Next Article in Journal
An Unexpected Reaction between Diaryliodonium Salts and DMSO
Next Article in Special Issue
Exploration of the Divergent Outcomes for the Nenitzescu Reaction of Piperazinone Enaminoesters
Previous Article in Journal
Specific Bifunctionalization on the Surface of Phosphorus Dendrimers Syntheses and Properties
Previous Article in Special Issue
A Molecular Electron Density Theory Study of the [3+2] Cycloaddition Reaction of Pseudo(mono)radical Azomethine Ylides with Phenyl Vinyl Sulphone
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dihydrooxazine Byproduct of a McMurry–Melton Reaction en Route to a Synthetic Bacteriochlorin

1
Department of Chemistry, North Carolina State University, Raleigh, NC 27695-8204, USA
2
Department of Materials Science and Engineering, School of Molecular Science and Engineering, Vidyasirimedhi Institute of Science and Technology, Wangchan, Rayong 21210, Thailand
*
Author to whom correspondence should be addressed.
Organics 2022, 3(3), 262-274; https://doi.org/10.3390/org3030019
Submission received: 5 June 2022 / Revised: 26 July 2022 / Accepted: 1 August 2022 / Published: 8 August 2022
(This article belongs to the Special Issue Chemistry of Heterocycles)

Abstract

:
A synthetic route to gem-dimethyl-substituted bacteriochlorins—models of native bacteriochlorophylls—relies on the formation of a dihydrodipyrrin precursor via a series of established reactions: van Leusen pyrrole formation, Vilsmeier formylation, Henry reaction, borohydride reduction, Michael addition, and McMurry–Melton pyrroline formation. The latter is the least known of the series. Here, the McMurry–Melton reaction of a 2-(6-oxo-2-nitrohexyl)pyrrole in the presence of TiCl3 and an ammonium acetate buffer formed the expected Δ1-pyrroline, as well as an unexpected polar, cyclic byproduct (a 5,6-dihydro-4H-1,2-oxazin-6-ol), each attached to the 2-methylpyrrole unit. Both species were characterized by single-crystal X-ray diffraction. The McMurry–Melton reaction is a type of intercepted Nef reaction (the transformation of a nitroalkyl motif into a carbonyl group), where both the Δ1-pyrroline and the dihydrooxazine derive from the reaction of the nitrogen derived from the nitro group upon complete or partial reductive deoxygenation, respectively, with the γ-keto group. The report also considers competing Nef and McMurry–Melton reactions, the nature of available TiCl3 reagents, and the use of ammonium acetate for buffering the TiCl3/HCl reagent.

1. Introduction

Bacteriochlorophyll a constitutes nature’s chief light absorber in the near-infrared (NIR) region and underpins anoxygenic bacterial photosynthesis (Chart 1) [1]. The ability to harvest NIR light offers many research and practical applications spanning energy sciences, photonics, and medicine. The core chromophore of bacteriochlorophyll a is a bacteriochlorin. To gain access to synthetically malleable bacteriochlorins, we have been working to develop routes to gem-dimethyl-substituted macrocycles that contain the bacteriochlorin chromophore. The rationale for the use of gem-dimethyl substituents is twofold: (1) enforce saturation at the β-positions of the pyrrolic rings; and (2) achieve more direct synthesis without a requirement for the control of stereochemical configuration, as would be the case for the trans-dialkyl substituted native bacteriochlorophyll a.
The synthesis of bacteriochlorins can be achieved by self-condensation of a dihydrodipyrrin-acetal (5a), as illustrated for bacteriochlorin BC-1 (Chart 1). During the course of the synthesis of BC-1, a polar byproduct was isolated in the final step of dihydrodipyrrin formation. The byproduct arises during the McMurry–Melton reaction wherein a Δ1-pyrroline is formed by TiCl3-mediated reductive deoxygenation of a 1-nitro-4-oxo-substituted alkane. In this paper, we first report the synthesis of BC-1 and then pivot to focus on the McMurry–Melton reaction byproduct.

2. Results and Discussion

A. Bacteriochlorin synthesis. The synthesis of bacteriochlorin BC-1 follows the established pathway [2,3] shown in Scheme 1. The van Leusen reaction [4,5] of methyl cinnamate with toluenesulfonylmethyl isocyanide (TosMIC) afforded the corresponding pyrrole as the crude product in estimated yield of 62%. The subsequent Vilsmeier formylation [6] gave aldehyde 1a in 64% yield, accompanied by the positional isomer 1b in 6.4% yield. Aldehyde 1a was subjected to the Henry reaction [7] with nitromethane to give the nitrovinyl product, which upon reduction with NaBH4 gave the 2-(2-nitroethyl)pyrrole 2 in 50% yield. The Michael addition [8] of 2 with the α,β-unsaturated ketone 3 [9] under solventless conditions [3] in the presence of DBU provided Michael adduct 4 in 68% yield. The McMurry–Melton reaction was carried out by treatment of 4 with NaOMe followed by TiCl3 in a buffer of saturated aqueous ammonium acetate. The expected dihydrodipyrrin 5a was obtained in 41% yield, accompanied by the unexpected and more polar byproduct 5b, a 5,6-dihydro-4H-1,2-oxazine, in 23% yield. The difference in polarity of 5a and 5b is substantial, with Rf values on silica TLC of 0.5 and 0.1, respectively. The Eastern–Western self-condensation [2,3] of dihydrodipyrrin 5a in the presence of BF3·O(Et)2 in CH2Cl2 gave 5-unsubstituted bacteriochlorin BC-1 in 37% yield and the 5-methoxybacteriochlorin MeOBC-1 in 6.1% yield.
The absorption spectra of the two bacteriochlorins are displayed in Figure 1. The spectra are typical of bacteriochlorins [10], with strong transitions in the NIR and near-ultraviolet region and a weak transition in the visible region. For BC-1, the long-wavelength (Qy) band is at 767 nm, whereas that for MeOBC-1 is at 748 nm, consistent with the known hypsochromic effect [2] of the 5-methoxy group.
B. The dihydrooxazine product of the McMurry–Melton reaction. The structure of 5b was first thought to be an acyclic oxime. Single-crystal X-ray determination, however, revealed the presence of the 5,6-dihydro-4H-1,2-oxazine motif (Figure 2). In 5b, the pyrrole and dihydrooxazine rings exhibit a butterfly arrangement, tilted by ~60° with respect to each other due to the intervening methylene unit. Slight disorder is observed for the phenyl substituent. The distance between the pyrrolic NH and the dihydrooxazine N (2.955 Å) is consistent with moderate intramolecular hydrogen bonding. The dihydrooxazine presents as a somewhat flattened chair structure due to the sp2-hybridized imine carbon, where the C=N and O–C bonds exhibit a dihedral angle of 22.96° (versus 60° in chair cyclohexane). The hydroxy group of the hemi-ketal is positioned axial, whereas the dimethoxymethyl group is equatorial; the H···O distance between the OH group and the two methoxy oxygen atoms is 2.539 and 3.103 Å. In short, the structure of the dihydrooxazine unit somewhat resembles that of fructopyranose, which is also a cyclic hemiketal. The crystal structure of 5b contains both R and S stereoisomers at the C3 position, confirming that 5b exists as a pair of enantiomers.
The structure of dihydrodipyrrin 5a is ordinary, containing nearly coplanar pyrrole and pyrroline rings, a partially rotated phenyl group with respect to the pyrrole plane (52.7°), and the ester carbonyl unit projected away from the adjacent phenyl moiety.
Although the crystal structure of 5b was unambiguous, we wanted to be able to rely on NMR spectra going forward to examine crude dihydrodipyrrin-forming reaction mixtures for the presence of analogous pyrrole–dihydrooxazines. The distinct structures of the precursor 4 and products 5a and 5b are readily seen in the 1H NMR spectra (Figure 3). There are four key differences between compounds 4 and 5b. (1) An apparent (but partially unresolved) doublet of doublets is observed for the proton (labeled “d”) attached to the stereogenic carbon bearing the nitro group of compound 4, a feature that is missing in 5b due to the C=N bond of the dihydrooxazine. (2) An intense singlet for the two methoxy groups of the acetal is observed for 4, whereas the two methoxy groups in 5b resonate with Δδ of ~0.15 ppm, consistent with the distinct environments (where one but not the other is likely hydrogen-bonded with the hydroxyl group) observed upon X-ray structural analysis. (3) The methylene protons (“a”) adjacent to the pyrrole in 4 are diastereotopic and resonate as an apparent doublet of doublets at 3.19 ppm (largely resolved) and 2.94 ppm (unresolved), whereas in 5b the “a” protons again are diastereotopic but resonate downfield (Δδ of ~0.34 and ~0.55 ppm) as a partially obscured doublet. (4) The methylene protons flanked by the gem-dimethyl group and the ketone (“b”) in 4 resonate as a multiplet centered at ~2.5 ppm, versus ~1.77 ppm in 5b due to the flanking hemiketal motif rather than the ketone. Both 4 and 5b are racemic.
The 1H NMR spectrum of dihydrodipyrrin 5a is simpler, with singlets due to the methine proton “a” and the methylene protons “b”. In all three compounds, the chemical shift of the dimethyl acetal methine proton (4.2 in 4, 4.0 in 5b, and 5.2 ppm in 5a) reflects differences in structure at the adjacent carbon. On the other hand, the resonance of the protons of the carbomethoxy group are nearly constant across the three compounds. The gem-dimethyl groups also give distinct resonances (not shown) for the three compounds, including 4 (0.98, 1.00 ppm), 5b (0.89, 1.07 ppm), and 5a (1.14 ppm). The observed distinctions in 1H NMR spectra should provide a convenient diagnostic for the future assessment of crude samples from dihydrodipyrrin-forming reactions.

3. Discussion

A key objective over the years has been to refine reaction pathways so as to achieve streamlined implementation and increased yields of bacteriochlorins, for which BC-1 and MeOBC-1 are prototypical. Here, the identification of an unexpected byproduct in the McMurry–Melton reaction, leading to the dihydrodipyrrin, prompts focus on this transformation. In this section, we describe the original report by McMurry and Melton, then overview extension of the reaction to form dihydrodipyrrins with emphasis on observations that point to the complexities of this ostensibly simple transformation. A goal of the following discussion is to develop a framework in support of future, more extensive studies.
The treatment of a nitroalkane with a base followed by aqueous acid constitutes the Nef reaction [11,12,13,14,15], first reported in 1894 (Scheme 2). The base causes formation of the nitronate, whereas the aqueous acid causes hydrolysis of the nitronate [14], forming the corresponding ketone or aldehyde. McMurry and Melton (1973) reported that a nitroalkane gave the ketone directly upon treatment with TiCl3 in aqueous acid (pH < 1) at room temperature [16]. This modification of the Nef reaction has been referred to previously as the McMurry method [14,15] or the McMurry–Melton method [17], to be distinguished from the better-known McMurry reaction [18] (low-valent titanium-induced deoxygenative dimerization of ketones to form an alkene), which was reported one year later [19].
In an effort to broaden the scope, McMurry and Melton also examined reactions at more neutral pH: the nitroalkane was first treated with one equivalent of NaOMe in methanol followed by an aqueous solution of TiCl3 buffered at pH 6 (with ammonium acetate) [16]. They stated [16], “Much to our surprise, however, several nitro compounds gave worse results under these neutral conditions than under acidic conditions. For example, nitro ketone [I] gave none of the expected diketone [heptane-2,5-dione, II] but gave rather, as the only isolable organic compound, the pyrroline [III] (20%).” The inadvertent formation of a Δ1-pyrroline (III, albeit without reported characterization data; Scheme 2) [16], was extended by Battersby and coworkers (1981) for the conversion of a 6-oxo-2-nitrohexylpyrrole (IV) to a tetrahydrodipyrrin (V) for use in the synthesis of an isobacteriochlorin [20]. Battersby and coworkers (1988) also employed the McMurry–Melton reaction to convert a 6-oxo-2-nitrohexylpyrrole (VI) to a dihydrodipyrrin (VII) for use in the synthesis of a chlorin (Scheme 2) [21]. We followed in the same manner to synthesize diverse dihydrodipyrrins, with the conversion of 4 to 5a being illustrative. The same route has been exploited to prepare Δ1-pyrrolines unrelated to tetrapyrrole macrocycles [22].
Although the McMurry–Melton reaction constitutes a cornerstone of the synthesis of chlorins and bacteriochlorins [2,3,9,23,24,25,26,27,28,29,30,31,32], anomalies in this domain are known. During the synthesis of 15N isotopologues of bacteriochlorins, we observed partial loss of the 15N label during the McMurry–Melton reaction upon reaction of VIII-[15N]-NO2 [28] (Scheme 3). The dihydrodipyrrin mixture was comprised of intact 15N-dihydrodipyrrin isotopologue IX-[15N] and the natural abundance dihydrodipyrrin IX-NA. The latter was proposed to originate from competing Nef and McMurry–Melton reactions. Thus, the Nef reaction of nitronate VIII-[15N]-nitronate yields the diketone X-keto along with the loss of a 15N species. Condensation with ammonia (from the ammonium acetate buffer) affords the imine or enamine X-enamine, which then undergoes intramolecular condensation in a Paal–Knorr-like process with the second carbonyl group to form IX-NA.
The dihydrooxazine 5b contains an oxime motif. Reduced nitro species, including nitroso and oxime groups, have been proposed as reaction intermediates by McMurry and Melton [16], and the oxime is considered an intermediate (if not a product) of the Nef reaction [13,14,15]. 5,6-Dihydro-4H-1,2-oxazines have been prepared previously [33,34,35], typically by reduction of the corresponding and well-studied 1,2-oxazine-N-oxides (for which multiple synthetic entry points are known [36]). To our knowledge, 5,6-dihydro-4H-1,2-oxazines have heretofore not been identified in a McMurry–Melton reaction. The mechanistic question of whether a dihydrooxazine N-oxide is formed here and then reduced, or the reduction occurs prior to cyclization to form the dihydrooxazine directly, requires additional study. Extensive work remains to understand the role of various reaction parameters on the McMurry–Melton reaction course and the yields of the desired Δ1-pyrroline versus the dihydrooxazine byproduct. Further discussion of the various McMurry–Melton reaction conditions is provided in Appendix A.

4. Conclusions

The isolation of dihydrooxazine 5b sheds light on the little-studied McMurry–Melton reaction. An unknown is whether analogous dihydrooxazine byproducts are generally formed in McMurry–Melton reactions and have heretofore not been identified given far greater polarity compared with the Δ1-pyrroline. The use of the 1H NMR spectral comparisons outlined in Figure 3 should facilitate the examination of crude reaction mixtures and thereby afford a broader understanding of the McMurry–Melton reaction.

5. Experimental Section

General. Commercial compounds were used as received. All solvents (anhydrous or reagent-grade) were used as received unless noted otherwise. THF was freshly distilled from sodium/benzophenone. Silica (40 μm average particle size) was used for column chromatography. Electrospray ionization mass spectrometry (ESI-MS) data generally enable accurate mass measurements, were obtained in the positive ion mode (unless noted otherwise) and are reported for the protonated molecular ion. 1H NMR and proton-decoupled 13C NMR spectra were collected at room temperature in CDCl3 unless noted otherwise and using instruments as indicated for each compound.
4-Methoxycarbonyl-3-phenylpyrrole-2-carboxaldehyde (1a). A solution of methyl cinnamate (5.00 g, 30.8 mmol) and TosMIC (7.22 g, 37.0 mmol) in dry DMF (100 mL) was treated with NaOH (1.85 g, 46.3 mmol), and stirring was continued under argon for 20 h at room temperature. The reaction mixture was quenched by the addition of saturated aqueous NH4Cl (200 mL), and then stirred for 30 min at room temperature. After H2O (100 mL) was added, the mixture was extracted with ethyl acetate (100 mL × 3). The combined organic extract was washed with H2O (100 mL) and brine (100 mL), dried over Na2SO4, and concentrated to dryness in vacuo. The residue was suspended in hexanes/CH2Cl2 (50 mL, 9:1). The suspension was sonicated and filtered, and the precipitate was rinsed with hexanes/CH2Cl2 (50 mL 9:1 × 3) to give 3-methoxycarbonyl-4-phenylpyrrole (3.83 g, 62%) as a brown solid. The Vilsmeier reagent was prepared by treatment of dry DMF (16 mL) with POCl3 (4.01 g, 26.2 mmol) at room temperature and stirring of the resulting mixture for 1 h. In a separate flask, a solution of crude 3-methoxycarbonyl-4-phenylpyrrole (3.83 g, 19.0 mmol) in dry DMF (162 mL) was treated with the freshly prepared Vilsmeier reagent at room temperature and stirring was continued for 3 h under argon. The reaction mixture was quenched by the addition of saturated aqueous NaOAc (100 mL), and then stirred for 30 min at room temperature. After H2O (200 mL) was added, the mixture was extracted with ethyl acetate (100 mL × 3). The combined organic extract was washed with H2O (300 mL × 2) and brine (100 mL), dried over Na2SO4, and concentrated to dryness in vacuo. Purification by column chromatography [silica gel, 100 g; hexanes/ethyl acetate (4:1)] afforded the title compound (2.81 g, 64%, Rf = 0.25 (hexanes/ethyl acetate = 2:1)) followed by the positional isomer 1b (0.281 mg, 6.4%, Rf = 0.30 (hexanes/ethyl acetate = 2:1)).
Data for 1a: white solid; mp 184–186 °C; 1H NMR (500 MHz, CDCl3) δ 3.74 (3H, s), 7.41–7.47 (5H, m), 7.72 (1H, dd, J = 3.5 Hz, J = 1 Hz), 9.38 (1H, d, J = 1 Hz), 9.79 (1H, brs); 13C-{1H} NMR (175 MHz, CDCl3) δ 51.4, 116.5, 128.0, 128.4, 130.0, 130.89, 130.90, 131.1, 137.3, 163.9, 181.0. HRMS (ESI) for C13H11NO3 (M + H+), calcd 230.0812, found 230.0811.
Data for 1b: white solid; mp 118–121 °C; 1H NMR (500 MHz, CDCl3) δ 3.80 (3H, s), 7.07 (1H, dd, J = 3 Hz, J = 1 Hz), 7.32–7.35 (1H, m), 7.36–7.42 (1H, m), 10.13 (1H, brs), 10.17 (1H, d, J = 1 Hz); 13C-{1H} NMR (175 MHz, CDCl3) δ 51.8, 119.5, 1234.0, 127.5, 128.1, 129.4, 130.0, 133.6, 133.7, 164.3, 182.3. HRMS (ESI) for C13H11NO3 (M + H+), calcd 230.08117, found 230.08069.
4-Methoxycarbonyl-2-(2-nitroethyl)-3-phenylpyrrole (2). A solution of aldehyde 1a (1.03 g, 4.49 mmol) in EtOH (45 mL) was treated with potassium acetate (0.610 g, 8.99 mmol) and methylamine hydrochloride (0.441 g, 4.49 mmol), and stirring was continued under argon for 40 min at room temperature. The solution was treated with nitromethane (1.00 mL, 18.6 mmol). After stirring for 24 h at room temperature, saturated aqueous NH4Cl (50 mL) was added. Stirring was continued for 30 min at room temperature, then H2O (200 mL) was added. The mixture was extracted with ethyl acetate (100 mL × 3). The combined organic extract was washed with H2O (50 mL) and brine (50 mL), dried over Na2SO4, and concentrated to dryness in vacuo. The residue (1.37 g) was dissolved in MeOH (50 mL), and the solution was cooled at 0 °C. The solution was treated portionwise with NaBH4 (205 mg, 5.42 mmol). The reaction mixture was stirred for 1 h at 0 °C, and then quenched by the addition of saturated aqueous NH4Cl (100 mL). After H2O was added, the mixture was extracted with ethyl acetate (100 mL × 3). The combined organic extract was washed with H2O (50 mL) and brine (50 mL), dried over Na2SO4, and concentrated to dryness in vacuo. Purification by column chromatography [silica gel 50 g; hexanes/ethyl acetate (2:1)] afforded a colorless film (0.61 g, 50% in 2 steps). 1H NMR (500 MHz, CDCl3) δ 3.20 (2H, t, J = 6.5 Hz), 3.67 (3H, s), 4.42 (2H, t, J = 6.5 Hz), 7.24–7.27 (2H, m), 7.29–7.34 (1H, m), 7.37–7.40 (3H, m), 8.61 (1H, brs); 13C-{1H} NMR (125 MHz, CDCl3) δ 23.5, 51.0, 75.2, 115.0, 124.0, 124.6, 125.3, 127.2, 128.1, 130.3, 134.3, 164.9. HRMS (ESI) for C14H14N2O4 (M + H+), calcd 275.1026, found 275.1023.
4-Methoxycarbonyl-2-(6,6-dimethoxy-5-oxo-3,3-dimethyl-2-nitrohexyl)-3-phenylpyrrole (4). A solution of pyrrole 2 (610. mg, 2.22 mmol) and Michael acceptor 3 [9] (520. mg, 3.29 mmol) in DBU (25 mL) was stirred for 22 h at room temperature under argon. Saturated aqueous NH4Cl (100 mL) was added, and the mixture was extracted with ethyl acetate (20 mL × 3). The combined organic extract was washed with H2O (10 mL) and brine (10 mL), dried over Na2SO4, and concentrated to dryness in vacuo. Purification by column chromatography [silica gel 25 g; hexanes/ethyl acetate (2:1)] afforded a pale-yellow film (652.1 mg, 68%). 1H NMR (600 MHz, CDCl3) δ 0.98 (3H, s), 1.00 (3H, s), 2.42 (1H, d, J = 18.6 Hz), 2.60 (1H, d, J = 18.6 Hz), 2.94 (1H, dd, estimated J~15.6 Hz, J~2.4 Hz), 3.19 (1H, dd, J = 15.6 Hz, J = 12 Hz), 3.37 (6H, s), 3.63 (3H, s), 4.26 (1H, s), 4.98 (1H, dd, estimated J~12 Hz, J~2.4 Hz), 7.23–7.25 (2H, m), 7.28–7.30 (1H, m), 7.32 (1H, d, J = 3 Hz), 7.35–7.38 (2H, m), 8.43 (1H, brs); 13C-{1H} NMR (150 MHz, CDCl3) δ 23.9, 24.1, 24.9, 36.6, 44.8, 50.9, 55.20, 55.24, 95.0, 104.7, 114.9, 124.0, 124.8, 125.3, 127.0, 128.0, 130.4, 134.3, 164.9, 203.4. HRMS (ESI) for C22H28N2O7 (M + H+), calcd 433.1969, found 433.1965.
8-Methoxycarbonyl-1-(1,1-dimethoxymethyl)-3,3-dimethyl-7-phenyl-2,3-dihydrodipyrrin (5a). In a first flask, a solution of the Michael adduct 4 (498 mg, 1.15 mmol) in distilled THF (25 mL) was treated with NaOMe (88.4 mg, 1.64 mmol), and stirring was continued under argon for 30 min at room temperature. In a second flask, TiCl3 (10.0 mL, 12 wt % TiCl3 in aqueous HCl solution, 15.6 mmol TiCl3) was treated with saturated aqueous NH4OAc solution until a pH of 6 was obtained (pH paper; ultimately 60 mL saturated aqueous NH4OAc solution was added), and stirring of the resulting cloudy, heterogeneous mixture was continued for 10 min at room temperature under argon. (Note that saturated aqueous NH4OAc solution is reported to contain 148 g NH4OAc/100 mL at 4 °C [37]). The contents of the first flask were added to the second flask. The reaction mixture under argon was stirred for 24 h at room temperature. After saturated aqueous NH4Cl (100 mL) was added, the mixture was extracted with ethyl acetate (30 mL × 3). The combined organic extract was washed with H2O (20 mL) and brine (20 mL), dried over Na2SO4, and concentrated to dryness in vacuo. Purification by column chromatography [silica gel 25 g; hexanes/ethyl acetate (2:1)] afforded the title compound (179 mg, 41%, Rf = 0.50 (hexanes/ethyl acetate = 1:1)) and byproduct 5b (111 mg, 23%, Rf = 0.10 (hexanes/ethyl acetate = 1:1)), each as a pale-yellow film.
Data for 5a: 1H NMR (700 MHz, CDCl3) δ 1.14 (6H, s), 2.62 (2H, s), 3.47 (6H, s), 3.69 (3H, s), 5.05 (1H, s), 5.78 (1H, s), 7.31–7.33 (1H, m), 7.37–7.41 (4H, m), 7.53 (1H, d, J = 3.5 Hz), 11.22 (1H, brs); 13C-{1H} NMR (175 MHz, CDCl3) δ 29.1, 40.5, 48.5, 50.9, 54.7, 102.7, 105.3, 114.1, 124.7, 125.5, 126.6, 127.7, 129.7, 131.0, 134.6, 161.6, 165.3, 175.6. HRMS (ESI) for C22H26N2O4 (M + H+), calcd 383.1965, found 383.1970.
Data for the byproduct 4-methoxycarbonyl-2-(6-(dimethoxymethyl)-(6-hydroxy-4,4-dimethyl-5,6-dihydro-4H-1,2-oxazin-3-yl)methyl)-3-phenylpyrrole (5b): 1H NMR (600 MHz, CDCl3) δ 0.90 (3H, s), 1.07 (3H, s), 1.71 (1H, d, J = 12 Hz), 1.84 (1H, d, J = 12 Hz), 3.49 (3H, s), 3.51 (1H, d, J = 16 Hz), 3.56 (1H, d, J = 16 Hz), 3.63 (3H, s), 3.66 (3H, s), 4.18 (1H, s), 7.29–7.30 (3H, m), 7.37–7.40 (3H, m), 9.16 (1H, brs); 13C-{1H} NMR (150 MHz, CDCl3) δ 16.7, 26.5, 26.7, 28.1, 30.7, 38.8, 50.8, 55.8, 58.2, 96.8, 105.6, 114.4, 122.8, 123.8, 126.1, 126.7, 130.6, 134.8, 164.9, 165.2. HRMS (ESI) for C22H28N2O6 (M + H+), calcd 417.2020, found 417.2016.
3,13-Dimethoxycarbonyl-8,8,18,18-tetramethyl-2,12-diphenylbacteriochlorin (BC-1). A solution of dihydropyrrin 5a (156.4 mg, 408.9 μmol) in dry CH2Cl2 (25 mL) was treated with BF3·OEt2 (0.16 mL, 1.5 mmol), and stirring under argon was continued for 24 h at room temperature. After saturated aqueous NaHCO3 (50 mL) was added, the mixture was extracted with CH2Cl2 (10 mL × 3). The combined organic extract was washed with H2O (10 mL) and brine (10 mL), dried over Na2SO4, and concentrated to dryness in vacuo. Purification by column chromatography [silica gel 50 g; hexanes/ethyl acetate (4:1)] afforded the title bacteriochlorin (48.2 mg, 37%) followed by the 5-methoxybacteriochlorin MeOBC-1 (8.3 mg, 6.1%).
Data for BC-1: purple solid; 1H NMR (500 MHz, CDCl3) δ –1.14 (2H, s), 1.84 (12H, s), 4.05 (6H, s), 4.42 (4H, s), 7.65–7.68 (2H, m), 7.72–7.75 (4H, m), 7.93–7.95 (4H, m), 8.45 (2H, s), 9.63 (2H, s). 13C-{1H} NMR (125 MHz, CDCl3) δ 29.9, 30.9, 46.1, 52.0, 91.1, 97.5, 120.0, 128.0, 132.1, 134.3, 134.6, 135.6, 138.6, 161.3, 166.8, 172.2. HRMS (ESI) for C40H38N4O4 (M + H+), calcd 639.2966, found 639.2955. λabs (toluene) 357, 382, 526, 767 nm.
Data for 5-methoxy-3,13-dimethoxycarbonyl-8,8,18,18-tetramethyl-2,12-diphenylbacteriochlorin (MeOBC-1): purple film; 1H NMR (700 MHz, CDCl3) δ –1.54 (1H, s), –1.25 (1H, s), 1.81 (6H, s), 1.86 (6H, s), 4.04 (3H, s), 4.13 (3H, s), 4.26 (3H, s), 4.36 (2H, s), 4.42 (2H, s), 7.63–7.67 (2H, m), 7.71–7.75 (4H, m), 7.92–7.93 (2H, m), 8.08-8.09 (2H, m), 8.44 (1H, s), 8.58 (1H, s), 9.57 (1H, s); 13C-{1H} NMR (175 MHz, CDCl3) δ 29.9, 30.8, 30.9, 45.6, 46.2, 48.0, 51.7, 51.9, 53.0, 64.5, 96.2, 98.1, 98.2, 119.1, 124.5, 127.9, 128.19, 128.23, 128.9, 131.8, 132.2, 133.0, 133.6, 134.3, 134.4, 134.9, 135.4, 135.7, 138.7, 156.5, 161.3, 166.8, 169.2, 172.7. HRMS (ESI) for C41H40N4O5 (M + H+), calcd 669.3072, found 668.3000. λabs (toluene) 361, 373, 528, 748 nm.
X-ray crystallography Single-crystal X-ray diffraction (SCXRD) data were collected using a Bruker D8 Venture single-crystal diffractometer at 100 K via the Psi and Omega scan strategy (MoKα = 0.71073 Å). The crystal data acquisition and refinement were achieved with use of APEX4 [38], SHELX2018 [39], and OLEX2 [40] programs. Suitable crystals of 5a and 5b for SCXRD analysis were obtained upon recrystallization in CH2Cl2/hexane. Compounds 5a and 5b are found in monoclinic systems with the P21/c and P21 space group, respectively. The crystallographic data and molecular packing are described in the Supplementary Materials (see Table S1).
pH simulation The simulation was carried out using fsolve from SciPy [41]. The parameters employed were as follows: pKa for NH4+ = 9.25; pKb for AcO = 9.25 (note that the pKa for AcOH is 4.75, thus the pKb = 14–4.75 = 9.25); and Kw = 10−14 (autoprotolysis of water). The script for the calculations is provided in the Supplementary Materials.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/org3030019/s1, Table of X-ray data for 5a and 5b, computer script for pH calculations, and NMR spectra.

Author Contributions

Conceptualization, J.S.L.; formal analysis, V.-P.T., H.J. and C.-Y.C.; funding acquisition, J.S.L.; investigation, N.M. and P.N.; project administration, J.S.L.; writing—original draft, V.-P.T. and J.S.L.; writing—review & editing, J.S.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a grant from the Chemical Sciences, Geosciences and Biosciences Division, Office of Basic Energy Sciences, of the U. S. Department of Energy (DE-FG02-05ER15661) and by Vidyasirimedhi Institute of Science and Technology (VISTEC).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

X-ray crystal data are deposited at the CCDC for 5a and 5b (CCDC 2176728 and 2176729). All other data are contained herein.

Acknowledgments

All mass spectrometry, NMR, and X-ray structural data were obtained in the Molecular Education, Technology, and Research Innovation Center (METRIC) at NC State University, except for 5a and 5b, for which the X-ray structural data and additional NMR data were obtained at VISTEC.

Conflicts of Interest

The authors declare no competing financial interest.

Appendix A

One challenge to comparative analysis of McMurry–Melton reactions is that the conditions applied have varied over a period of 50 years due to variation in composition of the commercially available TiCl3 reagent. The variation in TiCl3 reagent has been complemented by a range of buffer quantities, which together give diverse reaction conditions. McMurry and Melton reported use of 20 wt % TiCl3 in aqueous solution (pH 1) [16]. Examples of TiCl3 reagents from our own work spanning two decades are shown in Table A1.
Table A1. TiCl3 reagent compositions.
Table A1. TiCl3 reagent compositions.
TiCl3HClHCl/TiCl3 Molar RatioReferences
8.6 wt %28 wt %13.8[2,3,23,24]
solid--[3,25,26]
20 wt %3 wt %0.63[3,9,27,30]
12 wt %5–10 wt %1.76–3.53[29]
20 wt %2 N (~7 wt %)1.48[30,31,32]
12 wt %unspecified-[31]
10–15 wt %unspecified-this work
In the original McMurry–Melton paper, the ratio of nitro substrate: TiCl3: NH4OAc was 1:4:24 [16] (the requisite stoichiometric quantity of TiCl3, as Ti3+ is a one-electron donor, and the complete deoxygenation of the nitro group requires four electrons), whereas Battersby and coworkers employed a ratio of 1:6:16 [21]. A caveat concerning the use of these ratios as guides is that in both cases the quantity of any HCl in the reaction mixture was not reported. The buffer constituents required to achieve pH 6 depend very much on the quantity of HCl present in the TiCl3 reagent. Most recently, we generally have employed a ratio of 1 nitro substrate: 4–8 TiCl3: 300–400 NH4OAc and obtained the desired dihydrodipyrrin in yield of ~30–50% depending on the nature of the substituents [31,32]. Here, on the assumption of 10 wt % HCl, the ratio of nitro substrate: TiCl3: HCl: NH4OAc was 1: 6.8: 23.8: 1000. A consequence of such ratios is a somewhat dilute reaction mixture, which has ranged from ~6–50 mM across numerous nitro substrates [2,3,9,23,24,25,26,27,31,32]. Said differently, a lesser quantity of HCl can be buffered with commensurably diminished NH4OAc, whereupon a higher overall concentration of the nitro substrate can be employed.
The choice of a number of equivalents of NH4OAc to employ depends on several factors: (1) the known or assumed quantity of HCl in the TiCl3 reagent; (2) the liberation of up to 3 HCl from TiCl3 and H2O during the course of the reaction; (3) further acidification of the milieu upon possible oxidation of Ti3+ → Ti4+; and (4) the purity of the initial TiCl3 reagent. Lesser concerns include the number of equivalents of unreacted excess NaOMe carried over from the nitronate formation, and any buffering of HCl provided by the Δ1-pyrroline product and/or amine-containing byproducts. The consideration of buffering also must be cognizant of the non-ideal nature of the reaction mixture, which typically is a quite thick slurry due to the excess ammonium acetate in a mixed aqueous–organic solvent. In some McMurry–Melton reactions, the NH4OAc has been added as a solid to a solution comprised predominately of THF, but again the yields were generally <50% [3,26,29,30]. Although more exacting control over the reaction medium might be achieved with use of neat TiCl3, this reagent is pyrophoric and must be handled in an anaerobic environment. The calculation of pH in such complex mixtures is at best idealized.
The effect of the ratio of NH4OAc to HCl on the pH in an ideal solution is shown in Figure A1. The number of equivalents of NH4OAc and HCl refers to the nitro substrate, which is considered to be at 10 mM (nearly identical results are obtained at much higher concentrations). The use of 1000 equivalents of NH4OAc retains a solution pH value of slightly greater than 6 in the presence of 24 equivalents of HCl. Substantially fewer equivalents of NH4OAc would afford lower pH values, as shown in the graph.
Figure A1. The pH in an ideal solution for given quantities of NH4OAc (1000–25.6 equivalents) and HCl (0–100 equivalents). All equivalents are relative to the nitro substrate (set at 10 mM).
Figure A1. The pH in an ideal solution for given quantities of NH4OAc (1000–25.6 equivalents) and HCl (0–100 equivalents). All equivalents are relative to the nitro substrate (set at 10 mM).
Organics 03 00019 g0a1
The focus on pH 6 in the original McMurry–Melton paper originated from a desire to avoid the adverse effects of low pH on functional groups unrelated to the reacting entities, not on the yield per se [16]. For the reaction of 4 and analogues, care must be taken to avoid acidic hydrolysis of the dimethyl acetal unit. Looking forward, conditions that are only moderately acidic may prove tolerable and thus provide a suitable tradeoff of pH and reaction constituents. Moderately lower pH values also entail diminished NH4OAc, thereby enabling higher reaction concentrations and/or more homogeneous reaction mixtures. Understanding the interplay of the nature of the TiCl3 reagent and any accompanying HCl, desired pH, buffer quantity, reactant concentration, and reaction heterogeneity is a necessary prelude to systematic studies of the origin of dihydrodipyrrin versus dihydrooxazole species in the McMurry–Melton reaction.

References

  1. Scheer, H. An Overview of Chlorophylls and Bacteriochlorophylls. Biochemistry, Biophysics, Functions and Applications. In Chlorophylls and Bacteriochlorophylls. Biochemistry, Biophysics, Functions and Applications; Grimm, B., Porra, R.J., Rüdiger, W., Scheer, H., Eds.; Springer: Dordrecht, The Netherlands, 2006; Volume 25, pp. 1–26. [Google Scholar]
  2. Kim, H.-J.; Lindsey, J.S. De Novo Synthesis of Stable Tetrahydroporphyrinic Macrocycles: Bacteriochlorins and a Tetradehydrocorrin. J. Org. Chem. 2005, 70, 5475–5486. [Google Scholar] [CrossRef] [PubMed]
  3. Krayer, M.; Ptaszek, M.; Kim, H.-J.; Meneely, K.R.; Fan, D.; Secor, K.; Lindsey, J.S. Expanded Scope of Synthetic Bacteriochlorins via Improved Acid Catalysis Conditions and Diverse Dihydrodipyrrin-Acetals. J. Org. Chem. 2010, 75, 1016–1039. [Google Scholar] [CrossRef] [PubMed]
  4. van Leusen, A.M.; Siderius, H.; Hoogenboom, B.E.; van Leusen, D. A New and Simple Synthesis of the Pyrrole Ring System from Michael Acceptors and Tosylmethylisocyanides. Tetrahedron Lett. 1972, 13, 5337–5340. [Google Scholar] [CrossRef]
  5. Ma, Z.; Ma, Z.; Zhang, D. Synthesis of Multi-Substituted Pyrrole Derivatives Through [3 + 2] Cycloaddition with Tosylmethyl Isocyanides (TosMICs) and Electron-Deficient Compounds. Molecules 2018, 23, 2666. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Jones, G.; Stanforth, S.P. The Vilsmeier Reaction of Fully Conjugated Carbocycles and Heterocycles. Org. React. 1997, 49, 1–330. [Google Scholar]
  7. Luzzio, F.A. The Henry Reaction: Recent Examples. Tetrahedron 2001, 57, 915–945. [Google Scholar] [CrossRef]
  8. Tokoroyama, T. Discovery of the Michael Reaction. Eur. J. Org. Chem. 2010, 2009–2016. [Google Scholar] [CrossRef]
  9. Mass, O.; Lindsey, J.S. A trans-AB-Bacteriochlorin Building Block. J. Org. Chem. 2011, 76, 9478–9487. [Google Scholar] [CrossRef]
  10. Kobayashi, M.; Akiyama, M.; Kano, H.; Kise, H. Spectroscopy and Structure Determination. In Chlorophylls and Bacteriochlorophylls. Biochemistry, Biophysics, Functions and Applications; Springer: Dordrecht, The Netherlands, 2006; Volume 25, pp. 79–94. [Google Scholar]
  11. Noland, W.E. The Nef Reaction. Chem. Rev. 1955, 55, 137–155. [Google Scholar] [CrossRef]
  12. Wolfrom, M.L. John Ulric Nef 1862–1915 (Biographical Memoir). Natl. Acad. Sci. 1960, 34, 204–227. [Google Scholar]
  13. Pinnick, H.W. The Nef Reaction. Org. React. 1990, 38, 655–792. [Google Scholar]
  14. Ballini, R.; Petrini, M. Recent Synthetic Developments in the Nitro to Carbonyl Conversion (Nef Reaction). Tetrahedron 2004, 60, 1017−1047. [Google Scholar] [CrossRef]
  15. Ballini, R.; Petrini, M. The Nitro to Carbonyl Conversion (Nef Reaction): New Perspectives for a Classical Transformation. Adv. Synth. Catal. 2015, 357, 2371−2402. [Google Scholar] [CrossRef]
  16. McMurry, J.E.; Melton, J. A New Method for the Conversion of Nitro Groups into Carbonyls. J. Org. Chem. 1973, 38, 4367–4373. [Google Scholar] [CrossRef]
  17. Lindsey, J.S. De Novo Synthesis of Gem-Dialkyl Chlorophyll Analogues for Probing and Emulating our Green World. Chem. Rev. 2015, 115, 6534–6620. [Google Scholar] [CrossRef] [Green Version]
  18. McMurry, J.E. Carbonyl-Coupling Reactions Using Low-Valent Titanium. Chem. Rev. 1989, 89, 1513–1524. [Google Scholar] [CrossRef]
  19. McMurry, J.E.; Fleming, M.P. A New Method for the Reductive Coupling of Carbonyls to Olefins. Synthesis of β-Carotene. J. Am. Chem. Soc. 1974, 96, 4708–4709. [Google Scholar] [CrossRef]
  20. Harrison, P.J.; Fookes, C.J.R.; Battersby, A.R. Synthesis of the Isobacteriochlorin Macrocycle: A Photochemical Approach. J. Chem. Soc. Chem. Commun. 1981, 797–799. [Google Scholar] [CrossRef]
  21. Battersby, A.R.; Dutton, C.J.; Fookes, C.J.R. Synthetic Studies Relevant to Biosynthetic Research on Vitamin B12. Part 7. Synthesis of (±)-Bonellin Dimethyl Ester. J. Chem. Soc. Perkin Trans. 1988, 1, 1569–1576. [Google Scholar] [CrossRef]
  22. Marrec, O.; Christophe, C.; Billard, T.; Langlois, B.; Vors, J.-P.; Pazenok, S. Synthesis of β-Trifluoromethylated Δ1-Pyrrolines. Adv. Synth. Catal. 2010, 352, 2825–2830. [Google Scholar] [CrossRef]
  23. Strachan, J.-P.; O’Shea, D.F.; Balasubramanian, T.; Lindsey, J.S. Rational Synthesis of Meso-Substituted Chlorin Building Blocks. J. Org. Chem. 2000, 65, 3160–3172. [Google Scholar] [CrossRef]
  24. Kim, H.-J.; Dogutan, D.K.; Ptaszek, M.; Lindsey, J.S. Synthesis of Hydrodipyrrins Tailored for Reactivity at the 1- and 9-Positions. Tetrahedron 2007, 63, 37–55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Taniguchi, M.; Cramer, D.L.; Bhise, A.D.; Kee, H.L.; Bocian, D.F.; Holten, D.; Lindsey, J.S. Accessing the Near-Infrared Spectral Region with Stable, Synthetic, Wavelength-Tunable Bacteriochlorins. New J. Chem. 2008, 32, 947–958. [Google Scholar] [CrossRef]
  26. Borbas, K.E.; Ruzié, C.; Lindsey, J.S. Swallowtail Bacteriochlorins. Lipophilic Absorbers for the Near-Infrared. Org. Lett. 2008, 10, 1931–1934. [Google Scholar] [CrossRef]
  27. Aravindu, K.; Krayer, M.; Kim, H.-J.; Lindsey, J.S. Facile Synthesis of a B,D-Tetradehydrocorrin and Rearrangement to Bacteriochlorins. New J. Chem. 2011, 35, 1376–1384. [Google Scholar] [CrossRef]
  28. Chen, C.-Y.; Bocian, D.F.; Lindsey, J.S. Synthesis of 24 Bacteriochlorin Isotopologues, Each Containing a Symmetrical Pair of 13C or 15N Atoms in the Inner Core of the Macrocycle. J. Org. Chem. 2014, 79, 1001–1016. [Google Scholar] [CrossRef]
  29. Zhang, S.; Kim, H.-J.; Tang, Q.; Yang, E.; Bocian, D.F.; Holten, D.; Lindsey, J.S. Synthesis and Photophysical Characteristics of 2,3,12,13-Tetraalkylbacteriochlorins. New J. Chem. 2016, 40, 5942–5956. [Google Scholar] [CrossRef]
  30. Fujita, H.; Jing, H.; Krayer, M.; Allu, S.; Veeraraghavaiah, G.; Wu, Z.; Jiang, J.; Diers, J.R.; Magdaong, N.C.M.; Mandal, A.K.; et al. Annulated Bacteriochlorins for Near-Infrared Photophysical Studies. New J. Chem. 2019, 43, 7209–7232. [Google Scholar] [CrossRef]
  31. Jing, H.; Wang, P.; Chen, B.; Jiang, J.; Vairaprakash, P.; Liu, S.; Rong, J.; Chen, C.-Y.; Tran, V.-P.; Nalaoh, P.; et al. Synthesis of Bacteriochlorins Bearing Diverse β-Substituents. New J. Chem. 2022, 46, 5534–5555. [Google Scholar] [CrossRef]
  32. Sun, R.; Liu, M.; Wang, P.; Qin, Y.; Schnedermann, C.; Maher, A.G.; Zheng, S.-L.; Liu, S.; Chen, B.; Zhang, S.; et al. Synthesis and Properties of Metalated Tetradehydrocorrins. Inorg. Chem. 2022, in press. [Google Scholar] [CrossRef]
  33. Tabolin, A.A.; Lesiv, A.V.; Khomutova, Y.A.; Nelyubina, Y.V.; Ioffe, S.L. Rearrangement of 3-Alkylidene-2-siloxy-tetrahydro-1,2-oxazines (ASENA). A New Approach Toward the Synthesis of 3-α-Hydroxyalkyl-5, 6-dihydro-4H-1,2-oxazines. Tetrahedron 2009, 65, 4578–4592. [Google Scholar] [CrossRef]
  34. Tabolin, A.A.; Leviv, A.V.; Khomutova, Y.A.; Ioffe, S.L. Synthesis of α-Prolinols and 2-Amino-1,5-diols from Primary Nitroalkanes and Other Simple Precursors via Intermediacy of 5,6-Dihydro-4H-1,2-oxazines. Synthesis 2012, 44, 1898–1906. [Google Scholar] [CrossRef] [Green Version]
  35. Malykhin, R.S.; Golovanov, I.S.; Nelyubina, Y.V.; Ioffe, S.L.; Sukhorukov, A.Y. Construction of Saturated Oxazolo[3,2-b]oxazines via Tandem [3+2]-Cycloaddition/[1,3]-Rearrangement of Cyclic Nitronates and Ketenes. J. Org. Chem. 2021, 86, 16337–16348. [Google Scholar] [CrossRef] [PubMed]
  36. Tabolin, A.A.; Sukhorukov, A.Y.; Ioffe, S.L.; Dilman, A.D. Recent Advances in the Synthesis and Chemistry of Nitronates. Synthesis 2017, 49, 3255–3268. [Google Scholar] [CrossRef]
  37. Weast, R.C. (Ed.) CRC Handbook of Chemistry and Physics, 52nd ed.; The Chemical Rubber Co.: Cleveland, OH, USA, 1971–1972; p. B-64. [Google Scholar]
  38. Bruker (2009) APEX2, SAINT and SADABS; Bruker AXS Inc.: Madison, WI, USA, 2002.
  39. Available online: https://onlinelibrary.wiley.com/iucr/doi/10.1107/S2053229614024218 (accessed on 2 June 2022).
  40. Available online: http://scripts.iucr.org/cgi-bin/paper?S0021889808042726 (accessed on 2 June 2022).
  41. Available online: https://docs.scipy.org/doc/scipy/reference/generated/scipy.optimize.fsolve.html (accessed on 2 June 2022).
Chart 1. Native bacteriochlorophyll a and the target synthetic bacteriochlorin (BC-1) with its immediate dihydrodipyrrin precursor (5a).
Chart 1. Native bacteriochlorophyll a and the target synthetic bacteriochlorin (BC-1) with its immediate dihydrodipyrrin precursor (5a).
Organics 03 00019 ch001
Scheme 1. Synthesis of bacteriochlorins with formation of the unexpected dihydrooxazine species 5b.
Scheme 1. Synthesis of bacteriochlorins with formation of the unexpected dihydrooxazine species 5b.
Organics 03 00019 sch001
Figure 1. Absorption spectra of BC-1 (red) and MeOBC-1 (blue) in toluene at room temperature (normalized at the B band).
Figure 1. Absorption spectra of BC-1 (red) and MeOBC-1 (blue) in toluene at room temperature (normalized at the B band).
Organics 03 00019 g001
Figure 2. ORTEP diagrams of dihydrodipyrrin 5a (left) and pyrrole–dihydrooxazine 5b (right) (50% probability at 100 K).
Figure 2. ORTEP diagrams of dihydrodipyrrin 5a (left) and pyrrole–dihydrooxazine 5b (right) (50% probability at 100 K).
Organics 03 00019 g002
Figure 3. 1H NMR spectra in CDCl3 at room temperature of precursor 4 (top) and products 5b (middle) and 5a (bottom).
Figure 3. 1H NMR spectra in CDCl3 at room temperature of precursor 4 (top) and products 5b (middle) and 5a (bottom).
Organics 03 00019 g003
Scheme 2. Nef and McMurry–Melton reactions (top) and McMurry–Melton extensions by Battersby and coworkers (bottom).
Scheme 2. Nef and McMurry–Melton reactions (top) and McMurry–Melton extensions by Battersby and coworkers (bottom).
Organics 03 00019 sch002
Scheme 3. Competing McMurry–Melton and Nef reactions.
Scheme 3. Competing McMurry–Melton and Nef reactions.
Organics 03 00019 sch003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tran, V.-P.; Matsumoto, N.; Nalaoh, P.; Jing, H.; Chen, C.-Y.; Lindsey, J.S. Dihydrooxazine Byproduct of a McMurry–Melton Reaction en Route to a Synthetic Bacteriochlorin. Organics 2022, 3, 262-274. https://doi.org/10.3390/org3030019

AMA Style

Tran V-P, Matsumoto N, Nalaoh P, Jing H, Chen C-Y, Lindsey JS. Dihydrooxazine Byproduct of a McMurry–Melton Reaction en Route to a Synthetic Bacteriochlorin. Organics. 2022; 3(3):262-274. https://doi.org/10.3390/org3030019

Chicago/Turabian Style

Tran, Vy-Phuong, Nobuyuki Matsumoto, Phattananawee Nalaoh, Haoyu Jing, Chih-Yuan Chen, and Jonathan S. Lindsey. 2022. "Dihydrooxazine Byproduct of a McMurry–Melton Reaction en Route to a Synthetic Bacteriochlorin" Organics 3, no. 3: 262-274. https://doi.org/10.3390/org3030019

Article Metrics

Back to TopTop