Synthesis 2023; 55(22): 3777-3792
DOI: 10.1055/a-2126-3774
feature

Access to Symmetrical and Unsymmetrical Cyclobutanes via Template-Directed [2+2]-Photodimerization Reactions of Cinnamic Acids

Bilge Banu Yagci
a   Department of Chemistry, Faculty of Science, Bilkent University, Ankara 06800, Türkiye
,
Badar Munir
a   Department of Chemistry, Faculty of Science, Bilkent University, Ankara 06800, Türkiye
,
Yunus Zorlu
b   Department of Chemistry, Gebze Technical University, Gebze, Kocaeli 41400, Türkiye
,
a   Department of Chemistry, Faculty of Science, Bilkent University, Ankara 06800, Türkiye
c   UNAM, National Nanotechnology Research Center, Institute of Materials Science and Nanotechnology, Bilkent University, Ankara 06800, Türkiye
› Institutsangaben
Y.E.T. acknowledges financial support by the GEBIP Award of the Turkish Academy of Sciences.
 


Abstract

In this work, we have developed a general and broadly applicable template-directed photochemical [2+2]-cycloaddition reaction which provides access to a wide range of symmetrical and unsymmetrical cyclobutane products. The use of 1,8-dihydroxynaphthalene as a covalent template paved the way for successful and highly selective photochemical homodimerization and heterodimerization reactions in the solid state between cinnamic acid derivatives. Notably, the method works equally well with aryl- and heteroaryl-containing substrates leading to the formation of β-truxinic acid analogues as single diastereomers and in high yields (up to 99%).


#

Biographical Sketches

Zoom Image

Bilge Banu Yagci received her B.Sc. in chemistry from Bilkent University in 2018. Subsequently, she completed her M.Sc. studies under the supervision of Dr. Yunus Türkmen in 2021. Her research is focused on template-directed photochemical [2+2]-cycloaddition reactions of cinnamic acid derivatives for the diastereocontrolled synthesis of cyclobutane products.

Zoom Image

Badar Munir studied chemistry at the University of Sargodha (Pakistan) where he obtained his B.Sc. in Chemistry. Currently, he is doing his M.Sc. in Chemistry at Bilkent University (Türkiye) under the supervision of Dr. Yunus Emre Türkmen. His research is mainly focused on photochemical dimerization reactions to access carbocyclic compounds.

Zoom Image

Yunus Zorlu was born in İstanbul in 1982. He carried out his Ph.D. on water soluble symmetrical and asymmetrical phthalocyanine-based photosensitizers for photodynamic therapy of cancer. After studying inorganic chemistry, he was then appointed research scientist position in the field of structural chemistry in 2008, then associate professor in 2015 at Gebze Technical University. Since 2020, he has been an inorganic chemistry professor at Gebze Technical University. His primary research interests include the development of functional porous materials (metal-organic frameworks, hydrogen-bonded organic frameworks, and porous organic polymers) and non-covalent interactions (halogen and tetrel bondings).

Zoom Image

Yunus Emre Türkmen received his B.Sc. in 2005 at the Chemistry Department of Middle East Technical University (METU) with a minor study in biology. He finished his M.Sc. studies at the same institution under the supervision of Prof. Dr. Cihangir Tanyeli. He then moved to Chicago in 2006 and did his Ph.D. in the group of Prof. Dr. Viresh Rawal at the University of Chicago. Between 2013 and 2015, he worked as a postdoctoral fellow in the group of Prof. Dr. Varinder Aggarwal FRS at University of Bristol, UK. Since 2015, he has been working at the Chemistry Department of Bilkent University as a faculty member. His research interests include the development of new catalytic methods to access carbo- and heterocycles, and template-directed photochemical cycloaddition reactions. He has been a recipient of the BAGEP (Young Scientist Awards Program) award by the Science Academy, Türkiye, in 2019, and the GEBIP (Outstanding Young Scientists) award given by the Turkish Academy of Sciences in 2022.

Since early studies reported in the first half of the 20th century,[1] regio- and stereoselective photodimerization reactions of cinnamic acids have continued to be of interest to the synthetic community.[2] As a consequence of the natural occurrence of many substituted cinnamic acid derivatives, it is not surprising that cinnamic acid dimers constitute a prevalent motif among cyclobutane-containing natural products. Examples of biologically active members of this class of natural products such as eucommicin A,[3] itoside N,[4] caracasandiamide,[5] and incarvillateine[6] are shown in Figure [1]. The early observation that solution phase irradiation of cinnamic acids gave primarily E to Z isomerization paved the way for the advancement of studies on their solid state photochemical [2+2] cycloadditions.[7] Extensive pioneering work of Schmidt and co-workers on the crystal structure-photochemical reactivity relationship for a large number of trans-cinnamic acid derivatives resulted in the formulation of a number of principles currently known as the Schmidt criteria.[8] Based on these findings, it was concluded that irradiation of the α- and β-crystal polymorphs of cinnamic acids, which have distances of 3.6–4.1 Å between their parallelly oriented olefin centroids, resulted in the selective formation of α-truxillic acids (syn-head-to-tail dimers) and β-truxinic acids (syn-head-to-head dimers), respectively (Scheme [1a] and 1b).[8] [9] However, cinnamic acids having the γ-polymorphic structure in solid state have a distance of >4.7 Å between their olefin centroids, and as a result, they are photo-resistant under UV irradiation conditions (Scheme [1c]).[9]

Zoom Image
Figure 1 Examples of bioactive natural products having dimers of cinnamic acid derivatives in their structures

This is undoubtedly a powerful synthetic approach provided that the cinnamic acid derivative of interest has the aforementioned geometrical features in solid state, and thus is in agreement with the Schmidt criteria. Such photochemical cycloaddition reactions proceed generally with very high yields and diastereoselectivities.[7] [10] Moreover, this type of synthetic transformations is favored from a green chemistry perspective as they do not require the use of a solvent during the reaction.[11,12] Despite the advantages mentioned above, this approach does not provide a general solution to the selective photodimerization of any cinnamic acid derivative regardless of its solid-state structure. In this respect, several strategies were developed to achieve the selective photochemical [2+2] cycloadditions of a variety of olefin classes including cinnamic acids. One such strategy is the use of templates in solid state which involve the utilization of metal coordination[13] and salt formation[14] as well as a diverse range of non-covalent interactions such as hydrogen,[15] halogen,[16] and chalcogen bonding.[17] [18] Along these lines, a related second type of strategy is to utilize supramolecular host-guest complex formation[19] and non-covalent templates[20] for selective photochemical [2+2] cycloadditions in solution.[21] A particular advantage of these two approaches is that they operate in single synthetic reaction step as they do not involve the covalent binding of the reactants to the template. However, the first strategy requires the co-crystallization of the reactants with the template, which can be a limiting factor, whereas the second strategy requires a high binding constant between the reactants and the template in solution.

Zoom Image
Scheme 1 Photochemical reactivity of different polymorphs of cinnamic acids

In addition to the strategies described above, a third approach is the use of covalent templates which can be used in solid state or in solution. The use of certain diols as templates in the early studies of this field resulted in the formation of diastereomeric mixtures in the investigated photodimerization reactions.[22] In 1996, König and co-workers demonstrated that 1,2-bis(hydroxymethyl)benzene was an effective diol-based covalent template for the photochemical homodimerization of trans-cinnamic acid in solution.[23] In the seminal studies of Hopf and co-workers, [2.2]paracyclophane core was shown to be an effective, rigid scaffold satisfying the distance criterion for a successful [2+2]-cycloaddition reaction as described by Schmidt.[8] [24] [25] In 2010, Wolf and co-workers developed bisaniline 1 as a covalent template for the homodimerization of two cinnamic acid derivatives to afford symmetrical β-truxinic acids in high yield (Scheme [2a]).[26]

Within the past decade, visible light photocatalysis has emerged as a highly efficient method for [2+2]-cycloaddition reactions.[27] In 2017, Reiser and co-workers reported anti-head-to-head photodimerization of cinnamate esters giving δ-truxinates catalyzed by an Ir-based photocatalyst (Scheme [2b]).[28] However, when heterodimerization reactions were tested in this study via reactions of two different cinnamates, no appreciable selectivity was observed for heterodimerization over homodimerization. Furthermore, the [2+2]-cycloaddition reactions of cinnamic acids, cinnamates, and chalcones were investigated using a variety of other visible light photocatalytic systems.[29] Despite this rich history, the photochemical dimerization reactions of cinnamic acid derivatives were mainly restricted to homodimerization,[30] and a general method for the selective heterodimerization of this compound class had yet to be discovered.[31]

Zoom Image
Scheme 2 Selected examples of photochemical [2+2] cycloadditions of cinnamic acid derivatives

Table 1 Synthesis of the Symmetrical Diesters 6ag

Entry

Ar

Product

Yield (%)

1

Ph

6a a

95

2

6b a

87

3

6c

74

4

6d

81

5

6e

73

6

6f

81

7

6g

72

a The syntheses of these cycloadducts were reported previously.[32]

In 2021, we reported the use of 1,8-dihydroxynaphthalene (1,8-DHN, 5) as a highly effective covalent template for the selective photochemical homodimerization and heterodimerization reactions of cinnamic acids (Scheme [2c]).[32] It should be noted that, to the best of our knowledge, this method represents the first general solution for the selective heterodimerization reactions of this compound class. During our previous work on the hydrogen bonding properties of 1,8-DHN (5),[33] and its use in the synthesis of the fungal natural product daldiquinone,[34] we realized the potential of compound 5 to be utilized as an ideal covalent template for the photochemical [2+2]-cycloaddition reactions of cinnamic acids. Indeed, in addition to providing the right distance between the two reacting olefins (ca. 3.6 Å)[32] so as to ensure that Schmidt’s distance criterion (<4.2 Å) is satisfied, template 5 also enables the attachment of two different cinnamic acids selectively and sequentially (Scheme [2c]). We should note that, Beeler and co-workers recently described the use of catechol as a covalent tether in the photochemical heterodimerization reactions of cinnamic acids in solution, and showed the application of this method to the syntheses of the natural products piperarborenines C–E.[35] Herein, we report a full account of our studies on the selective and controlled homodimerization and heterodimerization reactions of cinnamic acids via the use of 1,8-DHN (5) as a covalent template, and extension of its scope to the selective photodimerization of heteroaromatic substrates (Scheme [2c]).

Our studies commenced with the investigation of the homodimerization reactions of cinnamic acid derivatives. The synthesis of the symmetrical diesters 6ag to be tested in the photochemical [2+2] cycloaddition was achieved via a DCC-mediated ester formation between 1,8-DHN (5) and the corresponding cinnamic acids 7 (Table [1]). We have previously shown that phenyl- and 4-methoxyphenyl-substituted diesters 6a and 6b could be prepared in high yields (95% and 87%, respectively) using this method (entries 1 and 2).[32] In the present work, diester 6c bearing the strongly electron-donating dioxolane moiety was synthesized in 74% yield (entry 3). In order to check the effect of an electron-withdrawing group on the aryl ring and assess the stability of an aryl halide group under photochemical reaction conditions, 4-chlorophenyl-substituted diester 6d was prepared in 81% yield (entry 4). Next, we turned our attention to the synthesis of diesters with heteroaromatic rings at the β-position of the acrylate moieties. To this end, diesters 6eg possessing thiophene, furan and N-methylpyrrole rings were synthesized successfully in 73%, 81%, and 72% yields, respectively (entries 5–7).

With the successful preparation of symmetrical diesters 6ag in hand, we next turned our attention to the construction of unsymmetrical diesters as photochemical cycloaddition precursors (Table [2]). Monoesters 8a and 8b were synthesized via deprotonation of 1,8-DHN (5) with NaH followed by treatment with one equivalent of the corresponding cinnamoyl chloride.[32] For the conversion of monoesters 8a and 8b to unsymmetrical diesters 6hp two methods were employed. Method A involves reacting monoesters 8a or 8b with the second cinnamic acid partner under DCC coupling conditions. Following this method, diesters 6h, 6i, 6m, and 6o were synthesized in 66–90% yields (entries 1, 2, 6, and 8).[32]

Table 2 Synthesis of the Unsymmetrical Diesters 6hp

Entry

Ar1

Ar2

Method, product, yield (%)

1

Ph

A, 6h a, 66

2

Ph

A, 6i a, 76

3

Ph

B, 6j, 86

4

Ph

B, 6k, 74

5

Ph

B, 6l, 72

6

Ph

A, 6m a, 90

7

Ph

B, 6n, 75

8

A, 6o a, 66

9

B, 6p, 77

a The syntheses of these cycloadducts were reported previously.[32]

However, for some substrates method B, which involves the reaction of the monoester with an acyl chloride under basic conditions, was found to work better. During these experiments, we realized that deprotonation of monoester 8a with one equivalent of NaH followed by treatment with one equivalent of acyl chloride of a different cinnamic acid led to the formation of a mixture of the targeted unsymmetrical diester and the symmetrical diester 6a. In control experiments, when monoester 8a was treated with DBU (0.1 equiv) or NaH (1.0 equiv) in the absence of additional cinnamoyl chloride, monoester 8a was observed to undergo a non-redox disproportionation to form diester 6a and 1,8-DHN (5) (Scheme [3]). Pleasingly, we found that this pathway could be prevented by a simple change in the addition order of reagents. In this way, when one equivalent of NaH was added to a 1:1 mixture of monoester 8a and (E)-3-(4-nitrophenyl)acryloyl chloride in THF, unsymmetrical diester 6j was obtained in 86% yield without the formation of any symmetrical diester 6a (Table [2], entry 3). This new protocol was successfully applied to the synthesis of other unsymmetrical diesters (6k, 6l, 6n, and 6p) shown in Table [2], which were isolated in high yields (72–77%).

Zoom Image
Scheme 3 Non-redox disproportionation of monoester 8a

The photochemical reactivity of the diesters 6 was examined using two different irradiation conditions (Table [3]). In method C, solid samples were irradiated for 8 h using a 400-W medium pressure mercury lamp, and samples were mixed thoroughly every two hours. Alternatively, in method D, a commercial nail dryer with four 9-W UV-A fluorescent lamps was used to irradiate diesters 6 in solid state for 16 h by mixing the samples every 4 h.[36] In general, both methods work comparably well with the majority of the tested cycloaddition precursors, albeit longer reaction times are required with method D. Please note that the substrates, which were investigated after our initial report,[32] were all tested with both irradiation methods. In this respect, the symmetrical template-bound β-truxinic acid esters 9ad bearing electron-donating and -withdrawing substituents were isolated in high yields (78–96%) and as single diastereomers (Table [3], entries 1–4). Diesters 6eg with heteroaromatic thiophene, furan, and N-methylpyrrole rings were also found to be excellent substrates under both irradiation conditions, and cycloadducts 9eg were obtained in 80–98% yields (entries 5–7).

Table 3Photochemical [2+2] Cycloadditions of Diesters 6

Entry

Product

Ar1

Ar2

Yield (%)

Method C

Method D

 1

9a a

Ph

Ph

95

92

 2

9b

78

 3

9c

93

96

 4

9d

89

90

 5

9e

96

98

 6

9f

86

98

 7

9g

95

80

 8

9h a

Ph

99

 9

9i a

Ph

88

10

9j

Ph

77

61

11

9k

Ph

88

86

12

9l

Ph

44

49

13

9m a

Ph

88

14

9n

Ph

86

96

15

9o a

93

16

9p

57

84

a The syntheses of these cycloadducts were reported previously.[32]

Previously, we had shown that unsymmetrical β-truxinates 9h and 9i, which have electron-rich 4-methoxyphenyl and electron-deficient 4-(trifluoromethyl)phenyl groups, could be obtained in 99% and 88% yields, respectively, using method C.[32] In the current work, we systematically examined the effect of the position of the NO2 group on the photochemical [2+2] cycloaddition. Whereas diesters 6j and 6k with NO2 groups at the para and meta positions were successful substrates providing cycloadducts 9j and 9k in 77% and 88% yields, respectively (entries 10 and 11), β-truxinate 9l with the NO2 group at the ortho position was isolated in a lower yield (44% with method C, and 49% with method D; entry 12). We were pleased to confirm the structure and relative stereochemistry of cycloadduct 9j by single-crystal X-ray analysis (Figure [2], CCDC 2265841). Mixed aryl-heteroaryl-substituted diesters 6m and 6n were also competent substrates affording the unsymmetrical cyclobutane diester products 9m and 9n in 88% and 96% yields, respectively (entries 13 and 14). The effect of having one electron-rich and one electron-deficient aryl ring on the substrate was tested with diester 6o which gave cycloadduct 9o in 93% yield (entry 15).[32] Finally, the presence of having two electron-rich aromatic groups was checked with diester 6p having furyl and 4-methoxyphenyl groups. For this substrate, method D was found to be superior affording unsymmetrical cycloadduct 9p in 84% yield, while method C resulted in a moderate yield of 57% (entry 16). Cycloadducts 9ap were observed to form as single diastereomers in all of the experiments shown in Table [3]. Finally, we checked the reversibility of the photochemical [2+2] cycloaddition under the irradiation conditions. To this end, we subjected a sample of pure cycloadduct 9a to irradiation using the Hg lamp (method C) for 4 h. At the end of this control experiment, cycloadduct 9a was found to be intact with no formation of diester 6a, which indicates that the photocycloaddition reaction is irreversible under these conditions.

Zoom Image
Figure 2 X-ray crystal structure of cycloadduct 9j

In order to check the possibility of a single crystal-to-single crystal transformation[37] during the photochemical [2+2] reaction, crystals of 6d were irradiated with UVA light using the nail dryer. While cycloadduct 9d was found to form with full conversion upon irradiation for 16 h as determined by NMR spectroscopy, the crystals were observed to crumble during the irradiation process resulting in the formation of product 9d in powder form (Figure S1). Regarding the interaction of diesters 6 with UV light, we have previously shown that the λmax values of diester 6a and 6o in their UV-vis spectra in CH2Cl2 are 279 and 283 nm (with a shoulder at 320 nm), respectively.[32] In order to acquire more relevant data related to their solid state photochemical reactivities, we recorded their diffuse reflectance UV-vis spectra in solid powder form, which exhibited absorption below 389 nm for 6a and 411 nm for 6o (Figure S2). This provides an explanation for the success of both irradiation sources (methods C and D) in the photochemical [2+2]-cycloaddition reactions of diesters 6 (Table [3]).

Zoom Image
Scheme 4 Synthesis of β-truxinic acid products 10 via hydrolysis of cycloadducts 9 under basic conditions. a The syntheses of these β-truxinic acid products were reported previously.[32] b Product 11 was obtained by the reaction of 9g with NaOMe in MeOH at 23 °C.

The synthesis of the β-truxinic acid products was achieved via a saponification-type basic hydrolysis of the template-bound cycloadducts 9. In this respect, treatment of diesters 9 with KOH in a mixture of THF and H2O at 23 °C followed by acidic workup afforded cyclobutanedicarboxylic acids 10 in high yields and as single diastereomers (Scheme [4]). The unsubstituted β-truxinic acid (10a) was isolated in quantitative yield along with complete recovery of the template (1,8-DHN, 5).[32] The method works successfully on diesters with electron-donating and withdrawing substituents affording β-truxinic acid analogues 10bd in 84–95% yields (Scheme [4]). Thienyl- and furyl-substituted symmetrical cyclobutanedicarboxylic acids 10e and 10f were obtained in 90% and 71% yields, respectively. Unexpectedly, we faced problems during the isolation of N-methylpyrrole-substituted dicarboxylic acid. In order to circumvent this problem, cycloadduct 9g was subjected to a transesterification reaction with the use of NaOMe in MeOH leading to the formation of cyclobutane diester 11 in 65% yield. Not surprisingly, the basic hydrolysis method works equally well with the heterodimerization products giving unsymmetrical β-truxinic acid products 10hp in 79–96% yields (Scheme [4]). Of particular note is the little variation of reaction yield depending on the position of the NO2 group in substrates 9jl. In these reactions, β-truxinic acid products 10j, 10k, and 10l were obtained in 91%, 91%, and 85% yields, respectively. Finally, mixed aryl-heteroaryl-substituted cycloadducts were also competent substrates, and hydrolysis products 10m, 10n, and 10p were isolated in 84–96% yields (Scheme [4]).

Zoom Image
Scheme 5 Conversion of cycloadduct 9n to 12 and 13

Next, we wanted to show the synthetic utility of the photocycloaddition reaction developed in this work via other transformations (Scheme [5]). First, cycloadduct 9n was reacted with NaOMe in methanol giving dimethyl diester product 12 in 93% yield. In a second experiment, reduction of the two ester groups of 9n with LiAlH4 afforded cyclobutanedimethanol 13 in 86% yield. In these reactions, both products were obtained as single diastereomers. Finally, we opted to check the [2+2] cycloaddition of diester 6a under visible light photocatalytic conditions.[29a] For this purpose, diester 6a was irradiated with blue LEDs (450 nm) in 1,4-dioxane in the presence of Ir(ppy)3 photocatalyst, and the reaction was observed to be complete within one hour (Scheme [6]). Whereas this reaction also gave β-truxinate 9a as the major cycloadduct, the product was found to consist of a mixture of three diastereomers (dr = 1:0.32:0.16). The same reaction was further tested in the absence of a photocatalyst but via irradiation using the medium-pressure mercury lamp in solution phase. In this experiment, when a solution of 6a in acetone was irradiated, the reaction was complete in 2 h and afforded again a diastereomeric mixture (dr = 1:0.66:0.13) with cycloadduct 9a being the major diastereomer. These results underscore the advantage of our solid-state photochemical cycloaddition protocol for achieving high levels of diastereoselectivity when 1,8-DHN (5) is used as a template.

Zoom Image
Scheme 6 Photocatalytic cycloaddition of diester 6a

In summary, we have developed a robust and selective method for the controlled photochemical homodimerization and heterodimerization reactions of cinnamic acids leading to the formation of symmetrical and unsymmetrical cyclobutanes. The method involves the use of 1,8-dihydroxynaphthalene (5) as a covalent template, that enables the positioning of the two reacting olefins within a distance of <4 Å which is in agreement with Schmidt’s distance criterion for a successful photochemical [2+2] cycloaddition in solid state. The photodimerization reactions work uniformly well with aryl- and heteroaryl-containing cinnamic acid derivatives affording cycloadducts in up to 99% yield and as single diastereomers. Cycloadducts were shown to be hydrolyzed easily under basic conditions at room temperature to yield symmetrical and unsymmetrical β-truxinic acids in excellent yields (71% to quant.). Research to render these photochemical [2+2] cycloadditions enantioselective is currently underway in our laboratory.

All air-sensitive solution phase reactions were run using oven-dried glassware under N2. Reactions were monitored by TLC using aluminum-backed plates pre-coated with silica gel (Silicycle, indicator: F-254; thickness: 200 μm). UV light and/or KMnO4 staining solution were used for the visualization of TLC plates. Flash column chromatographic separations were performed on Silicycle 40–63 μm (230–400 mesh) flash silica gel. 1H NMR (400 MHz) and 13C{1H} NMR (100 MHz) spectra were recorded using a Bruker Avance 400 spectrometer in CDCl3 and DMSO-d 6. Internal standard signal (TMS, δ = 0) or residual solvent signals (CDCl3 δ = 7.26, and DMSO-d 6 δ = 2.50 for 1H NMR; CDCl3­ δ = 77.16 and DMSO-d 6 δ = 39.52 for 13C{1H}-NMR spectra) were used for the calibration of 1H and 13C{1H} NMR spectra. Infrared (FTIR) spectra were recorded on a Bruker Alpha-Platinum-ATR spectrometer with only selected peaks reported. HRMS were performed using Agilent Technologies 6224 TOF LC/MS at UNAM-National Nanotechnology Research Center and Institute of Materials Science and Nanotechnology, Bilkent University. Single crystal X-ray diffraction analysis was performed at Gebze Technical University, Turkey. Photochemical experiments for method C were carried out using a reactor obtained from Photochemical Reactors Ltd., which consists of a 400-W medium-pressure mercury lamp (3040/PX0686) and a quartz double-walled immersion well with water cooling. Regular microscope slides made of soda-lime glass were used for the photochemical experiments in solid state. The slides were kept at ca. 4 cm away from the lamp during the experiments. Photochemical experiments for method D were carried out using a commercial UV gel nail dryer (Elle by Beurer, MPE58) fitted with four 9-W UVA (Philips PL-S, 365 nm) fluorescent lamps. Quartz microscope slides were used in method D, and were kept at ca. 4.5 cm away from the lamps during the experiments. All photochemical reactions in method C were performed inside a fully closed safety cabinet.[38] 1,8-Dihydroxynaphthalene (1,8-DHN, 5) was purchased from abcr Co. and used as received. Anhyd CH2Cl2 and THF were purchased from Acros Organics (AcroSeal®). All other commercially available reagents were used as received unless stated otherwise.


#

Synthesis of Symmetrical Diesters 6a–g; General Procedure I (Method A)

To a solution of 1,8-DHN (5; 1.0 equiv) in anhyd CH2Cl2 (10 mL) in a 100-mL round-bottomed flask were added sequentially trans-cinnamic acid derivative (2.0 equiv), DCC (2.0 equiv), and DMAP (20 mol%) at 23 °C under N2. The resulting cloudy, heterogeneous mixture was stirred at 23 °C for 24 h. TLC analysis indicated full consumption of 1,8-DHN (5). The mixture was quenched with H2O (10 mL), and the aqueous phase was extracted with CH2Cl2 (3 ×) The combined organic phases were dried (anhyd Na2SO4), filtered, and concentrated under vacuum. The crude reaction mixture was purified by flash column chromatography.


#

Naphthalene-1,8-diyl (2E,2′E)-Bis(3-(benzo[d][1,3]dioxol-5-yl)acrylate) (6c)

Diester 6c was synthesized using 1,8-DHN (5; 100 mg, 0.62 mmol), (E)-3-(benzo[d][1,3]dioxol-5-yl)acrylic acid (240 mg, 1.25 mmol), DCC (258 mg, 1.25 mmol), DMAP (15.3 mg, 0.125 mmol), and anhyd CH2Cl2 (10 mL) following General Procedure I. The crude product was purified by flash column chromatography (SiO2, CH2Cl2) to afford pure 6c (234 mg, 74% yield) as a white solid; mp 253–254 °C; Rf = 0.56 (CH2Cl2).

IR (ATR, film): 2921, 1728, 1703, 1631, 1603, 1501, 1452, 1369 cm–1.

1H NMR (DMSO-d 6, 400 MHz): δ = 7.97 (d, J = 8.3 Hz, 2 H), 7.65 (d, J = 15.9 Hz, 2 H), 7.59 (t, J = 7.9 Hz, 2 H), 7.29 (d, J = 7.5 Hz, 2 H), 7.02–7.00 (m, 4 H), 6.74 (d, J = 7.8 Hz, 2 H), 6.56 (d, J = 15.9 Hz, 2 H), 6.02 (s, 4 H).

13C{1H} NMR (DMSO-d 6, 100 MHz): δ = 165.5, 149.5, 147.7, 146.4, 144.9, 136.2, 128.0, 126.6, 126.4, 125.5, 121.1, 120.9, 114.8, 108.0, 106.4, 101.6.

HRMS (ESI): m/z [M + Na]+ calcd for C30H20NaO8: 531.1050; found: 531.1050.


#

Naphthalene-1,8-diyl (2E,2′E)-Bis(3-(4-chlorophenyl)acrylate) (6d)

Diester 6d was synthesized using 1,8-DHN (5; 100 mg, 0.62 mmol), (E)-3-(4-chlorophenyl)acrylic acid (228 mg, 1.25 mmol), DCC (258 mg, 1.25 mmol), DMAP (15.3 mg, 0.125 mmol), and anhyd CH2Cl2 (10 mL) following General Procedure I. The crude product was purified by flash column chromatography (SiO2, CH2Cl2/hexanes 1:1) to afford pure 6d (245 mg, 81% yield) as a white solid; mp 247–248 °C; Rf = 0.74 (CH2Cl2).

IR (ATR, film): 3082, 1729, 1658, 1640, 1512 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 7.84 (d, J = 8.3 Hz, 2 H), 7.76 (d, J = 16.1 Hz, 2 H), 7.51 (t, J = 7.9 Hz, 2 H), 7.21–7.18 (m, 6 H), 7.12 (d, J = 8.3 Hz, 4 H), 6.54 (d, J = 16.1 Hz, 2 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 165.9, 146.8, 145.4, 141.9, 136.9, 135.7, 129.3, 128.8, 127.4, 126.9, 126.1, 126.0, 120.713.

HRMS (ESI): m/z [M + Na]+ calcd for C28H18 35Cl2NaO4: 511.0474; found: 511.0469; calcd for C28H18 35Cl37ClNaO4: 513.0445; found: 513.0438.


#

Naphthalene-1,8-diyl (2E,2′E)-Bis(3-(thiophen-2-yl)acrylate) (6e)

Diester 6e was synthesized using 1,8-DHN (5; 50 mg, 0.31 mmol), (E)-3-(thiophen-2-yl)acrylic acid (96 mg, 0.62 mmol), DCC (129 mg, 0.62 mmol), DMAP (7.7 mg, 0.062 mmol), and anhyd CH2Cl2 (5 mL) following General Procedure I. The crude product was purified by flash column chromatography (SiO2, CH2Cl2/hexanes 1:1) to afford pure 6e (98 mg, 73% yield) as an orange solid; mp 208–209 °C; Rf = 0.17 (CH2Cl2/ hexanes 1:1).

IR (ATR, film): 3060, 1726, 1623, 1603, 1422, 1363 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 7.95 (d, J = 15.7 Hz, 2 H), 7.82 (d, J = 8.4 Hz, 2 H), 7.50 (t, J = 7.9 Hz, 2 H), 7.23 (d, J = 5.0 Hz, 2 H), 7.20 (d, J = 7.5 Hz, 2 H), 7.11 (d, J = 3.5 Hz, 2 H), 6.90 (app t, J = 4.6 Hz, 2 H), 6.42 (d, J = 15.7 Hz, 2 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 165.9, 145.3, 139.44, 139.38, 136.9, 131.5, 129.3, 128.0, 127.0, 126.2, 121.6, 120.7, 116.0.

HRMS (ESI): m/z [M + H]+ calcd for C24H17O4S2: 433.0563; found: 433.0562.


#

Naphthalene-1,8-diyl (2E,2′E)-Bis(3-(furan-2-yl)acrylate) (6f)

Diester 6f was synthesized using 1,8-DHN (5; 100 mg, 0.62 mmol), (E)-3-(furan-2-yl)acrylic acid (173 mg, 1.25 mmol), DCC (258 mg, 1.25 mmol), DMAP (15.3 mg, 0.125 mmol), and anhyd CH2Cl2 (10 mL) following General Procedure I. The crude product was purified by flash column chromatography (SiO2, 1% MeOH in CH2Cl2/hexanes 1:1) to afford pure 6f (200 mg, 81% yield) as a brown solid; mp 179–180 °C; Rf = 0.23 (CH2Cl2/hexanes 1:1).

IR (ATR, film): 3144, 3053, 1722, 1633, 1603, 1549, 1471, 1370 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 7.81 (d, J = 8.3 Hz, 2 H), 7.58 (d, J = 15.7 Hz, 2 H), 7.49 (t, J = 7.9 Hz, 2 H), 7.27 (s, 2 H), 7.20 (d, J = 7.5 Hz, 2 H), 6.53–6.49 (m, 4 H), 6.36 (app s, 2 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 166.0, 150.8, 145.34, 145.29, 136.9, 132.9, 126.9, 126.1, 121.5, 120.7, 115.5, 115.2, 112.3.

HRMS (ESI): m/z [M + H]+ calcd for C24H17O6: 401.1020; found: 401.1023.


#

Naphthalene-1,8-diyl (2E,2′E)-Bis(3-(1-methyl-1H-pyrrol-2-yl)acrylate) (6g)

Diester 6g was synthesized using 1,8-DHN (5; 85 mg, 0.53 mmol), (E)-3-(1-methyl-1H-pyrrol-2-yl)acrylic acid (200 mg, 1.32 mmol), DCC (252 mg, 1.22 mmol), DMAP (16 mg, 0.13 mmol), and anhyd CH2Cl2 (10 mL) following General Procedure I. The crude product was purified by flash column chromatography (SiO2, CH2Cl2/hexanes 1:1) to afford pure 6f (163 mg, 72% yield) as a red solid; mp 176–177 °C; Rf = 0.26 (CH2Cl2/hexanes 1:1).

IR (ATR, film): 2944, 1765, 1608, 1577, 1492 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 7.80 (dd, J = 8.3, 0.8 Hz, 2 H), 7.75 (d, J = 15.7 Hz, 2 H), 7.49 (t, J = 7.9 Hz, 2 H), 7.19 (dd, J = 7.5, 0.9 Hz, 2 H), 6.67 (t, J = 2.0 Hz, 2 H), 6.48 (dd, J = 3.9, 1.5 Hz, 2 H), 6.33 (d, J = 15.7 Hz, 2 H), 6.08 (dd, J = 3.8, 2.6 Hz, 2 H), 3.56 (s, 6 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 167.0, 145.7, 136.9, 134.4, 129.1, 128.0, 126.8, 126.1, 121.9, 120.7, 113.8, 111.4, 109.8, 34.5.

HRMS (ESI): m/z [M + Na]+ calcd for C26H22N2NaO4: 449.1472; found: 449.1471.


#

Synthesis of Unsymmetrical Diesters 6j–l, 6n, and 6p; General Procedure II (Method B)

In a 50-mL round-bottomed flask, monoester 8a or 8b (1 equiv) was dissolved in anhyd THF (10 mL) under N2, and the resulting solution was cooled to 0 °C in an ice bath. Cinnamoyl chloride derivative (1.0 equiv) was added, and the mixture was stirred for 5 min. Then, NaH (1.1 equiv, 60% in mineral oil) was added carefully. The ice bath was removed, and the mixture was stirred at 23 °C for 1.5 h. It was then quenched with sat. aq NH4Cl soln (10 mL), and the aqueous phase was extracted with CH2Cl2 (3 ×). The combined organic phases were dried (anhyd Na2SO4), filtered, and concentrated under vacuum. The crude reaction mixture was purified by flash column chromatography.


#

8-(Cinnamoyloxy)naphthalen-1-yl (E)-3-(4-Nitrophenyl)acrylate (6j)

Diester 6j was prepared using monoester 8a (90 mg, 0.31 mmol), NaH (13.7 mg, 0.34 mmol, 60% in mineral oil), (E)-3-(4-nitrophenyl)acryloyl chloride (66 mg, 0.31 mmol), and anhyd THF (15 mL) following General Procedure II. The crude product was purified by flash column chromatography (SiO2, CH2Cl2/hexanes 2:1) to afford pure 6j (125 mg, 86% yield) as a white solid; mp 222–223 °C; Rf = 0.33 (CH2Cl2/hexanes 1:1).

IR (ATR, film): 3058, 2930, 1731, 1634, 1600, 1512, 1344 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 7.89–7.81 (m, 6 H), 7.55–7.50 (m, 2 H), 7.35 (d, J = 8.7 Hz, 2 H), 7.30–7.28 (m, 3 H), 7.22 (d, J = 7.5 Hz, 2 H), 7.15 (t, J = 7.6 Hz, 2 H), 6.71 (d, J = 16.1 Hz, 1 H), 6.59 (d, J = 16.1 Hz, 1 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 165.8, 165.0, 148.6, 147.0, 145.2, 145.0, 143.8, 139.8, 137.0, 133.8, 131.0, 129.0, 128.6, 128.2, 127.3, 127.2, 126.4, 126.2, 124.0, 121.8, 121.3, 120.9, 120.7, 117.5.

HRMS (ESI): m/z [M + Na]+ calcd for C28H19NNaO6: 488.1105; found: 488.1101.


#

8-(Cinnamoyloxy)naphthalen-1-yl (E)-3-(3-Nitrophenyl)acrylate (6k)

Diester 6k was prepared using monoester 8a (72 mg, 0.25 mmol), NaH (11 mg, 0.27 mmol, 60% in mineral oil), (E)-3-(3-nitrophenyl)acryloyl chloride (55 mg, 0.26 mmol), and anhyd THF (12 mL) following General Procedure II. The crude product was purified by flash column chromatography (SiO2, 3% MeOH in CH2Cl2/hexanes 1:1) to afford pure 6k (85 mg, 74% yield) as a white solid; mp 228–229 °C; Rf = 0.35 (3% MeOH in CH2Cl2/hexanes 1:1).

IR (ATR, film): 3085, 3055, 1746, 1723, 1629, 1601, 1575, 1523, 1375, 1346, 1305 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 8.04–8.03 (m, 2 H), 7.85–7.84 (m, 4 H), 7.57–7.50 (m, 3 H), 7.31–7.26 (m, 3 H), 7.22–7.17 (m, 3 H), 7.07 (t, J = 7.5 Hz, 2 H), 6.72 (d, J = 16.0 Hz, 1 H), 6.60 (d, J = 16.0 Hz, 1 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 165.7, 165.1, 148.5, 146.9, 145.2, 145.1, 143.8, 137.0, 135.5, 133.7, 133.5, 130.8, 129.8, 128.9, 128.1, 127.3, 127.1, 126.4, 126.2, 124.8, 122.5, 121.4, 120.9, 120.7, 117.5.

HRMS (ESI): m/z [M + Na]+ calcd for C28H19NNaO6: 488.1105; found: 488.1095.


#

8-(Cinnamoyloxy)naphthalen-1-yl (E)-3-(2-Nitrophenyl)acrylate (6l)

Diester 6l was prepared using monoester 8a (65 mg, 0.22 mmol), NaH (9.9 mg, 0.25 mmol, 60% in mineral oil), (E)-3-(2-nitrophenyl)acryloyl chloride (50 mg, 0.24 mmol), and anhyd THF (5 mL) following General Procedure II. The crude product was purified by flash column chromatography (SiO2, CH2Cl2/hexanes 1:1) to afford pure 6l (75 mg, 72% yield) as a white solid; mp 198–199 °C; Rf = 0.21 (CH2Cl2/hexanes 1:1).

IR (ATR, film): 3064, 2961, 1918, 1717, 1635, 1603, 1570, 1517, 1335 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 8.27 (d, J = 15.8 Hz, 1 H), 7.91 (dd, J = 8.2, 1.0 Hz, 1 H), 7.86 (d, J = 15.7 Hz, 1 H), 7.84 (d, J = 8.3 Hz, 2 H), 7.54–7.50 (m, 2 H), 7.38–7.35 (m, 3 H), 7.30–7.11 (m, 7 H), 6.67 (d, J = 16.1 Hz, 1 H), 6.54 (d, J = 15.8 Hz, 1 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 165.7, 164.8, 148.3, 146.9, 145.3, 145.1, 141.9, 137.0, 134.0, 133.3, 130.8, 130.5, 129.9, 129.0, 128.8, 128.3, 127.2, 127.1, 126.3, 126.2, 125.0, 122.5, 121.4, 120.8, 120.7, 117.7.

HRMS (ESI): m/z [M + Na]+ calcd for C28H19NNaO6: 488.1105; found: 488.1092.


#

8-(Cinnamoyloxy)naphthalen-1-yl (E)-3-(Thiophen-2-yl)acrylate (6n)

Diester 6n was prepared using monoester 8a (88 mg, 0.30 mmol), NaH (13.3 mg, 0.33 mmol, 60% in mineral oil), (E)-3-(thiophen-2-yl)acryloyl chloride (55 mg, 0.32 mmol), and anhyd THF (5 mL) following General Procedure II. The crude product was purified by flash column chromatography (SiO2, CHCl3/hexanes 1:1) to afford pure 6n (96 mg, 75% yield) as a white solid; mp 208–209 °C; Rf = 0.16 (CH2Cl2/ hexanes 1:1).

IR (ATR, film): 3059, 1715, 1638, 1620, 1600, 1574, 1512 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 7.93 (d, J = 15.7 Hz, 1 H), 7.85 (d, J = 16.0 Hz, 1 H), 7.80 (d, J = 8.3 Hz, 2 H), 7.47 (t, J = 7.7 Hz, 2 H), 7.31 (d, J = 7.5 Hz, 2 H), 7.25 (t, J = 7.3 Hz, 1 H), 7.21–7.13 (m, 5 H), 7.04 (d, J = 3.3 Hz, 1 H), 6.83 (t, J = 4.3 Hz, 1 H), 6.62 (d, J = 16.0 Hz, 1 H), 6.40 (d, J = 15.7 Hz, 1 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 165.9, 165.8, 147.1, 145.28, 145.26, 139.3, 139.1, 136.9, 134.0, 131.5, 130.5, 129.3, 128.8, 128.3, 128.1, 126.9, 126.2, 121.5, 120.70, 120.67, 117.3, 115.9.

HRMS (ESI): m/z [M + Na]+ calcd for C26H18NaO4S: 449.0818; found: 449.0832.


#

8-(((E)-3-(Furan-2-yl)acryloyl)oxy)naphthalen-1-yl (E)-3-(4-Meth­oxyphenyl)acrylate (6p)

Diester 6p was prepared using monoester 8b (75 mg, 0.23 mmol), NaH (9.4 mg, 0.24 mmol, 60% in mineral oil), (E)-3-(furan-2-yl)acryloyl chloride (40 mg, 0.26 mmol), and anhyd THF (15 mL) following General Procedure II. The crude product was purified by flash column chromatography (SiO2, 1% MeOH in CH2Cl2/hexanes 1:1) to afford pure 6p (80 mg, 77% yield) as a pale brown solid; mp 138–140 °C; Rf = 0.30 (CH2Cl2/hexanes 1:1).

IR (ATR, film): 3059, 2932, 2839, 1729, 1635, 1601, 1512 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 7.82–7.79 (m, 3 H), 7.58 (d, J = 15.7 Hz, 1 H), 7.50 (t, J = 7.9 Hz, 1 H), 7.49 (t, J = 7.9 Hz, 1 H), 7.33 (d, J = 8.7 Hz, 2 H), 7.22–7.18 (m, 3 H), 6.75 (d, J = 8.7 Hz, 2 H), 6.52 (d, J = 4.3 Hz, 1 H), 6.49–6.47 (m, 2 H), 6.33 (dd, J = 3.3, 1.8 Hz, 1 H), 3.82 (s, 3 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 166.3, 166.1, 161.6, 150.7, 146.7, 145.4, 145.2, 136.9, 132.8, 130.0, 127.0, 126.9, 126.8, 126.2, 126.1, 121.6, 120.8, 120.7, 115.8, 115.2, 115.0, 114.3, 112.3, 55.5.

HRMS (ESI): m/z [M + Na]+ calcd for C27H20NaO6: 463.1152; found: 463.1166.


#

Photochemical Cycloaddition Reactions of 6; General Procedure III (Method C)

A solid powder sample of diester 6 was placed between two soda-lime glass microscope slides. The sample was irradiated with a 400-W broadband medium-pressure Hg lamp for 8 h inside a safety box. The solid powder was mixed with a spatula to ensure homogeneity every 2 h. At the end of 8 h, the sample was analyzed by 1H NMR spectroscopy.


#

Photochemical Cycloaddition Reactions of 6; General Procedure IV (Method D)

A solid powder sample of diester 6 was placed between two quartz glass microscope slides. The sample was irradiated inside a UV gel nail dryer, equipped with four 9-W UVA fluorescent bulbs, for 16 h. The solid powder was mixed with a spatula to ensure homogeneity every 4 h. At the end of 16 h, the sample was analyzed by 1H NMR spectroscopy.


#

(8aR,9S,10R,10aS)-9,10-Bis(benzo[d][1,3]dioxol-5-yl)-8a,9,10,10a-tetrahydrocyclobuta[g]naphtho[1,8-bc][1,5]dioxonine-8,11-dione (meso-9c)

Cycloadduct 9c was synthesized using diester 6c (24.9 mg, 0.049 mmol) following General Procedure III (irradiation time = 8 h) to give pure product (23.1 mg, 93% yield) as a brown solid.

Cycloadduct 9c was also synthesized using diester 6c (24.8 mg, 0.049 mmol) following General Procedure IV (irradiation time = 16 h). The crude product was purified by flash column chromatography (SiO2, CH2Cl2/hexanes 1:1) to afford pure 9c (23.9 mg, 96% yield) as an orange solid; mp 209–213 °C; Rf = 0.27 (CH2Cl2/hexanes 1:1).

IR (ATR, film): 3060, 2957, 2900, 1764, 1608, 1504, 1492, 1444 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 7.80 (d, J = 8.3 Hz, 2 H), 7.50 (t, J = 7.9 Hz, 2 H), 7.26 (d, J = 7.9 Hz, 2 H), 6.66 (d, J = 7.9 Hz, 2 H), 6.54 (d, J = 8.0 Hz, 2 H), 6.50 (s, 2 H), 5.88 (s, 4 H), 4.61 (app d, J = 6.0 Hz, 2 H), 4.11 (app d, J = 6.1 Hz, 2 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 170.0, 147.8, 146.4, 145.4, 137.1, 132.1, 127.1, 126.5, 121.09, 121.05, 119.6, 108.4, 108.2, 101.1, 45.2, 44.1.

HRMS (ESI): m/z [M + Na]+ calcd for C30H20NaO8: 531.1050; found: 531.1035.


#

(8aR,9S,10R,10aS)-9,10-Bis(4-chlorophenyl)-8a,9,10,10a-tetrahydrocyclobuta[g]naphtho[1,8-bc][1,5]dioxonine-8,11-dione (meso-9d)

Cycloadduct 9d was synthesized using diester 6d (25.1 mg, 0.051 mmol) following General Procedure III (irradiation time = 8 h) to give pure product (22.3 mg, 89% yield) as a white solid.

Cycloadduct 9d was also synthesized using diester 6d (25.5 mg, 0.052 mmol) following General Procedure IV (irradiation time = 16 h) to give pure product (23.0 mg, 90% yield); mp 216–218 °C; Rf = 0.41 (CH2Cl2).

IR (ATR, film): 3060, 2925, 1765, 1607, 1577, 1366 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 7.83 (d, J = 8.3 Hz, 2 H), 7.52 (t, J = 7.9 Hz, 2 H), 7.28 (d, J = 7.4 Hz, 2 H), 7.16 (d, J = 8.3 Hz, 4 H), 6.93 (d, J = 8.3 Hz, 4 H), 4.72 (app d, J = 5.9 Hz, 2 H), 4.17 (app d, J = 5.9 Hz, 2 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 169.7, 145.4, 137.1, 136.4, 132.9, 129.2, 128.7, 127.2, 126.5, 121.1, 119.5, 44.8, 43.6.

HRMS (ESI): m/z [M + Na]+ calcd for C28H18 35Cl2NaO4: 511.0474; found: 511.0461.


#

(8aR,9S,10R,10aS)-9,10-Di(thiophen-2-yl)-8a,9,10,10a-tetrahydrocyclobuta[g]naphtho[1,8-bc][1,5]dioxonine-8,11-dione (meso-9e)

Cycloadduct 9e was synthesized using diester 6e (19.2 mg, 0.044 mmol) following General Procedure III (irradiation time = 8 h) to give pure product (18.5 mg, 96% yield).

Cycloadduct 9e was also synthesized using diester 6e (25.2 mg, 0.058 mmol) following General Procedure IV (irradiation time = 16 h). The crude product was purified by flash column chromatography (SiO2, CH2Cl2/hexanes 1:1) to afford pure 9e (24.8 mg, 98% yield) as a pale-yellow solid; mp 193–195 °C; Rf = 0.45 (CH2Cl2/hexanes 1:1).

IR (ATR, film): 2926, 1763, 1608, 1577, 1365 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 7.83 (d, J = 8.3 Hz, 2 H), 7.52 (t, J = 7.8 Hz, 2 H), 7.29 (d, J = 7.4 Hz, 2 H), 7.15 (d, J = 5.1 Hz, 2 H), 6.91 (t, J = 4.0 Hz, 2 H), 6.84 (d, J = 3.1 Hz, 2 H), 4.90 (d, J = 5.5 Hz, 2 H), 4.23 (d, J = 5.5 Hz, 2 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 169.2, 145.3, 141.0, 137.1, 127.1, 126.9, 126.5, 125.8, 125.2, 121.1, 119.5, 46.8, 41.0.

HRMS (ESI): m/z [M + Na]+ calcd for C24H16NaO4S2: 455.0382; found: 455.0385.


#

(8aR,9S,10R,10aS)-9,10-Di(furan-2-yl)-8a,9,10,10a-tetrahydrocyclobuta[g]naphtho[1,8-bc][1,5]dioxonine-8,11-dione (meso-9f)

Cycloadduct 9f was synthesized using diester 6f (21.6 mg, 0.054 mmol) following General Procedure III (irradiation time = 8 h). The crude product was purified by flash column chromatography (SiO2, CH2Cl2/hexanes 1:1) to afford pure 9f (18.6 mg, 86% yield).

Cycloadduct 9f was also synthesized using diester 6f (25.7 mg, 0.064 mmol) following General Procedure IV (irradiation time = 16 h). The crude product was purified by flash column chromatography (SiO2, CH2Cl2/hexanes 1:1) to afford pure 9f (25.2 mg, 98% yield) as a brown-orange solid; mp 164–165 °C; Rf = 0.49 (CH2Cl2/hexanes 1:1).

IR (ATR, film): 2922, 1765, 1608, 1504, 1363 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 7.80 (d, J = 8.3 Hz, 2 H), 7.50 (t, J = 7.9 Hz, 2 H), 7.28–7.25 (m, 4 H), 6.25 (dd, J = 3.0, 1.8 Hz, 2 H), 6.04 (d, J = 3.1 Hz, 2 H), 4.61 (d, J = 5.8 Hz, 2 H), 4.28 (d, J = 5.7 Hz, 2 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 169.5, 152.0, 145.3, 142.2, 137.1, 127.1, 126.5, 121.1, 119.5, 110.5, 107.6, 44.3, 38.0.

HRMS (ESI): m/z [M + Na]+ calcd for C24H16NaO6: 423.0839; found: 423.0839.


#

(8aR,9S,10R,10aS)-9,10-Bis(1-methyl-1H-pyrrol-2-yl)-8a,9,10,10a-tetrahydrocyclobuta[g]naphtho[1,8-bc][1,5]dioxonine-8,11-dione (meso-9g)

Cycloadduct 9g was synthesized using diester 6g (25.3 mg, 0.059 mmol) following General Procedure III (irradiation time = 8 h). The crude product was purified by flash column chromatography (SiO2, CH2Cl2/hexanes 1:1) to afford pure 9g (24.0 mg, 95% yield) as a pale purple solid.

Cycloadduct 9g was also synthesized using diester 6g (18.7 mg, 0.044 mmol) following General Procedure IV (irradiation time = 16 h). The crude product was purified by flash column chromatography (SiO2, CH2Cl2/hexanes 1:1) to afford pure 9g (15.0 mg, 80% yield) as a purple solid; mp 224–227 °C; Rf = 0.68 (CH2Cl2/hexanes 1:1).

IR (ATR, film): 3061, 2916, 1732, 1633, 1605, 1576 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 7.81 (d, J = 8.3 Hz, 2 H), 7.51 (t, J = 7.8 Hz, 2 H), 7.26 (d, J = 8.1 Hz, 2 H), 6.53 (br s, 2 H), 6.04 (t, J = 2.7 Hz, 2 H), 5.79–5.74 (m, 2 H), 4.56 (app d, J = 5.3 Hz, 2 H), 4.08 (app d, J = 5.3 Hz, 2 H), 3.35 (s, 6 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 169.9, 145.5, 137.1, 129.8, 127.1, 126.5, 122.5, 121.1, 119.7, 107.4, 107.1, 46.2, 36.2, 33.7.

HRMS (ESI): m/z [M + Na]+ calcd for C26H22N2NaO4: 449.1472; found: 449.1462.


#

(8aR,9S,10R,10aS)-9-(4-Nitrophenyl)-10-phenyl-8a,9,10,10a-tetrahydrocyclobuta[g]naphtho[1,8-bc][1,5]dioxonine-8,11-dione (rac-9j)

Cycloadduct 9j was synthesized using diester 6j (24.3 mg, 0.052 mmol) following General Procedure III (irradiation time = 8 h). The crude product was purified by flash column chromatography (SiO2, CH2Cl2/hexanes 1:1) to afford pure 9j (18.8 mg, 77% yield) as a gray solid.

Cycloadduct 9j was also synthesized using diester 6j (23.8 mg, 0.051 mmol) following General Procedure IV (irradiation time = 16 h). The crude product was purified by flash column chromatography (SiO2, CH2Cl2/hexanes 1:1) to afford pure 9j (14.4 mg, 61% yield); mp 219–220 °C; Rf = 0.24 (CH2Cl2/hexanes 1:1).

IR (ATR, film): 2924, 2853, 1760, 1606, 1515, 1343 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 8.00 (d, J = 8.5 Hz, 2 H), 7.84 (d, J = 8.2 Hz, 2 H), 7.533 (t, J = 7.9 Hz, 1 H), 7.528 (t, J = 7.9 Hz, 1 H), 7.30 (d, J = 7.4 Hz, 2 H), 7.21–7.12 (m, 5 H), 7.02 (d, J = 7.1 Hz, 2 H), 4.88 (dd, J = 10.5, 5.5 Hz, 1 H), 4.82 (dd, J = 10.5, 4.9 Hz, 1 H), 4.31–4.24 (m, 2 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 169.62, 169.58, 146.7, 145.9, 145.3, 137.3, 137.1, 128.8, 128.7, 127.8, 127.5, 127.3, 127.2, 127.0, 126.5, 123.5, 121.1, 121.0, 119.4, 44.8, 44.5, 44.3, 44.2.

HRMS (ESI): m/z [M + H]+ calcd for C28H19NNaO6: 488.1105; found: 488.1115.


#

(8aR,9S,10R,10aS)-9-(3-Nitrophenyl)-10-phenyl-8a,9,10,10a-tetrahydrocyclobuta[g]naphtho[1,8-bc][1,5]dioxonine-8,11-dione (rac-9k)

Cycloadduct 9k was synthesized using diester 6k (17.0 mg, 0.037 mmol) following General Procedure III (irradiation time = 8 h). The crude product was purified by flash column chromatography (SiO2, CH2Cl2/hexanes 1:1) to afford pure 9k (14.9 mg, 88% yield) as a white solid.

Cycloadduct 9k was also synthesized using diester 6k (26.7 mg, 0.057 mmol) following General Procedure IV (irradiation time = 16 h). The crude product was purified by flash column chromatography (SiO2, CH2Cl2/hexanes 1:1) to afford pure 9k (23.0 mg, 86% yield); mp 185–187 °C; Rf = 0.24 (CH2Cl2/hexanes 1:1).

IR (ATR, film): 2954, 2921, 2852, 1761, 1746, 1606, 1526, 1347 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 7.98–7.95 (m, 1 H), 7.91 (s, 1 H), 7.84 (d, J = 8.3 Hz, 2 H), 7.53 (t, J = 7.9 Hz, 2 H), 7.33–7.30 (m, 4 H), 7.19 (t, J = 7.3 Hz, 2 H), 7.11 (t, J = 7.3 Hz, 1 H), 7.03 (d, J = 7.2 Hz, 2 H), 4.88 (dd, J = 9.9, 4.6 Hz, 1 H), 4.82 (dd, J = 9.6, 3.4 Hz, 1 H), 4.33–4.27 (m, 2 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 169.7, 169.6, 148.2, 145.3, 140.4, 137.1, 134.1, 129.2, 128.8, 127.9, 127.4, 127.23, 127.19, 126.6, 126.5, 122.7, 121.9, 121.12, 121.06, 119.4, 44.48, 44.46, 44.2, 44.0.

HRMS (ESI): m/z [M + Na]+ calcd for C28H19NNaO6: 488.1105; found: 488.1103.


#

(8aR,9S,10R,10aS)-9-(2-Nitrophenyl)-10-phenyl-8a,9,10,10a-tetrahydrocyclobuta[g]naphtho[1,8-bc][1,5]dioxonine-8,11-dione (rac-9l)

Cycloadduct 9l was synthesized using diester 6l (24.7 mg, 0.053 mmol) following General Procedure III (irradiation time = 8 h). The crude product was purified by flash column chromatography (SiO2, CH2Cl2/hexanes 1:1) to afford pure 9l (10.8 mg, 44% yield) as a pale-yellow solid.

Cycloadduct 9l was also synthesized using diester 6l (24.3 mg, 0.052 mmol) following General Procedure IV (irradiation time = 16 h). The crude product was purified by flash column chromatography (SiO2, CH2Cl2/hexanes 1:1) to afford pure 9l (11.8 mg, 49% yield); mp 191–193 °C; Rf = 0.53 (CH2Cl2/hexanes 1:1).

IR (ATR, film): 2922, 1761, 1606, 1574, 1525, 1359 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 7.83 (d, J = 8.3 Hz, 3 H), 7.59–7.50 (m, 3 H), 7.46 (d, J = 7.7 Hz, 1 H), 7.34–7.28 (m, 3 H), 7.15–7.08 (m, 3 H), 7.02 (d, J = 7.7 Hz, 2 H), 5.48 (t, J = 9.6 Hz, 1 H), 4.84 (dd, J = 10.2, 4.8 Hz, 1 H), 4.52 (t, J = 10.0 Hz, 1 H), 4.01 (dd, J = 10.8, 4.9 Hz, 1 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 170.0, 169.8, 148.2, 145.37, 145.35, 137.9, 137.1, 134.3, 133.4, 128.8, 128.6, 128.4, 127.9, 127.3, 127.2, 127.1, 126.6, 126.5, 125.4, 121.2, 121.1, 119.4, 45.2, 44.6, 44.1, 43.0.

HRMS (ESI): m/z [M + H]+ calcd for C28H20NO6: 466.1285; found: 466.1277.


#

(8aS,9S,10S,10aS)-9-Phenyl-10-(thiophen-2-yl)-8a,9,10,10a-tetrahydrocyclobuta[g]naphtho[1,8-bc][1,5]dioxonine-8,11-dione (rac-9n)

Cycloadduct 9n was synthesized using diester 6n (25.3 mg, 0.059 mmol) following General Procedure III (irradiation time = 8 h). The crude product was purified by flash column chromatography (SiO2, CH2Cl2/hexanes 1:1) to afford pure 9n (21.8 mg, 86% yield) as a white solid.

Cycloadduct 9n was also synthesized using diester 6n (25.7 mg, 0.060 mmol) following General Procedure IV (irradiation time = 16 h) to give pure product (24.8 mg, 96% yield); mp 202–203 °C; Rf = 0.30 (CH2Cl2/hexanes 1:1).

IR (ATR, film): 3057, 2925, 1752, 1606, 1576, 1497, 1365 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 7.81 (d, J = 8.3 Hz, 2 H), 7.51 (t, J = 7.8 Hz, 1 H), 7.50 (t, J = 7.8 Hz, 1 H), 7.29–7.21 (m, 4 H), 7.19–7.16 (m, 1 H), 7.11 (d, J = 7.3 Hz, 2 H), 7.05 (d, J = 5.0 Hz, 1 H), 6.82 (t, J = 4.2 Hz, 1 H), 6.73 (d, J = 3.0 Hz, 1 H), 4.91 (dd, J = 9.9, 5.9 Hz, 1 H), 4.76 (t, J = 8.8 Hz, 1 H), 4.33 (dd, J = 10.4, 7.8 Hz, 1 H), 4.14 (dd, J = 10.5, 5.9 Hz, 1 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 169.7, 169.6, 145.39, 145.37, 141.6, 137.6, 137.1, 128.4, 127.8, 127.14, 127.11, 126.5, 125.6, 124.9, 121.11, 121.05, 119.6, 47.0, 44.8, 44.4, 40.3.

HRMS (ESI): m/z [M + Na]+ calcd for C26H18NaO4S: 449.0818; found: 449.0832.


#

(8aR,9S,10R,10aS)-9-(Furan-2-yl)-10-(4-methoxyphenyl)-8a,9,10,10a-tetrahydrocyclobuta[g]naphtho[1,8-bc][1,5]dioxonine-8,11-dione (rac-9p)

Cycloadduct 9p was synthesized using diester 6p (25.8 mg, 0.059 mmol) following General Procedure III (irradiation time = 8 h). The crude product was purified by flash column chromatography (SiO2, CH2Cl2/hexanes 1:1) to afford pure 9p (14.7 mg, 57% yield).

Cycloadduct 9p was also synthesized using diester 6p (15.5 mg, 0.035 mmol) following General Procedure IV (irradiation time = 16 h). The crude product was purified by flash column chromatography (SiO2, CH2Cl2/hexanes 1:1) to afford pure 9p (13.0 mg, 84% yield) as a white solid; mp 154.8–155.3 °C; Rf = 0.32 (CH2Cl2/hexanes 1:1).

IR (ATR, film): 2929, 2838, 1765, 1609, 1515, 1364 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 7.81 (d, J = 7.8 Hz, 2 H), 7.51 (t, J = 7.8 Hz, 1 H), 7.50 (t, J = 7.8 Hz, 1 H), 7.29–7.25 (m, 2 H), 7.22 (m, 1 H), 7.03 (d, J = 8.7 Hz, 2 H), 6.77 (d, J = 8.5 Hz, 2 H), 6.19 (t, J = 2.2 Hz, 1 H), 6.03 (d, J = 3.1 Hz, 1 H), 4.66 (t, J = 9.2 Hz, 1 H), 4.57 (dd, J = 10.1, 4.8 Hz, 1 H), 4.33 (dd, J = 10.5, 8.5 Hz, 1 H), 4.19 (dd, J = 10.6, 4.8 Hz, 1 H), 3.76 (s, 3 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 170.2, 169.6, 158.6, 152.2, 145.4, 142.2, 137.1, 130.2, 128.5, 127.08, 127.05, 126.5, 121.1, 121.0, 119.6, 113.7, 110.4, 108.2, 55.3, 45.7, 43.8, 43.6, 38.5.

HRMS (ESI): m/z [M + Na]+ calcd for C27H20NaO6: 463.1152; found: 463.1160.


#

Hydrolysis of Cycloadducts 9; General Procedure V

To a solution of cycloadduct 9 (1.0 equiv) in THF (2.0 mL) in a 20-mL scintillation vial was added distilled water (1.0 mL) and KOH (19 equiv). The resulting mixture was stirred at 23 °C for 2 h and then quenched with 1.0 M HCl solution until pH 1–2. The aqueous phase was extracted with EtOAc (3 ×). The combined organic phases were dried (anhyd Na2SO4), filtered, and concentrated under vacuum. The crude mixture was purified by flash column chromatography.


#

(1R,2S,3R,4S)-3,4-Bis(benzo[d][1,3]dioxol-5-yl)cyclobutane-1,2-dicarboxylic Acid (meso-10c)

Dicarboxylic acid 10c was synthesized using cycloadduct 9c (24.2 mg, 0.048 mmol), KOH (51.2 mg, 0.91 mmol), THF (2 mL), and H2O (1 mL) following General Procedure V. The crude product was purified by flash column chromatography (SiO2, EtOAc/hexanes 1:1 → 0.5% AcOH in EtOAc/hexanes 1:1) to afford pure 10c (17.5 mg, 95% yield) as a dark brown solid; mp 151–152 °C; Rf = 0.11 (0.5% AcOH in EtOAc/hexanes 1:1).

IR (ATR, film): 3012, 2891, 1745, 1697, 1503, 1491, 1442 cm–1.

1H NMR (DMSO-d 6, 400 MHz): δ = 12.43 (br s, 2 H), 6.66 (d, J = 8.0 Hz, 2 H), 6.65 (d, J = 1.2 Hz, 2 H), 6.54 (d, J = 8.0, 1.2 Hz, 2 H), 5.88 (s, 4 H), 4.07 (d, J = 6.2 Hz, 2 H), 3.71 (d, J = 6.2 Hz, 2 H).

HRMS (ESI): m/z [M – H] calcd for C20H15O8: 383.0772; found: 383.0780.


#

(1R,2S,3R,4S)-3,4-Bis(4-chlorophenyl)cyclobutane-1,2-dicarboxylic Acid (meso-10d)

Dicarboxylic acid 10d was synthesized using cycloadduct 9d (22.3 mg, 0.046 mmol), KOH (48.5 mg, 0.86 mmol), THF (2 mL), and H2O (1 mL) following General Procedure V. The crude product was purified by flash column chromatography (SiO2, 0.5% AcOH in EtOAc/hexanes 1:1) to afford pure 10d (14.2 mg, 84% yield) as a pale brown solid; mp 178–180 °C; Rf = 0.15 (0.5% AcOH in EtOAc/hexanes 1:1). The 1H and 13C NMR spectroscopic data are in agreement with those reported in the literature.[30a]

IR (ATR, film): 3123 (br), 1710, 1493, 1425 cm–1.

1H NMR (DMSO-d 6, 400 MHz): δ = 12.45 (br s, 2 H), 7.16 (d, J = 8.2 Hz, 4 H), 7.07 (d, J = 8.0 Hz, 4 H), 4.22 (app d, J = 5.4 Hz, 2 H), 3.80 (app d, J = 5.5 Hz, 2 H).

13C{1H} NMR (DMSO-d 6, 100 MHz): δ = 173.8, 138.2, 130.7, 129.8, 127.7, 43.8, 42.4.

HRMS (ESI): m/z [M – H] calcd for C18H13 35Cl2O4: 363.0196; found: 363.0196; calcd for C18H13 35Cl37ClO4: 365.0167; found: 365.0166.


#

(1R,2S,3R,4S)-3,4-Di(thiophen-2-yl)cyclobutane-1,2-dicarboxylic Acid (meso-10e)

Dicarboxylic acid 10e was synthesized using cycloadduct 9e (24.8 mg, 0.057 mmol), KOH (48.5 mg, 0.86 mmol), THF (2 mL), and H2O (1 mL) following General Procedure V. The crude product was purified by flash column chromatography (SiO2, EtOAc/hexanes 1:1 → 0.5% AcOH in EtOAc/hexanes 1:1) to afford pure 10e (16.0 mg, 90% yield) as a pale brown solid; mp 173–175 °C; Rf = 0.13 (0.5% AcOH in EtOAc/hexanes 1:1).

IR (ATR, film): 2925, 1708, 1419, 1239 cm–1.

1H NMR (DMSO-d 6, 400 MHz): δ = 12.60 (br s, 2 H), 7.24 (d, J = 4.8 Hz, 2 H), 6.86–6.83 (m, 4 H), 4.31 (app d, J = 6.1 Hz, 2 H), 3.70 (app d, J = 6.1 Hz, 2 H).

13C{1H} NMR (DMSO-d 6, 100 MHz): δ = 173.1, 142.0, 126.6, 125.4, 124.9, 45.1, 40.7.

HRMS (ESI): m/z [M – H] calcd for C14H11O4S2: 307.0104; found: 307.0119.


#

(1R,2S,3R,4S)-3,4-Di(furan-2-yl)cyclobutane-1,2-dicarboxylic Acid (meso-10f)

Dicarboxylic acid 10f was synthesized using cycloadduct 9f (24.0 mg, 0.060 mmol), KOH (62.8 mg, 1.12 mmol), THF (2 mL), and H2O (1 mL) following General Procedure V. The crude product was purified by flash column chromatography (SiO2, 0.5% AcOH in EtOAc/hexanes 1:1) to afford pure 10f (11.8 mg, 71% yield) as a pale-yellow solid; mp 175–177 °C; Rf = 0.24 (0.5% AcOH in EtOAc/hexanes 1:1). The 1H NMR spectrum in DMSO-d 6 is in agreement with the spectrum reported in the literature.[30c]

IR (ATR, film): 2924, 1710, 1505, 1427 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 9.16 (br s, 2 H), 7.24 (app s, 2 H), 6.21 (app t, J = 1.4 Hz, 2 H), 5.96 (d, J = 2.8 Hz, 2 H), 4.29 (app d, J = 5.6 Hz, 2 H), 3.97 (app d, J = 5.4 Hz, 2 H).

1H NMR (DMSO-d 6, 400 MHz): δ = 12.61 (br s, 2 H), 7.41–7.40 (m, 2 H), 6.25 (dd, J = 3.0, 1.9 Hz, 2 H), 6.10 (d, J = 3.2 Hz, 2 H), 4.05 (app d, J = 6.1 Hz, 2 H), 3.66 (app d, J = 6.1 Hz, 2 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 179.2, 152.2, 142.1, 110.5, 107.4, 43.6, 38.5.

HRMS (ESI): m/z [M + Na]+ calcd for C14H12NaO6: 299.0526; found: 299.0535.


#

Dimethyl (1R,2S,3R,4S)-3,4-Bis(1-methyl-1H-pyrrol-2-yl)cyclobutane-1,2-dicarboxylate (meso-11)

NaOMe (3.0 mg, 0.056 mmol) was added to a solution of cycloadduct 9g (11.8 mg, 0.028 mmol) in MeOH (4 mL) at 23 °C. The resulting mixture was stirred at 23 °C for 3.5 h at which time TLC analysis indicated complete consumption of the reactant. MeOH was then evaporated under reduced pressure and the crude mixture was mixed with CDCl3­. Ionic components including the sodium salt of 5 did not dissolve in CDCl3. The supernatant liquid was transferred to another flask, and all volatiles were removed under reduced pressure to afford pure 11 (6.0 mg, 65% yield) as a blackish amorphous solid; Rf = 0.43 (CH2Cl2/hexanes 1:1).

IR (ATR, film): 2950, 2928, 1735, 1435, 1354, 1208 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 6.48–6.47 (m, 2 H), 5.98 (dd, J = 3.5, 2.8 Hz, 2 H), 5.61 (dd, J = 3.6, 1.7 Hz, 2 H), 4.20 (app d, J = 6.0 Hz, 2 H), 3.72 (s, 6 H), 3.69 (app d, J = 5.9 Hz, 2 H), 3.28 (s, 6 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 173.1, 130.5, 122.2, 107.1, 106.7, 52.3, 45.2, 36.8, 33.6.

HRMS (APCI): m/z [M + H]+ calcd for C18H23N2O4: 331.1652; found: 331.1666.


#

(1S,2R,3S,4R)-3-(4-Nitrophenyl)-4-phenylcyclobutane-1,2-dicarboxylic Acid (rac-10j)

Dicarboxylic acid 10j was synthesized using cycloadduct 9j (14.4 mg, 0.031 mmol), KOH (32.9 mg, 0.59 mmol), THF (2 mL), and H2O (1 mL) following General Procedure V. The crude product was purified by flash column chromatography (SiO2, EtOAc/hexanes 1:1 → 0.5% AcOH in EtOAc/hexanes 1:1) to afford pure 10j (9.6 mg, 91% yield) as a brownish amorphous solid.; Rf = 0.18 (0.5% AcOH in EtOAc/hexanes 1:1).

IR (ATR, film): 3030, 2925, 2854, 1707, 1601, 1517, 1425, 1345 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 10.13 (br s, 2 H), 7.96 (d, J = 8.4 Hz, 2 H), 7.17–7.11 (m, 3 H), 7.06 (d, J = 8.4 Hz, 2 H), 6.93 (d, J = 7.0 Hz, 2 H), 4.64 (t, J = 8.8 Hz, 1 H), 4.40 (dd, J = 9.6, 5.5 Hz, 1 H), 4.02 (t, J = 8.8 Hz, 1 H), 3.91 (dd, J = 9.2, 5.4 Hz, 1 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 179.5, 178.9, 146.7, 146.0, 137.2, 128.8, 128.5, 127.8, 127.5, 123.5, 45.0, 44.6, 44.1, 43.4.

HRMS (ESI): m/z [M + Na]+ calcd for C18H15NNaO6: 364.0792; found: 364.0791.


#

(1S,2R,3S,4R)-3-(3-Nitrophenyl)-4-phenylcyclobutane-1,2-dicarboxylic Acid (rac-10k)

Dicarboxylic acid 10k was synthesized using cycloadduct 9k (20.0 mg, 0.043 mmol), KOH (46 mg, 0.82 mmol), THF (2 mL), and H2O (1 mL) following General Procedure V. The crude product was purified by flash column chromatography (SiO2, 0.5% AcOH in EtOAc/hexanes 1:1) to afford pure 10k (13.4 mg, 91% yield) as a yellow amorphous solid; Rf = 0.17 (0.5% AcOH in EtOAc/hexanes 1:1).

IR (ATR, film): 3030, 2925, 2855, 1706, 1527, 1422, 1346 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 8.25 (br s, 2 H), 7.93 (d, J = 7.9 Hz, 1 H), 7.80 (s, 1 H), 7.29–7.25 (m, 1 H), 7.22 (d, J = 7.7 Hz, 1 H), 7.17–7.14 (m, 2 H), 7.11–7.07 (m, 1 H), 6.96 (d, J = 7.2 Hz, 2 H), 4.66 (t, J = 9.0 Hz, 1 H), 4.41 (dd, J = 10.0, 5.4 Hz, 1 H), 4.04 (t, J = 9.1 Hz, 1 H), 3.95 (dd, J = 10.0, 5.7 Hz, 1 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 179.5, 178.8, 148.1, 140.5, 137.2, 134.0, 129.1, 128.8, 127.8, 127.4, 122.6, 121.9, 44.9, 44.4, 43.9, 43.5.

HRMS (ESI): m/z [M – H] calcd for C18H14NO6: 340.0827; found: 340.0812.


#

(1S,2R,3S,4R)-3-(2-Nitrophenyl)-4-phenylcyclobutane-1,2-dicarboxylic Acid (rac-10l)

Dicarboxylic acid 10l was synthesized using cycloadduct 9l (20.5 mg, 0.044 mmol), KOH (46.9 mg, 0.84 mmol), THF (2 mL), and H2O (1 mL) following General Procedure V. The crude product was purified by flash column chromatography (SiO2, 0.5% AcOH in EtOAc/hexanes 1:1 → 5% MeOH in EtOAc) to afford pure 10l (12.7 mg, 85% yield) as a white solid; mp 182–185 °C; Rf = 0.25 (0.5% AcOH in EtOAc/hexanes 1:1).

IR (ATR, film): 3055, 3032, 2924, 1709, 1524, 1423, 1346 cm–1.

1H NMR (CD3OD, 400 MHz): δ = 7.76 (d, J = 8.1 Hz, 1 H), 7.51 (t, J = 7.5 Hz, 1 H), 7.43 (d, J = 7.6 Hz, 1 H), 7.26 (t, J = 7.6 Hz, 1 H), 7.08–7.00 (m, 5 H), 5.09 (t, J = 10.1 Hz, 1 H), 4.33 (t, J = 10.2 Hz, 1 H), 4.18 (dd, J = 10.0, 3.7 Hz, 1 H), 3.59 (dd, J = 10.0, 3.7 Hz, 1 H).

13C{1H} NMR (CD3OD, 100 MHz): δ = 176.6, 175.8, 149.8, 140.1, 136.1, 134.1, 130.5, 129.3, 129.2, 128.4, 127.8, 125.6, 47.0, 46.0, 44.2, 42.6.

HRMS (ESI): m/z [M – H] calcd for C18H14NO6: 340.0827; found: 340.0825.


#

(1S,2S,3S,4S)-3-Phenyl-4-(thiophen-2-yl)cyclobutane-1,2-dicarboxylic Acid (rac-10n)

Dicarboxylic acid 10n was synthesized using cycloadduct 9n (19.8 mg, 0.046 mmol), KOH (49.5 mg, 0.88 mmol), THF (2 mL), and H2O (1 mL) following General Procedure V. The crude product was purified by flash column chromatography (SiO2, 0.1% AcOH in EtOAc/hexanes 1:1) to afford pure 10n (13.5 mg, 96% yield) as a pale brown solid; mp 181–182 °C; Rf = 0.30 (0.5% AcOH in EtOAc/hexanes 1:1).

IR (ATR, film): 3031, 2922, 1698, 1413, 1255 cm–1.

1H NMR (CD3OD, 400 MHz): δ = 7.16–7.12 (m, 2 H), 7.09–7.06 (m, 3 H), 7.03 (dd, J = 5.1, 1.0 Hz, 1 H), 6.75 (dd, J = 5.0, 3.5 Hz, 1 H), 6.68 (d, J = 3.4 Hz, 1 H), 4.46 (dd, J = 9.8, 6.3 Hz, 1 H), 4.32 (t, J = 8.7 Hz, 1 H), 3.94 (dd, J = 9.7, 7.8 Hz, 1 H), 3.74 (dd, J = 9.9, 6.3 Hz, 1 H).

13C{1H} NMR (CD3OD, 100 MHz): δ = 176.1, 175.8, 143.9, 139.9, 129.0, 128.9, 127.5, 127.4, 126.2, 125.0, 47.3, 46.8, 44.3, 42.2.

HRMS (APCI): m/z [M – H] calcd for C16H13O4S: 301.0540; found: 301.0521.


#

(1S,2R,3S,4R)-3-(Furan-2-yl)-4-(4-methoxyphenyl)cyclobutane-1,2-dicarboxylic Acid (rac-10p)

Dicarboxylic acid 10p was synthesized using cycloadduct 9p (8.2 mg, 0.019 mmol), KOH (19.8 mg, 0.35 mmol), THF (2 mL), and H2O (1 mL) following General Procedure V. The crude product was purified by flash column chromatography (SiO2, EtOAc/hexanes 1:1 → 0.5% AcOH in EtOAc/hexanes 1:1) to afford pure 10p (5.5 mg, 93% yield) as a pale brown amorphous solid; Rf = 0.21 (0.5% AcOH in EtOAc/hexanes 1:1).

IR (ATR, film): 3418 (br), 2917, 2849, 1720, 1514, 1463, 1249 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 9.53 (br s, 2 H), 7.18 (s, 1 H), 6.94 (d, J = 8.4 Hz, 2 H), 6.72 (d, J = 8.4 Hz, 2 H), 6.16 (br app s, 1 H), 5.95 (d, J = 2.9 Hz, 1 H), 4.41 (t, J = 9.4 Hz, 1 H), 4.18 (dd, J = 9.5, 4.7 Hz, 1 H), 4.05 (t, J = 9.6 Hz, 1 H), 3.82 (dd, J = 9.9, 4.8 Hz, 1 H), 3.73 (s, 3 H).

13C{1H} NMR (DMSO-d 6, 100 MHz): δ = 173.46, 173.43, 157.7, 153.3, 141.9, 131.1, 128.4, 113.1, 110.2, 107.2, 54.9, 43.4, 43.2, 42.3, 38.4.

HRMS (ESI): m/z [M – H] calcd for C17H15O6: 315.0874; found: 315.0871.


#

Dimethyl (1S,2S,3S,4S)-3-Phenyl-4-(thiophen-2-yl)cyclobutane-1,2-dicarboxylate (rac-12)

NaOMe (7.9 mg, 0.15 mmol) was added to a solution of cycloadduct 9n (31.2 mg, 0.073 mmol) in MeOH (2 mL) at 23 °C. The resulting mixture was stirred at 23 °C for 2 h at which time TLC analysis indicated complete consumption of the reactant. MeOH was then evaporated under reduced pressure. The remaining crude oil was dissolved in EtOAc (10 mL), and to this solution water (10 mL) and brine (5 mL) were added. The two phases were partitioned in a separatory funnel. The aqueous phase was further extracted with EtOAc (3 × 15 mL). The combined organic phases were dried (anhyd Na2SO4), filtered, and concentrated under reduced pressure. The crude product was purified by flash column chromatography (SiO2, EtOAc/hexanes 1:1) to afford pure 12 (22.5 mg, 93% yield) as yellow oil; Rf = 0.72 (EtOAc/hexanes 1:1).

IR (ATR, film): 2951, 1736, 1435, 1264, 1207 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 7.20–7.13 (m, 3 H), 7.02 (d, J = 7.0 Hz, 2 H), 7.00 (d, J = 5.1 Hz, 1 H), 6.77 (dd, J = 5.0, 3.6 Hz, 1 H), 6.61 (d, J = 3.5 Hz, 1 H), 4.55 (dd, J = 9.8, 6.2 Hz, 1 H), 4.41 (t, J = 8.8 Hz, 1 H), 3.95 (dd, J = 10.0, 7.7 Hz, 1 H), 3.76 (s, 3 H), 3.74 (s, 3 H), 3.78–3.71 (m, 1 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 172.8, 172.6, 142.3, 138.1, 128.2, 127.8, 126.9, 126.6, 125.3, 124.5, 52.4, 52.3, 46.1, 45.6, 42.9, 41.1.

HRMS (ESI): m/z [M + Na]+ calcd for C18H18NaO4S: 353.0818; found: 353.0830.


#

((1S,2S,3S,4S)-3-Phenyl-4-(thiophen-2-yl)cyclobutane-1,2-diyl)dimethanol (rac-13)

LiAlH4 (25.8 mg, 0.68 mmol) was added to a solution of cycloadduct 9n (29.1 mg, 0.068 mmol) in anhyd THF (2 mL) at 0 °C under N2. The resulting gray-colored mixture was stirred at 23 °C for 3 h at which time TLC analysis indicated complete consumption of the reactant. The mixture was then quenched carefully with water, and gas formation was observed. The aqueous phase was extracted with EtOAc (3 ×). The combined organic phases were dried (anhyd Na2SO4), filtered, and concentrated under reduced pressure. The crude product was purified by flash column chromatography (SiO2, 5% MeOH in CH2Cl2) to afford pure 13 (16.0 mg, 86% yield) as a brown oil; Rf = 0.23 (5% MeOH in CH2Cl2).

IR (ATR, film): 3317 (br), 2926, 2872, 1496, 1451, 1264 cm–1.

1H NMR (CDCl3, 400 MHz): δ = 7.18–7.08 (m, 3 H), 7.04 (d, J = 7.0 Hz, 2 H), 6.96 (d, J = 5.1 Hz, 1 H), 6.75 (dd, J = 5.0, 3.5 Hz, 1 H), 6.57 (d, J = 3.3 Hz, 1 H), 4.08–3.99 (m, 2 H), 3.93–3.82 (m, 3 H), 3.63 (t, J = 8.5 Hz, 1 H), 3.41 (br s, 2 H), 3.29–3.22 (m, 1 H), 3.14–3.07 (m, 1 H).

13C{1H} NMR (CDCl3, 100 MHz): δ = 144.0, 139.4, 128.2, 128.1, 126.6, 126.4, 124.7, 123.8, 62.4, 62.2, 44.8, 43.6, 40.2, 40.1.

HRMS (ESI): m/z [M – H] calcd for C16H17O2S: 273.0955; found: 273.0972.


#
#

Conflict of Interest

The authors declare no conflict of interest.

Acknowledgment

The authors thank Eylül Çalıkyılmaz for providing assistance in the diffuse reflectance UV-vis spectroscopic measurements.

Supporting Information

  • References

  • 1 Bernstein HI, Quimby WC. J. Am. Chem. Soc. 1943; 65: 1845 ; and references therein
  • 2 Bassani DM. The Dimerization of Cinnamic Acid Derivatives . In CRC Handbook of Photochemistry and Photobiology, 2nd ed. Horspool WM, Lenci F. CRC Press; Boca Raton: 2004: 20-1-20-20
  • 3 Fujiwara A, Nishi M, Yoshida S, Hasegawa M, Yasuma C, Ryo A, Suzuki Y. Phytochemistry 2016; 122: 139
  • 4 Chai XY, Song YL, Xu ZR, Shi HM, Bai CC, Bi D, Wen J, Li FF, Tu PF. J. Nat. Prod. 2008; 71: 814
  • 5 Carmignani M, Volpe AR, Monache FD, Botta B, Espinal R, De Bonnevaux SC, De Luca C, Botta M, Corelli F, Tafi A, Ripanti G, Monache GD. J. Med. Chem. 1999; 42: 3116
  • 6 Huang B, Zhang F, Yu G, Song Y, Wang X, Wang M, Gong Z, Su R, Jia Y. J. Med. Chem. 2016; 59: 3953
    • 7a Cohen MD, Schmidt GM. J. J. Chem. Soc. 1964; 1996
    • 7b Cohen MD, Schmidt GM. J, Sonntag FI. J. Chem. Soc. 1964; 2000

    • Reviews:
    • 7c Sonoda Y. [2+2]-Photocycloadditions in the Solid State . In CRC Handbook of Photochemistry and Photobiology, 2nd ed. Horspool WM, Lenci F. CRC Press; Boca Raton: 2004: 73-1-73-15
    • 7d Biradha K, Santra R. Chem. Soc. Rev. 2013; 42: 950
    • 7e Ramamurthy V, Sivaguru J. Chem. Rev. 2016; 116: 9914
  • 8 Schmidt GM. J. Pure Appl. Chem. 1971; 27: 647
  • 9 Schmidt GM. J. J. Chem. Soc. 1964; 2014
  • 10 Song X, Gu J, Zhang E, Jiang Y, Xin M, Meng Y, Chan AS. C, Zou Y. ACS Sustainable Chem. Eng. 2022; 10: 16399
  • 11 Tanaka K, Toda F. Chem. Rev. 2000; 100: 1025
  • 12 Anastas PT, Warner JC. Green Chemistry: Theory and Practice . Oxford University Press; New York: 1998
    • 13a Laird RC, Sinnwell MA, Nguyen NP, Swenson DC, Mariappan SV. S, MacGillivray LR. Org. Lett. 2015; 17: 3233
    • 13b Liu D, Lang J.-P, Abrahams BF. Chem. Commun. 2013; 49: 2682
  • 14 Ito Y, Borecka B, Trotter J, Scheffer JR. Tetrahedron Lett. 1995; 36: 6083
    • 15a Ito Y, Hosomi H, Ohba S. Tetrahedron 2000; 56: 6833
    • 15b MacGillivray LR, Reid JL, Ripmeester JA. J. Am. Chem. Soc. 2000; 122: 7817
    • 15c Friščić T, MacGillivray LR. Chem. Commun. 2005; 5748
    • 15d Mei X, Liu S, Wolf C. Org. Lett. 2007; 9: 2729
    • 15e Bhogala BR, Captain B, Parthasarathy A, Ramamurthy V. J. Am. Chem. Soc. 2010; 132: 13434
    • 15f Campillo-Alvarado G, Brannan AD, Swenson DC, MacGillivray LR. Org. Lett. 2018; 20: 5490
    • 16a Caronna T, Liantonio R, Lagotheis TA, Metrangolo P, Pilati T, Resnati G. J. Am. Chem. Soc. 2004; 126: 4500
    • 16b Sinnwell MA, MacGillivray LR. Angew. Chem. Int. Ed. 2016; 55: 3477
    • 16c Quentin J, MacGillivray LR. ChemPhysChem 2020; 21: 154
  • 17 Alfuth J, Jeannin O, Fourmigué M. Angew. Chem. Int. Ed. 2022; 61: e202206249

    • Reviews:
    • 18a Gan M.-M, Yu J.-G, Wang Y.-Y, Han Y.-F. Cryst. Growth Des. 2018; 18: 553
    • 18b MacGillivray LR, Papaefstathiou GS, Friščić T, Hamilton TD, Bučar D.-K, Chu Q, Varshney DB, Georgiev IG. Acc. Chem. Res. 2008; 41: 280
    • 19a Pattabiraman M, Natarajan A, Kaanumalle LS, Ramamurthy V. Org. Lett. 2005; 7: 529
    • 19b Nguyen N, Clements AR, Pattabiraman M. New J. Chem. 2016; 40: 2433
    • 19c Kashyap A, Balraj V, Ramalingam V, Pattabiraman M. J. Photochem. Photobiol. A 2022; 425: 113695
    • 20a Bassani DM, Darcos V, Mahony S, Desvergne J.-P. J. Am. Chem. Soc. 2000; 122: 8795
    • 20b Müller C, Bauer A, Bach T. Angew. Chem. Int. Ed. 2009; 48: 6640
    • 20c Müller C, Bauer A, Maturi MM, Cuquerella MC, Miranda MA, Bach T. J. Am. Chem. Soc. 2011; 133: 16689
    • 20d Alonso R, Bach T. Angew. Chem. Int. Ed. 2014; 53: 4368
    • 20e Telmesani R, Park SH, Lynch-Colameta T, Beeler AB. Angew. Chem. Int. Ed. 2015; 54: 11521
    • 20f Brimioulle R, Lenhart D, Maturi MM, Bach T. Angew. Chem. Int. Ed. 2015; 54: 3872
    • 20g Telmesani R, White JA. H, Beeler AB. ChemPhotoChem 2018; 2: 865
    • 22a Kimura M, Shimoyama M, Morosawa S. J. Chem. Soc., Chem. Commun. 1991; 375
    • 22b Haag D, Scharf H.-D. J. Org. Chem. 1996; 61: 6127
  • 23 König B, Leue S, Horn C, Caudan A, Desvergne J.-P, Bouas-Laurent H. Liebigs Ann. 1996; 1231
    • 24a Greiving H, Hopf H, Jones PG, Bubenitschek P, Desvergne J.-P, Bouas-Laurent H. J. Chem. Soc., Chem. Commun. 1994; 1075
    • 24b Greiving H, Hopf H, Jones PG, Bubenitschek P, Desvergne J.-P, Bouas-Laurent H. Liebigs Ann. 1995; 1949
    • 24c Hopf H, Greiving H, Jones PG, Bubenitschek P. Angew. Chem. Int. Ed. Engl. 1995; 34: 685
  • 25 Zitt H, Dix I, Hopf H, Jones PG. Eur. J. Org. Chem. 2002; 2298
  • 26 Ghosn MW, Wolf C. J. Org. Chem. 2010; 75: 6653
    • 27a Sicignano M, Rodríguez RI, Alemán J. Eur. J. Org. Chem. 2021; 3303
    • 27b Daub ME, Jung H, Lee BJ, Won J, Baik M.-H, Yoon TP. J. Am. Chem. Soc. 2019; 141: 9543
  • 28 Pagire SK, Hossain A, Traub L, Kerres S, Reiser O. Chem. Commun. 2017; 53: 12072
    • 29a Lei T, Zhou C, Huang M.-Y, Zhao L.-M, Yang B, Ye C, Xiao H, Meng Q.-Y, Ramamurthy V, Tung C.-H, Wu L.-Z. Angew. Chem. Int. Ed. 2017; 56: 15407
    • 29b Jiang Y, Wang C, Rogers CR, Kodaimati MS, Weiss EA. Nat. Chem. 2019; 11: 1034
    • 29c Wu Q.-A, Chen F, Ren C.-R, Liu X.-F, Chen H, Xu L.-X, Yu X.-C, Luo S.-P. Org. Biomol. Chem. 2020; 18: 3707
    • 29d Wang J.-S, Wu K, Yin C, Li K, Huang Y, Ruan J, Feng X, Hu P, Su C.-Y. Nat. Commun. 2020; 11: 4675
    • 29e Jiang Y, López-Arteaga R, Weiss EA. J. Am. Chem. Soc. 2022; 144: 3782

      For recent examples, see:
    • 30a Nguyen TB, Al-Mourabit A. Photochem. Photobiol. Sci. 2016; 15: 1115
    • 30b Amjaour H, Wang Z, Mabin M, Puttkammer J, Busch S, Chu QR. Chem. Commun. 2019; 55: 214
    • 30c Wang ZD, Elliott Q, Wang Z, Setien RA, Puttkammer J, Ugrinov A, Lee J, Webster DC, Chu QR. ACS Sustainable Chem. Eng. 2018; 6: 8136
  • 31 Ito Y. Solid-State Organic Photochemistry of Mixed Molecular Crystals. In Molecular and Supramolecular Photochemistry, Vol. 3. Ramamurthy V, Schanze KS. Marcel Dekker; New York: 1999: 1
  • 32 Yagci BB, Zorlu Y, Türkmen YE. J. Org. Chem. 2021; 86: 13118
    • 33a Mammadova F, Hamarat B, Ahmadli D, Şahin O, Bozkaya U, Türkmen YE. ChemistrySelect 2020; 5: 13387
    • 33b Türkmen YE. Turk. J. Chem. 2018; 42: 1398
  • 34 Ahmadli D, Türkmen YE. Tetrahedron Lett. 2022; 100: 153877
  • 35 Lenihan JM, Mailloux MJ, Beeler AB. Org. Process Res. Dev. 2022; 26: 1812
  • 36 Aung T, Liberko CA. J. Chem. Educ. 2014; 91: 939
  • 37 Chaudhary A, Mohammad A, Mobin AM. Cryst. Growth Des. 2017; 17: 2893
  • 38 Please see the Supporting Information of ref. 32 for illustrations of instrumental setups of the photochemical experiments.

Corresponding Author

Yunus Emre Türkmen
UNAM, National Nanotechnology Research Center, Institute of Materials Science and Nanotechnology, Bilkent University
Ankara 06800
Türkiye   

Publikationsverlauf

Eingereicht: 31. Mai 2023

Angenommen nach Revision: 10. Juli 2023

Accepted Manuscript online:
10. Juli 2023

Artikel online veröffentlicht:
16. August 2023

© 2023. Thieme. All rights reserved

Georg Thieme Verlag KG
Rüdigerstraße 14, 70469 Stuttgart, Germany

  • References

  • 1 Bernstein HI, Quimby WC. J. Am. Chem. Soc. 1943; 65: 1845 ; and references therein
  • 2 Bassani DM. The Dimerization of Cinnamic Acid Derivatives . In CRC Handbook of Photochemistry and Photobiology, 2nd ed. Horspool WM, Lenci F. CRC Press; Boca Raton: 2004: 20-1-20-20
  • 3 Fujiwara A, Nishi M, Yoshida S, Hasegawa M, Yasuma C, Ryo A, Suzuki Y. Phytochemistry 2016; 122: 139
  • 4 Chai XY, Song YL, Xu ZR, Shi HM, Bai CC, Bi D, Wen J, Li FF, Tu PF. J. Nat. Prod. 2008; 71: 814
  • 5 Carmignani M, Volpe AR, Monache FD, Botta B, Espinal R, De Bonnevaux SC, De Luca C, Botta M, Corelli F, Tafi A, Ripanti G, Monache GD. J. Med. Chem. 1999; 42: 3116
  • 6 Huang B, Zhang F, Yu G, Song Y, Wang X, Wang M, Gong Z, Su R, Jia Y. J. Med. Chem. 2016; 59: 3953
    • 7a Cohen MD, Schmidt GM. J. J. Chem. Soc. 1964; 1996
    • 7b Cohen MD, Schmidt GM. J, Sonntag FI. J. Chem. Soc. 1964; 2000

    • Reviews:
    • 7c Sonoda Y. [2+2]-Photocycloadditions in the Solid State . In CRC Handbook of Photochemistry and Photobiology, 2nd ed. Horspool WM, Lenci F. CRC Press; Boca Raton: 2004: 73-1-73-15
    • 7d Biradha K, Santra R. Chem. Soc. Rev. 2013; 42: 950
    • 7e Ramamurthy V, Sivaguru J. Chem. Rev. 2016; 116: 9914
  • 8 Schmidt GM. J. Pure Appl. Chem. 1971; 27: 647
  • 9 Schmidt GM. J. J. Chem. Soc. 1964; 2014
  • 10 Song X, Gu J, Zhang E, Jiang Y, Xin M, Meng Y, Chan AS. C, Zou Y. ACS Sustainable Chem. Eng. 2022; 10: 16399
  • 11 Tanaka K, Toda F. Chem. Rev. 2000; 100: 1025
  • 12 Anastas PT, Warner JC. Green Chemistry: Theory and Practice . Oxford University Press; New York: 1998
    • 13a Laird RC, Sinnwell MA, Nguyen NP, Swenson DC, Mariappan SV. S, MacGillivray LR. Org. Lett. 2015; 17: 3233
    • 13b Liu D, Lang J.-P, Abrahams BF. Chem. Commun. 2013; 49: 2682
  • 14 Ito Y, Borecka B, Trotter J, Scheffer JR. Tetrahedron Lett. 1995; 36: 6083
    • 15a Ito Y, Hosomi H, Ohba S. Tetrahedron 2000; 56: 6833
    • 15b MacGillivray LR, Reid JL, Ripmeester JA. J. Am. Chem. Soc. 2000; 122: 7817
    • 15c Friščić T, MacGillivray LR. Chem. Commun. 2005; 5748
    • 15d Mei X, Liu S, Wolf C. Org. Lett. 2007; 9: 2729
    • 15e Bhogala BR, Captain B, Parthasarathy A, Ramamurthy V. J. Am. Chem. Soc. 2010; 132: 13434
    • 15f Campillo-Alvarado G, Brannan AD, Swenson DC, MacGillivray LR. Org. Lett. 2018; 20: 5490
    • 16a Caronna T, Liantonio R, Lagotheis TA, Metrangolo P, Pilati T, Resnati G. J. Am. Chem. Soc. 2004; 126: 4500
    • 16b Sinnwell MA, MacGillivray LR. Angew. Chem. Int. Ed. 2016; 55: 3477
    • 16c Quentin J, MacGillivray LR. ChemPhysChem 2020; 21: 154
  • 17 Alfuth J, Jeannin O, Fourmigué M. Angew. Chem. Int. Ed. 2022; 61: e202206249

    • Reviews:
    • 18a Gan M.-M, Yu J.-G, Wang Y.-Y, Han Y.-F. Cryst. Growth Des. 2018; 18: 553
    • 18b MacGillivray LR, Papaefstathiou GS, Friščić T, Hamilton TD, Bučar D.-K, Chu Q, Varshney DB, Georgiev IG. Acc. Chem. Res. 2008; 41: 280
    • 19a Pattabiraman M, Natarajan A, Kaanumalle LS, Ramamurthy V. Org. Lett. 2005; 7: 529
    • 19b Nguyen N, Clements AR, Pattabiraman M. New J. Chem. 2016; 40: 2433
    • 19c Kashyap A, Balraj V, Ramalingam V, Pattabiraman M. J. Photochem. Photobiol. A 2022; 425: 113695
    • 20a Bassani DM, Darcos V, Mahony S, Desvergne J.-P. J. Am. Chem. Soc. 2000; 122: 8795
    • 20b Müller C, Bauer A, Bach T. Angew. Chem. Int. Ed. 2009; 48: 6640
    • 20c Müller C, Bauer A, Maturi MM, Cuquerella MC, Miranda MA, Bach T. J. Am. Chem. Soc. 2011; 133: 16689
    • 20d Alonso R, Bach T. Angew. Chem. Int. Ed. 2014; 53: 4368
    • 20e Telmesani R, Park SH, Lynch-Colameta T, Beeler AB. Angew. Chem. Int. Ed. 2015; 54: 11521
    • 20f Brimioulle R, Lenhart D, Maturi MM, Bach T. Angew. Chem. Int. Ed. 2015; 54: 3872
    • 20g Telmesani R, White JA. H, Beeler AB. ChemPhotoChem 2018; 2: 865
    • 22a Kimura M, Shimoyama M, Morosawa S. J. Chem. Soc., Chem. Commun. 1991; 375
    • 22b Haag D, Scharf H.-D. J. Org. Chem. 1996; 61: 6127
  • 23 König B, Leue S, Horn C, Caudan A, Desvergne J.-P, Bouas-Laurent H. Liebigs Ann. 1996; 1231
    • 24a Greiving H, Hopf H, Jones PG, Bubenitschek P, Desvergne J.-P, Bouas-Laurent H. J. Chem. Soc., Chem. Commun. 1994; 1075
    • 24b Greiving H, Hopf H, Jones PG, Bubenitschek P, Desvergne J.-P, Bouas-Laurent H. Liebigs Ann. 1995; 1949
    • 24c Hopf H, Greiving H, Jones PG, Bubenitschek P. Angew. Chem. Int. Ed. Engl. 1995; 34: 685
  • 25 Zitt H, Dix I, Hopf H, Jones PG. Eur. J. Org. Chem. 2002; 2298
  • 26 Ghosn MW, Wolf C. J. Org. Chem. 2010; 75: 6653
    • 27a Sicignano M, Rodríguez RI, Alemán J. Eur. J. Org. Chem. 2021; 3303
    • 27b Daub ME, Jung H, Lee BJ, Won J, Baik M.-H, Yoon TP. J. Am. Chem. Soc. 2019; 141: 9543
  • 28 Pagire SK, Hossain A, Traub L, Kerres S, Reiser O. Chem. Commun. 2017; 53: 12072
    • 29a Lei T, Zhou C, Huang M.-Y, Zhao L.-M, Yang B, Ye C, Xiao H, Meng Q.-Y, Ramamurthy V, Tung C.-H, Wu L.-Z. Angew. Chem. Int. Ed. 2017; 56: 15407
    • 29b Jiang Y, Wang C, Rogers CR, Kodaimati MS, Weiss EA. Nat. Chem. 2019; 11: 1034
    • 29c Wu Q.-A, Chen F, Ren C.-R, Liu X.-F, Chen H, Xu L.-X, Yu X.-C, Luo S.-P. Org. Biomol. Chem. 2020; 18: 3707
    • 29d Wang J.-S, Wu K, Yin C, Li K, Huang Y, Ruan J, Feng X, Hu P, Su C.-Y. Nat. Commun. 2020; 11: 4675
    • 29e Jiang Y, López-Arteaga R, Weiss EA. J. Am. Chem. Soc. 2022; 144: 3782

      For recent examples, see:
    • 30a Nguyen TB, Al-Mourabit A. Photochem. Photobiol. Sci. 2016; 15: 1115
    • 30b Amjaour H, Wang Z, Mabin M, Puttkammer J, Busch S, Chu QR. Chem. Commun. 2019; 55: 214
    • 30c Wang ZD, Elliott Q, Wang Z, Setien RA, Puttkammer J, Ugrinov A, Lee J, Webster DC, Chu QR. ACS Sustainable Chem. Eng. 2018; 6: 8136
  • 31 Ito Y. Solid-State Organic Photochemistry of Mixed Molecular Crystals. In Molecular and Supramolecular Photochemistry, Vol. 3. Ramamurthy V, Schanze KS. Marcel Dekker; New York: 1999: 1
  • 32 Yagci BB, Zorlu Y, Türkmen YE. J. Org. Chem. 2021; 86: 13118
    • 33a Mammadova F, Hamarat B, Ahmadli D, Şahin O, Bozkaya U, Türkmen YE. ChemistrySelect 2020; 5: 13387
    • 33b Türkmen YE. Turk. J. Chem. 2018; 42: 1398
  • 34 Ahmadli D, Türkmen YE. Tetrahedron Lett. 2022; 100: 153877
  • 35 Lenihan JM, Mailloux MJ, Beeler AB. Org. Process Res. Dev. 2022; 26: 1812
  • 36 Aung T, Liberko CA. J. Chem. Educ. 2014; 91: 939
  • 37 Chaudhary A, Mohammad A, Mobin AM. Cryst. Growth Des. 2017; 17: 2893
  • 38 Please see the Supporting Information of ref. 32 for illustrations of instrumental setups of the photochemical experiments.

Zoom Image
Zoom Image
Zoom Image
Zoom Image
Zoom Image
Figure 1 Examples of bioactive natural products having dimers of cinnamic acid derivatives in their structures
Zoom Image
Scheme 1 Photochemical reactivity of different polymorphs of cinnamic acids
Zoom Image
Scheme 2 Selected examples of photochemical [2+2] cycloadditions of cinnamic acid derivatives
Zoom Image
Scheme 3 Non-redox disproportionation of monoester 8a
Zoom Image
Figure 2 X-ray crystal structure of cycloadduct 9j
Zoom Image
Scheme 4 Synthesis of β-truxinic acid products 10 via hydrolysis of cycloadducts 9 under basic conditions. a The syntheses of these β-truxinic acid products were reported previously.[32] b Product 11 was obtained by the reaction of 9g with NaOMe in MeOH at 23 °C.
Zoom Image
Scheme 5 Conversion of cycloadduct 9n to 12 and 13
Zoom Image
Scheme 6 Photocatalytic cycloaddition of diester 6a