Birch-Type Reduction of Arenes in 2-Propanol ... - ACS Publications

Jul 2, 2019 - thermal expansion of the fluid and gas (air) (conditions B). These results indicated ... glass tube filled with different reaction mixtu...
0 downloads 0 Views 2MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: ACS Omega 2019, 4, 11522−11531

http://pubs.acs.org/journal/acsodf

Birch-Type Reduction of Arenes in 2‑Propanol Catalyzed by ZeroValent Iron and Platinum on Carbon Yoshinari Sawama,*,† Kazuho Ban,† Kazuhiro Akutsu-Suyama,‡ Hiroki Nakata,† Misato Mori,† Tsuyoshi Yamada,† Takahiro Kawajiri,† Naoki Yasukawa,† Kwihwan Park,† Yasunari Monguchi,†,∥ Yukio Takagi,§ Masatoshi Yoshimura,§ and Hironao Sajiki*,† †

Laboratory of Organic Chemistry, Gifu Pharmaceutical University, 1-25-4, Daigaku-nishi, Gifu 501-1196, Japan Neutron Science and Technology Center, Comprehensive Research Organization for Science and Society (CROSS), 162-1 Shirakata, Tokai-Mura, Naka-gun, Ibaraki 319-1106, Japan § Catalyst Development Center, N. E. Chemcat Corporation, 678 Ipponmatsu, Numazu, Shizuoka 410-0314, Japan

Downloaded via 85.202.195.119 on August 14, 2019 at 11:51:16 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



S Supporting Information *

ABSTRACT: Catalytic arene reduction was effectively realized by heating in 2propanol/water in the presence of Pt on carbon (Pt/C) and metallic Fe. 2Propanol acted as a hydrogen source, obviating the need for flammable (and hence, dangerous and hard-to-handle) hydrogen gas, while metallic Fe acted as an essential co-catalyst to promote reduction. The chemical states of Pt and Fe in the reaction mixture were determined by X-ray absorption near-edge structure analysis, and the obtained results were used to suggest a plausible reaction mechanism, implying that catalytic reduction involved Pt- and Fe-mediated single-electron transfer and the dehydrogenation of 2-propanol.

1. INTRODUCTION Arene reduction is a useful method of preparing cyclohexane derivatives, which are utilized as functional materials,1 liquid organic hydrogen carriers in the energy field,2 etc. However, the reduction (hydrogenation) of resonance-stabilized arenes generally requires the use of excess H2 gas,3 the industrial production of which from fossil fuels such as CH4 is accompanied by CO2 emission. Moreover, the transportation, storage, and usage of H2 gas are strictly regulated by law.4 Therefore, reduction not involving the utilization of H2 is obviously advantageous from the viewpoint of safety, ease of handling, and avoidance of special equipment usage. Therefore, catalytic transfer hydrogenation using biomass bearing a large amount of alcohol moiety and its analogue-like partial structure within the molecule has been eagerly investigated,5 and bimetallic Fe- and Pd-catalyzed reductions have also been reported.6 We have continuously developed methods of heterogeneously catalyzed dehydrogenation of organic substrates (e.g., alcohols7 and cyclohexane derivatives8) and investigated ways of mechanochemical H2 generation from water,9a,b alkanes,9c and ether9c to avoid CO2 emission. Furthermore, we demonstrated that in situ generated H2 can be directly utilized for reduction,9,10 as exemplified by Pt on carbon (Pt/C)-catalyzed arene reduction in 2-propanol (2PrOH) as a solvent and hydrogen source (Scheme 1, A).10b Unexpectedly, although this reduction proceeded smoothly in a sealed stainless steel tube, it could not be realized in a sealed glass tube. Our recent screening of metallic additives revealed that zero-valent Fe is an essential co-catalyst for arene reduction (Scheme 1, B). Although a bimetallic system © 2019 American Chemical Society

Scheme 1. Pt/C−Fe-Catalyzed Arene Reduction in 2Propanol as a Solvent and Hydrogen Source

containing Pt and Fe was previously utilized to control the catalyst activity in the hydrogenation of ethylene, 11a cinnamaldehyde,11b and other aldehydes,11c,d all of these reductions could be realized in the presence of Pt alone. Hydrogenations promoted by bimetallic catalysts comprising Fe and Pt group metals other than Pt (Rh,12a Pd,11c,12b,c and Ru11c) were also reported, and the performances of these catalysts were heavily reliant on the reaction conditions. For example, Pt−Fe11d and Pd−Fe12b catalysts inhibited the hydrogenation of furan and benzene rings. Herein, we demonstrate that Pt/C-catalyzed arene reduction can be switched on/off by the presence/absence of Fe and perform X-ray absorption near edge spectroscopy (XANES) analysis to suggest a mechanistic explanation. The metal-containing reaction apparatus, such as a sealed stainless steel tube, etc., is generally utilized in the reactions under high-pressure and Received: April 19, 2019 Accepted: June 20, 2019 Published: July 2, 2019 11522

DOI: 10.1021/acsomega.9b01130 ACS Omega 2019, 4, 11522−11531

ACS Omega

Article

high-temperature conditions. In contrast with the Pt/C- and Fe-catalyzed arene reduction, H−D exchange reaction of arenes could proceed without the reduction of arene moiety in the glass flask under the similar reaction conditions using Pt/C in 2-PrOH and D2O instead of H2O.13 Therefore, the present phenomenon is valuable to alert the chemists about the effect of reaction apparatus.

Table 2. Effect of Solvent on Arene Reduction Efficiency

2. RESULTS AND DISCUSSION n-Heptylbenzene (1a, 0.5 mmol) could not be catalytically reduced by 3 h of heating at 100 °C in a sealed glass tube containing Pt/C, 2-PrOH (4 mL), and H2O (2 mL) (Table 1,

1 2 3 4 5 6 7 8 9 10 11c 12d 13e

catalyst 10% 10% 10% 10% 10% 10% 10% 10% 10%

Pt/C Pt/C Pt/C Pt/C Pt/C Pt/C Pt/C Pd/C Rh/C

10% Pt/C 10% Pt/C 10% Pt/C

additive

1a/2aa (mol/mol)

Fe Zn Co Ni, Cr, Cu, Al, or Mg FeO Fe2O3 Fe Fe Fe Fe Fe Fe

no reaction 0/100 (98)b 76/24 89/11 no reaction 33/67 25/75 no reaction no reaction no reaction 47/53 no reaction 0/100 (97)b

solvent

1a/2aa (mol/mol)

1 2 3 4 5c

2-PrOH (4 mL)/H2O (2 mL) 1-PrOH (4 mL)/H2O (2 mL) t-BuOH (4 mL)/H2O (2 mL) H2O (6 mL) H2O (6 mL)

0/100 (98)b no reaction no reaction no reaction 11/89

a

Ratio was determined by 1H NMR. bIsolated yield. cUnder H2 atmosphere.

Table 1. Effect of Metal Co-catalyst on the Reduction of nHeptylbenzene

entry

entry

also completely performed using 3 mL of 2-PrOH, further decrease of 2-PrOH usage to 2 mL caused the incomplete reduction within 3 h (see the Supporting Information). On the other hand, the reduction was completely inhibited by the use of 1-PrOH or t-BuOH as co-solvents (Table 2, entries 2 and 3). Furthermore, pure H2O could not be employed as a hydrogen source (entry 4). Importantly, reduction in the 2PrOH−H2O mixture was more effective than reduction in pure H2O under H2 atmosphere (entry 5). The arene nuclei of cyclohexylbenzene (1b), t-butylbenzene (1c), and acetanilide (1d) effectively underwent Pt/C−Fecatalyzed reduction in 2-PrOH/H 2 O to produce the corresponding cyclohexane derivatives (2b−2d, Scheme 2). Scheme 2. Substrate Scope of the Reduction Reaction

Ratio was determined by 1H NMR. bIsolated yield. cAt 80 °C. dAt 50 °C. e7.4 mmol of 1a (1.3 g) was used as a substrate. a

entry 1). On the other hand, this reduction proceeded in the presence of the zero-valent Fe (5 mol %) as a main component of stainless steel to produce the corresponding cyclohexane derivative (2a) in excellent isolated yield (98%). The above reduction also proceeded in the presence of Zn and Co as cocatalysts, although lower conversions were obtained (entries 3 and 4). The use of Ni and Cr as other major stainless steel components was ineffective, as was the case for Cu, Al, and Mg (entry 5). Although 2a could be obtained in the presence of iron oxides (FeO and Fe2O3) in moderate yields (entries 6 and 7, respectively), these reductions were incomplete within 3 h. Heterogeneous platinum-group metal catalysts other than Pt/ C (e.g., Pd/C and Rh/C) were ineffective for the above reduction even in the presence of Fe (entries 8 and 9), and no conversion was observed when Fe was used as the only catalyst (entry 10). The deference of reactivity of Pt/C, Pd/C, or Rh/ C cannot be explained. Furthermore, the reduction efficiency decreased with decreasing reaction temperature (entry 11, 80 °C), and no reduction was observed at 50 °C (entry 12). Notably, gram-scale reduction of 1a (1.3 g, 7.4 mmol) afforded 2a in 97% isolated yield (entry 13). The above reduction could be carried out in a 2-PrOH (4 mL)−H2O mixture7a (Table 1, entry 2; Table 2, entry 1), which allowed one to decrease the risk posed by the pyrophoric nature of Pt/C. While the arene reduction was

a

10% Pt/C (10 mol %) and Fe0 (10 mol %) were used.

Anthracene (1e) was partially reduced to 1,2,3,4,5,6,7,8octahydroanthracene (2e) in good yield. Toluene (1f) could be smoothly transformed to methylcyclohexane (2f), which is viewed as a promising hydrogen storage material for creating a hydrogen society,2 without the external addition of hydrogen. In all cases, substrates (1b−1f) were completely reduced without any byproduct formation. Although the Pt/C−Fe-catalyzed reduction of stilbene (3) hardly proceeded in 2-PrOH/H2O at 25 or 40 °C (Table 3, entries 1 and 2, respectively), the olefin functionality of 3 was selectively reduced at 60 °C without the reduction of arene nuclei to produce 1,2-diphenylethane (4) in 98% isolated yield (entry 3). However, olefin reduction was hardly catalyzed at 60 °C by Pt/C in the absence of Fe (entry 4). On the other hand, both olefin and arene functionalities of 3 were reduced in the presence of Pt/C and Fe (5 mol % each) at 100 °C (12 h) to afford 1,2-dicyclohexylethane (5) in quantitative yield (entry 5). 7-Tetradecene (6), bearing no aromatic nucleus within the molecule, also efficiently underwent Pt/C−Fe-catalyzed 11523

DOI: 10.1021/acsomega.9b01130 ACS Omega 2019, 4, 11522−11531

ACS Omega

Article

Table 3. Reduction of Stilbene under Different Conditions

yield (%) entry

temp. (°C)

recovered 3

4

5

1 2 3 4a 5

25 40 60 60 100

100 75 0 96 0

0 18 98 3 0

0 0 0 0 99

a

Without Fe.

reduction of the olefin functionality at 60 °C to give tetradecane (7) in 88% yield, while 11% of 6 was recovered (Scheme 3). Complete reduction was observed at 100 °C, furnishing 7 in 97% isolated yield. Meanwhile, in the absence of Fe powder, 6 was completely recovered at 60 and 100 °C. Scheme 3. Reduction of 7-Tetradecene under Different Conditions

Figure 1. Time profile of the internal pressure inside a 50 mL sealed glass tube filled with different reaction mixtures.

Table 4. Gaseous Products Produced under the Conditions of Figure 1 after 6 h internal gas (mmol)

Subsequently, we compared the hydrogenation activities of Pt/C−Fe, Pt/C, and Fe in 1-PrOH (poor hydrogen source; Table 2, entry 2) and H2O at 100 °C under H2 (1 atm, 1 h; Scheme 4). In the case of Pt/C−Fe (5 mol % Pt and Fe each), Scheme 4. Comparison of Pt/C−Fe, Pt/C, and Fe Hydrogenation Activities

a

entry

conditions

H2

C3H8

1 2a 3 4 5 6

A B C D E F

0.33 not detected not detected trace 0.29 not detected

0.57 not detected not detected 116.2 2.7 not detected

For 3 h.

thermal conductivity detection (GC−TCD). The reduction of 1a (0.5 mmol) in 2-PrOH (4 mL)−H2O (2 mL) in the presence of 10% Pt/C (5 mol %: 0.025 mmol) and Fe (5 mol %: 0.025 mmol) at 100 °C was complete within 3 h (Table 1, entry 2). The internal pressure after 3 h reached ∼2 atm, and no further increase was observed after the completion of arene reduction (Figure 1, conditions A), with the yield of H2 after 6 h equaling only 0.33 mmol (Table 4, entry 1). Heating of the 2-PrOH-H2O mixture only resulted in a slight increase of the internal pressure up to ∼2 atm due to simple vaporization and thermal expansion of the fluid and gas (air) (conditions B). These results indicated that the developed arene reduction could safely proceed without the generation of excessive amounts of high-pressure flammable H2 gas. Since Fe did not catalyze H2 generation in 2-PrOH/H2O (conditions C), H2 generation by corrosion (oxidation) of Fe in the aqueous solution was ruled out.14 The simple use of Pt/C in 2-PrOH/ H2O resulted in the generation of a massive amount of propane (116.2 mmol) by Pt/C-catalyzed hydrodeoxygenation,15 while only traces of H2 were concomitantly produced (conditions D). On the other hand, the generation of propane

11% of 1a was reduced, and 89% of 1a was recovered, whereas 2a was obtained in 17% yield when Pt/C was used on its own, and no hydrogenation was observed when Fe was employed on its own. These results indicated that under the conditions of the Pt/C−Fe-catalyzed reaction, Fe does not facilitate Pt/Ccatalyzed hydrogenation, with H2 gas generated in situ by the Pt/C-catalyzed dehydrogenation of 2-PrOH (Table 1, entry 2). Thus, as has previously been reported,11 the accelerating effect of combining Pt and Fe on hydrogenation was not observed in the present arene reduction. Figure 1 shows the time profiles of the internal pressure in a 50 mL sealed glass tube containing 2-PrOH/H2O at 100 °C under various conditions, and Table 4 lists the total yields of the in situ generated H2 determined by gas chromatography 11524

DOI: 10.1021/acsomega.9b01130 ACS Omega 2019, 4, 11522−11531

ACS Omega

Article

the Supporting Information). Similarly, the analysis of Pt LIIIedge XANES spectra (Figure 4a) in comparison with the absorption peak energy (absorbance = 0.5)17 revealed that the absorption peaks of the samples A and B shifted to higher energies (i.e., to higher oxidation states) than that of zerovalent Pt metal (Figure 4b).18 As a result, the valence of Pt in both A and B samples after reduction was determined to be 1.51 and 1.32, respectively, which indicated that both A and B reactions were triggered by electron transfer from Pt and Fe metals to the aromatic ring to form partially oxidized Pt and Fe ions. Two reaction mechanisms were proposed based on these findings (Scheme 5). First, the Pt/C-catalyzed dehydrogenation of 2-PrOH produces H2, which can be consumed in the traditional Pt/C-catalyzed arene (1) hydrogenation3 and acetone (route a). In situ generated H2 rapidly reacts with acetone to afford 2-PrOH, which suppresses the accumulation of excess H2 in the vessel. Although the role of Fe is unclear, its addition promotes the dehydrogenation of 2-PrOH (Table 4, conditions A vs F) and suppresses the hydrodeoxygenation of 2-PrOH to propane, precluding the consumption of H2 required for arene hydrogenation (Table 4, conditions D vs E). As shown in Table 2, entries 2 and 3, arene reduction cannot be achieved when 2-PrOH is substituted for 1-PrOH or t-BuOH. This behavior is ascribed to the fact that Pt-group metal/C-catalyzed dehydrogenation of primary alcohols is accompanied by several side reactions such as decarbonylation of intermediate aldehyde derivatives,7b which can retard arene reduction, while tertiary alcohols without α-protons cannot act as H2 sources. Therefore, Pt/C−Fe-catalyzed arene reduction effectively proceeds only in 2-PrOH. Alternatively, arene reduction can take place via single electron transfer (SET) analogously to the Birch reduction (route b). SET from zero-valent Pt metal18 to the arene nucleus of the charge transfer complex (A) formed from 1, and Pt/C gives a radical anion intermediate (B) that is immediately protonated by 2-PrOH or H2O to form C. The subsequent second SET and protonation afford a 1,3-cyclohexadiene intermediate (E) complexed by PtII. PtII smoothly accepts electrons from Fe0 to regenerate Pt0 (F) and produce FeII, which is immediately reduced to Fe0 by H2 generated via Pt/ C-catalyzed dehydrogenation of 2-PrOH (route a). Repetition of SET and protonation processes afford cyclohexane product 2 and regenerated Pt0 (catalytic cycle 1). The PtI intermediate C can also be reduced to Pt0 by Fe0 to produce intermediate G and FeI. Subsequent SET within G and the protonation of anion intermediate H furnishes a diene−PtI complex (I). PtI can then be reduced by FeI to generate a diene−Pt0 complex J

could be suppressed by the use of Fe powder along with Pt/C (conditions E). Furthermore, the addition of 1a to the solvent mixture containing Pt/C but no Fe powder suppressed the generation of C3H8 and H2, while no reduction of the aromatic nucleus was observed (conditions F). Probably, Pt metal was strongly coordinated by arene nuclei, which inhibited the Pt/ C-catalyzed dehydrogenation and hydrodeoxygenation of 2PrOH. To elucidate the mechanism of arene reduction, the mean valence of Fe in reaction mixtures was analyzed by XANES. As noted in the Introduction section, the Pt/C-catalyzed reduction of 1a proceeded only in a stainless steel vessel, while the addition of Fe (5 mol %) allowed this reduction to be performed in a glass tube. Figure 2 shows the X-ray

Figure 2. X-ray fluorescence profiles of the mixtures A and B (see Scheme 1).

fluorescence profiles of the reaction mixtures A and B (Scheme 1) after 3 h of heating, revealing that these mixtures contained 4.4 and 5.0 mmol Fe, respectively, as determined from the peak integral values. These findings clearly confirmed the leaching of Fe from stainless steel during the reaction. The valences of Fe and Pt in the mixtures A and B after reduction were determined from Fe K-edge (Figure 3) and Pt LIII-edge (Figure 4) XANES spectra of these mixtures using Fe0 powder, Pt0 foil, FeCl3, and PtCl4 as standards (detailed descriptions are provided in the Supporting Information). Compared with each energy based on the absorption edge energy shift (normalized absorbance = 0.5 in Figure 3a), the Fe K-edge absorptions of the samples A and B were shifted to higher energies than that of the zero-valent Fe,16 and the mean valence of Fe in these samples was estimated as 1.57 and 1.50, respectively (Figure 3b). X-ray absorption fine structure (XAFS) and polarized neutron reflectivity (PNR) analyses also strongly supported the generation of oxidized Fe ions during the reduction (see

Figure 3. (a) Normalized Fe K-edge XANES profiles. (b) Plot of Fe K-edge absorption edge energy shift vs the valence of Fe. Reaction conditions A and B are described in Scheme 1. 11525

DOI: 10.1021/acsomega.9b01130 ACS Omega 2019, 4, 11522−11531

ACS Omega

Article

Figure 4. (a) Normalized Pt LIII-edge XANES profiles. (b) Plot of Pt LIII-edge absorption edge energy shift vs the valence of Pt. Reaction conditions A and B are described in Scheme 1.

corresponding diene intermediate (K) was smoothly converted to toluene (8; 0.99 mmol) as the sole product via the dehydrogenation of K under conditions of 2-PrOH-free aqueous Pt/C catalysis in both the presence and absence of Fe (Scheme 6).19 K2PtCl4 and PtO2 also catalyzed the

Scheme 5. Proposed Mechanisms or Pt/C−Fe-Catalyzed Arene Reduction

Scheme 6. Pt-Catalyzed Dehydrogenative Regeneration of Toluene (8) from Cyclohexadiene Derivative 9

dehydrogenation of 9, which resulted in the formation of 8. Furthermore, the use of 1-PrOH or t-BuOH instead of 2PrOH should afford diene intermediates equivalent to K, since in these cases, Fe0 cannot be regenerated from FeII because of the poor (or absence of) H2 generation ability of these alcohols. The regeneration of 1 from K also causes the complete recovery of 1 when 1-PrOH and t-BuOH are used as solvents (Table 2, entries 2 and 3). Meanwhile, the (Pt/C (5 mol %) + 2-PrOH)-mediated reduction of the olefin functionality of stilbene (3) in the absence of Fe proceeded to an insignificant extent to produce 1,2-diphenylethane (4, 3% yield) via the SET mechanism not involving redox shuttling between PtI/PtII and Pt0 (Table 3, entry 4). This finding was ascribed to the fact that the dehydrogenation of the saturated ethylene function of 4 to 3 was not catalyzed by Pt/C in H2O (Scheme 7). The consecutive generation of appropriate quantities of H2 via Pt/C−Fe-catalyzed dehydrogenation of 2-PrOH for the regeneration of Fe0 from FeII was required to

(catalytic cycle-2), and the resulting FeII species are reduced to Fe0 by H2 derived from 2-PrOH. The mean valence of Pt (1.32 and 1.51) determined by XANES analysis supports the operation of catalytic cycle 1, which involves redox shuttling between Pt0 and PtII. The release of PtII from complex E in the absence of catalytic Fe powder should afford traces of diene intermediate K. Notably, 1a was completely recovered under the conditions of Pt/C catalysis in the absence of Fe (Table 1, entry 1). On the other hand, unsaturated cyclohexane derivatives such as cyclohexadiene and cyclohexene easily undergo Pt-group metal-catalyzed dehydrogenative aromatization in H2O to form the corresponding benzene derivatives.8a Therefore, diene intermediate K might be immediately transformed to 1 via Pt-catalyzed dehydrogenation. Indeed, a mixture of toluene (8; 0.95 mmol) and 1-methyl-1,4cyclohexadiene (9; 0.05 mmol) as a model of the

Scheme 7. Results of 1,2-Diphenylethane (4) Reduction under the Conditions of Pt/C Catalysis

11526

DOI: 10.1021/acsomega.9b01130 ACS Omega 2019, 4, 11522−11531

ACS Omega

Article

Scheme 8. Reaction Using Deuterated Solventsa

a

Italic number means the deuterium content

3. CONCLUSIONS Pt/C−Fe-catalyzed arene reduction with 2-PrOH as a hydrogen source under ambient-pressure conditions was developed. The Pt/C-catalyzed in situ production of H2 via the dehydrogenation of 2-PrOH was shown to play a critical role in the regeneration of the composite catalyst to promote arene reduction. Since only small amounts of H2 accumulated in the reaction vessel after reduction, the present method was concluded to be useful from the viewpoint of safety. The presence of a catalytic amount of Fe was found to be crucial for the external hydrogen-free reduction of aromatic nuclei in aqueous 2-PrOH.

maintain the catalytic cycle for effective arene reduction, although a FeI, FeII, and FeIII cycle for the repeatable regeneration of Pt0 cannot be ruled out. Route b seems to be preferred over route a, since (a) reduction with 2-PrOH as a hydrogen source is faster than that with excess amount of H2 gas (Table 2, entry 1 vs 5) and (b) a small amount of H2 is generated by the use of Pt/C and Fe in 2-PrOH/H2O (Table 4, entry 5). Whereas the traditional Birch reduction of arenes by stoichiometric amounts of Na in liquid NH3 gives 1,4-cyclohexadienes via SET followed by protonation, the present reduction method directly affords cyclohexane derivatives.20 The H−D exchange reaction of acetanilide (1d) could proceed in the presence of Pt/C in a 2-PrOH and D2O-mixed solvent without arene reduction to give the deuterated acetanilide (1da-dn) (Scheme 8, top).13a Under the present arene reduction conditions using Pt/C and Fe as catalysts in D2O instead of H2O, both H−D exchange reaction and arene reduction were carried out to give the deuterium-labeled cyclohexane derivatives (2db-dn) with moderate D content (middle). Meanwhile, the use of 2-PrOD-d8 instead of 2-PrOH reduced the reaction efficiencies of both arene reduction and H−D exchange reaction (bottom). Probably, the Pt/Ccatalyzed dehydrogenation of 2-PrOD-d8 was suppressed due to the isotopic effect causing a stronger C−D bond at the αposition of 2-PrOD-d8, and the enough amount of deuterium, which is essential to regenerate the zero-valent catalyst (Scheme 5), was not produced. The low D contents of 2dcdn indicated that the protonation (Scheme 5) is mainly derived from water. The similar Pt/C-catalyzed reaction conditions using 2-PrOH and D2O may induce the H−D exchange reaction of hydrocarbon, such as cyclohexane moiety.13d Namely, the effect using deuterated solvents cannot be clearly explained, since three reactions (H−D exchange reactions of arene and hydrocarbon and arene reduction) can proceed under Pt/C- and Fe-catalyzed reaction conditions. At least, it is certain that the additional Fe is required to facilitate the desired Pt/C-catalyzed arene reduction using 2-PrOH.

4. EXPERIMENTAL SECTION 4.1. General Information. Ten percent Pt/C, Pd/C, Rh/ C, and Ru/C were supplied by the N. E. Chemcat Corporation (Tokyo, Japan). 1-PrOH, 2-PrOH, t-BuOH, and water were purchased from commercial sources and used without further purification. All of the commercially available reagents were used without further purification. Reactions were carried out in 50 mL glass sealed tube [tinyclave steel (Büchiglasuster, Switzerland)] or 15 × 150 m/m test tube. 1H NMR spectra were recorded on a JEOL ECZ 400 or ECA 500 spectrometer at room temperature in CDCl3 as a solvent and internal standard (1H NMR: δ = 7.26 for CDCl3) with tetramethylsilane as a further internal standard. A JMS Q1000 GC [7890A gas chromatography (Agilent Technologies, USA) equipped with a JEOL MK II mass spectrometer (JEOL Co., Ltd., Japan)] and an Inert Cap5MS/sil capillary column (30 × 0.25 mm i.d., 0.25 μm film thickness; GL Science, Japan) were used for toluene and methylcyclohexane analyses. GC-3200 [gas chromatography equipped with thermal conductivity detector (GC/TCD; GL Science, Japan)] was used for gas analysis with Molecular Sieve 5 Å (60/80 mesh) packed column (3 × 2.2 mm i.d., 1/8 inch: GL Science, Japan) for H2 analysis, Porapak Q (80/100 mesh) packed column (2 × 2.2 mm i.d., 1/8 inch: GL Science, Japan) for C3H8 analysis. EXAFS measurements of solutions were performed using the fluorescence mode at the BL11S2 Hard X-ray EXAFS beamline station in the Aichi SR. 11527

DOI: 10.1021/acsomega.9b01130 ACS Omega 2019, 4, 11522−11531

ACS Omega

Article

amounts of H2 and C3H8 were detected from the analysis of the headspace gas by GC/TCD. Helium was employed as a carrier gas at 220 kPa (molecular sieve 5 Å) and 140 kPa (Porapak Q). The injector and detector temperatures were 100 °C. The column temperature was programmed to 70 °C. One milliliter of sample gas was injected. The products were identified by their GC/TCD retention times in comparison to those of authentic commercial samples. Retention times (min) were 1.3 (H2, Molecular Sieve 5 Å) and 13.7 (C3H8, Porapack Q). 4.4. General Procedure for Arene Reduction under Hydrogen Condition (Scheme 4). A 50 mL sealed glass tube was sequentially charged with n-heptylbenzene (1a; 88.3 mg, 0.50 mmol), Fe0 (1.4 mg, 0.025 mmol, 5 mol %), 10% Pt/ C (48.8 mg, 0.025 mmol, 5 mol %), H2O (2 mL), and 1-PrOH (4 mL). The inside air was replaced with H2 (balloon) in five vacuum/argon cycles, and the suspension was stirred at 100 °C. After stirring for 1 h, the mixture was cooled to room temperature and filtered through a membrane filter (Millipore, Millex-LH, 0.2 μm) to remove the catalysts. The filtrate was extracted with Et2O (20 mL), diluted with H2O (20 mL), and then the aqueous layer was further extracted with Et2O (3 × 10 mL). The organic layer was dried over anhydrous MgSO4, filtered, and concentrated in vacuo to give the cyclohexane product. 4.5. General Procedure for the Dehydrogenation of Cyclohexadiene Derivative (9) Mixed with Small Amount of Arene (8) (Scheme 6, Equation 1). A 30 mL co-plug test tube was sequentially charged with toluene (8; 100 μL, 0.95 mmol), 1-methyl-1,4-cyclohexadiene (9; 5.6 μL, 0.05 mmol), Fe0 (2.8 mg, 0.050 mmol), 10% Pt/C (97.6 mg, 0.050 mmol) and H2O (6 mL), and the suspension was stirred at 100 °C. After stirring for 1 h, the mixture was cooled to room temperature and anisole (100 μL) added as an internal standard. The mixture was filtered through a membrane filter (Millipore, Millex-LH, 0.2 μm) to remove the catalysts. The filtrate was transferred to a 100 mL volumetric flask with Et2O. Twenty microliters of the sample was dissolve in 1.5 mL of Et2O. Compound 9 was detected by GC/MS. 4.6. General Procedure for Dehydrogenation of 1,2Diphenylethane (4) Mixed with Small Amount of Stilbene (3) (Scheme 6, Equation 2). A 50 mL sealed glass tube was sequentially charged with stilbene (3; 171.2 mg, 0.95 mmol), 1,2-diphenylethane (4; 9.1 mg, 0.05 mmol), Fe powder (2.8 mg, 0.050 mmol), 10% Pt/C (97.6 mg, 0.050 mmol), and H2O (6 mL), and the suspension was stirred at 100 °C. After stirring for 1 h, the mixture was cooled to room temperature and filtered through a membrane filter (Millipore, Millex-LH, 0.2 μm) to remove the catalysts. The filtrate was extracted with Et2O (20 mL), diluted with H2O (20 mL), and the aqueous layer was further extracted with Et2O (3 × 10 mL). The organic layer was dried over anhydrous MgSO4, filtered, and concentrated in vacuo to give the products. 4.7. Spectroscopic Data of the Products. 4.7.1. nHeptylcyclohexane (2a) in Tables 1 and 2. n-Heptylbenzene (1a; 88.3 mg, 0.50 mmol), Fe0 (5 mol %), 10% Pt/C (5 mol %), H2O (2 mL), and 2-PrOH (4 mL) were used, and the reaction was carried out according to the general procedure shown in Section 4.2.1 to give 2a (92.2 mg, 0.49 mmol) in 98% yield. Scale-up study (Table 1, entry 13): A 200 mL two-neck round-bottom flask was sequentially charged with n-heptylbenzene (1a; 1.30 g, 7.40 mmol), Fe powder (21.3 mg, 0.38 mmol, 5 mol %), 10% Pt/C (731.4 mg, 0.37 mmol, 5 mol %),

The incident X-rays were monitored by an ionization chamber (14 cm in length) filled with N2 gas. 4.2. General Procedure for Arene Reduction (Tables 1−3; Schemes 2 and 3). 4.2.1. Reduction of nHeptylbenzene, Cyclohexylbenzene, tert-Butylbenzene, or Anthracene. A 50 mL sealed glass tube was sequentially charged with arene (0.50 mmol), Fe0 (1.4 mg, 0.025 mmol, 5 mol %), 10% Pt/C (48.8 mg, 0.025 mmol, 5 mol %), H2O (2 mL), and 2-PrOH (4 mL), and the suspension was stirred at 100 °C (Caution: the addition of 2-PrOH in the absence of H2O causes the ignition of Pt/C). After stirring for a specific reaction time, the mixture was cooled to room temperature and filtered through a membrane filter (Millipore, Millex-LH, 0.2 μm) to remove catalysts. The filtrate was extracted with Et2O (20 mL) diluted with H2O (20 mL), and the aqueous layer was further extracted with Et2O (3 × 10 mL). The organic layer was dried over anhydrous MgSO4, filtered, and concentrated in vacuo to give the cyclohexane product. 4.2.2. Reduction of Toluene. A 50 mL sealed glass tube was sequentially charged with toluene (1f; 46.1 mg, 0.50 mmol), Fe0 (2.8 g, 0.050 mmol, 10 mol %), 10% Pt/C (97.6 g, 0.050 mmol, 10 mol %), H2O (2 mL), and 2-PrOH (4 mL), and the suspension was stirred at 100 °C (Caution: the addition of 2PrOH in the absence of H2O causes the ignition of Pt/C). After stirring for 3 h, the mixture was cooled to room temperature and anisole added (50 μL) as an internal standard. The mixture was filtered through a membrane filter (Millipore, Millex-LH, 0.2 μm) to remove catalysts. The filtrate was transferred to a 50 mL volumetric flask with Et2O. Twenty microliters of the sample was dissolve in 1.5 mL of Et2O. 2f was determined by GC-MS and yield was 82%. Helium was employed as a carrier gas at the flow rate of 1.3 mL/min. The injector and detector temperatures were 280 °C. The column temperature was programmed to ramp from 35 °C (1 min hold) to 110 °C (3 min hold) at the rate of 5 °C/min and then to 280 °C (1.5 min hold) at the rate of 30 °C/min. One microliter of the sample solution was injected (split; 1:10). The products were identified by their GC/MS retention times in comparison to those of authentic commercial samples. Retention times (min) were 3.53 (methylcyclohexane), 4.14 (toluene), and 8.40 (internal standard) (Anisole). 4.2.3. Reduction of Stilbene and 7-Tetradecene. A test tube was sequentially charged with alkene (0.25 mmol), Fe0 (0.7 mg, 0.013 mmol, 5 mol %), 10% Pt/C (24.4 mg, 0.013 mmol, 5 mol %), H2O (1 mL), and 2-PrOH (2 mL), and the suspension was stirred at 100 °C (Caution: the addition of 2PrOH in the absence of H2O causes the ignition of Pt/C). After stirring for a specific reaction time, the mixture was cooled to room temperature and filtered through a membrane filter (Millipore, Millex-LH, 0.2 μm) to remove the catalysts. The filtrate was extracted with Et2O (20 mL), diluted with H2O (20 mL), and the aqueous layer was further extracted with Et2O (3 × 10 mL). The organic layer was dried over anhydrous MgSO4, filtered, and concentrated in vacuo to give the alkane product. 4.3. Analysis of the Internal Gas Components (Figure 1 and Table 4). A 50 mL sealed glass tube was sequentially charged with n-heptylbenzene (1a; 88.3 mg, 0.50 mmol), Fe0 (1.4 mg, 0.025 mmol, 5 mol %), 10% Pt/C (48.8 mg, 0.025 mmol, 5 mol %), H2O (2 mL), and 2-PrOH (4 mL), and the suspension was stirred at 100 °C. Pressure gauge on the glass sealed tube was checked at 5, 10, 30 min, 1, 2, 3, 4, 5, and 6 h. After 6 h, the mixture was cooled to room temperature and the 11528

DOI: 10.1021/acsomega.9b01130 ACS Omega 2019, 4, 11522−11531

ACS Omega

Article

0.84 (m, 4H). 1H NMR spectrum of the product was identical to that of the ref 10b. 4.7.8. Tetradecane (7) in Scheme 3. 7-Tetradecene (6; 49.1 mg, 0.25 mmol), Fe powder (5 mol %), 10% Pt/C (5 mol %), H2O (1 mL), and 2-PrOH (2 mL) were used, and the reaction was carried out according to the general procedure shown in Section 4.2.3 to give 7 (47.1 mg, 0.24 mmol) in 97% yield. Colorless oil; 1H NMR (400 MHz, CDCl3): δ 1.32−1.26 (m, 24H), 0.90−0.86 (m, 6H). 1H NMR spectrum of the product was identical to that of the ref 22.

H2O (15 mL), and 2-PrOH (30 mL). The air inside was replaced with argon (balloon) by five vacuum/argon cycles, and the suspension was stirred at 100 °C. After stirring for 4 h, the mixture was cooled to room temperature and filtered through celite pad to remove catalysts. The filtrate was extracted with Et2O (40 mL) and H2O (40 mL), and then the aqueous layer was further extracted with Et2O (3 × 20 mL). The organic layer was dried over anhydrous MgSO4, filtered, and concentrated in vacuo to give 2a (1.32 mg, 7.21 mmol) in 97% yield. Colorless oil; 1H NMR (400 MHz, CDCl3): δ 1.75−1.67 (m, 5H), 1.31−1.12 (m, 16H), 0.94−0.85 (m, 5H). 1 H NMR spectrum of the product was identical to that of the ref 10b. 4.7.2. Bicyclohexyl (2b) in Scheme 2. Cyclohexylbenzene (1b; 80.1 mg, 0.50 mmol), Fe0 (1.4 mg), 10% Pt/C (5 mol %), H2O (2 mL), and 2-PrOH (4 mL) were used, and the reaction was carried out according to the general procedure shown in Section 4.2.1 to give 2b (75.1 mg, 0.45 mmol) in 90% yield. Colorless oil; 1H NMR (500 MHz, CDCl3): δ 1.74−1.62 (m, 10H), 1.24−0.89 (m, 12H). 1H NMR spectrum of the product was identical to that of the ref 10b. 4.7.3. tert-Butylcyclohexane (2c) in Scheme 2. tertButylbenzene (1c; 67.1 mg, 0.50 mmol), Fe0 (5 mol %), 10% Pt/C (5 mol %), H2O (2 mL), and 2-PrOH (4 mL) were used, and the reaction was carried out according to the general procedure shown in Section 4.2.1 to give 2c (53.5 mg, 0.382mmol) in 78% yield. Colorless oil; 1H NMR (500 MHz, CDCl3): δ 1.76−1.63 (m, 5H), 1.26−0.84 (m, 15H). 1H NMR spectrum of the product was identical to that of the ref 3s. 4.7.4. N-Cyclohexylacetamide (2d) in Scheme 2. Acetanilide (1d; 67.6 mg, 0.50 mmol), Fe0 (5 mol %), 10% Pt/C (5 mol %), H2O (2 mL), and 2-PrOH (4 mL) were used, and the reaction was carried out according to the general procedure shown in Section 4.2.1 to give 2d (66.1 mg, 0.47 mmol) in 94% yield. Colorless solid; 1H NMR (500 MHz, CDCl3): δ 6.09 (brs, 1H), 3.68−3.63 (m, 1H), 1.88−1.81 (m, 5H), 1.64− 1.52 (m, 3H), 1.29−1.02 (m, 5H). 1H NMR spectrum of the product was identical to that of the ref 10b. 4.7.5. 1,2,3,4,5,6,7,8-Octahydroanthracene (2e) in Scheme 2. Anthracene (1e; 89.1 mg, 0.50 mmol), Fe0 (5 mol %), 10% Pt/C (5 mol %), H2O (2 mL), and 2-PrOH (4 mL) were used, and the reaction was carried out according to the general procedure shown in Section 4.2.1 to give 2e (85.6 mg, 0.46 mmol) in 92% yield. Colorless solid; 1H NMR (500 MHz, CDCl3): δ 6.82 (s, 2H), 2.73 (m, 8H), 1.80 (m, 8H). 1H NMR spectrum of the product was identical to that of the ref 10b. 4.7.6. 1,2-Diphenylethane (4) in Table 3, Entry 3. Stilbene (3; 45.1 mg, 0.25 mmol), Fe powder (5 mol %), 10% Pt/C (5 mol %), H2O (1 mL), and 2-PrOH (2 mL) were used, and the reaction was carried out according to the general procedure shown in Section 4.2.3 to give 4 (44.7 mg, 0.25 mmol) in 98% yield. Colorless solid; 1H NMR (400 MHz, CDCl3): δ 7.33− 7.20 (m, 10H), 2.95 (s, 4H). 1H NMR spectrum of the product was identical to that of the ref 21. 4.7.7. 1,2-Dicyclohexylethane (5) in Table 3, Entry 4. Stilbene (3; 45.1 mg, 0.25 mmol), Fe powder (5 mol %), 10% Pt/C (5 mol %), H2O (1 mL), and 2-PrOH (2 mL) were used, and the reaction was carried out according to the general procedure shown in Section 4.2.3 to give 5 (48.1 mg, 0.25 mmol) in 99% yield. Colorless oil; 1H NMR (400 MHz, CDCl3): δ 1.70−1.62 (m, 10H), 1.26−1.12 (m, 12H), 0.89−



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.9b01130. XANES, EXAFS, and PNR analysis, XPS of fresh 10% Pd/C, mechanism study, and spectroscopic data of the products (PDF)



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (Y.S.). *E-mail: [email protected]. Phone/Fax: (+81)-58-230- 8109 (H.S.). ORCID

Yoshinari Sawama: 0000-0002-9923-2412 Kazuhiro Akutsu-Suyama: 0000-0002-4797-6604 Tsuyoshi Yamada: 0000-0002-6048-5578 Yasunari Monguchi: 0000-0002-2141-3192 Hironao Sajiki: 0000-0003-2792-6826 Present Address ∥

Laboratory of Organic Chemistry, Daiichi University of Pharmacy, 22-1 Tamagawa-cho, Minami-ku, Fukuoka 8158511, Japan (Y.M.).

Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This study was partially supported by a Grant-in-Aid for Scientific Research (C) from the Japan Society for the Promotion of Science (JSPS: 16K08169 for Y.S.) and the Sasakawa Scientific Research Grant from The Japan Science Society (2019−3032 for K.B.). XAFS measurements were performed at the Knowledge Hub Aichi of the Aichi Synchrotron Radiation Center (proposal nos. 201705053 and 201802072). We thank technical staff for their kind support with XAFS spectra acquisition. PNR experiments were performed at the BL17 SHARAKU of J-PARC MLF under proposal no. 2017I0017. We thank N. E. Chemcat Corp. for their kind gift of catalysts. We would like to thank Editage (www.editage.jp) for English language editing.



REFERENCES

(1) (a) Singh, V.; Iyer, S. R.; Pal, S. Recent Approaches Towards Synthesis of cis-Decalins. Tetrahedron 2005, 61, 9197−9231. (b) Weissermel, K.; Arpe, H. J. Industrial Organic Chemistry; VCH: New York, 1993. (2) (a) Alhumaidan, F.; Cresswell, D.; Garforth, A. Hydrogen Storage in Liquid Organic Hydride: Producing Hydrogen Catalytically from Methylcyclohexane. Energy Fuels 2011, 25, 4217−4234. 11529

DOI: 10.1021/acsomega.9b01130 ACS Omega 2019, 4, 11522−11531

ACS Omega

Article

(b) Preuster, P.; Alekseev, A.; Wasserscheid, P. Hydrogen Storage Technologies for Future Energy Systems. Annu. Rev. Chem. Biomol. Eng. 2017, 8, 445−471. (3) Arene reduction using H2; Recent selected papers: (a) Wiesenfeldt, M. P.; Nairoukh, Z.; Daloton, T.; Glorius, F. Selective Arene Hydrogenation Provides Direct Access to Saturated Carbo- and Heterocycles. Angew. Chem., Int. Ed. 2019, DOI: 10.1002/ anie.201814471. (b) Tran, B. L.; Fulton, J. L.; Linehan, J. C.; Lercher, J. A.; Bullock, R. M. Rh (CAAC)-Catalyzed Arene Hydrogenation: Evidence for Nanocatalysis and Sterically Controlled Site-Selective Hydrogenation. ACS Catal. 2018, 8, 8441−8449. (c) Ohtaka, A.; Kawase, M.; Aihara, S.; Miyamoto, Y.; Terada, A.; Nakamura, K.; Hamasaka, G.; Uozumi, Y.; Shinagawa, T.; Shimomura, O.; Nomura, R. Poly(tetrafluoroethylene)-Stabilized Metal Nanoparticles: Preparation and Evaluation of Catalytic Activity for Suzuki, Heck, and Arene Hydrogenation in Water. ACS Omega 2018, 3, 10066−10073. (d) Miyamura, H.; Suzuki, A.; Yasukawa, T.; Kobayashi, S. Polysilane-Immobilized Rh−Pt Bimetallic Nanoparticles as Powerful Arene Hydrogenation Catalysts: Synthesis, Reactions under Batch and Flow Conditions and Reaction Mechanism. J. Am. Chem. Soc. 2018, 140, 11325−11334. (e) Joannou, M. V.; Bezdek, M. J.; Chirik, P. J. Pyridine(diimine) Molybdenum-Catalyzed Hydrogenation of Arenes and Hindered Olefins: Insights into Precatalyst Activation and Deactivation Pathways. ACS Catal. 2018, 8, 5276− 5285. (f) Ghosh, S.; Jagirdar, B. R. Effect of the Crystallographic Phase of Ruthenium Nanosponges on Arene and Substituted-Arene Hydrogenation Activity. ChemCatChem 2018, 10, 3086−3095. (g) Ji, P.; Song, Y.; Drake, T.; Veroneau, S. S.; Lin, Z.; Pan, X.; Lin, W. Titanium(III)-Oxo Clusters in a Metal−Organic Framework Support Single-Site Co(II)-Hydride Catalysts for Arene Hydrogenation. J. Am. Chem. Soc. 2018, 140, 433−440. (h) Tang, N.; Cong, Y.; Shang, Q.; Wu, C.; Xu, G.; Wang, X. Coordinatively Unsaturated Al3+ Sites Anchored Subnanometric Ruthenium Catalyst for Hydrogenation of Aromatics. ACS Catal. 2017, 7, 5987−5991. (i) Cui, X.; Surkus, A.-E.; Junge, K.; Topf, C.; Radnik, J.; Kreyenschulte, C.; Beller, M. Highly selective hydrogenation of arenes using nanostructured ruthenium catalysts modified with a carbon−nitrogen matrix. Nat. Commun. 2016, 7, No. 11326. (j) Karakhanov, E. A.; Maximov, A. L.; Zolotukhina, A. V.; Terenina, M. V.; Vutolkina, A. V. Nanoheterogeneous ruthenium-containing catalysts based on dendrimers in the hydrogenation of aromatic compounds under two-phase conditions. Pet. Chem. 2016, 56, 491−502. (k) Ibrahim, M.; Poreddy, R.; Philippot, K.; Riisager, A.; Garcia-Suarez, E. J. Chemoselective hydrogenation of arenes by PVP supported Rh nanoparticles. Dalton Trans. 2016, 45, 19368−19373. (l) Morioka, Y.; Matsuoka, A.; Binder, K.; Knappett, B. R.; Wheatley, A. E. H.; Naka, H. Selective hydrogenation of arenes to cyclohexanes in water catalyzed by chitinsupported ruthenium nanoparticles. Catal. Sci. Technol. 2016, 6, 5801−5805. (m) Pélisson, C.-H.; Denicourt-Nowicki, A.; Roucoux, A. Magnetically Retrievable Rh(0) Nanocomposite as Relevant Catalyst for Mild Hydrogenation of Functionalized Arenes in Water. ACS Sustainable Chem. Eng. 2016, 4, 1834−1839. (n) Shi, J.; Zhao, M.; Wang, Y.; Fu, J.; Lu, X.; Hou, Z. Upgrading of aromatic compounds in bio-oil over ultrathin graphene encapsulated Ru nanoparticles. J. Mater. Chem. A 2016, 4, 5842−5848. (o) Sun, B.; Süss-Fink, G. Ruthenium-catalyzed hydrogenation of aromatic amino acids in aqueous solution. J. Organomet. Chem. 2016, 81−86. (p) Baghbanian, S. M.; Farhang, M.; Vahdat, S. M.; Tajbakhsh, M. Hydrogenation of arenes, nitroarenes, and alkenes catalyzed by rhodium nanoparticles supported on natural nanozeolite clinoptilolite. J. Mol. Catal. A: Chem. 2015, 128−136. (q) Martínez-Prieto, L. M.; Urbaneja, C.; Palma, P.; Cámpora, J.; Philippot, K.; Chaudret, B. A betaine adduct of Nheterocyclic carbene and carbodiimide, an efficient ligand to produce ultra-small ruthenium nanoparticles. Chem. Commun. 2015, 51, 4647−4650. (r) Kang, X.; Zhang, J.; Shang, W.; Wu, T.; Zhang, P.; Han, B.; Wu, Z.; Mo, G.; Xing, X. One-Step Synthesis of Highly Efficient Nanocatalysts on the Supports with Hierarchical Pores Using Porous Ionic Liquid-Water Gel. J. Am. Chem. Soc. 2014, 136, 3768− 3771. (s) Maegawa, T.; Akashi, A.; Yaguchi, K.; Iwasaki, Y.;

Shigetsura, M.; Monguchi, Y.; Sajiki, H. Efficient and Practical Arene Hydrogenation by Heterogeneous Catalysts under Mild Conditions. Chem. - Eur. J. 2009, 15, 6953−6963. (t) Maegawa, T.; Akashi, A.; Sajiki, H. A Mild and Facile Method for Complete Hydrogenation of Aromatic Nuclei in Water. Synlett 2006, 9, 1440− 1442. (4) For selected papers, see: (a) Jie, X.; Gonzalez-Cortes, S.; Xiao, T.; Wang, J.; Yao, B.; Slocombe, D. R.; Al-Megren, H. A.; Dilworth, J. R.; Thomas, J. M.; Edwards, P. P. Rapid Production of High-Purity Hydrogen Fuel through Microwave-Promoted Deep Catalytic Dehydrogenation of Liquid Alkanes with Abundant Metals. Angew. Chem., Int. Ed. 2017, 56, 10170−10173. (b) Gonzalez-Cortes, S.; Slocombe, D. R.; Xiao, T.; Aldawsari, A.; Yao, B.; Kuznetsov, V. L.; Leberti, E.; Kirkland, A. I.; Alkinani, M. S.; Al-Megren, H. A.; Thomas, J. M.; Edwards, P. P. Wax: A benign hydrogen-storage material that rapidly releases H2-rich gases through microwaveassisted catalytic decomposition. Sci. Rep. 2016, 6, No. 35315. (c) Susanti, R. F.; Dianningrum, L. W.; Yum, T.; Kim, Y.; Lee, Y.-W.; Kim, J. High-yield hydrogen production by supercritical water gasification of various feedstocks: Alcohols, glucose, glycerol and long-chain alkanes. Chem. Eng. Res. Des. 2014, 92, 1834−1844. (d) Ahmed, S.; Aitani, A.; Rahman, F.; Al-Dawood, A.; Al-Muhaish, F. Decomposition of hydrocarbons to hydrogen and carbon. Appl. Catal., A 2009, 359, 1−24. (5) (a) Espro, C.; Gumina, B.; Szumelda, T.; Paone, E.; Mauriello, F. Catalytic transfer hydrogenolysis as an effective tool for the reductive upgrading of cellulose, hemicellulose, lignin, and their derived molecules. Catalysts 2018, 3, 313. (b) Wang, D.; Astruc, D. The golden age of transfer hydrogenation. Chem. Rev. 2015, 115, 6621− 6686. (c) Gilkey, M. J.; Xu, B. Heterogeneous catalytic transfer hydrogenation as an effective pathway in biomass upgrading. ACS Catal. 2016, 6, 1420−1436. (6) (a) Paone, E.; Espro, C.; Pietropaolo, R.; Mauriello, F. Selective arene production form transfer hydrogenolysis of benzyl phenyl ether promoted by the coprecipitated Pd/Fe3O4 catalyst. Catal. Sci. Technol. 2016, 6, 7937−7941. (b) Espro, C.; Gumina, B.; Paone, E.; Mauriello, F. Upgrading Lignocellulosic Biomasses: Hydrogenolysis of Platform Derived Molecules Promoted by Heterogeneous Pd-Fe Catalysts. Catalysts 2017, 7, 78. (7) (a) Sawama, Y.; Morita, K.; Yamada, T.; Nagata, S.; Yabe, Y.; Monguchi, Y.; Sajiki, H. Rhodium-on-carbon catalyzed hydrogen scavenger- and oxidant-free dehydrogenation of alcohols in aqueous media. Green Chem. 2014, 16, 3439−3443. (b) Sawama, Y.; Morita, K.; Asai, S.; Kozawa, M.; Tadokoro, S.; Nakajima, J.; Monguchi, Y.; Sajiki, H. Palladium on Carbon-Catalyzed Aqueous Transformation of Primary Alcohols to Carboxylic Acids Based on Dehydrogenation under Mildly Reduced Pressure. Adv. Synth. Catal. 2015, 357, 1205− 1210. (8) (a) Yasukawa, N.; Yokoyama, H.; Masuda, M.; Monguchi, Y.; Sajiki, H.; Sawama, Y. Highly-functionalized arene synthesis based on palladium on carbon-catalyzed aqueous dehydrogenation of cyclohexadienes and cyclohexenes. Green Chem. 2018, 20, 1213−1217. (b) Ichikawa, T.; Matsuo, T.; Tachikawa, T.; Yamada, T.; Yoshimura, T.; Yoshimura, M.; Takagi, Y.; Sawama, Y.; Sugiyama, J.; Monguchi, Y.; Sajiki, H. Microwave-Mediated Site-Selective Heating of SphericalCarbon-Bead-Supported Platinum for the Continuous, Efficient Catalytic Dehydrogenative Aromatization of Saturated Cyclic Hydrocarbons. ACS Sustainable Chem. Eng. 2019, 7, 3052−3061. (9) (a) Sawama, Y.; Niikawa, M.; Yabe, Y.; Goto, R.; Kawajiri, T.; Marumoto, T.; Takahashi, T.; Itoh, M.; Sasai, Y.; Yamauchi, Y.; Kondo, S.; Kuzuya, M.; Itoh, M.; Monguchi, Y.; Sajiki, H. StainlessSteel-Mediated Quantitative Hydrogen Generation from Water under Ball Milling Conditions. ACS Sustainable Chem. Eng. 2015, 3, 683− 689. (b) Sawama, Y.; Kawajiri, T.; Niikawa, M.; Goto, R.; Yabe, Y.; Takahashi, T.; Marumoto, T.; Itoh, M.; Kimura, Y.; Monguchi, Y.; Kondo, S.; Sajiki, H. Stainless-Steel Ball-Milling Method for Hydro-/ Deutero-genation using H2O/D2O as a Hydrogen/Deuterium Source. ChemSusChem 2015, 8, 3773−3776. (c) Sawama, Y.; Yasukawa, N.; Ban, K.; Goto, R.; Niikawa, M.; Monguchi, Y.; Itoh, M.; Sajiki, H. 11530

DOI: 10.1021/acsomega.9b01130 ACS Omega 2019, 4, 11522−11531

ACS Omega

Article

Stainless Steel-Mediated Hydrogen Generation from Alkanes and Diethyl Ether and Its Application for Arene Reduction. Org. Lett. 2018, 20, 2892−2896. (10) (a) Sawama, Y.; Yabe, Y.; Shigetsura, M.; Yamada, T.; Nagata, S.; Fujiwara, Y.; Maegawa, T.; Monguchi, Y.; Sajiki, H. Platinum on Carbon-Catalyzed Hydrodefluorination of Fluoroarenes using Isopropyl Alcohol-Water-Sodium Carbonate Combination. Adv. Synth. Catal. 2012, 354, 777−782. (b) Sawama, Y.; Mori, M.; Yamada, T.; Monguchi, Y.; Sajiki, H. Hydrogen Self-Sufficient Arene Reduction to Cyclohexane Derivatives Using a Combination of Platinum on Carbon and 2-Propanol. Adv. Synth. Catal. 2015, 357, 3667−3670. (11) (a) Wang, H.; Krier, J. M.; Zhu, Z.; Melaet, G.; Wang, Y.; Kennedy, G.; Alayoglu, S.; An, K.; Somorjai, G. A. Promotion of Hydrogenation of Organic Molecules by Incorporating Iron into Platinum Nanoparticle Catalysts: Displacement of Inactive Reaction Intermediates. ACS Catal. 2013, 3, 2371−2375. (b) Liu, Z.; Tan, X.; Li, J.; Lv, C. Easy synthesis of bimetal PtFe-containing ordered mesoporous carbons and their use as catalysts for selective cinnamaldehyde hydrogenation. New J. Chem. 2013, 37, 1350− 1357. (c) Lee, J.; Kim, Y. T.; Huber, G. W. Aqueous-phase hydrogenation and hydrodeoxygenation of biomass-derived oxygenates with bimetallic catalysts. Green Chem. 2014, 16, 708−718. (d) Liu, L.; Lou, H.; Chen, M. Selective hydrogenation of furfural over Pt based and Pd based bimetallic catalysts supported on modified multiwalled carbon nanotubes (MWNT). Appl. Catal., A 2018, 550, 1−10. (12) (a) Nakamula, I.; Yamanoi, Y.; Imaoka, T.; Yamamoto, K.; Nishihara, H. A Uniform Bimetallic Rhodium/Iron Nanoparticle Catalyst for the Hydrogenation of Olefins and Nitroarenes. Angew. Chem., Int. Ed. 2011, 50, 5830−5833. (b) Kim, J. K.; Lee, J. K.; Kang, K. H.; Lee, J. W.; Song, I. K. Catalytic decomposition of phenethyl phenyl ether to aromatics over Pd−Fe bimetallic catalysts supported on ordered mesoporous carbon. J. Mol. Catal. A: Chem. 2015, 410, 184−192. (c) Wang, R.; Tang, T.; Huang, K.; Zou, M.; Tao, X.; Yin, H.; Lin, Z.; Dang, Z.; Li, G. Debromination of polybrominated biphenyls (PBBs) by zero valent metals and iron-based bimetallic particles: Mechanisms, pathways and predicting descriptor. Chem. Eng. J. 2018, 351, 773−781. (13) (a) Sawama, Y.; Yamada, T.; Yabe, Y.; Morita, K.; Shibata, K.; Shigetsura, M.; Monguchi, Y.; Sajiki, H. Platinum on carbon-catalyzed H-D exchange reaction of aromatic nuclei due to isopropyl alcoholmediated self-activation of platinum metal in deuterium oxide. Adv. Synth. Catal. 2013, 355, 1529−1539. (b) Sawama, Y.; Park, K.; Yamada, T.; Sajiki, H. New gateways to the platinum group metalcatalyzed direct deuterium-labeling method utilizing hydroten as a catalyst activator. Chem. Pharm. Bull. 2018, 66, 21−28. (c) Sawama, Y.; Nakano, A.; Matsuda, T.; Kawajiri, T.; Yamada, T.; Sajiki, H. H-D exchange deuteration of arenes at room temperature. Org. Process Res. Dev. 2019, 23, 648−653. (d) Yamada, T.; Sawama, Y.; Shibata, K.; Morita, K.; Monguchi, Y.; Sajiki, H. Multiple deuteration of alkanes synergistically-catalyzed by platinum and rhodium on carbon as a mixed catalytic system. RSC Adv. 2015, 5, 13727−13732. (14) Zhang, L.; Zhou, M.; Shao, L.; Wang, W.; Fan, K.; Qin, Q. Reactions of Fe with H2O and FeO with H2. A Combined Matrix Isolation FTIR and Theoretical Study. J. Phys. Chem. A 2001, 105, 6998−7003. (15) (a) Peng, B.; Zhao, C.; Mejía-Centeno, I.; Fuentes, G. A.; Jentys, A.; Lercher, J. A. Comparison of kinetics and reaction pathways for hydrodeoxygenation of C3 alcohols on Pt/Al2O3. Catal. Today 2012, 183, 3−9. (b) Nakagawa, Y.; Mori, K.; Chen, K.; Amada, Y.; Tamura, M.; Tomishige, K. Hydrogenolysis of CO bond over Remodified Ir catalyst in alkane solvent. Appl. Catal., A 2013, 468, 418− 425. (c) Zhu, S.; Hao, S.; Zheng, H.; Mo, T.; Li, Y.; Zhu, Y. One-step hydrogenolysis of glycerol to biopropanols over Pt−H4SiW12O40/ ZrO2 catalysts. Green Chem. 2012, 14, 2607−2616. (16) Arčon, I.; Kolar, J.; Kodre, A.; Hanžel, D.; Strlič, M. XANES analysis of Fe valence in iron gall inks. X-Ray Spectrom. 2007, 36, 199−205.

(17) Tomita, A.; Shimizu, K.; Kato, K.; Akita, T.; Tai, Y. Mechanism of Low-Temperature CO Oxidation on Pt/Fe-Containing Alumina Catalysts Pretreated with Water. J. Phys. Chem. C 2013, 117, 1268− 1277. (18) X-ray photoelectron spectroscopy (XPS) analysis of fresh 10% Pt/C indicated that zero valent of platinum was supported on carbon. See Supporting Information. (19) Although Pt/C-catalyzed dehydrogenation of 9 in the absence of 8 provided 8 as a dehydrogenated product, methylcyclohexane was also produced by the hydrogenation of 9 using H2 in situ-generated via the dehydrogenation process of 8 to 9. See Supporting Information. (20) (a) Birch, A. J. Reduction by Dissolving Metals. J. Chem. Soc. 1944, 430−436. (b) Zimmerman, H. E. A Mechanistic Analysis of the Birch Reduction. Acc. Chem. Res. 2012, 45, 164−170. (21) Le Bailly, B. A. F.; Greenhalgh, M. D.; Thomas, S. P. Ironcatalysed, hydride-mediated reductive cross-coupling of vinyl halides and Grignard reagents. Chem. Commun. 2012, 48, 1580−1582. (22) Cai, Y.; Qian, X.; Gosmini, C. Cobalt-Catalyzed Csp3-Csp3 Homocoupling. Adv. Synth. Catal. 2016, 358, 2427−2430.

11531

DOI: 10.1021/acsomega.9b01130 ACS Omega 2019, 4, 11522−11531