H5OH175K2S2O830 W CFLC2H5OH196DDQ30 W CFLC2H5OH37PhI(OAc)230 W CFLC2H5OH58O230 W red LEDsC2H5OH69O230 W green LEDsC2H5OH6810O230 W blue LEDsC2H5OH7411O230 W purple LEDsC2H5OH8312O230 W purple LEDsCH3OH1113O230 W purple LEDsDMSO814O230 W purple LEDsPhMe1115O230 W purple LEDsCH3CN9216cO230 W purple LEDsCH3CN7117—30 W purple LEDsC2H5OH018O2—C2H5OH0Open in a separate windowaReaction conditions: 1a (0.2 mmol), 2a (0.1 mmol), oxidant (1.5 equiv.), solvent (1.5 mL), rt, 24 h.bIsolated yields based on 2a.cNH4SCN was used instead of KSCN.With the optimized conditions, we explored the substrate scope of this visible-light-induced cyanation reaction between N-aryltetrahydroisoquinoline (1) and potassium thiocyanate (2a) (Scheme 2). The N-aryltetrahydroisoquinolines with either an electron-withdrawing or electron-donating group reacted well with potassium thiocyanate to afford the desired products in moderate to good yields. Specifically, when the benzene ring attached to the nitrogen atom had an electron donating group (Me, Et, and Ph) at the para position, the desired products were in 79%, 86% and 62% yields, respectively (3b–d). In the cases of having an electron-withdrawing group (F, Cl, Br, CN, CF3, and OCF3) attached to the benzene ring at para position, the desired products were obtained in moderate to good yields, and the highest one was up to 88% (3e–j). Besides, ortho- and meta-substituted N-aryltetrahydroisoquinolines also exhibited high reactivity, and the target products were obtained in moderate yields (3k–n).Open in a separate windowScheme 2Substrate scope for cyanation of N-aryltetrahydroisoquinolines (1) with potassium thiocyanate (2a)a. aConditions: 1 (0.2 mmol), 2a (0.1 mmol), CH3CN (1.5 mL), 24 h, isolated yields of column chromatography based on 2a. bThe yield was based on 1 mmol of KSCN. cNo product was detected by GC-MS.What''s more, the substrates with two identical or different substituents on the phenyl ring were also suitable for this reaction, and the yields of the products vary from 60% to 80% (3o–q). Regrettably, dialkyl substituted aromatic amine (such as N,N-dimethylaniline) was not suitable for this reaction (3r).To gain reasonable insight into the reaction mechanism, we conducted the following control experiments (Scheme 3). When 2,2,6,6-tetramethyl-1-piperidinyloxy (TEMPO) and 2,6-di-tert-butyl-4-methylphenol (BHT), both are radical quenchers, was added to the reaction under the standard conditions, the product 3a was obtained in 3% and 9% yield, respectively, detected by GC. The results reveal that the reaction may proceed with the participation of free radicals.Open in a separate windowScheme 3Control experiments.Based on our experimental results and those of previous reports,19,21,32,33 a mechanism for this visible-light-induced oxidative cyanation reaction is proposed as illustrated in Scheme 4. Initially, substrate 1 is transformed into cationic radical I in the presence of visible light and oxygen, and anionic superoxide radical O2˙ is generated simultaneously. Then, cationic radical I reacts with O2˙ to generate radical II which loses an electron to form intermediate V. Under the action of HOO, thiocyanate is converted to cyano anion which then reacts with intermediate V to generate the desired product 3. At the same time, the intermediate II is combined with HOO˙ to form III, and the latter III is dehydrated to form a by-product IV.Open in a separate windowScheme 4Possible mechanism.  相似文献   

3.
Correction: Droplet microfluidics: fundamentals and its advanced applications     
Somayeh Sohrabi  Nour Kassir  Mostafa Keshavarz Moraveji 《RSC advances》2020,10(54):32843
Correction for ‘Droplet microfluidics: fundamentals and its advanced applications’ by Somayeh Sohrabi et al., RSC Adv., 2020, 10, 27560–27574, DOI: 10.1039/D0RA04566G.

The authors regret the omission of a funding acknowledgement in the original article. This acknowledgement is given below.The authors would like to acknowledge the financial support of the Iran National Science Foundation (INSF), grant number 98017171.In addition, the authors regret that incorrect reference numbers were given in Geometry and materialContinuous phaseSize/μmFrequency/HzRef. in original article reference listRef. in this CorrectionWater in oilChannel array in siliconKerosene with monolaurate21∼5300 (est.)— 1 T-junction in acrylated urethaneDecane, tetradecane, and hexadecane with Span 8010 to 3520 to 80— 2 T-junction in PMMAHigh oleic sunflower oil100 to 35010 to 2500— 3 T-junction in PDMSC14F12 with (C6F13)(CH2)2OH7.5 nl (plug flow)255 4 Shear-focusing in PDMSOleic acid13 to 35 (satellites <100 nm)15–10049 5 Oil in waterChannel array in siliconWater with SDS22.5∼5300 (est.)— 1 Sheath flow in glass capillaryWater with SDS2 to 200100 to 10 000— 6 Gas in liquidFlow-focusing in PDMSWater with Tween 2010 to 1000>100 000— 7 Shear-focusing in PDMSWater with phospholipids5 to 50>1 000 000— 8 Liquid in airDEP on hydrophobic insulatorAir10 pl∼8 (est.)57 9 EWOD on hydrophobic insulatorAir∼700 nl∼1 (est.)28 10 Open in a separate windowThe Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

4.
ortho-Naphthoquinone-catalyzed aerobic oxidation of amines to fused pyrimidin-4(3H)-ones: a convergent synthetic route to bouchardatine and sildenafil     
Kyeongha Kim  Hun Young Kim  Kyungsoo Oh 《RSC advances》2020,10(52):31101
A facile access to fused pyrimidin-4(3H)-one derivatives has been established by using the metal-free ortho-naphthoquinone-catalyzed aerobic cross-coupling reactions of amines. The utilization of two readily available amines allowed a direct coupling strategy to quinazolinone natural product, bouchardatine, as well as sildenafil (Viagra™) in a highly convergent manner.

Fused pyrimidin-4(3H)-one derivatives have been accessed by using the ortho-naphthoquinone-catalyzed aerobic cross-coupling reactions of amines.

N-Heterocyclic compounds with a pyrimidin-4(3H)-one core constitute a large number of natural products and biologically active molecules. For example, quinazolinone alkaloids possess a phenyl-fused pyrimidin-4(3H)-one structure and display a wide spectrum of pharmacological activities (Scheme 1a).1 Sildenafil (Viagra™), a potent and selective inhibitor of type 5 phosphodiesterases on smooth muscle cell, is based on the pyrazole-fused pyrimidin-4(3H)-one structure and marketed for erectile dysfunction.2 The synthetic approaches to the phenyl-fused pyrimidin-4(3H)-ones, quinazolinones, typically involve the condensation between anthranilamides and aldehydes to give aminal intermediates that in turn oxidized to quinazolinones under oxidation conditions (Scheme 1b). The oxidation catalysts include Cu,3 Fe,4 Ga,5 Ir,6 Mn,7 iodine,8 peroxide,9 however the aerobic oxidation of aminal intermediates is also known at 150 °C.10 The utilization of alcohols also effects the one-pot synthesis of quinazolinones through in situ oxidation to aldehydes in the presence of Fe,11 Ir,12 Mn,13 Ni,14 Pd,15 Ru,16 V,17 Zn,18 and iodine catalysts.19 Other precursors to aldehydes have been also identified using alkynes,20 benzoic acids,21 indoles,22 α-keto acid salts,23 β-keto esters,24 styrenes,25 sulfoxides,26 and toluenes.27 Non-aldehyde approaches to quinazolinones have been also demonstrated in the cross coupling of amidines,28 amines,29 benzamides,30 isocyanides,31 and nitriles.32 In 2013, the Nguyen group disclosed the synthesis of four quinazolinones, utilizing the autooxidation of benzylamines to imines that subsequently condensed with anthranilamides.33 While a closed system at 150 °C was necessary, the use of 40 mol% AcOH without solvent provided the quinazolinones in 46–75% yields. The Nguyen group also developed the FeCl3·6H2O-catalyzed condensation of 2-nitroanilines and benzylamines in the presence of 20 mol% of S8, where six quinazolinones were obtained in 68–75% yields.34 While the cross condensation of anthranilamides and benzylamines was accomplished, there exists a significant knowledge gap due to the limited substrate scope combined with less optimal reaction conditions (i.e. high reaction temperature, closed system in neat conditions and excess use of amines). In addition, it is not entirely clear if the cross condensation of amide-containing amines and benzylamines would work for other fused pyrimidin-4(3H)-one derivatives.35 To address such shortcomings in the cross condensation of amines, the ortho-naphthoquinone (o-NQ)-catalyzed aerobic cross amination strategy was investigated (Scheme 1c).36 Herein, we report a highly general approach to fused pyrimidin-4(3H)-one derivatives in the presence of o-NQ catalyst, culminating to the direct aerobic coupling of two amines to bouchardatine and sildenafil.Open in a separate windowScheme 1Biologically active fused pyrimidin-4(3H)-one derivatives and their synthetic methods.Given that the o-NQ-catalyzed aerobic cross coupling of benzylamines and aniline derivatives such as o-phenylenediamines provided a facile approach to heterocyclic compounds including benzoimidazoles,37 the use of anthranilamide 1a and benzylamine 2a was examined as a model study (38 The reaction in DMSO lowered down the ratio of 1a and 2a from 1.0 : 1.5 to 1.0 : 1.2 (entry 12) and the catalyst loading to 5 mol% without affecting the overall reaction efficiency (entry 13). The control experiments confirmed the critical roles of both o-NQ1 and TFA (entries 15–18), and the reaction utilized molecular oxygen as a terminal oxidant (entries 19 and 20). Piecing together the experimental data, the employment of 5 mol% o-NQ1 and 20 mol% TFA in DMSO at 100 °C was selected for further studies.Optimization of o-NQ-catalyzed aerobic cross-coupling of amines to quinazolinonea
EntryCat. (mol%)Solvent T (°C)Yieldb (%)
1 o-NQ1 (10)CH3CN803a, 11
2 o-NQ1 (10)/TFA (20)CH3CN803a, 83
3 o-NQ2 (10)/TFA (20)CH3CN803a, 13
4 o-NQ3 (10)/TFA (20)CH3CN803a, 66
5 o-NQ4 (10)/TFA (20)CH3CN803a, 32
6 o-NQ1 (10)/TFA (20)MeOH653a, 5
7 o-NQ1 (10)/TFA (20)EtOH783a, 31
8 o-NQ1 (10)/TFA (20)DMF1504a, >95
9c o-NQ1 (10)/TFA (20)DMSO1504a, >95
10 o-NQ1 (10)/TFA (20)DMSO1004a, >95
11 o-NQ1 (10)/TFA (20)DMSO803a, 30
12d o-NQ1 (10)/TFA (20)DMSO1004a, >95
13d o-NQ1 (5)/TFA (20)DMSO1004a, >95 (93)
14 o-NQ1 (5)/TFA (10)DMSO1003a/4a, 45/50
15DMSO1003a, 10
16 o-NQ1 (10)DMSO1003a/4a, 34/51
17TFA (20)DMSO100NR
18 o-NQ1 (5)/AcOH (20)DMSO1003a, 25
19e o-NQ1 (5)/TFA (20)DMSO1003a, 33
20f o-NQ1 (5)/TFA (20)DMSO100NR
Open in a separate windowaReaction using 1a (0.20 mmol), 2a (0.30 mmol), and o-NQ in solvent (0.2 M) under O2 balloon for 24 h.bYields based on internal standard and isolated yield in parentheses.cReaction for 6 h.dUse of 2a (0.24 mmol, 1.2 equiv.).eReaction under air.fReaction under argon. NR = no reaction.The optimized aerobic cross-coupling condition was applied to a variety of benzylamine derivatives (Scheme 2). In general, the electronic and steric characters of benzylamines did not significantly affect the formation of quinazolinones (4a–4m). However, the use of halogen-substituted and dimethoxy-substituted benzylamines led to the slightly lower yields of quinazolinones (4e, 4f and 4m) in 58–75% yields. In addition, the current aerobic cross-coupling reaction tolerated the furanyl and thiophenyl moieties, where the corresponding quinazolinones (4o and 4p) were obtained in 54% and 75% yields, respectively.Open in a separate windowScheme 2Substrate scope of aerobic oxidation to quinazolinones (areaction for 36 h, breaction for 12 h).Further extension of the current aerobic cross-coupling reactions of amines is illustrated in Scheme 3. Thus, an array of substituted anthranilamides was readily employed to give the fused pyrimidin-4(3H)-one derivatives (4q–4x) in 61–84% yields. In particular, the N-substituted anthranilamides also participated in the current aerobic cross-coupling reaction in excellent yields (4y–4za). While the use of 3-amino-2-naphthamide led to the corresponding quinazolinone 4zb in 46% yield, the synthetic advantage of the current method was well demonstrated in the preparation of heteroaryl fused pyrimidin-4(3H)-one derivatives (4zc–4zh), where a variety of heterocyclic amines were successfully utilized in a tandem sequence of aerobic oxidation processes.Open in a separate windowScheme 3Further substrate scope for fused pyrimidin-4(3H)-one derivatives (areaction at 120 °C, breaction at 140 °C).The mechanistic rationale of the aerobic cross-coupling reactions of amines is depicted in Scheme 4. Thus, the benzylamine 2a is condensed with the o-NQ1 catalyst to give the naphthol-imine species A.37a While the nucleophilic attack of 2a to the naphthol-imine A is favored due to the low nucleophilicity of the anthranilamide 1a, the presence of TFA promotes the cross-coupling between naphthol-imine A and anthranilamide 1a to give the naphthol-aminal B1. This process releases the hetero-coupled imine 3a′ and naphthol-amine C. The use of TFA promotes the intramolecular Mannich cyclization of imine 3a′, leading to the aminal 3a that in turn converts to the desired fused pyrimidin-4(3H)-one 4a with the help of o-NQ1 catalyst and molecular oxygen. Alternatively, the naphthol-imine A can produce the homocoupled imine 2a′ and the naphthol-amine Cvia the naphthol-aminal B2 through the nucleophilic attack of benzylamine 2a. The conversion of the naphthol-amine C to o-NQ1 catalyst is effected upon exposure to oxygen atmosphere.38 The homocoupled imine 2a′ undergoes hydrolysis at >80 °C to the benzaldehyde and benzylamine 2a that in turn re-enters the catalytic cycle.37b Our experimental observation of the homocoupled imine 2a′ by the 1H NMR and TLC analysis during the reaction supports the involvement of 2a′. However, the major pathway to the fused pyrimidin-4(3H)-one 4a appears to involve the naphthol-aminal B1 since the use of benzaldehyde instead of benzylamine 2a under the optimized conditions only led to the 80% conversion. The o-NQ1 catalyst without added TFA provided a mixture of aminal 3a and quinazolinone 4a in 34% and 51% yields, respectively (39Open in a separate windowScheme 4Mechanistic rationale for aerobic cross coupling reaction of amines.The synthetic utility of the aerobic cross-coupling strategy is demonstrated in the synthesis of quinazolinone alkaloids and sildenafil (Scheme 5). The direct cross coupling of a commercially available pyrazole amine 1o and benzylamine 2r afforded a highly convergent synthetic approach to sildenafil.40 Likewise, the employment of anthranilamide 1a and 2-(aminomethyl)indole 2s provided the desired quinazolinone 5 in 70% yield, and the subsequent formylation under the Zeng''s conditions41 paved a way to the total synthesis of bouchardatine. In addition, while the basicity of quinolin-2-ylmethanamine 2t required an excess of TFA, the corresponding quinazolinone 6 was obtained in 60% yield under the optimized conditions. The conversion of 6 to the luotonin natural products has been reported by the Argade group and others.42Open in a separate windowScheme 5Synthetic utilization to quinazolinone alkaloids and sildenafil.  相似文献   

5.
Rapid and halide compatible synthesis of 2-N-substituted indazolone derivatives via photochemical cyclization in aqueous media     
Hui-Jun Nie  An-Di Guo  Hai-Xia Lin  Xiao-Hua Chen 《RSC advances》2019,9(23):13249
Indazolone derivatives exhibit a wide range of biological and pharmaceutical properties. We report a rapid and efficient approach to provide structurally diverse 2-N-substituted indazolones via photochemical cyclization in aqueous media at room temperature. This straightforward protocol is halide compatible for the synthesis of halogenated indazolones bearing a broad scope of substrates, which suggests a new avenue of great importance to medicinal chemistry.

A straightforward protocol for the rapid construction of privileged indazolone architectures suggests a new avenue of great importance to medicinal chemistry.

The indazolone ring system constitutes the core structural element found in a large family of nitrogen heterocycles as exemplified by those shown in Scheme 1.1 Indazolone derivatives have been receiving much attention due to their promising pharmacological activities. Given the unique bioactive core skeleton, indazolone derivatives exhibit a wide range of biological and pharmaceutical properties such as antiviral and antibacterial activities (1–4),2 new prototypes for antichagasic drugs (5),3 antihyperglycemic properties (6),4 TRPV1 receptor antagonists for analgesics (7),5 anti-flammatory agents (8),6 angiotensin II receptor antagonists (9),7 highly potent CDKs inhibitors for anticancer (10)8 and so on.9 The privileged indazolone structures have high potential as core components for the development of related compounds leading to medicinal agents.Open in a separate windowScheme 1Biologically active molecules containing indazolone skeletons.Due to their versatility in pharmaceutical applications, many synthetic approaches have been developed for the construction of indazolone skeletons (Scheme 2), including CuO-mediated coupling of 2-haloarylcarboxylic acids with methylhydrazine,10 cyclization of N-aryl-o-nitrobenzamides through Ti(vi) reagent or Zn(ii) reagent,11 Cu(i)-mediated intramolecular C–N bond formation or a base-mediated intramolecular SNAr reaction of 2-halobenzohydrazides,12 Cu(i)-catalyzed oxidative C–N cross-coupling and dehydrogenative N–N formation sequence,13 Rh-catalyzed C–H activation/C–N bond formation and Cu-catalyzed N–N bond formation between azides and arylimidates,14 Friedel–Crafts cyclization of N-isocyanates using Masked N-isocyanate precursors,15 PIFA-mediated intramolecular oxidative N–N bond formation by trapping of N-acylnitrenium intermediates,16 and recently reported reaction of o-nitrobenzyl alcohol with primary amines in basic conditions.17 These approaches are complementary providing avenue to access various substitution patterns,18 however most methods rely on the requirements for transition-metal catalysts. In fact, the procedures for synthesis indazolone skeletons from Friedel–Crafts cyclization of N-isocyanates and Davis–Beirut derived reaction still suffer from harsh reaction conditions such as high reaction temperature (i.e. more than 150 °C or 20 equiv. of KOH at 100 °C for 24 h).15,17b Very recently, one photochemical route was reported for preparation of indazolone skeletons from o-nitrobenzyl alcohols and primary amines,19 however, this approach still need long reaction time (24 hours) and halogen substituted substrate could not be compatible in the reaction conditions.19b Thus, the efficient and general methods tolerating a wide scope of readily available starting materials for synthesis of indazolones without a transition-metal catalyst involved are still in great demand.Open in a separate windowScheme 2Representative approaches for the preparation of indazolone skeletons. o-Nitrobenzyl alcohol derivatives have shown many applications in material science and chemical biology area as a photolabile protecting group (Scheme 3a).20 Upon UV light-activation, o-nitrobenzyl alcohol derivatives generate corresponding aryl-nitroso compounds via photoisomerization.21 Based on the distinguishing feature of highly reactive of these photogenerated intermediates, we assumed that the reaction conditions would be crucial for the photoisomerization,20,22 thus the reactive intermediates should spontaneously and rapidly form indazolone structures via cyclization in the presence of primary amines in suitable reaction conditions (Scheme 3b). Herein, we report a rapid and efficient approach to provide structural diversity 2-N-substituted indazolones via photochemical cyclization in aqueous media at room temperature. This photochemical cyclization reaction is halide compatible for synthesis of halogen substituted indazolones, bearing a broad scope of substrates. This straightforward protocol for rapid construction of halogenated indazolone architectures suggests a new avenue of great importance to medicinal chemistry.Open in a separate windowScheme 3Synthesis of indazolone derivatives via photochemical cyclization.The initial investigation to develop a method for synthesis of indazolone derivatives via photochemical cyclization started with 4-(hydroxymethyl)-3-nitro-N-propylbenzamide 11 and heptan-1-amine 12 upon UV light-activation in methanol, smoothly leading to the formation of indazolone 13 in 52% yield ( EntrySolvent11 : 12Time (h)Yieldf (%)1MeOH2.5 : 13522THF2.5 : 13583 n-BuOH2.5 : 13574CH3CN2.5 : 13615CH3CN : H2O = 3 : 12.5 : 13676CH3CN : PBS = 3 : 12.5 : 1361 7 n-BuOH : H 2 O = 3:1 2.5 : 1 3 82 8 n-BuOH : PBS = 3 : 12.5 : 13569MeOH : H2O = 3 : 12.5 : 133810i-PrOH : H2O = 3 : 12.5 : 136311 tBuOH : H2O = 3 : 12.5 : 135512THF : H2O = 3 : 12.5 : 134913DMF : H2O = 3 : 12.5 : 134514Dioxane : H2O = 3 : 12.5 : 136715b n-BuOH : H2O = 3 : 12.5 : 13<1016c n-BuOH : H2O = 3 : 12.5 : 132217d n-BuOH : H2O = 3 : 11.5 : 134518e n-BuOH : H2O = 3 : 12.5 : 132819 n-BuOH : H2O = 3 : 12.5 : 168520PBS2.5 : 131921 n-BuOH : H2O = 3 : 11 : 2.5346Open in a separate windowaReaction conditions: heptan-1-amine (12, 0.3 mmol), 4-(hydroxymethyl)-3-nitro-N-propylbenzamide (11, 0.75 mmol), solvent 6 mL, exposed to UV lamp with 365 nm, at R.T.bUV lamp with 254 nm.cUsing blue light.dThe ratio of 11/12 = 1.5/1.eReaction carried out at 50 °C.fIsolated yield.With the optimal conditions in hand, we investigated the generality for the scope of o-nitrobenzyl alcohols and primary amines ( Open in a separate windowaReaction conditions: primary amines (15, 0.3 mmol), o-nitrobenzyl alcohols derivatives (14, 0.75 mmol), solvent 6 mL, isolated yield.Many drug candidates and drugs are halogenated structures. In drug discovery, insertion of halogen atoms on hit or lead compounds was predominantly performed, with the aim to exploit their steric effects and structure–activity relationship, to form halogen bonds in ligand–target complexes, to optimize the ADME/T property.23 Given the versatility of halogen atom on bioactive molecules, we next investigated the halogen substituted o-nitrobenzyl alcohols as starting materials for construction of indazolone skeletons (17b as well as aniline failed to give product in the recently reported photochemical approach (see ESI, Fig. S1).19b Of note, in the recently reported photochemical approach, the reaction of chloride substituted o-nitrobenzyl alcohol with alkylamine gave indazolone with low yield, possibly because photocleavage of aryl halide bond is involved in that reaction conditions.19b These outcomes are significant in view of the challenges in construction of indazolone skeletons, in which additional halogen substitution on substrates is incompatible for indazolones synthesis.10,19b Importantly, our reaction condition is compatible with halide substrates, suggests a new protocol of importance to photochemical reactions, in which dehalogenation of aryl halide is known to be radical-mediated and exist in some reaction conditions.24Scope of halogen substituted o-nitrobenzyl alcohols for indazolone formationa
Open in a separate windowaReaction conditions: primary amines (15, 0.3 mmol), o-nitrobenzyl alcohols derivatives (14, 0.75 mmol), solvent 6 mL, isolated yield.With developed and optimized protocol, we rapid synthesis of indazolones 1 and 2 with antiviral and antibacterial activities with good to excellent yields,2 in aqueous media at room temperature for 3 hours (Scheme 4), respectively. This rapid and efficient access to the privileged indazolone architectures will have great usefulness in medicinal chemistry.Open in a separate windowScheme 4Straightforward synthesis of indazolones (1 and 2). Reaction conditions: primary amines (0.3 mmol), o-nitrobenzyl alcohol (0.75 mmol), isolated yield.Finally, we could detect the aryl-nitroso compound as photogenerated intermediate on UPLC-MS analysis (see ESI, Fig. S2). The proposed mechanism for this photochemical cyclization begins by generation of the aryl-nitroso compound which can rapidly undergo cyclization with primary amines, subsequent for dehydration and tautomerization (see ESI, Fig. S3).17In summary, we report a photochemical cyclization approach to provide 2-N-substituted indazolones up to 99% yield with structural diversity from the reaction of o-nitrobenzyl alcohols and primary amines in aqueous media at room temperature. This photochemical cyclization reaction is rapid and halide compatible for synthesis of halogenated indazolones, bearing a broad scope of substrates, while previous reported photochemical reaction has met less success.19b In addition, our reaction condition is compatible with halide substrates, suggests a new protocol of importance to photochemical reactions, in which photocleavage of aryl halide bonds is exist in some reaction conditions. The current transformation enabling rapid and efficient access to the privileged indazolone architectures has great usefulness in medicinal chemistry and diversity-oriented synthesis, thus will provide promising candidates for chemical biology research and drug discovery.  相似文献   

6.
Pd-catalyzed [3 + 2] cycloaddition of vinylcyclopropanes with 1-azadienes: synthesis of 4-cyclopentylbenzo[e][1,2,3]oxathiazine 2,2-dioxides     
Yan Lin  Qijun Wang  Yang Wu  Chang Wang  Hao Jia  Cheng Zhang  Jiaxing Huang  Hongchao Guo 《RSC advances》2018,8(71):40798
The palladium-catalyzed [3 + 2] cycloaddition of vinylcyclopropanes and 1-azadienes has been developed under mild reaction conditions, giving the multisubstituted cyclopentane derivatives in good to excellent yields with moderate to good diastereoselectivities. The relative configuration of both diastereomers of the products have been determined through X-ray crystallographic diffraction.

Pd-catalyzed [3 + 2] cycloaddition of vinylcyclopropanes with 1-azadienes gave highly functionalized cyclopentane derivatives in high yields.

The cyclopentane framework is ubiquitous in nature and it is also an important structural moiety in many pharmaceuticals, agrochemicals, and materials.1 The development of simple, fast and efficient synthetic methods for highly substituted cyclopentanes has attracted much attention. Among various methods for synthesis of cyclopentane structure, the cycloaddition reaction is a very attractive one.2Under palladium catalysis conditions, vinylcyclopropane derivatives underwent a ring-opening reaction to generate the zwitterionic allylpalladium intermediate, which reacted with carbon–carbon, carbon–oxygen, carbon–nitrogen double bonds and diazo compounds to provide a variety of five-membered cyclic compounds.3 In the past decades, this type of annulation reactions has emerged as a powerful tool for the synthesis of carbocyclic and heterocyclic compounds.2 Diverse substrates including isocyanates,4 aldehydes,5 isatins,6 3-diazooxindoles,7 electron-deficient alkenes such as para-quinone methides,8 α,β-unsaturated aldehydes,9 β,γ-unsaturated α-keto esters,10 nitroolefins,11 azlactone- and Meldrum''s acid alkylidenes,12 and α-nucleobase substituted acrylates,13 have been exploited in palladium-catalyzed [3 + 2] cycloadditions of vinyl cyclopropane, delivering biologically interesting functionalized heterocyclic compounds and cyclopentane derivatives. In 2015, Liu and He reported a palladium-catalyzed [3 + 2] cycloaddition of vinyl cyclopropane and α,β-unsaturated imines generated in situ from aryl sulfonyl indoles, providing the optically enriched spirocyclopentane-1,3′-indolenines with high diastereoselectivity.14 This is the only example where α,β-unsaturated imines were employed in Pd-catalyzed [3 + 2] cycloaddition reaction of vinyl cyclopropane.As a type of α,β-unsaturated imines, cyclic 1-azadienes such as (E)-4-styrylbenzo[e][1,2,3]oxathiazine 2,2-dioxides 2 are easily accessible and stable (Scheme 1). In particular, they contain the sulfonate-moiety, which is an interesting biologically important motif, and has a great potential in the synthesis of bioactive molecules.15 The cyclic 1-azadienes have been used in a series of annulation reactions such as [2 + n],16–18 [3 + n],19 and [4 + n]20 annulation reactions. Based on the electron-deficient nature of the carbon–carbon double bond in these cyclic 1-azadienes, in 2016, Chen and Ouyang developed cinchona-derived tertiary amine-catalyzed asymmetric [3 + 2] annulation of isatin-derived Morita–Baylis–Hillman (MBH) carbonates with cyclic 1-azadienes to form spirooxindole.17 In the same year, our group demonstrated a phosphine-catalyzed [3 + 2] annulation of MBH carbonates with cyclic 1-azadienes.18 Considering that the 1-azadiene 2 is an electron-deficient alkene with good reactivity, its reaction with a zwitterionic π-allyl Pd complex formed via ring-opening of vinylcyclopropane may be feasible. However, this type of cyclic 1-azadienes have never been used in Pd-catalyzed annulation reactions involving vinylcyclopropanes. As our continuing interest on cycloaddition reactions,21 herein we disclose a [3 + 2] cycloaddition of palladium-catalyzed vinyl cyclopropane with cyclic 1-azadienes to afford the multisubstituted cyclopentane derivatives (Scheme 1).Open in a separate windowScheme 1[3 + 2] Cycloaddition of 1,3-zwitterions with cyclic 1-azadienes.We carried out an initial screening with 2-vinylcyclopropane-1,1-dicarbonitrile 1a and (E)-4-styrylbenzo[e][1,2,3]oxathiazine 2,2-dioxide 2a in CH2Cl2 (DCM) at room temperature in the presence of Pd2(dba)3·CHCl3 (2.5 mol%) (22 Under the optimized reaction conditions, we attempted to develop asymmetric variant of this reaction. Unfortunately, the two enantiomers of the product could not be resolved at present stage.23Optimization of the reaction conditionsa
Entry x L (y mol%)Solventt/hYield (%)3aa : 3aa′b
12.5DCM24965 : 1
22.5PPh3 (10)DCM5813 : 1
32.5dppp (5.0)DCM5975 : 1
42.5DPEphos (5)DCM5892 : 1
52.5dppbz (5)DCM15956 : 1
62.5Xantphos (5)DCM0.5966 : 1
71.0Xantphos (2)DCM0.5866 : 1
82.5Xantphos (5)Toluene0.5976 : 1
92.5Xantphos (5)THF0.5926 : 1
102.5Xantphos (5)DCE1.5994 : 1
112.5Xantphos (5)MeCN5685 : 1
Open in a separate windowaUnless otherwise stated, all reactions were carried out with 1a (0.12 mmol), 2a (0.10 mmol) and catalyst in solvent (2 mL) at room temperature.bDetermined by isolated yield.Having the optimized reaction condition in hand, the generality of Pd-catalyzed [3 + 2] cycloaddition of 2-vinylcyclopropane-1,1-dicarbonitrile 1a was scrutinized by using a series of cyclic 1-azadienes 2b–2n ( EntryR1 in 1R2, R3 in 23Yield (%)drb1CN (1a)H, Ph (2a)3aa966 : 12CN (1a)H, 2-FC6H4 (2b)3ab984 : 13CN (1a)H, 3-FC6H4 (2c)3ac954 : 14CN (1a)H, 4-FC6H4 (2d)3ad982 : 15CN (1a)H, 4-ClC6H4 (2e)3ae813 : 16CN (1a)H, 4-BrC6H4 (2f)3af803 : 17CN (1a)H, 2-MeC6H4 (2g)3ag992 : 18CN (1a)H, 3-MeC6H4 (2h)3ah953 : 19CN (1a)H, 4-MeC6H4 (2i)3ai953 : 110CN (1a)H, 2-OMeC6H4 (2j)3aj992 : 111CN (1a)H, 3-OMeC6H4 (2k)3ak943 : 112CN (1a)H, 4-OMeC6H4 (2l)3al992 : 113CN (1a)H, 3,4-OMe2C6H3 (2m)3am802 : 114CN (1a)6-Me, Ph (2n)3an782 : 115cCO2Me (1b)H, Ph (2a)3ba862 : 1Open in a separate windowaUnless otherwise stated, all reactions were carried out with 1 (0.18 mmol), 2 (0.15 mmol) and catalyst in CH2Cl2 (3 mL) at room temperature for 30 minutes.bDetermined by 1H NMR analysis.cAfter 24 h, the starting material was completely consumed (monitored by TLC).To further demonstrate the reaction to be a practical tool for the synthesis of polysubstituted cyclopentane derivatives, the reaction was carried out on the gram scale. We were satisfied to found that when decreasing the loading of palladium/ligand to 0.5%/1.0%, the reaction still worked very efficiently and completed in one hour to provide the product 3aa in 92% yield with 7 : 1 dr (Scheme 2).Open in a separate windowScheme 2Reaction on the gram scale.  相似文献   

7.
Metal free C-3 chalcogenation (sulfenylation and selenylation) of 4H-pyrido[1,2-a]pyrimidin-4-ones     
Prasanjit Ghosh  Gautam Chhetri  Sajal Das 《RSC advances》2021,11(17):10258
An expeditious metal free C-3 chalcogenation of 4H-pyrido[1,2-a]pyrimidin-4-one has been devised to synthesize diversely orchestrated 3-ArS/ArSe derivatives in high yields (up to 95%). This operationally simple reaction proceeds under mild reaction conditions, can be executed in gram scale, and also highlights broad functional group tolerance. Preliminary experimental investigation suggests a radical mechanistic pathway for these transformations.

We have discovered an unprecedented metal-free route for the direct C-3 chalcogenation of various 4H-pyrido[1,2-a]pyrimidin-4-ones under mild conditions in good to excellent yields.

Organosulfur species and derivatives thereof, have garnered a prominent position in contemporary organic synthesis.1 They also have widespread applications in pharmaceuticals, bioactive compounds and polymer materials.2 In addition, carbon-selenium-carbon skeletons are high value core structures for their extensive use as therapeutically active agents, such as antioxidant, antihypertensive, antimicrobial, antibacterial, antiviral, anticancer agents etc.3 Moreover, organoselenium species have been identified as non-toxic compounds.4 Some of the biologically active C–S/C–Se linkage containing scaffolds is outlined in Fig. 1. Consequently, numerous efforts have been devoted to develop facile and reliable methods for the installation of a sulfenyl/selenyl group into organic frameworks.5 Apart from the cross-coupling approach, an alternative protocol for the transition-metal-catalyzed C–S bond formation via C–H bond functionalization has been established, which is known as a sulfenylation reaction.6 In this reaction, aryl sulfonyl hydrazides,7 arylsulfonyl chlorides,8 sulfinic acids,9 and sodium sulfinates10, thiols11 are mostly used as the sulfenylating agents. Although, these methods are advantageous, certain limitations comprising the use of metal catalyst and toxic reagents still persist. Complete removal of trace amounts of transition-metal residues from the anticipated bioactive products is quite challenging task, and this contamination of a transition metal also inhibits the sustainable development.12 Despite these accomplishments, development of an efficient and practical method for transition-metal-free C–S/Se bond forming reactions using thiols/diselenide as the sulfenylation/selenation reagent is an attractive and synthetically desirable.Open in a separate windowFig. 1Representative examples of some biologically active 4H-pyrido[1,2-a]pyrimidin-4-one and diarylsulfide/diselenide scaffold.On the other hand, N-fused bicyclic heterocycles13 has received enormous interest from synthetic chemists as well as medicinal researchers due to their profound impact in agrochemicals, pharmaceuticals and material sciences.14 In this family, 4H-pyrido[1,2-a]pyrimidin-4-one (Fig. 1) exhibits versatile biological activities,15 such as CXCR3 antagonism,16 HLE inhibition,17 MexAB-OprM specific efflux pump inhibition,18 potent 5-HT6 antagonists,19 and acetylcholinesterase inhibition.20 Meanwhile, Pd catalyzed direct arylation and alkenylation of 4H-pyrido[1,2-a]pyrimidin-4-one through C–H bond functionalization has already been reported in the literature.21 Rather, only a single report for the insertion of –SAr group in 4H-pyrido[1,2-a]pyrimidin-4-one molecule using sulfonyl hydrazides as thiol surrogates is documented by Wang et al.22 Nevertheless, this protocol is effective at elevated temperature. Based on our research interests on the structural diversification of heterocyclic scaffolds, we recently reported different methodologies for the metal free direct C–H bond functionalization.23 Herein, we envisaged to disclose a straightforward and efficient protocol of sulfenylation/selenylation for 4H-pyrido[1,2-a]pyrimidin-4-one in the presence of iodine under mild conditions. Pleasingly, several thiols/organodiselenides are smoothly coupled with 4H-pyrido[1,2-a]pyrimidin-4-one and furnished the desired anticipated products in good to excellent yields (Scheme 1).Open in a separate windowScheme 1Previous approaches and the present route of C–H bond functionalization of 4H-pyrido[1,2-a]pyrimidin-4-ones.We commenced our studies with the optimization of the sulfenylation reaction where 2-phenyl-4H-pyrido[1,2-a]pyrimidin-4-one (1a) and thiophenol were used as a model coupling partner ( EntryReagent (equiv.)Oxidant (equiv.)Temperature (°C)Solvent (ml)Time (h)Yieldb (%)1NaI (3)TBHP (3)100DMSO24NR2NaI (3)TBHP (3)100CH3CN24NR3TBAI (2)K2S2O8 (2)70H2O24504TBAI (1)K2S2O8 (2)70CH3CN12675I2 (1)K2S2O8 (2)70CH3CN12916KI (1)K2S2O8 (2)70CH3CN12NR7NaI (1)K2S2O8 (2)70CH3CN12NR8NH4I (1)K2S2O8 (2)70CH3CN12609NIS (1)K2S2O8 (2)70CH3CN126310I2 (1)TBHP (2)70CH3CN128511I2 (1)DTBP (2)70CH3CN12NR12I2 (1)TBPB (2)70CH3CN12NR13I2 (1)H2O2 (2)70CH3CN127114I2 (1)K2S2O8 (2)50CH3CN125315I2 (1)K2S2O8 (2)30CH3CN12NR16I2 (1)—70CH3CN12NR 17 I 2 (50 mol%) K 2 S 2 O 8 (2) 70 CH 3 CN 12 92 18I2 (50 mol%)K2S2O8 (2)70Toluene123519I2 (50 mol%)K2S2O8 (2)70Dioxane124120I2 (50 mol%)K2S2O8 (2)70DCE126621I2 (50 mol%)K2S2O8 (2)70EtOH128122I2 (50 mol%)K2S2O8 (2)70DMF12NROpen in a separate windowaReaction condition: 2-phenyl substituted 4H-pyrido[1,2-a]pyrimidin-4-ones (0.125 mmol, 1 equiv.), benzene thiol (0.1875 mmol, 1.5 equiv.), inducer (equiv./mol%), solvent (2 ml), oxidant (3 equiv.).bIsolated yields based on the reactants 1a, the reaction was run for 12–24 h.Having assimilated the robust reaction conditions for the C–S coupling of 4H-pyrido[1,2-a]pyrimidin-4-one, we sought to explore the scope and general applicability of this protocol (Scheme 1; entries 2b–2h). Noticeably, a crucial effect in the product yield was surveyed with the substituents present at the benzene thiol. 4-Methoxybenzene thiol furnished much lower yield of the corresponding coupled product (2b), compared to electron-withdrawing group, probably due to the generation of a more stable dimer [disulphide]. However, the yield of the anticipated product [2b] could further be enhanced upon/on using stoichiometric amount of catalyst (I2). To our delight, maximum productivity of the product was obtained in the case of F-substituted benzenethiol compare to the other halogens. Notably, ortho bromo-substituted benzene thiol delivered in a higher yield of the corresponding product (2h) than the corresponding chloro derivative (2c). 2,5-Dimethylbenzene thiol was endured under the current reaction conditions to provide 88% yield of 2e. Importantly, the bulkier naphthalene thiol also effectively participated in this transformation to give 82% yield of the C–S coupled product (2g). For adorning the synthetic potentiality further, we investigated the reactivity of various thiols with diverse 4H-pyrido[1,2-a]pyrimidin-4-one. Employment of both electron-neutral (–Me) and electron-deficient (–Cl) functional group substituted parent scaffold provided synthetically useful yields of the desired sulfenylated products with a wide spectrum of benzene thiols (entries 2i–2p). In this context, a suitable choice of benzene thiols is also important, since electronic bias plays a pivotal role in this transformation (entries 2i–2p). Comparatively, a higher yield of the desired ArS derivatives was always obtained in the presence of an electron-withdrawing group (–F, –Cl) at the para position of benzene thiol (entries 2k, 2m, 2n and 2p). Exposure of 2-alkyl substituted 4H-pyrido[1,2-a]pyrimidin-4-one with thiophenol and 4-chlorothiophenol was also fruitful to give intended products in acceptable yields (entries 2q and 2r). Unfortunately, benzyl thiol, 1-pentane thiol and heterocyclic congener of thiol (2-mercapto benzimidazole) did not respond under the optimal reaction conditions (entries 2s, 2t and 2u). Especially, upscale synthesis of 2a was also achieved, illuminating potential capabilities to assemble specialized 3-ArS substituted 4H-pyrido-[1,2-a]pyrimidin-4-ones. It was remarkable that PhSSPh was also amenable instead of PhSH with this catalytic system.Scope of different substituted 4H-pyrido[1,2-a]pyrimidin-4-ones and thiol derivatives for I2 mediated sulfenylationa
Open in a separate windowaReaction condition: substituted 4H-pyrido[1,2-a]pyrimidin-4-ones (0.125 mmol, 1 equiv.), thiol (0.1875 mmol, 1.5 equiv.), I2 (50 mol%), MeCN (2 ml), K2S2O8 (2 equiv.).bIsolated yields based on the reactants 1, the reaction was run for 12 h.cYield at 1 g scale.dPhSSPh was used instead of PhSH.e1 equiv. of I2 was used.With the successful establishment of a straightforward and practical protocol for the C–S coupling reaction, we next investigated the broadness of the selenylation reaction between 2-substituted-4H-pyrido[1,2-a]pyrimidin-4-one and organodiselenides for the selective installation of the –SeAr group at the C-3 position of parent precursor ( Open in a separate windowaReaction condition: various 4H-pyrido[1,2-a]pyrimidin-4-ones (0.125 mmol, 1 equiv.), organo diselenides (0.1875 mmol, 1.5 equiv.), I2 (50 mol%), MeCN (2 ml), K2S2O8 (2 equiv.).bIsolated yields based on the reactants 1, the reaction was run for 12 h.c1 equiv. of I2 was used.To comprehend the plausible reaction mechanism, we executed the sulfenylation reaction under inert atmosphere (N2) and isolated 76% yield of the desired product 2a (Scheme 2; eqn (1)). This observation revealed that aerial oxygen was not only the sole oxidant for this transformation. Additionally, the reaction of 2-phenyl-4H-pyrido[1,2-a]pyrimidin-4-one and iodine in presence potassium persulfate did not afford the corresponding iodo derivative (Scheme 2; eqn (2)). The result unambiguously confirmed that iodinated derivative of 4H-pyrido[1,2-a]pyrimidin-4-one was not involved in the catalytic cycle. Furthermore, the presence of stoichiometric amount of radical scavengers (TEMPO, BHT and 1,1-diphenyl ethylene) inhibited the reactivity, refuting the involvement of non-radical pathway in the reaction mechanism (Scheme 2; eqn (3) and (4)). In addition, we have trapped the in situ generated radical intermediate (PhSe˙) and isolated the compound 4 in reasonable yield (Scheme 2; eqn (4)).Open in a separate windowScheme 2Mechanistic studies.On the basis of these findings and previous literature reports,24 a plausible mechanistic pathway is elaborated in Scheme 3. Presumably, this sulfenylation/selenylation strategy involve an initial generation of the thiyl radical or selenyl radical species A (˙SY/˙SeY, Y = R) in presence of persulfate (S2O82−) or sulfate radical anion (SO4˙). Subsequently, the reactive sulphur/selenyl radical intermediates A coupled with 2-phenyl-4H-pyrido[1,2-a]pyrimidin-4-one substrate leading to a formation of next intermediate 1ab. Then, it underwent further oxidation by sulfate radical anion via a SET mechanism to generate a cationic intermediate 1ac which could be stabilized by resonance to 1ac′. Lastly, the final coupled product (2a/3a) was formed with the liberation of H2 species.Open in a separate windowScheme 3Plausible mechanism (radical pathway).In summary, we have developed an efficient and straightforward transformative tool for regioselective chalcogenation of 4H-pyrido[1,2-a]pyrimidin-4-one under mild conditions. The protocol tolerated diverse common organic functional groups and resulted in good to excellent yields of the desired sulfenylated/selenylated products. Our methodology is operationally simple, regioselective, scalable and avoid the use any expensive metal catalyst. This present protocol opens a new avenue for the direct and convenient chalcogenation of 4H-pyrido[1,2-a]pyrimidin-4-one. Further, C–H bond functionalization reactions on 4H-pyrido[1,2-a]pyrimidin-4-one are currently underway in our laboratory and these observations will be forthcoming.  相似文献   

8.
Rh(iii)-catalyzed regioselective C–H activation dialkenylation/annulation cascade for rapid access to 6H-isoindolo[2,1-a]indole     
Fangming Zhang  Xin Zhu  Bo Luo  Chengming Wang 《RSC advances》2021,11(41):25194
6H-isoindolo[2,1-a]indoles were accessed via a Rh(iii)-catalyzed N–H free indole directed C–H activation dialkenylation/annulation cascade in moderate to excellent yields. This protocol also features: reaction procedures that are insensitive to air and moisture, excellent regioselectivity and good functional group tolerance.

6H-isoindolo[2,1-a]indoles were accessed via a Rh(iii)-catalyzed N–H free indole directed C–H activation dialkenylation/annulation cascade in moderate to excellent yields.

The indole motif, which widely occurs in many natural products, pharmaceuticals and other functional molecules (Fig. 1), is evidently one of the most important skeletons.1 Over the past few decades, people have developed numerous synthetic routes towards indole.2 Among them, direct modification of indoles by transition-metal catalysis through a C–H activation strategy is quite interesting and is undoubtedly of great significance in consideration of atom economy and step simplicity, thus attracting much attention from both academia and industry.2b,3,4Open in a separate windowFig. 1Compounds containing indole motif.Transition-metal catalyzed oxidative cross-coupling via a C–H activation pathway eliminating the need for preactivated reaction partners, has become one of the most powerful tools for molecule manipulation.2c,5 Since the pioneering work of Murai,6a Fujiwara and Moritani,6b,c this research area has undergone rapid developments. Generally, in order to achieve good reactivity and controlled selectivity, a directing group is usually needed.7 In this regard, various directing groups have been gradually designed and reported, including amides,8a amines,8b alcohols,8c carboxylic acids,8d esters,8e ketones,8f aldehydes,8g triazenes8h and N–H free indoles.9Meanwhile, in contrast to the well-reported C–H arylation of indoles,10 direct selective alkenylation of 2-position of N–H free indole is still limited and challenging due to the electrophilic nature of the C-2 position.11 In 2005, Gaunt et al. reported Pd-catalyzed selective C-2 alkenylation of indoles,11b but the reaction efficiency is low and high catalyst loading is required (Scheme 1a). Another most effective strategy is pre-installing a directing group into the indole structure to control the selectivity (Scheme 1b).12 For example, Miura et al. disclosed Pd-catalyzed C-2 alkenylation reaction using carboxylic acid as a blocking and directing group.12a Carretero and co-workers explored N-pyridylsulfonyl as a directing group to functionalize indole at the C-2 position employing excess of alkenes in the presence of Pd(ii) catalyst.12bOpen in a separate windowScheme 1Transition-metal catalyzed C-2 alkenylation of indole.In the above-mentioned examples, high catalyst loading11b (as much as 20 mol%, Scheme 2a) or pre-installed directing groups were generally required.12 These directing groups, possessing functional groups containing a metal-binding heteroatom, remain part of the products after reaction. Such groups can rarely be conveniently removed under ambient conditions or undergo versatile cyclization reactions,12c-f which have greatly limited the structural diversity of the products and subsequent applications to complex molecule synthesis. Therefore, the need for exploration of traceless or easily removable directing groups that can address these drawbacks remains urgent. Recently, Huang et al. reported a N–H indole directed sequential cascade olefination/cyclization reaction (Scheme 1c).13 However, the substrate scope of this transformation is limited, stepwise operation procedure is required, thus make this process tedious. Herein, we reported a Rh-catalyzed N–H free indole directed dialkenylation followed by an intramolecular cascade cyclization reaction leading to the efficient synthesis of 6H-isoindolo[2,1-a]indole.Open in a separate windowScheme 2Reaction under air.Our initial study was carried out by examining 2-phenyl indole 1a and ethyl acrylate 2a in the presence of [RhCp*Cl2]2 and Cu(OAc)2·H2O in DMF under argon atmosphere ( EntrySolventCatalystAdditiveYield1 o-xylene[Cp*RhCl2]2Na2CO310%2DMF[Cp*RhCl2]2—54%3b t-AmylOH[Cp*RhCl2]2AgSbF6Trace4DMF[Cp*RhCl2]2CsOAc55%5DMF[Cp*RhCl2]2K3PO4·3H2O23%6DMF[Cp*RhCl2]2 t-BuOK—7cDMF[Cp*RhCl2]2CsOAcTrace8dDMF[Cp*RhCl2]2CsOAcTrace9eDMF—CsOAc—10DMF[RuCl2(p-cymene)]2CsOAcTrace11DMFRhCl(PPh3)3CsOAcTrace12DMF[Cp*RhCl2]2Cu(acac)2Trace13fDMF[Cp*RhCl2]2CsOAc86%Open in a separate windowaReaction on a 0.2 mmol scale, using 1a (1.0 equiv.), 2a (1.5 equiv.), additive (2.0 equiv.), Cu(OAc)2·H2O (2.0 equiv.), [TM] (3 mol%), solvent (1.5 mL), under N2, 21 h, isolated yield.bAdditive (0.15 equiv.).cWithout oxidant.d10 mol% [Cu] under O2.eWithout [Rh].fUsing 2a (2.5 equiv.), [Rh] (5 mol%), 41 h.With the optimized conditions in hand, we next tend to investigate the scope of this transformation (14 also smoothly participated in this reaction. Currently, no electron-biased alkenes (for example: 1-octene) failed to afford the desired products for some unknown reasons.Substrates scopea
Open in a separate windowaIsolated yield.bAs a mixture (1 : 1.7).cMono/di = 1 : 2.4.The robustness of this Rh-catalyzed indole derivatization method was further examined under air instead of argon. Similar synthetic efficiency was got, thus further proving the practicality of this transformation (Scheme 2). It is worth dating that this dual C–H activation/annulation process is also insensitive to moisture. Commercially available solvent and reagents were directly used as received and were well-compatible in this reaction without any further purifications, which additionally expands the practical application of this conversion.Next, we try to investigate the regioselectivity of this reaction. As it is showed in Scheme 3, N-(2-phenyl-1H-indol-5-yl)pivalamide 1q can smoothly undergo this conversion and only afforded the desired 6H-isoindolo[2,1-a]indole product 3v and no amide group directed C–H alkenylation product was detected,15 thus indicating the excellent selectivity of the C–H activation/annulation process.Open in a separate windowScheme 3Regioselectivity of different directing groups.In order to shed lights on the mechanism of this transformation, a series of experiments were carried out. First, possible active rhodium catalytic species A was synthesized according to previous reports (Scheme 4a).13 Next, catalytic reaction of rhodium complex A with ethyl acrylate 2a under standard reaction conditions gave rise to desired product 3a in 53% yield (Scheme 4b), which indicate complex A possibly involves in this reaction. Thirdly, only mono- and di- alkenylation products were obtained in the absence of base while the cyclization product 3a produced by further extra addition of CsOAc (Scheme 4b). Moreover, the mono-alkenylation product 3a′′ can also be smoothly converted into the dialkenylation product 3a′ upon further treatments (Scheme 4b).Open in a separate windowScheme 4Mechanism study.Finally, we proposed a mechanism for this transformation based on above experiments and reported literatures.9,13,16,17 First, [{Cp*RhCl2}2] dissociates and generates the active catalyst species [Cp*Rh(OAc)2] in the presence of cesium acetate.9d,16 C–H activation of 2-phenyl indole 1a by Rh(iii) A produces rhodacyclic complex B,9 followed by olefin insertion and reductive elimination to afford 3a′′. One more this process gave rise to 3a′ and generated Rh(i), which can be further oxidized into Rh(iii) by copper acetate and fulfilled the catalytic cycle.9,17 A base promoted intramolecular aza-Michael addition of 3a′ finally produced the cascade cyclization product 3a (Scheme 5)Open in a separate windowScheme 5Proposed mechanism.In summary, we have reported a Rh-catalyzed N–H free indole directed C–H activation dialkenylation and cascade cyclization reaction. This strategy provides a general functionalization route of simple 2-aryl indole leading to the efficient synthesis of 6H-isoindolo[2,1-a]indoles. Excellent regioselectivity was obtained with substrates containing amide directing group. Further exploration of the synthetic utility of this chemistry is currently underway in our laboratory and will be reported in due course.  相似文献   

9.
A mild one-pot synthesis of 2-iminothiazolines from thioureas and 1-bromo-1-nitroalkenes     
Yuan Xu  Xin Ge  Yuhan Zhang  Hongbin Zhang  Xue-Wei Liu 《RSC advances》2021,11(4):2221
A mild method to access functionalized 2-iminothiazolines in a facile and efficient manner has been developed. The reaction started from 1,3-disubstituted thioureas and 1-bromo-1-nitroalkenes in the presence of triethylamine in THF and proceeded smoothly in air to afford 2-iminothiazoline derivatives in moderate to good yields.

A mild method to access functionalized 2-iminothiazolines in a facile and efficient manner has been developed.

Functionalized 2-iminothiazoline has been an important building block in organic chemistry.1 Its derivatives show significant pharmacological activities such as bactericidal and fungicidal activity.2 In addition, they are found in drug applications for treatment of allergies, hypertension, inflammations, bacterial and HIV infections.3 Pifithrin (Pft-α) (Fig. 1), isolated by screening of chemical libraries using 2-iminothiazoline skeleton, is a lead compound for p53 inactivation and has received increasing attention due to its possible applications in therapy of Alzheimer''s disease, Parkinson''s disease, stroke and other pathologies related to various signalling pathways.4 Hence, its utility and applicability are widely recognized in organic and biological areas.Open in a separate windowFig. 1Pharmacologically important molecules consisting of 2-iminothiazoline core structure. (a) p53 inactivator; (b) skin whitening agent; (c) anti-inflammatory agent.The classical synthesis of 2-aminothiazole moiety involves the Hantzsch condensation reaction of thioureas and α-haloketones.5 Birkinshaw et al. reported the synthesis of N-alkylated imino-thiazolines by replacing thioureas with mono-N-substituted thioureas.6 Also, several alternative strategies have been devised, which include synthesis of highly functionalized thiazoles and 2-iminothiazolines by replacing α-haloketone with 2,2-dicyano-3,3-bis(trifluoromethyl)oxirane7 and 2-chlorooxirane,8 treatment of α-bromoketimines with potassium thiocyanate,9 reaction of N-monoalkylated thioureas with 3-bromomethyl-2-cyanocinnamonitrile,10 cycloadditions followed by elimination of 5-imino-1,2,4-thiazolidin-3-ones with enamines and ester enolate,11 ring transformation of 1-arylmethyl-2-(thiocyanomethyl)aziridines in the presence of TiCl4 and acyl chloride,12 reaction of N-propargylaniline with acyl isothiocyanates.13 Less general approaches towards the synthesis of these compounds involve the reaction of ketone either with N-alkyl rhodanamine or bisbenzyl formamidine disulfide14 or the reaction of α-chloroketones with thiosemicarbazide in an acidic medium,15 condensation of α-haloketones with N-benzoyl-N′-arylthioureas or N,N′-disubstituted thioureas.16Although some of the methods used for preparing 2-iminothiazolines are convenient and effective, most procedures reported in literatures require arduous preparation of precursor substrates or harsh reaction conditions. Till now, only a few procedures on the one-pot synthesis of 2-iminothiazoline from N,N′-dialkylthiourea and in situ generated α-bromoketones have been reported.17 Herein, we reported a novel and efficient methodology for the synthesis of 2-imino-5-nitrothiazolines using 1,3-diarylthioureas and 1-bromo-1-nitroalkenes as starting materials.The β-bromo-β-nitrostyrenes, with β-disubstituted styrene structure, showed versatile reactivity as a trifunctional synthon. Previous literatures showed their activity as Michael acceptors and [3 + 2] and [4 + 2] cycloaddition partners.18 In our preliminary experiments, the reaction of 1,3-diphenylthiourea and β-bromo-β-nitrostyrene 1a was studied, while the latter one could easily be prepared according to the reported procedure.18e We found that when these two reactants were treated with base such as K2CO3 in THF at room temperature under atmospheric air, a red crystalline product was obtained ( EntryaBaseSolventTemperature (°C)Time (h)Yield (%)1K2CO3THFrt24622bK2CO3THF7010603Et3NTHFrt24724bEt3NTHF7010655DBUTHFrt24636bDBUTHF7010587KHCO3THFrt24558bKHCO3THF7010609DIPEATHFrt246810NoneTHFrt242911cEt3NTHFrt246412dEt3NTHFrt245813eEt3NTHFrt241714Et3NCH2Cl2rt244515Et3NToluenert244216bEt3NToluene110540Open in a separate windowaReactions were performed with β-bromo-β-nitrostyrene 1a (0.10 mmol) and 1,3-diphenylthiourea 2a (0.11 mmol) with base (0.02 mmol) in the indicated solvent (2.0 mL) under atmospheric air.bβ-Bromo-β-nitrostyrene was completely consumed.cReaction was carried out with 0.04 mmol of base.dReaction was carried out with 0.01 mmol of base.eReaction was carried out with 0.1 mmol of base.This structure was later confirmed by single crystal X-ray analysis as shown in Fig. 2. When the reaction temperature was increased to 70 °C, the reaction was completed in shorter time. However, the yield was lower due to an increase of side products (Open in a separate windowFig. 2X-ray crystallography of compound 3a.Several other bases were tested such as K2CO3, Et3N, DBU, KHCO3, DIPEA, and Et3N was found to give the best results ( Open in a separate windowaReactions were performed with 1-bromo-1-nitroalkenes 1a–1n (0.10 mmol) and 1,3-diarylthioureas 2 (0.11 mmol) with Et3N (0.02 mmol) in THF (2.0 mL) at room temperature in the air for 24 hours.A plausible mechanism for this reaction has been proposed as shown in Scheme 1. Initially, a typical Michael addition happens with the β-bromo-β-nitrostyrene 1a,19 which is initiated by the attack of lone pair of nitrogen atom in 1,3-diphenylthiourea 2a, affording intermediate II. The successive tautomerism of thiourea structure and nucleophilic substitution in intermediate II generates five-membered ring intermediate III. Deprotonation of intermediate III by preceding bromide ion affords intermediate IV which subsequently undergoes aromatization with aid of atmospheric oxygen,20 yielding final product 3a.Open in a separate windowScheme 1Plausible reaction pathway.  相似文献   

10.
Enantioselective Nozaki–Hiyama–Kishi allylation-lactonization for the syntheses of 3-substituted phthalides     
Sharad V. Kumbhar  Chinpiao Chen 《RSC advances》2018,8(72):41355
We report the synthesis of bipyridyl ligands and these ligands were applied in the chromium-catalyzed enantioselective Nozaki–Hiyama–Kishi allylation of substituted (2-ethoxycarbonyl)benzaldehydes and subsequently lactonized to synthesize phthalides with an optimal enantioselectivity of 99%. This approach was applied to synthesize (S)-cytosporone E at a 71% yield in three steps.

A chromium-catalyzed enantioselective Nozaki–Hiyama–Kishi allylation of substituted (2-ethoxycarbonyl)benzaldehydes and subsequent lactonization to synthesize phthalides with an optimal enantioselectivity of 99%.

The addition of organochromium reagents to carbonyl compounds was pioneered by Nozaki et al.1 and is a versatile reaction forming carbon–carbon bonds for use with various building blocks to synthesize complex natural products and pharmaceuticals.2 Organochromium reagents are formed in situ through the oxidative addition of Cr(ii) species to allyl, propargyl halides, and triflates (aryl, vinyl halides, and triflates requiring catalytic amounts of Ni(ii) salt)3a,b and can be added to aldehydes and ketones to produce the corresponding alcohols.3 The major disadvantage of the original NHK allylation method is the excessive amount of Cr(ii) salts necessary for the formation of the organochromium reagent.1,4 Fürstner and Shi5 reported NHK allylation that requires only catalytic amounts of active Cr(ii) species; Mn is used as the reducing agent, and trimethylchlorosilane prolongs the catalytic cycle. These improvements have driven chemists to improve the enantioselectivity of this reaction. Several chiral ligands have been used as enantioselective catalysts for the allylation of aldehydes.6 A chiral bipyridyldiol was used as the efficient enantioselective catalyst in an asymmetric NHK reaction to provide the corresponding homoallylic alcohol and simultaneously form lactone in high enantiomeric excess (90% ee).6n The successful application of these ligand systems motivated us to investigate the NHK allylation of (2-ethoxycarbonyl)benzaldehydes and simultaneous lactonization to produce phthalides. This approach was then employed to enantioselectively synthesize (S)-cytosporone E.Pineno-bipyridyl alcohols are readily obtainable from (+)-α-pinene in four steps (Scheme 1); these steps include the stereoselective alkylation of chiral bipyridine with ketones to obtain desired products 4–7.6o In a previous work, we used benzaldehyde, allyl chloride, and ligand 6 under Fürstner''s reaction conditions to obtain the corresponding product with 70% yield and 75% ee.6n In the same study, a bipyridyldiol was applied under Fürstner''s reaction conditions to prepare phthalide 8 with 92% yield and 90% ee. To improve the enantioselectivity of the NHK reaction using ligand 6 under Fürstner''s reaction conditions, we obtained the optimal reaction conditions using benzaldehyde, allyl bromide, CrCl3 (10 mol%), ligand 6 (30 mol%), and Et3N (60 mol%) to provide a product with 83% yield and 90% ee.6oOpen in a separate windowScheme 1Ligand synthesis.6oInitial evaluation of non-symmetrical bipyridine ligands (4–7) in the Cr(ii)-catalyzed addition of allyl bromide to (2-ethoxycarbonyl)-3,5-dimethoxybenzaldehyde revealed that ligand 6 has excellent reactivity and enantioselectivity () Reactions performed at 0 °C in THF afforded better results than all of screened conditions ( EntryLigandSolventTemp [°C]Yielda [%]eeb [%]14THFr.t.868325THFr.t.898936THFr.t.899347THFr.t.859054THF0888765THF0898976THF0869787THF0879296CH2Cl2r.t.5886106DMFr.t.3684116Toluener.t.NR—126Etherr.t.NR—136CH2Cl206889146DMF01582156Toluene0NR—166Ether0NR—Open in a separate windowaIsolated yield.bThe ee% were determined on a Chiracel OJ HPLC column.Ligands 4 and 5, which had less steric hindrance, showed slightly lower enantioselectivity than ligands 6 and 7. The reaction conducted at room temperature resulted in a product with less enantioselectivity than the reaction at 0 °C. ).Catalytic enantioselective synthesis of phthalides
EntryProductYielda [%]eeb [%]
1 88797
2 98594
3 108999
4 118696
5 129097
6 137696
7 147096
8 159098
9 167997
10 177798
Open in a separate windowaIsolated yields.bee% were determined on a Chiracel OJ HPLC column after recrystallization.This highly enantioselective catalyst was explored to determine its promise as an enantioselective catalyst of NHK allylation–lactonization. A natural product, (R)-cytosporone E, was synthesized through asymmetric allylation–lactonization. Ohzeki and Mori synthesized (S)-cytosporone E with 11.5% yield (nine steps) and 45% ee using AD-mix-β® as an asymmetric dihydroxylation reagent.7 The enantiomers of a tris(tert-butyldimethylsilyl)-protected derivative was separated by preparative HPLC on Chiralcel® OD to obtain pure (S)- cytosporone E (98.4% ee). Wyatt et al. synthesized racemic cytosporone E at 41.4% yield (seven steps).8 Therefore, asymmetric NHK allylation–lactonization was used in this study as an efficient allylation reaction using ligand 6 as the enantioselective catalyst to provide the corresponding homoallylic alcohol and simultaneously form lactone 10 in a high yield (89%) and enantiomeric excess (98.8%). Lactone 10 was converted into compound 18 through treatment with 9-BBN, Cs2CO3(aq), bromobutane, Pd(OAc)2, and PCy3 (Scheme 2). Finally, the methoxyl groups of 18 were cleaved using BBr3 to obtain (S)-cytosporone E (19) at 71% yield (three steps).Open in a separate windowScheme 2Synthesis of (S)-cytosporone E.In summary, a chiral Cr(ii) complex prepared using bipyridine alcohol and CrCl3 efficiently catalyzed the enantioselective NHK allylation–lactonization of substituted (2-ethoxycarbonyl)-benzaldehyde to produce phthalides with excellent enantioselectivity (up to 99% ee). This approach was used to synthesize (S)-cytosporone E and obtained excellent yield and enantioselectivity.  相似文献   

11.
Dual C–H activation: Rh(iii)-catalyzed cascade π-extended annulation of 2-arylindole with benzoquinone     
Qijing Zhang  Qianrong Li  Chengming Wang 《RSC advances》2021,11(21):13030
A rhodium-catalyzed, N–H free indole directed cyclization reaction of benzoquinone via a dual C–H activation strategy is disclosed. This protocol has a good functional group tolerance and affords useful indole-fused heterocylces. Besides, it is insensitive to moisture, commercially available solvent can be directly used and work quite well for this transformation.

A Rh-catalyzed cascade annulation of N–H free 2-arylindole with benzoquinone via dual C–H activation strategy was reported.

Quinones are widely distributed in nature, and commonly occur in bacteria, flowering plants and arthropods (Fig. 1). They have a wide range of applications, including diverse important pharmacological properties, involvement in redox reactions and development for advanced electrochemical energy storage.1 Among varied reported quinones, benzoquinone (BQ) is the simplest and most important one. It has been well reported that BQ has a significant and unique role in oxidative palladium(ii)-catalyzed coupling reactions.2 The chemistry of benzoquinone has been extensively explored in detail, including nucleophilic addition and cycloaddition reactions, photochemistry and oxidative coupling.1b,c,2 Although great achievements have been obtained, only a few examples are disclosed about BQ as a reactant applying to transition-metal catalyzed C–H functionalization.1e Among the examples reported, cyclization or BQ direct functionalization products were mainly afforded (Scheme 1a).Open in a separate windowFig. 1Selected examples of bioactive molecules containing the benzoquinone moiety.Open in a separate windowScheme 1Transition-metal catalyzed C–H functionalization of BQ.Transition-metal catalyzed C–H functionalization has undergone great progresses in the past two decades.3 In order to get a better reactivity and controlled selectivity, a directing group is usually needed for this process. Therefore, various directing groups have been developed.4 However, many of them (e.g. various nitrogen-containing heterocycles) remained parts of products after reaction, therefore increasing the procedures and difficulty for structure further modification and manipulation.5 As a result, it is highly demanded to explore traceless or easily removable directing groups.6 In this context, N–H free indole moiety has gradually emerged as a versatile functionalizable directing group in transition-metal catalyzed cyclization reaction.7On the other hand, although the above-mentioned great breakthrough obtained in C–H functionalization, there are few examples reported for dual C–H activation reactions.8 During our research program exploring transition-metal catalysis and heterocyclic synthesis,9 we intended to prepare the indole-containing heterocycles based on the consideration of their potential biological activity. Herein, we report a rhodium-catalyzed N–H free indole directed annulation reaction with BQ through dual C–H activation strategy (Scheme 1b).Our initial study was carried out by examining 2-phenyl indole 1a and benzoquinone 2a in the presence of [{Cp*RhCl2}2] and Cu(OAc)2·H2O in commercial available N,N-dimethylformamide under argon atmosphere. To our delight, the desired 9H-dibenzo[a,c]carbazol-3-ol product 3a was isolated in 55% yield ( EntrySolventCatalystAdditiveYield1DMF[Cp*RhCl2]2Cu(OAc)2·H2O55%2DMF[Cp*RhCl2]2—<5%3DMF—Cu(OAc)2·H2O—4 t-Amyl-OH[Cp*RhCl2]2Cu(OAc)2·H2O—5DMAc[Cp*RhCl2]2Cu(OAc)2·H2O<5%6DMSO[Cp*RhCl2]2Cu(OAc)2·H2OTrace7DMF[RuCl2(p-cymene)]2Cu(OAc)2·H2O—8DMFPd(OAc)2Cu(OAc)2·H2O—9DMFRhCl(PPh3)3Cu(OAc)2·H2O<5%10DMF[Cp*RhCl2]2AgOAc—11DMF[Cp*RhCl2]2Ag2O—12DMF[Cp*RhCl2]2Cu(acac)2Trace 13 bb , cc DMF/DCE[Cp*RhCl2]2Cu(OAc)2·H2O 84% Open in a separate windowaReaction on a 0.2 mmol scale, using 1a (1.0 equiv.), 2a (1.0 equiv.), additive (2.0 equiv.), CsOAc (2.0 equiv.), [TM] (5 mol%), solvent (1.0 mL), under N2, isolated yield.b1a (1.5 equiv.), solvent (0.3 M).cNaOAc was used instead of CsOAc.With the optimized conditions in hand, we next tend to examine the substrates scope of this reaction. Various 2-aryl indoles with electron-rich substituted groups were tested and worked well for this reaction (10 Finally, an interesting S, N-fused heterocycle 3m was obtained when 2-thienyl indole was employed. Other derivatives of benzoquinone such as 1,4-naphthaquinone or methyl-p-benzoquinone currently failed to produce the related cyclization products with proper yields.Substrates scopea
Open in a separate windowaCondition A: 2-aryl indole (1.5 equiv.), BQ (1.0 equiv.), [Rh] (5 mol%), Cu(OAc)2·H2O (2.1 equiv.), NaOAc (2.0 equiv.), DMF/DCE(1.5 mL, 2 : 1), 100 °C.bCondiiton B: 2-aryl indole (1.0 equiv.), BQ (2.0 equiv.), [Rh] (5 mol%), Cu(OAc)2·H2O (2.1 equiv.), NaOAc (2.0 equiv.), DMF/DCE(1.5 mL, 2 : 1), 60 °C.In addition, this method allows quick access to a number of functional heterocycles (Scheme 2).7g,11 For example, the hydroxyl group can be easily removed to afford 9H-dibenzo[a,c]carbazole 4a which can be further converted into organic electroluminescent element 5avia reported methods.11Open in a separate windowScheme 2Diversity of the product.Finally, we proposed a mechanism for this transformation (Scheme 3) based on reported literatures.7,9ac,12 First, [{Cp*RhCl2}2] dissociates and delivers the active catalyst monomer [Cp*Rh(OAc)2] with the assistance of copper acetate and sodium acetate.9ac C–H activation of 2-phenyl indole by Rh(iii) produces rhodacyclic intermediate A,7 followed by insertion of benzoquinone affording intermediate B, which can be transformed into Cvia two folds protonation and fulfills the catalytic cycle. The final product 3a can be easily accessed via intramolecular condensation of C.7gOpen in a separate windowScheme 3Proposed mechanism.In conclusion, we have developed a Rh(iii)-catalyzed traceless directed dual C–H activation of 2-aryl indole and annulation with benzoquinone affording indole-fused heterocycles. The protocol is applicable to a wide range of indole derivatives, affording related products in middle to good yields. Further exploration of the synthetic utilities of this chemistry and detailed mechanistic study are currently in progress in our lab and will be reported in due course.  相似文献   

12.
Sodium borohydride-nickel chloride hexahydrate in EtOH/PEG-400 as an efficient and recyclable catalytic system for the reduction of alkenes     
Kaoxue Li  Chuanchao Liu  Kang Wang  Yang Ren  Fahui Li 《RSC advances》2018,8(14):7761
An efficient, safe and one-pot convenient catalytic system has been developed for the reduction of alkenes using NaBH4–NiCl2·6H2O in EtOH/PEG-400 under mild conditions. In this catalytic system, a variety of alkenes (including trisubstituted alkene α-pinene) were well reduced and the Ni catalyst could be recycled.

An efficient, safe and one-pot convenient catalytic system has been developed for the reduction of alkenes using NaBH4–NiCl2·6H2O in EtOH/PEG-400 under mild conditions.

The reduction of alkenes is an important transformation in organic synthesis, widely used particularly in petrochemical, pharmaceutical, and fine chemical processes. Traditionally, direct hydrogenation,1–4 catalytic hydrogen transfer 5–7and hydride reduction methods8–10 have been employed for the reduction of alkenes. Among the reported methods, the utility of sodium borohydride for the reduction of simple alkenes first described by Brown in 1962 is well known.11 In recent years, several related catalytic systems on modification of Brown''s approach have been developed. These catalytic systems include NaBH4/NiCl2·6H2O/moist alumina in hexane,12 InCl3–NaBH4 reagent system,13 NaBH4/RuCl3 under aqueous conditions,14 NaBH4/CH3COOH in the presence of Pd/C15 and NaBH4-RANEY® nickel system in water.16 Nevertheless, most of these systems require costly transition metal catalyst, long reaction time and a large excess of NaBH4. In addition to this, little has been done to recycle the catalyst for the reduction of alkenes using NaBH4. Thus, a very simple, efficient and recyclable system for the reduction of alkenes by NaBH4 would be highly desirable.In the reduction of alkenes by NaBH4, the in situ generated metal nanoparticles (NPs) from the combination of appropriate metal salts and NaBH4 catalyze the reduction of alkene.15 In general, metal NPs tend to agglomerate during the catalytic processes and therefore need to be protected by stabilizers.17 Castro et al. ever noticed the aggregation of NPs after just one time in the reduction of alkenes by NaBH4 without use of a stabilizer.18 Immobilized nanoparticles (NPs) on insoluble solid supports were generally used for this process in the past literature.12,15,16 One significant example was that Takashi Morimoto and coworkers reached 90% yield of ethylbenzene within 3 h using NiCl2·6H2O on moist alumina reducted by NaBH4 in hexane at 30 °C.12 Unfortunately, heterogeneous catalysts of NPS on solid supports are often more inert than corresponding soluble NPs catalysts.19 In view of the above, we wanted to explore the use of soluble NPs generated in situ for this process. PEG-400 is known to be an excellent dispersion agent and stabilizer for soluble metal NPs.20,21 Abdul Rahman Mohamed et al. found that iron metal NPs formed in an ethanol-PEG-400 solution displayed a more uniform distribution.22 In this work, we introduce NaBH4/NiCl2·6H2O in EtOH/PEG-400 to the reduction of alkenes, to the best of our knowledge, the system is novel for the reaction. The novel system was expected to show the following advantages (Scheme 1): (a) NaBH4 not only reduces Ni2+ to soluble Ni (0) NPs in situ, but also serves as hydrogen donation for the reduction of alkenes with ethanol; (b) the in situ generated soluble Ni (0) NPs catalyst stabilized by PEG-400 is stable, efficient and recyclable for the reduction of alkenes.Open in a separate windowScheme 1Reduction of alkenes using NaBH4–NiCl2·6H2O in EtOH/PEG-400 system.We first chose the reduction of styrene as model reaction (12,15,23,24 This is possibly because the hydrogen source for the reduction can be sufficiently derived from the B–H of NaBH4 and the O–H of ethanol in our catalytic system (ideally 1 molar of NaBH4 can reduce 4 molar of alkenes with ethanol), as recently reported by Bai et al. for the semihydrogenation of alkynes with NaBH4 in methanol.25 The activity of the in situ generated Ni (0) NPs catalyst in EtOH/PEG-400 (3/2 ratio) had also been compared with the commercial RANEY® nickel catalyst. The results revealed that the in situ generated Ni (0) NPs exhibited a higher activity for the reduction of styrene (entries 4, 20). Thus, we can conclude that the EtOH/PEG-400 using NiCl2·6H2O–NaBH4 is a very simple and efficient system for the reduction of styrene.Optimization of reaction conditions for the reduction of styrenea
EntrySolvent (v/v)NaBH4 (equiv.)Yieldb (%)
1EtOH0.576
2PEG-4000.541
3EtOH-PEG400(4/1)0.584
4EtOH-PEG400(3/2)0.599
5EtOH-PEG400(2/3)0.592
6EtOH-PEG400(1/4)0.579
7MeOH-PEG400(3/2)0.598
81-Propanol-PEG400(3/2)0.580
91-Butanol-PEG400(3/2)0.560
102-Propanol-PEG400(3/2)0.518
11H2O-PEG400(3/2)0.526
12Ethyl acetate-PEG400(3/2)0.523
13Toluene-PEG400(3/2)0.521
14Cyclohexane-PEG400(3/2)0.526
15HCOOH-PEG400(3/2)0.5<1
16CH3COOH-PEG400(3/2)0.5<1
17EtOH-PEG400(3/2)0.7589
18EtOH-PEG400(3/2)1.079
19EtOH-PEG400(3/2)1.2568
20EtOH-PEG400(3/2)0.554c
Open in a separate windowaReaction conditions: N2 atmosphere, 30 °C, NiCl2·6H2O 0.25 mmol, solvent (5 mL), styrene 5 mmol.bGC yield.cCatalyzed by RANEY® nickel.The in situ generated Ni (0) NPs in EtOH/PEG-400 were characterized by UV-vis, XPS after model reaction. Fig. 1a showed the UV-vis spectrum of nickel chloride hexahydrate in EtOH/PEG-400 before and after reaction. Apparently, a broad band at 250–270 nm appeared after reaction, which indicated the formation of Ni (0) NPs.26 XPS spectra (Fig. 1b) showed that Ni 2p3/2 peak at approximately 852.8 eV and Ni 2p1/2 peak at 870.9 eV, respectively, indicating the generation of Ni (0) NPs.27Open in a separate windowFig. 1(a) UV-vis spectra of the solution of NiCl2·6H2O in EtOH/PEG-400 before and after the reduction of styrene using NaBH4. (b) XPS of in situ generated Ni NPs.The motive to use PEG-400 as a stabilizer in ethanol was the possibility to protect and recycle the Ni (0) NPs catalyst. After model reaction, the catalyst could be separated by simple extracting with n-heptane followed by decantation and reused directly without further purification. The results for the reuse of the Ni catalyst were shown in 28,29 However, prolonging the reaction time from 15 min to 120 min could still keep a high yield of ethylbenzene. Notably, the activities of Ni (0) NPs from second run still higher than the corresponding heterogeneous Ni NPS on moist alumina reported by Takashi Morimoto et al.12Recycle of the catalysta
RunTime/minYieldb (%)
11599
212099
312097
Open in a separate windowaReaction conditions: N2 atmosphere, 30 °C, NiCl2·6H2O 0.25 mmol, VEtOH/VPEG-400 = 3 : 2 (5 mL), styrene 5 mmol, NaBH4 2.5 mmol.bGC yield.To explore the reason why the catalytic activity decreased after first run, the leaching, size and distribution of Ni NPs were characterized after reaction. The leached Ni species were checked by ICP-AES and found to be only 0.3%. HRTEM image showed that Ni (0) NPs were well-dispersed with an average diameter of 3–5 nm (Fig. 2), indicating the good dispersion and stabilization of the Ni (0) NPs in EtOH/PEG-400 system. Therefore, the reason for the decrease of catalyst could hardly be explained by Ni leaching or the aggregate of the Ni (0) NPs.Open in a separate windowFig. 2HRTEM image of in situ generated Ni NPs after model reaction.Meanwhile, control experiments were performed between in situ generated Ni (0) NPs in the second run and preformed Ni (0) NPs catalyst (see the ESI for details)for the reduction of styrene (Fig. 3). Obviously, the activities of the two kinds of Ni (0) NPs catalysts for the model reaction were almost equivalent. This promoted us to infer Ni (0) NPs had been changed from in situ generated to preformed catalyst, which may be the main reason for the reduced activity of the catalyst in subsequent runs during recycling. The details are under investigation.Open in a separate windowFig. 3Control experiments between in situ generated Ni (0) NPs in the second run and preformed Ni (0) NPs catalyst for the model reaction.The scope of this catalytic system was also examined for the reduction of various olefins (ref. 30) and NaBH4/RuCl3 system,14,31 which were inert for the reduction of α-pinene) was also reduced without any difficulty (entry 14).Reduction of alkenes with NaBH4/NiCl2·6H2O in EtOH/PEG-400 systema
EntrySubstrateTime (min)Yieldb (%)
11-Hexene1598
21-Octene1596
31-Decene1595
41-Dodecene1594
5Styrene1599
64-Methylstyrene15100
7Allyl phenyl ether15100
8 trans-Anethole12094
9β-Pinene24096
10Norbornene15100
11Cyclopentene3096
12Cyclohexene12099
131,5-Cyclooctadiene30094(72 : 28)c
14α-Pinene30091
Open in a separate windowaReaction conditions: N2 atmosphere, 30 °C, NiCl2·6H2O 0.25 mmol, VEtOH/VPEG-400 = 3 : 2 (5 mL), alkenes 5 mmol, NaBH4 2.5 mmol.bGC yield.cRatio of cyclooctene/cyclooctane.  相似文献   

13.
Organic dye-catalyzed radical ring expansion reaction     
Masato Deguchi  Akitoshi Fujiya  Eiji Yamaguchi  Norihiro Tada  Bunji Uno  Akichika Itoh 《RSC advances》2018,8(28):15825
Herein, we reported an attractive method for synthesizing medium-sized rings that are catalyzed by erythrosine B under fluorescent light irradiation. This synthetic approach featured mild conditions, a facile procedure, a broad substrate scope, and moderate-to-good yields.

Herein, we reported an attractive method for synthesizing medium-sized rings that are catalyzed by erythrosine B under fluorescent light irradiation.

Medium-sized rings are present in numerous important natural products and pharmaceuticals.1 Therefore, several researchers have strived to develop various methods, including olefin metathesis,2 fragmentation,3 and pericyclic reactions4 (Scheme 1(a)), for synthesizing these materials. Nevertheless, unfavorable transannular interactions and entropic factors typical of rings of this size make this task quite challenging.5 Thus, we believe that new methods need to be developed to solve these problems.Open in a separate windowScheme 1(a) Examples of conventional approaches for the synthesis of medium-sized rings. (b) Beckwith–Dowd ring expansion reaction.More than 30 years ago, the Beckwith–Dowd ring expansion reaction6 was introduced as a novel method to form medium-sized rings (Scheme 1(b)). This approach can be an attractive alternative to conventional strategies for synthesizing the abovementioned structures. Actually, the fact that no new unsaturated C–C bonds are formed as part of this method renders hydrogenation processes unnecessary, thereby minimizing the impact of transformation on the substrate. Certainly, this approach could be applicable to the synthesis of natural products.7 However, the original method requires reagents, e.g., azobisisobutyronitrile (AIBN) or tributyltin hydride (Bu3SnH), that are difficult to handle. Some related methods that employ other reagents, including Sm,8 Zn or In,9 B12–TiO2 hybrid catalyst,10 silane,11 and amines,12 have been reported. Recently, a few synthetic approaches based on photo-induced reactions have been developed.12b,12d,13 A previous study reported that even α-(ω-carboxyalkyl) β-keto esters could be employed as relevant substrates.13In this context, we aimed to establish a more efficient and facile method than conventional ones to prepare a broad range of medium-sized rings via a ring expansion reaction. We continuously investigated various photo-initiated reactions by employing a photosensitizer and a fluorescent lamp as a light source.14 For example, the CDC cross-coupling reaction14a and 1,3-dipolar cycloaddition/aromatization reaction14d under photooxidative reaction conditions have been reported. While investigating these reactions, we found that organic dyes induced electron transfer from amine substrates to molecular oxygen.14a,14d Thus, we envisioned that C–halogen bonds can be cleaved in the presence of a sacrificial amine using photosensitized substrates as a springboard. Herein, we reported a convenient and environmentally friendly method that employs a photo-induced reaction as part of the Beckwith–Dowd ring expansion route to synthesize medium-sized rings.We initiated our study with the optimization of the reaction conditions (15 This result can be explained by two facts: (a) the maximum absorption of EB (∼520 nm)16 coincides with the wavelength of the fluorescent lamp and (b) iPr2NEt is an effective reductive quencher (Eox(iPr2NEt˙+/iPr2NEt) = +0.68 V vs. SCE; e.g., Eox(NEt3˙+/NEt3) = +0.99 V vs. SCE).17a,17b In fact, when we employed other combinations of photocatalysts, amines, and solvents, 2a was obtained in a relatively lower yield. Furthermore, in our experiments, we recovered 1a and/or its undesired hydrogenation product 1a′ from the reaction mixture in substantial amounts (entries 2–9).Optimization of the reaction conditions
EntryPhotocatalystSolventAmine2aa (%)1a′a (%)1aa (%)
1Erythrosine B (EB)DMSO iPr2NEt78(83)510
2Eosin YDMSO iPr2NEt50180
3AQN-2-ClDMSO iPr2NEt121063
4EBDMSOEt3N512622
5EBDMSO1-Methyl imidazole111574
6EBDMSO iPr2NH362140
7EBMeCN iPr2NEt70026
8EBDMF iPr2NEt181518
9EBCHCl3 iPr2NEt07914
Open in a separate windowaYields are determined by 1H NMR spectroscopy using 1,1,2,2-tetrachloroethane as an internal standard. The number in parentheses denotes isolated yield.With the optimized reaction conditions in hand, we next explored the substrate scope of this transformation ( SubstrateProductYieldsa (%) 83% 82% 77% 67% 64% 67% 71%b 71%cOpen in a separate windowaYields shown in the table are all pure, isolated yields. The underlined percentages are the yields when the reactions were performed under air atmosphere.b3 equiv. of LiOH added.c5 mol% of Ag2CO3 added, reaction time is 40 h. Scheme 2 summarizes the results of additional experiments conducted to investigate the substrate scope. We tested a substrate with an iodopropyl side chain, 1i, and found that the addition of a silver salt and extension of the reaction time led to the formation of a tertiary alcohol with a fused ring, 3i, instead of the corresponding cyclononanone (eqn (1)). Additionally, we found that this reaction is applicable to substrates with ketones in open-chain moieties, such as 1j and 1k. Particularly, although 2j was volatile, and we needed to treat it carefully during isolation, these substrates gave the rearranged product in high yield (eqn (2) and (3)).Open in a separate windowScheme 2Applying the reaction conditions on substrates with three-carbon lateral chain and acyclic β-keto esters.Next, we performed some experiments to elucidate the reaction mechanism (Scheme 3). We observed that no ring expansion product 2a was obtained when we omitted EB or iPr2NEt or photo-irradiation from the fluorescent lamps. Additionally, if we added 1 equiv. of TEMPO (2,2,6,6-tetramethylpiperidine 1-oxyl free radical), the desired product was obtained in low yield and the presence of a TEMPO-adduct was detected.18 This result confirms the idea that this reaction proceeds via a radical pathway.Open in a separate windowScheme 3Results of the control experiments.Based on these results, previous reports on the Beckwith–Dowd ring expansion reaction,6 and recent reports of radical reactions set off by C–halogen bond cleavage,17b,19 we inferred a plausible mechanistic pathway for the reaction, which is depicted in Scheme 4. The first step in the C–I bond cleavage reaction of 1a is photo-induced single electron transfer (SET) from the excited state of EB. In the experimental conditions employed, EB is considered to be a monoanionic (EB) or dianionic (EB2−) species because its commercially available disodium salt (EBNa2) is added to the reaction mixture containing DIPEA without acidic treatment. To gain further insight into this SET step, we calculated the Gibbs energies involved in the reaction between 1a and EB in a solution environment (solvent: DMSO; ε = 46.826) simulated by the IEF-PCM model using the M06-2X functional.20 The standard Gibbs energy change (ΔG°) calculated for the cleavage reaction of 1a with EB2− (1a + EB2− → 4a + I + EB˙) and with EB (1a + EB → 4a + I + EB˙) in the absence of light irradiation from the fluorescent lamp is 1.72 and 1.27 eV, respectively, as shown in Scheme 5. This indicates that these reactions are thermodynamically uphill electron-transfer processes. On the other hand, it is well known that the ΔG° values for excited EB2−(EB2−*)/EB˙ and excited EB(EB*/EB˙) redox pairs are given by the amended Rehm–Weller equation with the excited-state energies (E) as ΔG°(EB2−*/EB˙) = ΔG°(EB2−/EB˙) − E and ΔG°(EB*/EB˙) = ΔG°(EB/EB˙) − E, respectively.21,22 The reported excited-state energy of EB is 2.34 eV;16a thus, the ΔG° values for the two mentioned C–I bond cleavage reactions (1a + EB2−* → 4a + I + EB˙ and 1a + EB* → 4a + I + EB˙) involving photo-induced SET are −0.62 and −1.07 eV, respectively, which indicate that the E value is sufficiently high for both bond cleavage reactions to be thermodynamically feasible. These results indicate that the cleavage reaction of 1 is governed by an exergonic C–I bond cleavage mechanism involving SET from EB2−* or EB*. The primary radical 4 thus generated attacks the ketone, and then a rapid cyclopropane ring opening occurs. Meanwhile, the oxidized EB is reduced by the sacrificial amine, and the catalytic cycle of EB is completed. No particular H donor was necessary for this reaction to proceed; therefore, we assumed that an “extra” hydrogen atom of the product is derived from iPr2NEt. It believed that abstraction of a hydrogen atom from the trialkylammounium radical cation could occur,23 which strongly supports the above mechanism.Open in a separate windowScheme 4Plausible reaction mechanism.Open in a separate windowScheme 5Standard Gibbs energies calculated for the species involved in the C–I bond cleavage reaction with EB2− (1) and EB (2) using the M06-2X/PCM method with 6-31G (d) basis sets for hydrogen, carbon, and oxygen and the MIDI! basis set for iodine. Free energy corrections are made at standard conditions of 1 atm and 298.15 K.  相似文献   

14.
Detection of nucleic acids and other low abundance components in native bone and osteosarcoma extracellular matrix by isotope enrichment and DNP-enhanced NMR     
Ieva Goldberga  Rui Li  Wing Ying Chow  David G. Reid  Ulyana Bashtanova  Rakesh Rajan  Anna Puszkarska  Hartmut Oschkinat  Melinda J. Duer 《RSC advances》2019,9(46):26686
  相似文献   

15.
TfOH mediated intermolecular electrocyclization for the synthesis of pyrazolines and its application in alkaloid synthesis     
Sesuraj Babiola Annes  Pothiappan Vairaprakash  Subburethinam Ramesh 《RSC advances》2018,8(53):30071
TfOH mediated easy access to interesting pyrazolines starting from an aldehyde, phenylhydrazine and styrene has been developed. The scope of this synthetic methodology has been explored by synthesizing various 1,3,5-trisubstituted pyrazolines in very good yields with very high regioselectivity. The origin of regioselectivity has been explained by comparing the stability of possible intermediate carbocations. The synthetic utility of a green solvent has been explored by synthesizing some of pyrazolines in a DES medium. The synthetic application of the present methodology is employed in the synthesis of a pyrazoline alkaloid.

A metal free and green synthetic methodology employing aldehydes, phenylhydrazine and styrene mediated by TfOH has been developed to access 1,3,5-trisubstituted pyrazolines. The synthetic application of the methodology is demonstrated in the synthesis of a pyrazoline alkaloid.

Pyrazoles and pyrazolines are known to exhibit interesting biological and photo-physical behaviours. The biological activities of pyrazoles/pyrazolines have been discussed in several reviews.1–4 A few representative examples of biologically and medicinally important pyrazoles and pyrazolines are given in Fig. 1.5–7 In particular, considerable interest has been focused on 1,3,5-trisubstituted pyrazoline derivatives due their potential pharmacological activities including (i) antitubercular activity against the H37Rv strain of Mycobacterium,9 (ii) antiproliferative activity,8,10 (iii) antibacterial activity,11–13 (iv) antiobesity effect in an animal model of the potent cannabinoid CB1 receptor antagonist,14 (v) pre-emergent herbicide activity against various kinds of weeds,15 and (vi) ACE-inhibitory activity with 0.123 mM IC504.16 Pyrazoline derivatives show enhanced biological activity compared with their corresponding pyrazoles.17 In addition, the pyrazoline motif is known to exhibit photo-luminescent behaviour due to intra-molecular charge transfer (ICT) in the excited state and also shows hole transport behaviour.18–24Open in a separate windowFig. 1Representative examples of medicinally important 1,3,5-trisubstituted pyrazolines and pyrazoles.Various synthetic approaches have been developed to access these biologically important 1,3,5-trisubstituted pyrazoline/pyrazole compounds.25–28 The most general synthetic approach proceeding via a reaction of 1,3-dicarbonyl compounds with arylhydrazines results in poor regioselectivity.3,29,30 A synthetic method which employs appropriate chalcones and arylhydrazines has been considered to be the most widely accepted method for accessing pyrazolines, but this method falls behind due to a greater number of synthetic steps involved.31–34 Recently Wang et al. disclosed a methodology proceeding via a three component [3 + 2] cycloaddition using 20 mol% Cu(OTf)2 at elevated temperature.35 The development of synthetic methodology to prepare active pharmaceutical ingredients (APIs) would preferably involve (i) high regioselectivity, (ii) diversity in substrates, (iii) the least number of synthetic steps involved and (iv) attaining target compounds, free of metal traces. Hence, a regioselective tandem one-pot intermolecular electrocyclization reaction under metal free condition to access 1,3,5-trisubstituted pyrazolines would be of great importance.In this regard, we are disclosing a general, metal free and green synthetic methodology to access diverse pyrazoline derivatives with various functionalities including –NO2, –OH and aliphatic groups using arylhydrazines, aldehydes and styrenes. We have obtained the corresponding pyrazoline products in very good yields with enhanced regioselectivity. In addition, we have employed this synthetic methodology in the synthesis of a pyrazoline alkaloid (1,5-diphenyl-3-styryl-2-pyrazoline).In our initial studies to attempt metal free conditions, iodine mediated intermolecular electrocyclization of tolualdehyde 1a, phenylhydrazine 2a and styrene 3 was explored (l-proline, CH(OMe)3, TFA, p-TSA and MeSO3H (Scheme 1).36,37 The use of other solvents such as H2O, DMF, DMSO and ethanol did not result in product formation, as TfOH might be deactivated by these solvents (see ESI, Table S1).Optimization of reaction conditions in the synthesis of pyrazoline 4aa
EntryAdditive (equiv.)SolventTime (h)Yieldb (%)
1I2 (0.2)CH3CN2425
2I2 (1.0)CH3CN2434
3I2 (1.0)Toluene2440
4I2 (1.0)H2O2435
5I2 (1.0)EtOAc24Trace
6PhI(OAc)2 (0.2)CH3CN24ND
7NaI (1.0)CH3CN24ND
8NBS (1.0)CH3CN24ND
9CAN (1.0)CH3CN24ND
10 l-Proline (0.3)CH3CN24ND
11CH(OMe)3 (1.0)CH3CN24ND
12 TfOH (1.0) CH 3 CN 7 82
13TFA (1.0)CH3CN24ND
14p-TSA (1.0)CH3CN2418
15MeSO3H (1.0)CH3CN2413
Open in a separate windowaA solution of tolualdehyde 1a (1.0 mmol) and phenylhydrazine 2a (1.0 mmol) in solvent (1.0 mL) was treated with additive followed by styrene 3 (1.0 mmol) and stirred.bIsolated yield; ND = not detected.Open in a separate windowScheme 1Origin of regioselectivity in the formation of pyrazoline 4a.We have observed very good regioselectivity in the formation of pyrazolines mediated by TfOH. Mechanistically, the carbocation I-1 can interact with either of the alkenyl carbons in styrene resulting in two types of cationic intermediates: (i) the most stable benzylic carbocation I-2 and (ii) the least stable primary carbocation I-3. The stability of the benzylic carbocation I-2 over the primary carbocation I-3 directs the reaction pathway towards the formation of a 5-phenyl substituted product 4a over a 4-phenyl substituted product 5a (Scheme 1).The scope of this synthetic methodology has been studied by varying the substrates. Both aromatic and aliphatic aldehydes reacted well under the reaction conditions and their corresponding pyrazolines were obtained in yields of up to 86% ( EntryR =Ar =ProductYieldb (%)14-Me–C6H4– 1aC6H5– 2a4a8224-MeO–C6H4– 1bC6H5– 2a4b7333-MeO–C6H4– 1cC6H5– 2a4c574C6H5– 1dC6H5– 2a4d6354-Br–C6H4– 1eC6H5– 2a4e6164-Cl–C6H4– 1fC6H5– 2a4f5874-F–C6H4– 1gC6H5– 2a4g3984-NO2–C6H4– 1hC6H5– 2a4h2492-NO2–C6H4– 1iC6H5– 2a4i45102-HO–C6H4– 1jC6H5– 2a4j8211CH3CH2CH2– 1kC6H5– 2a4k8612(CH3)2CH– 1lC6H5– 2a4l58134-Me–C6H4– 1a4-Me–C6H4– 2b4m82Open in a separate windowaA solution of aldehyde 1a–l (1.0 mmol) and arylhydrazine 2a, 2b (1.0 mmol) in CH3CN (1.0 mL) was treated with TfOH followed by styrene 3 (1.0 mmol) and stirred.bIsolated yield.Using this methodology, we have attempted the synthesis of bis-pyrazoline 6 from terephthalaldehyde 1m. The bis-pyrazoline 6 was obtained in very low yield, which is also in accordance with the proposed mechanism depicted in Scheme 1. The formation bis-pyrazoline 6 was expected to proceed via a bis-benzylic cation I-4 (Scheme 2), which is highly destabilized by conjugation between two cationic sites. Thus, formation of the bis-benzylic cation I-4 is less probable and resulted in the observed poor yield of bis-pyrazoline 6 (18%) even after 48 h of reaction time. Bis-pyrazoline 6 was structurally characterized by NMR spectroscopy and high resolution mass spectrometry (HRMS) analysis.Open in a separate windowScheme 2Synthesis of bis-pyrazoline 7 from terephthalaldehyde.We wanted to apply this synthetic methodology in the synthesis of a pyrazoline alkaloid. Many species of the genus Euphorbia are medicinally important and known to treat various ailments including gonorrhea, skin diseases, gastrointestinal disorders, migraines and anaphylaxis.38–41 Alkaloid 7 (1,5-diphenyl-3-styryl-2-pyrazoline) was isolated from aerial parts of Euphorbia guyoniana.42 We have adopted TfOH mediated electrocyclization to access alkaloid 7 starting from cinnamaldehyde 1n, phenylhydrazine 2a and styrene 3. In this reaction, the formation of pyrazole 8via intramolecular electrocyclization is expected to compete with the formation of the desired alkaloid 7via intermolecular electrocyclization (Scheme 3). Interestingly, the reaction with cinnamaldehyde did not undergo intramolecular cyclization and the required alkaloid 7 was obtained in 82% yield.Open in a separate windowScheme 3Application in alkaloid synthesis: synthesis of alkaloid 7.Recently, deep eutectic solvents (DES) have been used as green solvents in various applications.43–45 We have explored the use of a few DES including the eutectic mixture of (i) ChCl : urea, (ii) ChCl : PTSA, (iii) ChCl : TfOH and (iv) ChCl : glycol in our synthetic methodology to access trisubstituted pyrazolines (see ESI, Table S2). The eutectic mixture ChCl : PTSA was found to be effective among the DES screened and corresponding 1,3,5-trisubstituted pyrazolines were obtained in yields of up to 61%, without the use of any additives ( EntryR =Ar =ProductYieldb (%)14-Me–C6H4– 1aC6H5– 2a4a4522-HO–C6H4– 1jC6H5– 2a4j613CH3CH2CH2– 1kC6H5– 2a4k454C6H5–CH Created by potrace 1.16, written by Peter Selinger 2001-2019 CH– 1n4-Me–C6H4– 2bAlkaloid 754Open in a separate windowaA mixture of aldehyde (1.0 mmol) arylhydrazine (1.0 mmol) and styrene (1 mmol) in DES (1.0 mL) stirred at 30 °C for 24 h.bIsolated yield.In conclusion, we have developed a metal free TfOH or DES mediated intermolecular electrocyclization of aldehydes, hydrazine and styrene to generate 1,3,5-trisubstituted pyrazolines. The observed very high regioselectivity of the reaction has been rationalized by the stability of the proposed benzylic cation intermediate. The substrate scope has been studied and various pyrazolines have been obtained in moderate to very good yields. The obtained yields were in accordance with the stability of the proposed intermediate carbocations obtained by the electronic effects of substitutions. Working towards greener synthesis, we have screened various deep eutectic mixtures as media for this transformation. The DES ChCl : PTSA was found to be effective in mediating pyrazoline formation and the corresponding pyrazolines were obtained in moderate yields in the absence of any other catalysts. The application of this synthetic methodology has been demonstrated in the synthesis of a natural product, alkaloid 7 obtained from aerial parts of Euphorbia guyoniana.  相似文献   

16.
I2-mediated Csp2–P bond formation via tandem cyclization of o-alkynylphenyl isothiocyanates with organophosphorus esters     
Yang Liu  Wenjin Wu  Xiaoyan Sang  Yu Xia  Guojian Fang  Wenyan Hao 《RSC advances》2022,12(28):18072
A novel, I2-mediated tandem cyclization of o-alkynylphenyl isothiocyanates with organophosphorus esters has been developed under mild conditions. Different kinds of 4H-benzo[d][1,3]thiazin-2-ylphosphonate could be synthesized in moderate to excellent yields. This method has the advantages of easy access to raw materials, free-metal catalyst, simple operation, high yield and high functional group tolerance.

A highly efficient molecular-iodine-catalyzed cascade cyclization reaction has been developed, creating a series of 4H-benzo[d][1,3]thiazin-2-yl phosphonates in moderate to excellent yields. This approach benefits from metal-free catalysts and available raw materials.

As an important class of organic products, organophosphorus compounds have received considerable attention because they have broad application in the field of materials science,1 medicinal chemistry,2 organic synthesis,3 natural products,4 and ligand chemistry.5 Phosphorus-containing compounds are valuable precursors of many biologically active molecules which can act as antibiotics,6 anti-tumor agents7 and enzyme inhibitors.8 Traditionally, the preparation of organophosphorus compounds relies on a transition-metal-catalyzed cross-coupling of phosphine reagents with electrophilic aryl halides (Ar-X),9 aryl boronic acids (Ar-B),10 aryl diazonium salts (Ar-N),11 and so on.12 Recently, the construction of a Csp2–P bond on heterocycles is another powerful method to synthesize the organophosphorus compounds.13 For instance, the Duan group and the Ackermann group reported a Ag-mediated C–H/P–H functionalization method to construct a Csp2–P bond by using arylphosphine oxides and internal alkynes as the substrates13a,b (Scheme 1a). In 2014, Studer and co-workers reported a pioneering radical cascade reaction for the synthesis of 6-phosphorylated phenanthridines from 2-isocyanobiphenyls and diphenylphosphine oxides (Scheme 1b).13c Before long, Ji and Lu''s group described two similar radical process with excess of PhI(OAc)2 or K2S2O8 as the oxidant (Scheme 1c).13d,e Recently, Liang and co-works developed two cases of cascade functionalization to construct phosphorylated heterocycles via the ionic pathway (Scheme 1d and 1e).13f,g Meanwhile, Li and coworkers reported a Mn(ii)-promoted tandem cyclization reaction of 2-biaryl isothiocyanates with phosphine oxides which went through the same mechanism(Scheme 1f).13h Despite the usefulness of the above methods, common problems, such as complex reaction substrates, relatively high temperature, excess amounts of oxidants, limited their applications. Furthermore, transition metals are required in these reactions, thereby resulting in limitations in reactants. Therefore, the development of a simple and transition-metal-free method for the formation of the Csp2–P bond from easily prepared starting materials is highly desirable.Open in a separate windowScheme 1Synthesis of P-containing heterocycles through Csp2–P bond formation. o-Alkynylphenyl isothiocyanates are easily prepared organic synthons with versatile chemical reactivity,14 and they could be used as electrophiles,15 nucleophiles,16 and radical receptors17 due to the N Created by potrace 1.16, written by Peter Selinger 2001-2019 C Created by potrace 1.16, written by Peter Selinger 2001-2019 S moiety in the structure. Recently, the rapid development of the transition-metal-catalyzed cascade cycloaddition of o-alkynylphenyl isothiocyanates with various nucleophiles provides a new and powerful synthetic strategy to synthesize different heterocycles. Very recently, we have developed a tandem cyclization process for the synthesis of 4H-benzo[d][1,3]thiazin-2-yl)phosphonates by using this strategy.18 As part of our continuing interest in the transformation of o-alkynylphenyl isothiocyanates,19 we describe herein a novel I2-promotedtandem reaction to construct Csp2–P bond from o-alkynylphenyl isothiocyanates and phosphine oxides(Scheme 1g).The starting o-alkynylphenyl isothiocyanates were prepared via the Sonogashira coupling of 2-iodoanilines with terminal alkynes,20 followed by the treatment with thiophosgene according to the literature procedure.21 We commenced our studies with the reaction of o-phenylethynylphenyl isothiocyanate (1a, 0.2 mmol) and diethyl phosphonate (2a, 0.6 mmol) in the presence of I2 (0.5 equiv.) as the catalyst, 8-diazabicyclo[5,4,0]undec-7-ene (DBU, 3.0 equiv.) as the base, in dichloromethane (DCM, 2 mL) at 0 °C for 12 h in air atmosphere. Gratifyingly, the desired product diethyl (Z)-(4-benzylidene-4H-benzo[d][1,3] thiazin-2-yl)phosphonate 3a was obtained in 75% yield (Schemes 1–4)Optimization of the reaction conditionsa
EntryCatalystBaseSolventTemp.Yield (%)b
1I2DBUDCM0 °C75
2KIDBUDCM0 °C64
3NaIDBUDCM0 °C50
4ZnI2DBUDCM0 °C55
5I2DCM0 °CNR
6I2DABCODCM0 °CTrace
7I2KOAcDCM0 °CNR
8I2NaOAcDCM0 °CNR
9I2K2HPO4DCM0 °CNR
10I2Cs2CO3DCM0 °CTrace
11I2NaOHDCM0 °CTrace
12I2DBUDCE0 °C28
13I2DBUCHCl30 °C32
14I2DBUDMF0 °CTrace
15I2DBU1,4-Dioxane0 °C50
16I2DBUMeCN0 °C35
17I2DBUToluene0 °C84
18I2DBUToluene25 °C70
19I2DBUToluene40 °C52
20I2DBUToluene80 °C42
21I2DBUToluene−10 °C66
Open in a separate windowaReaction was performed with 1a (0.2 mmol), 2a (0.6 mmol), catalyst (0.1 mmol), base (0.6 mmol), in solvent (2 mL) for 12 h.bIsolated yield based on o-phenylethynylphenyl isothiocyanate 1a.Open in a separate windowScheme 2The reaction of 2-isothiocyanato-3-(phenylethynyl)pyridine with 2a.Open in a separate windowScheme 3Two control experiments for mechanism.Open in a separate windowScheme 4Proposed mechanism.In order to further demonstrate the substrate scope, different o-alkynyl phenylisothiocyanates were then explored and the results are summarized in 19 no desired products were obtained when the R2 group in the substrate o-alkynylphenyl isothiocyanates 1 was an alkyl group, such as n-butyl, t-butyl, and n-hexyl. Similarly, when the R2 group in the substrate o-alkynylphenyl isothiocyanates 1 was the cyclopropyl group, the desired desired 4H-benzo[d][1,3]thiazin-2-yl phosphonate 3g was obtained in 70% yield. On the other hand, the reactions of o-alkynylphenyl isothiocyanates bearing various substituents such as fluoro, chloro, bromo, trifluoromethyl, methyl and methoxy groups on the aryl rings at the R1 position, regardless of their electronic properties and substitution positions, gave the desired products 3h–3r in moderate to good yields. Particularly, p-Br substituted 1j and 1o appeared excellent reactivity and the corresponding products 3j and 3o were obtained in 92% and 90% yield, respectively. In order to further expand the substrate scope, we moved on to examine the P-reagents under the optimal conditions. The reaction of dimethyl phosphate and diphenylphosphine oxide with 1a under the standard conditions, the corresponding products dimethyl (Z)-(4-benzylidene-4H-benzo[d][1,3]thiazin-2-yl)phosphonate and (Z)-(4-benzylidene-4H-benzo[d][1,3]thiazin-2-yl)diphenyl phosphine oxide (3s and 3t) were obtained in 81% yield and 79% yield, respectively. It is noteworthy that the corresponding target products (3u and 3v) with excellent yield (96% and 99% yield) are obtained when we replace diphenylphosphine oxide with di-p-tolylphosphine oxide and bis(4-methoxyphenyl)phosphine oxide. Similarly, all products were uniformly formed as the Z-isomer, which might be due to a kinetic effect according to Baldwin''s rules and a smaller steric effect compared to the E-isomer.22Tandem cyclization of o-alkynylphenyl isothiocyanates with diphenylphosphinesa,b
Open in a separate windowaReactions were performed with o-alkynylphenyl isothiocyanates 1 (0.2 mmol), phosphite or diphenylphosphines 2 (0.6 mmol), I2 (0.1 mmol), DBU (0.6 mmol), in toluene (2 mL) under 0 °C for 12 h.bIsolated yield based on o-phenylethynylphenyl isothiocyanate 1.Next, we examined the reaction of 2-isothiocyanato-3-(phenylethynyl)pyridine 1x with 2a under the standard conditions (Scheme 2). Not surprisingly, the corresponding product 3x diethyl (Z)-(4-benzylidene-4H-pyrido[2,3-d][1,3]thiazin-2-yl) phosphonate was obtained in 47% yield.Two control experiments were carried out to obtain some mechanism insight into the reaction. Firstly, 3.0 equiv. of 2,2,6,6 tetramethylpiperidine N-oxide (TEMPO) was added in the reaction of o-alkynylphenyl isothiocyanate 1a with diethyl phosphonate 2a, and product 3a could be isolated in 75% yield (Scheme 3, ,1).1). Similarly, the yield of 3a was not influenced when we added 3.0 equiv. of 2,6-di-tert-butyl-4-methylphenol (BHT) in the reaction (Scheme 3, ,2).2). These results probably suggested that the reaction may not follow a radical pathway.Base on the above results and previous reports,23 a possible mechanism was proposed for this reaction (Scheme 4). Firstly, in the presence of DBU as the base, the nucleophilic addition of P-anion to an isothiocyanate moiety in compound 1 would occur to produce the intermediate A. Then, Intermediate A could undergo isomerization to afford intermediate B. Next, molecular iodine serves as a π-acid to react with triple bond, giving iodocyclized intermediates C which followed by an intramolecular nucleophilic addition to give the intermediate D. Finally, intermediate D underwent the protodeiodination to give the target product 3.  相似文献   

17.
Access to 1-amino-3,4-dihydroisoquinolines via palladium-catalyzed C–H bond aminoimidoylation reaction from functionalized isocyanides     
Zhuang Xiong  Panyuan Cai  Yingshuang Mei  Jian Wang 《RSC advances》2019,9(72):42072
Efficient access to 1-amino-3,4-dihydroisoquinolines, through palladium-catalyzed intramolecular C–H bond aminoimidoylation of α-benzyl-α-isocyanoacetates, has been developed. Consecutive isocyanide insertion and C–H bond activation were realized with C–N and C–C bonds formation in one step under redox neutral conditions, employing O-benzoyl hydroxylamines as electrophilic amino sources.

Efficient and practice access to 1-amino-3,4-dihydroisoquinolines, through palladium-catalyzed intramolecular C–H bond aminoimidoylation of α-benzyl-α-isocyanoacetates, has been developed.

As a class of cyclic amidine compounds, 1-amino-3,4-dihydroisoquinolines are essential six-membered heterocycles which can be found in numerous natural products, functional chemicals and pharmaceutical agents (Scheme 1).1 For example, piperazine tethered dihydroisoquinoline and benzocyclic amide I was found to be an efficient cardiovascular agent.1a Phenylalanine-derived and arylamine or cyclic aliphatic-amine-containing 1-amino-3,4-dihydroisoquinolines (II–V) show outstanding BACE-1 inhibition, kinase VEGFR-2 inhibition, phosphodiesterase IV inhibition and hydrogen ion-sodium exchanger NHE-3 inhibition activities, respectively.1be Traditional preparation of such 1-amino-3,4-dihydroisoquinoline scaffolds is mainly based on nucleophilic substitution of readily prepared 3,4-dihydroisoquinolines2 and Bischler-Napieralski reaction starting from urea.3 Most of these methods suffer from multiple-step synthesis, harsh conditions or the use of toxic reagents. Thus, developing an efficient and practical method for the construction of 1-amino-3,4-dihydroisoquinoline derivatives remain an attractive synthetic task.Open in a separate windowScheme 1(a) Representative bioactive molecules containing the 1-amino-3,4-dihydroisoquinoline moiety. (b) Traditional synthetic methods.In recent decades, isocyanides have been extensively studied in multicomponent reactions,4 imidoyl radical-involved transformations5 and transition-metal-catalyzed insertion reactions6 due to their diverse and unique reactivities. A variety of acyclic imine derivatives were synthesized via palladium-catalyzed imidoylation reactions.7 For the construction of biorelevant nitrogen-containing heterocycles via Pd-catalyzed non-functionalized isocyanide insertion reactions, a nucleophilic group is usually preinstalled to the substrates containing C–X (C–H) bonds, followed by Pd-catalyzed oxidative addition (C–H bond activation)/isocyanide insertion/ligand substitution/reductive elimination sequence.8 For example, Maes and coworkers developed an efficient palladium-catalyzed synthesis of 4-aminoquinazolines from N-(2-bromoaryl)amidines using amidine as an internal nucleophile.8a Access to the same scaffold through palladium-catalyzed amidine-directed C–H bond imidoylation/cyclization reaction was reported by Zhu group.8c Isocyanide acted as a C1 synthon in these cyclization reactions, with the terminal carbon of isocyanide being used in the ring formation.9 Considering the diversity of R group on the nitrogen atom of isocyanide, the so-called functionalized isocyanide strategy was widely developed as well, with various N-heterocycles such as oxazoles,10 indoles,11 phenanthridines,12 9H-pyrido[3,4-b]indoles13 and nonaromatic azepines14 being synthesized efficiently. Zhu and coworkers developed the first asymmetric C–H bond imidoylation reaction from easily accessible α,α-dibenzyl-α-isocyanoacetates with chiral quaternary-carbon-containing 3,4-dihydroisoquinolines being generated (Scheme 2a).15 Recently, a catalytic system controlled regioselective C–H bond imidoylation was realized for the selective synthesis of 5 or 6-membered cyclic imines (Scheme 2b).16 Aryl halides were mostly used in these transformations while N–O compounds were rarely reported as electrophiles in Pd-catalyzed isocyanide insertion reactions.17 Zhu group reported the synthesis of phenanthridin-6-amine derivatives from O-benzoyl hydroxylamines and commonly used biaryl isocyanides.17c However, a methyl or methoxyl carbonyl group was needed at the ortho position of isocyano group, which critically limited the application of the reaction. Herein, we developed an efficient and practical synthesis of 1-amino-3,4-dihydroisoquinolines, through palladium-catalyzed intramolecular C–H bond aminoimidoylation reaction of α-benzyl-α-isocyanoacetates. Consecutive isocyanide insertion and C–H bond activation were realized with C–N and C–C bonds being constructed in one step, employing O-benzoyl hydroxylamines as electrophilic amino sources.Open in a separate windowScheme 2Palladium-catalyzed imidoylative cyclization of α-benzyl-ethylacetates.The reaction conditions were optimized with ethyl 2-benzyl-2-isocyano-3-phenylpropanoate (1a) and morpholino benzoate (2a) as model substrates, and ethyl 3-benzyl-1-morpholino-3,4-dihydroisoquinoline-3-carboxylate (3a) was obtained in 34% yield in the presence of Pd(OAc)2 (1.5 equiv.), PPh3 (20 mol%), Cs2CO3 (1.0 equiv.) and PivOH (0.6 equiv.) in toluene at 80 °C (entry 1, EntryCatalystLigandSolventYieldb1Pd(OAc)2PPh3Toluene342Pd(OAc)2PPh3THFTrace3Pd(OAc)2PPh3DCE294Pd(OAc)2PPh3Dioxane765Pd(OAc)2PPh3MeCNTrace6Pd(OAc)2PPh3DMSOTrace7Pd(OPiv)2PPh3Dioxane318Pd(TFA)2PPh3Dioxane269PdCl2(MeCN)2PPh3Dioxane3811Pd(PPh3)4Dioxane5612Pd(OAc)2BINAPDioxane3013Pd(OAc)2dppbDioxane2614Pd(OAc)2dpppDioxane2815Pd(OAc)2XantPhosDioxane4216Pd(OAc)2SPhosDioxane3417Pd(OAc)2XPhosDioxane2418cPd(OAc)2PPh3Dioxane84 (80)d19cNonePPh3Dioxane020cPd(OAc)2NoneDioxane5621c,ePd(OAc)2PPh3Dioxane022c,fPd(OAc)2PPh3Dioxane4123c,gPd(OAc)2PPh3Dioxane2424c,hPd(OAc)2PPh3Dioxane3125c,iPd(OAc)2PPh3Dioxane61Open in a separate windowaReaction conditions: 1a (0.1 mmol), 2a (0.15 mmol), Pd-catalyst (0.01 mmol, 10 mol%), ligand (0.02 mmol, 20 mol%), Cs2CO3 (0.1 mmol, 1.0 equiv.), PivOH (0.06 mmol, 0.6 equiv.), 80 °C, Ar. A solution of 1a was added via a syringe pump within 1 h.bNMR yield with 1-iodo-4-methoxybenzene as an internal standard.c T = 110 °C.dIsolated yield.eWithout Cs2CO3 and PivOH.fCsOPiv was used as the base.gK2CO3 was used as the base.hNa2CO3 was used as the base.i5 mol% of catalyst was used.With the optimized conditions in hand, the scope of functionalized isocyanide was investigated first, using morpholino benzoate 2a as the coupling partner (Scheme 3). Both electron rich and withdrawing substituents such as Me, t-Bu, F, Cl, CF3, CO2Me and NO2 at the para position of aromatic ring can tolerate the reaction conditions, and generate the corresponding products in good to excellent yields (3a–3h). Isocyanides bearing ortho substituents on the phenyl ring underwent the reaction smoothly, generating the products 3i and 3j in 58% and 82% yields respectively. Even substrates bearing two substituents were well suited to the reaction conditions, affording the aminoimidoylation products without loss of efficiency (3k–3l). Compound 3m was delivered with excellent regioselectivity in 63% yield when m-Cl substituted starting material was employed in the reaction, which indicated the hindrance sensitivity of this reaction. Modifying the ester moiety in 1 to a methyl ester or an amide did not impede the reaction, giving 3n and 3o with excellent yields. Changing one of the benzyl groups to alkyl substituents didn''t decrease the reaction efficiency, generating products 3p and 3q in moderate yields.Open in a separate windowScheme 3Scope of isocyanides. Reaction conditions: 1 (0.1 mmol), 2a (0.15 mmol), Pd(OAc)2 (0.01 mmol, 10 mol%), PPh3 (0.02 mmol, 20 mol%), Cs2CO3 (0.1 mmol, 1.0 equiv.), PivOH (0.06 mmol, 0.6 equiv.), 110 °C, Ar. A solution of 1 was added via a syringe pump within 1 h. Isolated yields.Then we moved our attention to the investigation of substrate scope of aminating partners (Scheme 4). It was found that various O-benzoyl hydroxylamines derived from piperidines could be smoothly employed in the reaction (4a–4e). Substituents such as Me, CO2Et, OMe and ketal group on the piperidines were well tolerated (4b–4e). Aminating reagent derived from Boc-protected piperazine was also compatible with this reaction, generating the product 4f in 99% yield. Fortunately, acyclic aminated product 4g was obtained as well in moderate yield.Open in a separate windowScheme 4Substrate scope of aminating agents. Reaction conditions: 1a (0.1 mmol), 2 (0.15 mmol), Pd(OAc)2 (0.01 mmol, 10 mol%), PPh3 (0.02 mmol, 20 mol%), Cs2CO3 (0.1 mmol, 1.0 equiv.), PivOH (0.06 mmol, 0.6 equiv.), 110 °C, Ar. A solution of 1a was added via a syringe pump within 1 h. Isolated yields.When ethyl 2-isocyano-2,3-diphenylpropanoate was conducted in the reaction with 2a, the 3,4-dihydroisoquinoline product 6 was generated exclusively, albeit in only 37% yield (Scheme 5a). To examine if isoindoline could be afforded by this aminoimidoylation reaction, ethyl 2-isocyano-4-methyl-2-phenylpentanoate (7) was selected as the substrate with 2a (Scheme 5b). However, no desired product was generated under the standard conditions, showing different site-selectivity with the previous imidoylation reaction developed by Zhu and coworkers.16Open in a separate windowScheme 5Reaction conditions: 5(7) (0.1 mmol), 2a (0.15 mmol), Pd(OAc)2 (0.01 mmol, 10 mol%), PPh3 (0.02 mmol, 20 mol%), Cs2CO3 (0.1 mmol, 1.0 equiv.), PivOH (0.06 mmol, 0.6 equiv.), 110 °C, Ar. A solution of 5(7) was added via a syringe pump within 1 h. Isolated yields.Gram-scale preparation of 3a was carried out smoothly under standard conditions (Scheme 6a). The reduction reaction of 3a was carried out using LiAlH4 and the corresponding product 8 was generated in 70% yield (Scheme 6b). The ester group was successfully transformed to tertiary alcohol (9) in moderate yield with Grignard reagent.Open in a separate windowScheme 6Gram-scale preparation and diversifications of 3a.In order to get further insight into the reaction mechanism, a radical trapping reaction with 0.1 equivalents of TEMPO was carried out (Scheme 7a). The product 3a was generated without loss of efficiency, which indicated that the radical-involved pathway could be ruled out. A plausible mechanism was then proposed in Scheme 7b. The reaction was initiated with oxidative addition of O-benzoyl hydroxylamine 2a to the in situ generated Pd(0) species. Migratory isocyanide insertion to intermediate B afforded aminoimidoyl Pd(ii) intermediate C. The following intramolecular C–H bond activation and reductive elimination viaE yielded the final product 3a and regenerated Pd(0) to complete the catalytic cycle.Open in a separate windowScheme 7Proposed mechanism.In conclusion, we have developed an efficient and practical method to prepare 1-amino-3,4-dihydroisoquinolines via palladium-catalyzed intramolecular C–H bond aminoimidoylation of α-benzyl-α-isocyanoacetates under redox-neutral conditions. Consecutive isocyanide insertion and C–H bond activation were realized with C–N and C–C bonds formation in one step, employing O-benzoyl hydroxylamines as electrophilic amino sources. A broad range of substrates were applicable to the reaction in good to excellent yields.  相似文献   

18.
Diastereoselective synthesis of chroman bearing spirobenzofuranone scaffolds via oxa-Michael/1,6-conjugated addition of para-quinone methides with benzofuranone-type olefins     
Hongmei Qin  Qimei Xie  Long He 《RSC advances》2022,12(26):16684
A simple and convenient cyclization of ortho-hydroxyphenyl-substituted para-quinone methides with benzofuran-2-one type active olefins via oxa-Michael/1,6-conjugated addition has been developed, which afforded an easy access to enriched functionalized chroman-spirobenzofuran-2-one scaffolds with good to excellent yields (up to 90%) and diastereoselectivities (up to >19 : 1 dr). This reaction provided an efficient method for constructing desired spirocyclic compounds combining both well-known heterocyclic pharmacophores chroman and benzofuran-2-one.

Highly diastereoselective synthesis of spirocyclic compounds combining both well-known heterocyclic pharmacophores chroman and benzofuran-2-one.

The chroman framework represents a privileged heterocyclic core commonly found within a wide variety of biologically active natural products1 and synthetic compounds of medicinal interest (Fig. 1).2 Owing to the wide application of these heterocyclic molecules, over the past few decades, numerous efforts have been devoted to the efficient synthesis of chroman nucleus motifs.3,4 In particular, incorporating chroman into spiro-bridged and spiro-fused heterocyclic systems is appealing due to its fascinating molecular architecture and proven biological activity.5 Among the existing methods, the [4 + m] cycloaddition of para-quinone methides (p-QMs) is found to be an efficient pathway to access these valuable spirocyclic skeletons.6 For instance, the Enders group synthesized functionalized chromans with an oxindole motif by the asymmetric organocatalytic domino oxa-Michael/1,6-addition reaction.7 After that, the Hao,8a Peng,8b Shi,8c,d Zhou,8e Liang8f and Wang8g groups developed convenient methods to construct chromans bearing spirocyclic skeletons from p-QMs, respectively. Despite all these shining achievements, however, it is still very challenging to simply and conveniently construct chromans bearing quaternary carbon spirals for organic chemical or drug discovery among these [4 + m] cycloaddition reactions.Open in a separate windowFig. 1Representative chroman compounds.Benzofuran-2-(3H)-ones as one of the important oxygen-containing heterocycles that exist in a broad array of natural products9 and potential medicines.10 The streamlined synthesis of benzofuran-2-ones pose considerable challenge due to their quaternary carbon centers at the C-3 position,11 especially those featuring relatively congested spirocyclic motifs represent challenging synthetic targets.12,13 In our continuous interests in developing efficient method for the synthesis of spirocyclic compounds based on cyclization reaction,14 we wish to report a cycloaddition of para-quinone methides with benzofuranone derived olefins, affording the spiro-cycloadducts in good to excellent yields and diastereoselectivities. This cyclization features the simultaneous formation of chroman and spirobenzofuran-2-one skeletons in a single step (Scheme 1), which may be potentially applied as pharmaceutical agents.Open in a separate windowScheme 1Strategy for the synthesis of chroman-spirobenzofuran-2-one.We initiated our investigations with the readily available ortho-hydroxyphenyl-substituted para-quinone methides 1a and 3-benzylidenebenzofuran-2-one 2a in toluene at room temperature in the presence of base. Unfortunately, no desired chroman derivatives bearing spirobenzofuranone scaffolds 3a was isolated in the presence of 2 eq. Na2CO3 after stirring at room temperature for 48 h (entry 1, EntryBaseSolventYieldb (%)Drc1Na2CO3Toluene——2DMAPToluene——3K2CO3Toluene253 : 14CsFToluene525 : 15Cs2CO3Toluene706 : 16Cs2CO3THF68>19 : 17Cs2CO3Et2O608 : 18Cs2CO3Dioxane537 : 19Cs2CO3CH3CN50>19 : 110Cs2CO3DMF3115 : 111dCs2CO3THF65>19 : 112eCs2CO3THF75>19 : 113fCs2CO3THF64>19 : 1Open in a separate windowaReaction conditions: p-QMs 1a (0.1 mmol), benzofuranones 2a (0.12 mmol) and base (0.2 mmol) in 2 mL of solvent for 6–48 h.bIsolated yields.cDetermined by crude 1H NMR analysis.dPerformed at 30 °C.ePerformed at 10 °C.fPerformed at 0 °C.With the optimized reaction conditions in hand, we explored the substrate scope of this cyclization reaction with a selection of benzofuranones. The results are shown in EntryR3Yieldb (%)Drc1C6H53a75>19 : 122-ClC6H43b69>19 : 132-BrC6H43c6715 : 142-CH3C6H43d64>19 : 153-CH3C6H43e75>19 : 163-MeOC6H43f72>19 : 173-NO2C6H43g71>19 : 183-BrC6H43h75>19 : 194-CH3C6H43i7310 : 1104-MeOC6H43j86>19 : 1114-CF3C6H43k8212 : 1124-NO2C6H43l8512 : 1134-ClC6H43m83>19 : 1144-BrC6H43n8112 : 1152-Naphthyl3o74>19 : 1163,4,5-(OMe)3C6H23p8812 : 1173-Pyridinyl3q79>19 : 1183-Thiophenyl3r80>19 : 1Open in a separate windowaReaction conditions: p-QMs 1a (0.1 mmol), benzofuranones 2 (0.12 mmol) and Cs2CO3 (0.2 mmol) in 2 mL of THF.bIsolated yields for 4–48 h.cDetermined by crude 1H NMR analysis.Subsequently, the generality of this cyclization reaction was further evaluated through varying p-QMs 1. As shown in EntryR4Yieldb (%)Drc15-Cl4a73>19 : 125-Br4b65>19 : 135-CH34c90>19 : 145-OCH34d8010 : 154-CH34e72>19 : 164-OCH34f6610 : 1Open in a separate windowaReaction conditions: p-QMs 1 (0.1 mmol), benzofuranones 2a (0.12 mmol) and Cs2CO3 (0.2 mmol) in 2 mL of THF for 6–48 h.bIsolated yields.cDetermined by crude 1H NMR analysis.The structure and relative configuration of 3a were determined by HRMS, NMR spectroscopy and single-crystal X-ray analysis.15 The relative configuration of other cycloadducts were tentatively assigned by analogy (Fig. 2 and see ESI).Open in a separate windowFig. 2X-ray crystal structure of 3a.  相似文献   

19.
Copper-catalyzed three-component reaction to synthesize polysubstituted imidazo[1,2-a]pyridines     
Zitong Zhou  Danyang Luo  Guanrong Li  Zhongtao Yang  Liao Cui  Weiguang Yang 《RSC advances》2022,12(31):20199
An efficient three-component one-pot and operationally simple cascade of 2-aminopyridines with sulfonyl azides and terminal ynones is reported, providing a variety of polysubstituted imidazo[1,2-a]pyridine derivatives in moderate to excellent yields. In particular, the reaction goes a through CuAAC/ring-cleavage process and forms a highly active intermediate α-acyl-N-sulfonyl ketenimine with base free.

Three-component one-pot synthesis of polysubstituted imidazo[1,2-a]pyridine derivatives through a base free CuAAC/ring-cleavage process.

Polysubstituted imidazo[1,2-a]pyridines are well established as privileged scaffolds which are commonly encountered in many bioactive natural products and biological molecules that may be good drug candidates.1 Most imidazo[1,2-a]pyridines possess various biological activities, like antibacterial,2 antiinflammatory,3 antiviral,4 and anticancer.5 Some of the imidazo[1,2-a]pyridine derivatives are commercially available drugs, including Saripidem,6 Alpidem,7 Zolpidem,8 Zolimidine,9 Miroprofen10 and drug candidates GSK812397 (Fig. 1).11 Therefore, the development of novel methods for the synthesis of these imidazo[1,2-a]pyridines is important in the field of synthetic organic and pharmaceutical chemistry.Open in a separate windowFig. 1Some imidazo[1,2-a]pyridine drugs or drug candidates.In the past few years, reactions utilizing Cu,12 Pd,13 Mn,14 TEMPO-mediated,15 I2 (ref. 16) and a few other catalysts17 have provided attractive and valuable routes for the construction of imidazo[1,2-a]pyridines. However, most reactions can only produce monosubstituted imidazo[1,2-a]pyridines or halogenated intermediates (Scheme 1a)18 which can undergo one more steps of coupling reaction leading to polysubstituted products. Therefore, developing one-pot synthetic reactions will provide a direct and powerful tool to meet these challenges. To the best of our knowledge, imidazo[1,2-a]pyridines can be synthesized from 2-aminopyridines, terminal alkyne and aldehyde in a three-component coupling reaction, catalyzed by copper, in one pot (Scheme 1b).19 However, aldehydes are unstable and easily oxidized. They are environmentally unfriendly for synthesis or complex procedures. Under this background, the development of multicomponent one-pot synthetic strategies for the preparation of polysubstituted imidazo[1,2-a]pyridines still remains highly desirable.Open in a separate windowScheme 1Synthesis of polysubstituted imidazo[1,2-a]pyridines.Previous studies reported that the copper-catalyzed multicomponent reactions (MCRs) of sulfonyl azides, terminal alkynes and other components (CuAAC/ring-cleavage reaction) has been applied to synthesize numerous oxygen- and nitrogen-containing heterocyclic compounds.20 However, the reaction generally carried out under strong base conditions, and limited the application of some substrates, such as terminal ynones, which will take a self-condensation under the base conditions.21 Thus, the neutral or weak acidic conditions have developed by our group and the terminal ynones successfully used in CuAAC/ring-cleavage reaction to form a highly active intermediate α-acyl-N-sulfonyl ketenimines.22 Accordingly, an efficient one-pot and operationally three-component reaction of 2-aminopyridines, sulfonyl azides and terminal ynones is reported (Scheme 1c).Our initial study began with an examination of the synthesis of imidazo[1,2-a]pyridine 4a from 2-aminopyridine (1a), ethyl propiolate (2a) and p-tosyl azide (3a). Initial screenings involved using CuI as catalyst and no additive with a variety of solvents in a range of standard solvents. These results revealed that the desired conversion could be effected in most solvents ( EntryCat.SolventYieldb (%) 4a1CuICHCl3742CuIDCE773CuIToluene784CuIMeCN805CuITHF626CuI1,4-Dioxane447CuIDMSO268CuIDMF359 CuI EtOH 83 10CuClEtOH7511CuBrEtOH7312CuBr2EtOH7013Cu(OAc)2EtOH5014Cu(OTf)2EtOH3215AgOAcEtOHndc16CuIEtOH80d17CuIEtOH76eOpen in a separate windowaReaction conditions: 1a (1.0 mmol), cat. (10 mol%) in the solvent (3 mL) was added 2a (1.5 mmol) and 3a (1.5 mmol) stirring at 80 °C for 6 h.bIsolated yields.cnd = not detected the target product.dThe reaction temperature was 70 °C.eThe temperature was 90 °C.Under the optimized conditions ( Open in a separate windowaUnless otherwise noted, the reaction conditions were as follow: 1 (1.0 mmol), CuI (10 mol%) in the MeCN (3 mL) was added 2 (1.5 mmol), 3 (1.5 mmol) stirring at 80 °C for 6 h.Next, the scope and limitation of the terminal ynone 2 and sulfonyl azide 3 substrates were tested. It is noteworthy that the sulfonyl azide substrates showed slight influences on this reaction. With R3 changed by aromatic or aliphatic substituents, such as –Ph, –(4-ClC6H4), –(4-CF3C6H4), –(4-OMeC6H4), –Me and –n-Bu, the reaction could smoothly give the anticipated products (4f–4n) in comparable yields. The substrates R2 bearing the –OMe, –OtBu, –Me and other alkyl group also can obtain 4o–4r in good yields.In order to broaden the suitability of substrates, we also investigated other terminal alkynes, such as aryl acetylenes. The experiments revealed that some aryl acetylene such as 3-methyl phenylacetylene can obtain imidazo[1,2-a]pyridine 4u with low yield of 26% and an uncyclized linear product 4v (Scheme 2a). Most aryl acetylenes such as 4-methoxy phenylacetylene only obtain uncyclized products (Scheme 2b). It shows that the reactivity of terminal ynones is higher than that of traditional terminal alkyne.Open in a separate windowScheme 2Investigation the reaction of 2-aminopyridine (1a), aryl acetylenes (2f, 2g) and p-tosyl azide (3a).None of the product imidazo[1,2-a]pyridines 4a–4r have been reported previously, which were subject to full spectroscopic characterization (see ESI for details) and the derived data were in complete accord with the assigned structures. And 4a was confirmed by single-crystal X-ray analysis (Fig. 2).Open in a separate windowFig. 2Single-crystal X-ray analysis of 4a (CCDC 2121234).A possible reaction pathway for the formation of imidazo[1,2-a]pyridine (4a) from precursors 1a, 2a and 3a is shown in Scheme 3. Thus, in keeping with earlier proposals,19,21 the substrates 2a and 3a are expected to react, in the presence of the copper(i) catalyst to form the metallated triazole A through the CuAAC procedure. Then, the complex A undergo a ring-cleavage rearrangement leading to a highly active intermediate N-sulfonyl-α-acylketenimine B. This last species B is captured by 1avia nucleophilic addition to generate the intermediate C, which deliver the observed product 4a by intramolecular oxidative coupling similar to literature.23 Otherwise, due to the poor activity, most of the traditional terminal alkynes involved in the reaction will stop in the intermediate C leading the uncyclized products.Open in a separate windowScheme 3Plausible reaction mechanism.In summary, we have developed an original approach for the synthesis of polysubstituted imidazo[1,2-a]pyridines from a mixture of the corresponding 2-aminopyridines, sulfonyl azides and terminal ynones, through CuAAC/ring-cleavage process and generated a highly active intermediate α-acyl-N-sulfonyl ketenimines. More detailed novel reactions and the investigation of new applications of this intermediate are now being undertaken in our laboratory.  相似文献   

20.
Direct phosphorylation of benzylic C–H bonds under transition metal-free conditions forming sp3C–P bonds     
Qiang Li  Chang-Qiu Zhao  Tieqiao Chen  Li-Biao Han 《RSC advances》2022,12(29):18441
Direct phosphorylation of benzylic C–H bonds was achieved in a biphasic system under transition metal-free conditions. A selective radical/radical sp3C–H/P(O)–H cross coupling was proposed, and various substituted toluenes were applicable. The transformation provided a promising method for constructing sp3C–P bonds.

Direct phosphorylation of benzylic C–H bonds with secondary phosphine oxides was first realized. The reaction was performed in organo/aqueous biphasic system and under transition metal-free conditions, proceeding via the cross dehydrogenative coupling.

To construct C–P bonds is of great significance in modern organic synthesis,1 because organophosphorus compounds play varied roles in medical,2,3 materials,4 and synthetic chemistry fields.5 Traditionally, the C–P bonds were formed from P–Cl species via nucleophilic substitutions with organometallic reagents,6 P–OR species via Michaelis–Arbusov reactions,7 or P–H species via the alkylation in the present of a base or a transition metal.8Over the past decades, cross dehydrogenative coupling reactions (CDC reactions) have become a powerful and atom-economic methodology for constructing chemical bonds.9 By using this strategy, C–H bonds can couple with Z–H bonds without prefunctionalization and thus short-cut the synthetic procedures (Scheme 1a).Open in a separate windowScheme 1Cross dehydrogenative coupling reactions and direct phosphorylation of benzylic C–H bonds.A similar construction of C–P bonds via CDC was also realized.10,11 Among these methods, the phosphorylation of sp3C–H having an adjacent N or O atom, or the carbonyl group was well-developed.11 Relatively, the formation of benzylic sp3C–P bond was less reported,11w which was mainly limited to the sp3C–H of xanthene or 8-methylquinoline (Scheme 1b).12 In these reported processes, transition metal catalysts or photo-, electro-catalysts were usually involved,11v and an excess of P(O)–H compounds was usually employed.11,13 To the best of our knowledge, the phosphorylation of non-active benzylic C–H bonds has scarcely been reported.Considering both benzylic and phosphorus radicals could be generated by oxidation,14 which might subsequently couple, the phosphorylation of benzylic sp3C–H bonds would be achieved (Scheme 1c). Herein, we disclosed the construction of benzylic C–P bonds from toluene and P–H species. The reaction was carried out under transition metal-free reaction conditions,15 and exhibited high regio-selectivity. The aromatic C–H remained intact during the reaction.We began our investigation by exploring the reaction of toluene 1a and diphenylphosphine oxide 2a in the presence of an oxidant (16 Other persulfates could also be used as an oxidant albeit with decreasing yields ( EntryOxidantToluene/water (v/v)Additive3a yieldb (%)3a/4c1K2S2O81 : 0—Trace—2K2S2O81 : 1—1216 : 843K2S2O81 : 1SDS3726 : 744Na2S2O81 : 1SDS3127 : 735(NH4)2S2O81 : 1SDS2723 : 776Oxone1 : 1SDSNone—7—1 : 1SDSNone—8—1 : 1—None—9K2S2O81 : 1SDBS4136 : 6410K2S2O81 : 2SDBS4628 : 7211K2S2O81 : 3SDBS3524 : 7612dK2S2O81 : 2SDBS4632 : 6813eK2S2O81 : 2SDBS2621 : 7914fK2S2O81 : 2SDBS2926 : 7415gK2S2O81 : 2SDBS4838 : 6216g,hK2S2O81 : 2SDBS0—Open in a separate windowaReaction condition: 1a (1 mL), 2a (0.2 mmol), oxidant (2 equiv.), additive (1 equiv.) and H2O, 120 °C, 3 h. under N2.bGC yields using n-dodecane as an internal standard.cThe ratio of 3a/4 was determined by GC analysis.d50 mol% SDBS was used.e20 mol% SDBS was used.fAt 100 °C.g1 (0.8 mL) and H2O (1.6 mL), 3 equiv. K2S2O8 was used, 120 °C for 15 min.h3 equiv. TEMPO was added.Based on the above results, we can easily find that a serious amount of 1,2-diphenylethane 4 was formed. These results suggest that the homocoupling rate of 1a was very quick. Thus, the choice of the phase transfer reagent and oxidant is the key to cross-coupling of toluene and diphenylphosphine oxide in this biphasic solvent system.Excessive P–H species were usually employed in reported CDC reactions to form C–P bonds, because of their facile oxidation.13 In our procedure, toluene was excessive, thus the yields were calculated based on 2a. The yields looked like low, which did not indicate the poorer conversion rate of P–H species. With the optimized conditions in hand, the substrate scope of the CDC reactions was explored ( Open in a separate windowaReaction conditions: 1 (0.8 mL), 2 (0.2 mmol), K2S2O8 (3 equiv.), SDBS (50%) and H2O (1.6 mL), 120 °C, under N2, the reactions were monitored by TLC and/or GC until 2 work out.b1 mmol scale, 30 min.c130 °C.d100 °C.In addition to toluene, o-xylene, m-xylene, p-xylene, mesitylene, and 1,2,4,5-tetramethylbenzene all coupled with 2a to give the expected 3b–f in moderate yields. Methoxy substituted toluene gave relatively lower yields of 3g and 3h. para-Halo substituted toluene exhibited good reactivity, furnishing the coupling products 3i and 3j in moderate yields. Comparing to toluene, a decreasing order of reactivity was observed for ethyl benzene (3l, 30% yield), isobutyl benzene (3m, 27% yield), isopropyl benzene (3n, <10% yield), and diphenyl methane (3o, trace). The order was probably controlled by the steric hindrance around the benzylic carbon. 1-Methylnaphthalene and 2-methylnaphthalene also gave low yields (3p and 3q). However, 2-methylquinoline served well and coupled with 2a, affording the product 3r in 49% yield, which could be ascribed to the activation of the nitrogen atom. Besides of 2a, diaryl phosphine oxides having methyl, F, and Cl substituents could also be employed as the substrate, producing 3s–3u in moderate yields under similar reaction conditions.Although the mechanism of the direct phosphorylation of benzyl C–H bond in aqueous solution is not quite clear, some aspects could be grasped based on experimental results. Firstly, the desired product 3a was not detected when the 2,2,6,6-tetramethyl-1-piperidinyloxy (TEMPO) was added, implying that a radical pathway might be possible in this reaction (entry 16). Secondly, the reaction only occurred in aqueous phase, and relied on the presence of PTC (phase transfer catalyst), as seen in that 3a was difficultly formed in entries 1 and 2 of 17We supposed 1 is converted to benzyl radical 5 by SO4 radical that is generated via hemolytic cleavage of potassium persulfates.18 Meanwhile, phosphorus radical 6 is similarly formed from 2. Because potassium persulfate was water soluble, both 1 and 2 had to be transferred into aqueous solution to react with potassium persulfates.19 This proposal is also in accord with the experimental results that no products of sp2C–H phosphorylation are detected, which are the main products in the previous radical systems.20 Finally, cross couplings between 5 and 6 produce 3 (Scheme 2).Open in a separate windowScheme 2Proposed mechanism for the direct phosphorylation of benzylic C–H bonds under transition metal-free reaction conditions.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号