Next Article in Journal
The Use of Zeolites for VOCs Abatement by Combining Non-Thermal Plasma, Adsorption, and/or Catalysis: A Review
Previous Article in Journal
Effect of Mono-, Di-, and Triethylene Glycol on the Activity of Phosphate-Doped NiMo/Al2O3 Hydrotreating Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Synthesis of 1,3-Butadiene and Its 2-Substituted Monomers for Synthetic Rubbers

1
Key Laboratory of High-Performance Synthetic Rubber and Its Composite Materials, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Renmin Road, Changchun 130022, China
2
Phytochemistry, College of Life Sciences, Jiangxi Normal University, Nanchang 330022, China
3
Applied Chemistry, University of Science and Technology of China, Jinzhai Road, Hefei 230026, China
*
Author to whom correspondence should be addressed.
Catalysts 2019, 9(1), 97; https://doi.org/10.3390/catal9010097
Submission received: 17 November 2018 / Revised: 7 January 2019 / Accepted: 14 January 2019 / Published: 17 January 2019

Abstract

:
Synthetic rubbers fabricated from 1,3-butadiene (BD) and its substituted monomers have been extensively used in tires, toughened plastics, and many other products owing to the easy polymerization/copolymerization of these monomers and the high stability of the resulting material in manufacturing operations and large-scale productions. The need for synthetic rubbers with increased environmental friendliness or endurance in harsh environments has motivated remarkable progress in the synthesis of BD and its substituted monomers in recent years. We review these developments with an emphasis on the reactive routes, the products, and the synthetic strategies with a scaling potential. We present reagents that are primarily from bio-derivatives, including ethanol, C4 alcohols, unsaturated alcohols, and tetrahydrofuran; the major products of BD and isoprene; and the by-products, activities, and selectivity of the reaction. Different catalyst systems are also compared. Further, substituted monomers with rigid, polar, or sterically repulsive groups, the purpose of which is to enhance thermal, mechanical, and interface properties, are also exhaustively reviewed. The synthetic strategies using BD and its substituted monomers have great potential to satisfy the increasing demand for better-performing synthetic rubbers at the laboratory scale; the laboratory-scale results are promising, but a big gap still exists between current progress and large scalability.

Graphical Abstract

1. Introduction

“Rubber” was coined by a chemist named Joseph Priestley in 1770. It has since come to be regarded as the “wheel” of civilization. Today, the term “rubber” is related to any material characterized by properties similar to those of natural rubber (NR). Attributed to the well-known importance of rubber since World War II, much effort has gone into producing rubbers by various synthetic routes [1]. In 2017, around 15.05 billion kilograms of synthetic rubbers were produced worldwide. The most prevalent synthetic rubber is styrene-butadiene rubber (SBR) which is derived from a copolymer of 1,3-butadiene (BD) and styrene. SBR has good aging stability and abrasion resistance [2], and it occupies about half of the market for car tires. Also widely used in tires is butadiene rubber (BR); it is made solely from the BD monomer and shows excellent wear resistance, good cold resistance, lower heat generation, and dynamic performance. Unfortunately, BR still falls short of meeting the goal of producing rubbery materials that are more environmentally friendly or robust to harsh environments. For example, BR needs to be blended with SBR and/or NR when it is used for sidewall and tread. A lot of effort has been made to clarify the nature of BR and its blends (such as with SBR and NR), and significant progress has been achieved in the regulation of the structure and performance of the polymer [3,4,5].
When used in tires, the elastic properties of synthetic rubbers made from 1,3-dienes play a primary role and are enormously affected by the microstructures of the repeating units, especially the presence or absence of a side substituent [6]. For example, cis-1,4-polybutadiene shows flexibility and excellent strong deformation tolerance [7,8] due to its low glass transition temperature (Tg) of −94 °C, which is much lower than that of the cis-1,4-polyisoprene with a –CH3 group at 2-C (from −64 to −70 °C) [9]. The latter, known as isoprene rubber (IR, the major component of NR), has excellent overall rubbery properties and is widely used as a synthetic substitute for NR without blending with other polymers. Therefore, IR can be used for carcass and off-the-road tires without blending with other rubbery materials. It can be easily concluded that the side substituent of 1,3-butadienes plays a crucial role in tuning the properties of the resulting materials.
However, serious problems are encountered when these polydienes are used for the production of mechanical engineering goods, such as heavy-lorry tires and aircraft tires, due to their inadequate mechanical properties (e.g., tensile strength, elongation at break, and tear resistance) [10]. Therefore, substituted butadienes are used to form polydienes with some BD derivatives, such as 2-phenylbutadiene, making it likely that a solution to these problems will be found. To this end, a functionalized 1,3-diene containing a rigid group is necessary.
On the other hand, it is well-known that tires are made by blending synthetic rubbers with inorganic fillers, such as carbon black and silica. Unfortunately, polydienes containing solely carbon and hydrogen always exhibit poor surface properties and low adhesive properties. Therefore, in the manufacturing of tires, fillers such as hydrophilic silica are completely incompatible with synthetic rubbers, as the resulting tires have poor tear resistance, abrasion resistance, and static toughness properties [11]. To solve this problem, considerable attention has been paid to the synthesis of polar-functionalized dienes because it is very difficult to modify inorganic fillers [9]. The incorporation of these kinds of monomers into copolymers and/or homopolymers can significantly improve the existing properties of synthetic rubbers. For example, nitrile butadiene rubber (NBR) can enhance the wear resistance of tires. Additionally, NBR is also widely used for seals, hoses, and gloves. In this respect, the polar-functionalized 1,3-diene attracts a lot of attention.
Accordingly, considering the entropy-elasticity of synthetic rubbers, various side substituents in 1,3-butadienes affect the properties of the resulting materials. To attain rubbers with tunable properties for meeting different requirements, introducing designed substituents into 1,3-butadiene is a promising strategy; it is also a great challenge, but this makes the success all the more rewarding. Functionalized monomers are also used to enhance the properties of synthetic rubbers, either by monopolymerization or copolymerization. In the present review, we highlight the following: (1) the production of BD and IP, which are in large demand in industry; (2) the synthesis of 2-substituted 1,3-butadienes with rigid groups, such as adamantly, phenyl, and its derivatives, which can be used to enhance the thermal and mechanical properties of the synthetic rubbers; and (3) the synthesis of polar-functionalized 1,3-butadienes containing O, Si, and N atoms and their unique properties for special applications. BD and IP are the basic monomers of the synthetic rubbers consumed widely in the automotive industry. Both BD and IP are mainly supplied by ethylene plants as by-products of naphtha cracking; taking sustainability into account, the alternative routes are urgently required [12]. The synthetic processes of substituted 1,3-butadienes always suffer from complicated operations under rigorous conditions.

2. Synthesis of 1,3-butadiene

At present, more than 95% of the world’s total BD is produced from either naphtha cracking or n-butane dehydrogenation [13,14]. Unfortunately, these routes require expensive extractive distillation with a low selectivity (for BD) of around 4–5%. Besides that, the shale gas revolution has led to a decrease in the production of naphtha-based ethylene, which could reduce the production of BD too [15,16]. Additionally, the global supply of BD faces trouble due to the variation in chemical feedstock and the unstable price of petrochemicals. In Japan, 3 out of 15 ethylene production plants have been shut down in the 3 years before June 2016. [17] Therefore, the industry urgently calls for the development of an alternative process to produce BD without depending on petroleum.
The first production of BD from ethanol, carried out by Ipatiev in Russia in 1903, had a very low yield of 1.5% [18], and great improvements were achieved by Lebedev’s group using an undisclosed composition [19]. A high yield and a BD selectivity of up to 70% under atmospheric pressure at 350 °C were reported. Interessen-Gemeinschaft Farbenindustrie AG (IG Farben) reported that magnesia catalysts promoted by Cr or Co exhibited a BD yield of 60% [20]. Subsequently, the ethanol-to-BD route was abandoned in the US and most European countries as this technology was deemed poorly competitive compared with the production of BD from petrochemical sources [21]. Currently, a return to the old ethanol-to-BD production route is gaining renewed interest because of its remarkable capacity for producing bio-ethanol and potential environmental benefits [22]. BD formation from ethanol has been studied intensively, and several possible BD formation proposals have been debated [19,23,24,25,26,27,28,29,30,31,32,33,34,35]. A reaction network proposed by Quattlebaum et al. is presented in Figure 1 [30]. Obviously, BD formation involves dehydrogenation, aldol condensation, hydrogenation, and a dehydration reaction, which implies that various active sites such as basic, acidic, and metallic species are required for an efficient catalyst in this reaction. Unfortunately, it seems very difficult to develop these multifunctional catalysts. Benefiting from an understanding of the reaction pathway, a more efficient two-step process employing acetaldehyde/ethanol as feedstock was developed (Figure 2), but there is still a lot of work to do to realize a BD production process with a yield higher than 90%. Although extensive studies have been performed over the past 3 years [36,37,38,39,40,41,42,43,44,45,46,47] (Table 1), as well as several published reviews [17,48,49,50,51,52,53], the development of efficient processes and robust catalysts still remains a challenge and issues related to carbon efficiency, catalyst cost, toxicity, feedstock tolerance and backend optimization need further investigation [49].
Next to ethanol, alternative bio-derived C4 alcohols, such as 1,3-butanediol (1,3-BDO), 2,3-butanediol (2,3-BDO), and 1,4-butanediol (1,4-BDO), are regarded as promising feedstocks to produce BD, and their use in BD synthesis has been demonstrated (Figure 3). Makshina et al. [49] summarized this part based on many research efforts as well as reviews. They have mainly concentrated on the dehydration and deoxydehydration of these C4 alcohols to BD from the aspects of mechanistic and catalytic chemistry. Very recently, Jing et al. [55] studied the dehydration of 1,3-BDO to BD over aluminosilicates. The best result was achieved over H–ZSM–5 (silica/alumina = 260) with a BD yield of 60% at 300 °C. They pointed out that the good catalytic performance of the catalysts is a result of the presence of Brønsted acid sites with weak and medium strength. Accompanied by the unavoidable formation of carbon deposits, the conversion decreases from 100% to 80% after a 102 h run. Subsequently, Al–doped SBA–15 catalysts were applied for this reaction [56]. The highest BD yield, 59%, was achieved over Al–SBA–15 (silica/alumina = 76), with stable activity in long-term runs. Notably, the operation is performed at a relatively low temperature (200 °C).
Zeng et al. [57] studied the direct conversion of 2,3-BDO to BD over alumina catalysts with a wide range of temperatures from 240 to 450 °C. They clarified that 3-buten-2-ol (3B2OL) is a key intermediate in the dehydration of 2,3-BDO to BD by DFT, and 3B2OL can be easily converted to methyl ethyl ketone (MVK) rather than BD in the presence of sodium-modified alumina catalysts because of the natural basicity of the catalysts. The experimental results suggest that a higher selectivity to BD is obtained under a higher reaction temperature. Fang et al. [58] conducted the dehydration of 1,3-BDO to BD over a series of Ce@MOR catalysts with a flower-like morphology at 350 °C. Ce@MOR with Si/Ce at an atomic ratio of 50 presents a high 1,3-BDO conversion of 100% but a poor selectivity of 46% to BD. The high activity of Ce@MOR is attributed to the high density of Brønsted acid sites with medium strength. Nguyena et al. [59] reported the dehydration of 2,3-BDO to BD over a GdPO4 catalyst which resulted in the complete conversion of 2,3-BDO and a selectivity of 58% to BD at 300 °C. Both the fresh and regenerated GdPO4 displayed relatively stable catalytic activity in a 50 h run. Kim et al. [60] carried out the direct dehydration of 2,3-BDO over silica-supported sodium phosphates catalysts at 400 °C. These findings show that the conversion of 2,3-BDO can be easily tuned up to 100%, while the selectivity to BD never exceeds 70%. All of these studies confirm that Brønsted acid sites with weak and medium strength are a key factor in the conversion of butanediols (BDOs) to BD. A great improvement was observed by Tsukamoto [61], who investigated the dehydration of 2,3-BDO to BD over SiO2-supported CsH2PO4 catalysts at 350 °C. The highest conversion of 2,3-BDO (>99.9%) and an excellent selectivity to BD (91.9%) were obtained over 10% CsH2PO4, with a slight decrease over an 8 h run. This impressive catalytic activity is ascribed to a combination of the proper acid–base sites of Cs phosphate and the large ionic radius of Cs+, which is significantly different from Na+ [57].
However, the direct dehydration of C4 alcohols into BD always needs a high reaction temperature. As BD can be readily obtained via the dehydration of unsaturated C4 alcohols (UOLs) over solid acid catalysts, a novel path way to produce BD from C4 alcohols via the dehydration of C4 alcohols to their corresponding UOLs was developed, and it is performed at a relatively low temperature. Sato’s group has done a lot of work in this field. Duan et al. have reviewed the dehydration of BDOs to UOLs in connection with their further dehydration to BD [17].
Not long ago, Duan et al. [62] investigated the dehydration of 2,3-BDO to 3B2OL over modified monoclinic ZrO2 catalysts, suggesting that BaO/ZrO2, SrO/ZrO2, and CaO/ZrO2 have good catalytic capacities. The optimized data were observed over BaO–0.0452–800 with a 2,3-BDO conversion of 72.4% and a 3B2OL selectivity of 74.4% during the early stage of 5 h at 350 °C. Unfortunately, a decrease in catalytic activity occurred over a long-term operation of 25 h. This is due to the formation of Perovskite compounds, such as BaZrO3, which are inactive for the reaction. The modified monoclinic ZrO2 also presents high catalytic activity for the dehydration of 1,4-BDO to 3-buten-1-ol (3B1OL), especially 7–CaO+2–ZrO2/m–ZrO2–800, which produces an excellent 1,4-BDO conversion of 95.2% and a high 3B1OL selectivity of 89.3% at 350 °C [63]. Then, the dehydration of UOLs to BD was conducted by Sun et al. [64] over solid acid catalysts, like SiO2–Al2O3, Al2O3, TiO2, and ZrO2, and basic rare earth metal oxides, including Yb2O3 and CeO2. This suggests that acid catalysts have a higher activity for the dehydration of 2B1OL and 3B2OL than the rare earth metal oxides. The 10SiO2/Al2O3 catalyst shows the highest 2-buten-1-ol (2B1OL) conversion (92.5%) and selectivity (75.0%) at 260 °C, but there is a rapid deactivation. Modified 10SiO2/Al2O3 with Ag can enhance the catalytic activity and inhibit the formation of carbon deposits. The best result of the dehydration of 2B1OL was achieved over 5Ag–SiO2/Al2O3, displaying a BD selectivity of 94.6% at a conversion of 99.1% at 260 °C. Also, 5Ag–SiO2/Al2O3 efficiently inhibits catalytic deactivation, with a slight decrease in the 2B1OL conversion (from 99.4% to 96.2%) and BD selectivity (from 95.0% to 87.1%) after 10 h. In contrast, the rare earth metal oxides, such as CeO2, show poor activity for the dehydration of 2B1OL and 3B2OL to BD, but they have a remarkable activity in the dehydration of 3B1OL to BD, which is attributed to the weak basicity and the redox nature of CeO2 via Ce4+–Ce3+. The synthesis of BD from 3B1OL over the acid catalysts shows a much lower BD yield resulting from the C–C cleavage of 3B1OL to propylene. Thus, it is concluded that the dehydration mechanism of 3B1OL to BD is quite different from the dehydration of 2B1OL and 3B2OL.
Therefore, Wang et al. [65] studied the production of BD by the dehydration of 3B1OL and 1,4-butanediol over rare earth oxides, such as Lu2O3 and Yb2O3. The optimal result of the dehydration of 3B1OL was obtained with a high conversion of 99.6% and a good BD selectivity of 96.7% over the Yb2O3 catalyst at 340 °C. Accordingly, the direct dehydration of 1,4-BDO to BD over Yb2O3 also exhibited excellent a catalytic performance with a yield of 96.6% at 360 °C. Thus, producing BD from bio-derived C4 alcohols and especially unsaturated C4 alcohols via the dehydration process is a promising route, although a large amount of work still needs to be done, especially for improving catalyst life for industry usage.
Tetrahydrofuran (THF) can also easily generate 1,4-BDO, which comes from five-carbon sugars such as xylose or furfural [66,67,68,69]. Therefore, Abdelrahman et al. [70] developed a novel way to synthesize BD via the catalytic ring-opening dehydration of THF, which is conducted in the presence of a solid acid catalyst with a high selectivity of 85–99% to BD at a high THF conversion of 89% (Figure 4). These encouraging data were obtained at 400 °C with a space velocity of 0.2 s−1. Unfortunately, the conversion of THF significantly decreases to 8.9% with a turning space velocity of up to 9.7 s−1. Therefore, a large change in this route must be implemented to improve the BD formation capacity.

3. Synthesis of Isoprene

Isoprene (IP), also called 2-methyl-1,3-butadiene, is well-known as an irreplaceable monomer of synthetic natural rubber. It is a colorless liquid with a characteristic odor, and it is highly significant for producing stereoregular polyisoprene because of the ability to purposefully manipulate its properties. The huge demand for this kind of material inspires the rapid development of the synthesis of isoprene (Figure 5). The main methods used to produce isoprene in the industry have been reported elsewhere [71]. In this review, the focus is on a promising route to produce isoprene using isobutene and formaldehyde as starting materials via one step; this has attracted great interest both in the industry and in academia. This process is conducted in the presence of solid acid catalysts, such as zeolites, silver, oxide catalysts, phosphates, and heteropolyacid catalysts. Moreover, the recent development of the isoprene synthesis using another method is also presented.
Dumitriu et al. studied the reaction over zeolite catalysts [75,76]. The strength of the acid sites is suggested to play a crucial role in isoprene synthesis. The weak Brønsted acid sites on zeolites are proved to be the most efficient sites. In this respect, the highest selectivity, 99–100%, was obtained over boralites and ferrisilicates. Nonetheless, these results were obtained in a pulse catalytic reactor, which cannot be compared with those obtained in continuous flow systems.
Ag–Sb catalysts have been employed as catalysts in bed reactors and resulted in good activity, with a conversion of around 81% (for formaldehyde) and selectivity of about 70% (for isoprene). However, the activity of Ag–Sb catalysts decreases quickly and presents poor stability [77]. CuSO4/SiO2 catalysts were developed which show a formaldehyde conversion of about 63% and selectivity of about 65% to isoprene. It has been suggested that doping CuSO4/SiO2 with basic oxides to decrease the acidity can greatly improve their activity. A formaldehyde conversion of 87% and selectivity of about 65% to isoprene were obtained over CuSO4/SiO2 doped with MgO. However, the short lifetime of the CuSO4/SiO2 catalysts impedes their application in the industry, and this deficiency can be attributed to the decreasing number of acid sites resulting from the rapid formation of carbon deposits and the destruction of the CuSO4 structure [78,79].
Krzywicki et al. [80] first reported the reaction of formaldehyde and isobutene over Al2O3–H3PO4 catalysts. A high H3PO4 content favors isoprene production owing to the increased number of acid sites. However, the phosphate catalysts still displayed low activity, with an isoprene yield of 22%. Ai et al. [81] tried to use the oxides of transition metals to improve the activity of the phosphate catalysts. MoO3-, WO3-, and V2O5–based phosphate catalysts were employed for this reaction, using tert-butyl alcohol as a raw material instead of isobutene. V2O5–P2O5 (P/V = 1.06) exhibited outstanding activity, with an isoprene yield of about 60%. The author pointed out that the basic sites also play a certain role in the synthesis of isoprene. Although the phosphate catalysts show a rather excellent activity in both conversion and selectivity, they are characterized by rapid deactivation due to the formation of large carbon deposits.
Therefore, catalysts with longer lifetimes and an understanding of the influence of carbon deposition are urgently needed. Recently, Sushkevich et al. [82] conducted the condensation of formaldehyde with isobutene over a Nb2O5–P2O5 catalyst. This catalyst is characterized by a much longer lifetime and stable formaldehyde conversion, producing a steady isoprene yield of 57% within more than 30 h of stream. Nb2O5–P2O5 shows high resistance to deactivation, which is ascribed to the in situ restoration of Brønsted acid sites in the presence of water during the reaction. Moreover, spent Nb2O5–P2O5 can be completely regenerated by calcination at 500 °C in an air flow. It was verified that Lewis acid sites are responsible for the side reaction, which led to a low selectivity of 62.6%. Therefore, supported Keggin-type heteropolyacids catalysts with weak Brønsted sites are being developed [83,84]. Most importantly, Ivanova et al. clarified that unsaturated branched carbonaceous species in carbon deposits can generate active sites for the condensation of formaldehyde with isobutene to produce isoprene [85]. To further identify the effect of carbon deposition on the synthesis of isoprene, our group conducted the reaction over MoO3–P2O5 [86]. 4,4-Dimethyldioxane-1,3 (DMD) has been demonstrated to be an important intermediate which can readily form on acid sites over a catalyst and further be converted into isoprene in the presence of carbonaceous species. In this regard, the formation of carbon deposits can give rise to a synergetic effect between acid species and carbonaceous species for isoprene synthesis. Slight carbon deposition is able to shorten the induction period of the reaction, while a large carbon deposition leads to catalyst deactivation because it attaches to Brønsted acid sites. For the above reasons, a simple way to restore the reactivity of deactivated phosphate catalysts was developed with the assistance of the carbon deposition [87].
Moreover, additional studies on the synthesis of isoprene by alternative routes have been reported. Songsiri et al. [72,73] synthesized isoprene using heteropolyacids as catalysts in a liquid-phase system under moderate conditions in which carbon deposits can be inhibited. Methyl tertiary-butyl ether (MTBE) was used as the source of isobutene. MTBE is more available because its use as an additive in gasoline has been prohibited in developed countries, such as the United States, Japan, and Western Europe. High formaldehyde conversion (>95.8%) and isoprene yield (68.3–77.5%) were achieved over a Cs-exchanged silicotungstic acid catalyst. Unfortunately, it is still a challenge to completely separate the catalysts from the reaction system, and a large amount of organic solvent is employed in this process, which could result in environmental problems.
Considering environmental concerns and sustained availability, the biosynthesis of isoprene was developed based on metabolic engineering. However, the necessary productivity has not been achieved with microbes [88]. It was necessary to develop an alternative method for the sustainable production of isoprene at a large scale. Therefore, it is a promising and competitive route to convert biomass-derived feedstocks into isoprene via chemical processes. Itaconic acid from glucose [89] was employed as raw material by Abdelrahman [74]. Firstly, itaconic acid was converted into 3-methyl-tetrahydrofuran (3-MTHF) with a yield of about 80% in a liquid phase using Pd–Re/C as a catalyst. Then the dehydration of 3-MTHF was performed in a gas phase over a P-SPP (self-pillared pentasil) catalyst with an isoprene selectivity of 70% at a 20–25% conversion. For this process, much work is needed to improve the yield of isoprene, to understand the catalytic pathway, and to develop a more efficient catalysts. Extensive research on IP synthesis is urgently required, and the exploration of novel platform chemicals faces challenges, but success would be greatly rewarding.
1,3-Pentadiene (piperylene) has similar properties to isoprene and is a large-scale by-product accompanying the production of isoprene. It is widely used in the production of sticky tapes, adhesives, and plastics. In Russia, 1,3-pentadiene is also used to produce liquid rubber, which is intensively used in paint and varnish industry [90,91,92]. Moreover, many studies have focused on the potential application of poly(1,3-pentadiene) in the domain of elastomers. All of these factors inspire wide interest to study the production of 1,3-pentadiene [93,94,95]. From the view of sustainable development, many efforts have been made to develop alternative technologies for producing 1,3-pentadiene based on biomass-derived feedstocks [96,97,98,99]. Recently, Sun et al. [100] developed a novel two-step synthetic approach using xylitol as a resource (Figure 6). Xylitol is industrially produced by extracting hardwoods or corncobs. In the first step, the deoxydehydration (DODH) of xylitol was conducted at 235 °C in the presence of formic acid to give 2,4-pentadien-1-ol, 1-formate (2E), with an optimized yield of 62.9%, followed by deoxygenation to obtain 1,3-pentadiene over a Pd/C catalyst at 100 °C. The total yield of 1,3-pentadiene reached 51.8%. Kumbhalkar et al. [101] employed 2-methyltetrahydrofuran (2-MTHF) which can be easily obtained from biomass-derived intermediates (levulinic acid and furfural, etc.), as the starting material to produce 1,3-pentadiene in a continuous system. This reaction is conducted at 350 °C, and the catalyst undergoes deactivation from the formation of carbon deposits in the presence of dienes. Therefore, a decrease in the conversion of 2-MTHF (from 100% to 77%) and yield of 1,3-pentadiene (from 67.8% to 51.8%) was observed over a period of 58 h.

4. Rigid-Group Functionalized 1,3-butadienes

4.1. Synthesis of 2-(1-adamantyl)-1,3-butadiene

Due to its highly symmetrical and rigid tricyclic structure, adamantane features chemical and thermal stability, hydrophobic character, and UV transparency. It has been introduced into polymers to improve their thermal and mechanical properties [102,103,104,105,106,107,108,109,110,111]. Therefore, introduction of adamantyl into 1,3-dienes is explored as presented in Figure 7 [112,113].
Firstly, methylmagnesium iodide is prepared by adding methyl iodide and magnesium into diethyl ether. 1-Adamantanecarboxylic acid (Figure 7a) and thionyl chloride are charged into a vessel, and the mixture is heated to 80 °C and then left for 2 h with stirring. After the reaction, thionyl chloride is removed to afford a white solid of 1-adamantanecarbonyl chloride (Figure 7b) under reduced pressure. Then, diethyl ether and CuCl are added to 1-adamantanecarbonyl chloride under a N2 atmosphere, followed by the dropwise addition of methylmagnesium iodide at 0 °C under a N2 atmosphere. After stirring at room temperature for 30 min, the reaction is quenched with 2 N HCl. The mixture is extracted with diethyl ether. Then, the organic layer is dried over anhydrous MgSO4. After the removal of the solvent, the product (Figure 7c) is obtained. In an ice/water bath under a N2 atmosphere, the THF solution of vinyl bromide is added dropwise to magnesium, which is activated with 1,2-dibromoethane and stirred at this temperature for another 1 h, then at room temperature for 1 h. Prior to the addition of the THF solution of the product in Figure 7c, the reaction mixture is cooled down to 0 °C. Finally, stirring at room temperature overnight is maintained. The reaction is quenched with 2 N HCl, and diethyl ether is used to extract the reaction system. After the evaporation of the solution, the product (Figure 7d) is obtained. The solution of the product (Figure 7d) in benzene refluxes for 2 h in the presence of p-toluenesulfonic acid. After cooling, the organic layer is separated and dried. The solvent is removed to afford the final product (Figure 7e).

4.2. Synthesis of 2-phenyl-1,3-butadienes

2-Phenyl-1,3-butadiene (2-PB) can be regarded as either an α-vinyl-substituted styrene or 2-phenyl-substituted 1,3-butadiene. Therefore, 2-PB is a unique monomer because of the importance of styrene and 1,3-butadiene in synthetic rubbers. Moreover, with a phenyl group replacing methyl, Poly(2-PB) exhibits distinctive properties compared with polyisoprene. Moreover, Poly(2-PB) can also be used to produce high-temperature thermoplastics with a Tg as high as 325 °C [114]. Since the successful synthesis of 2-PB by Marvel, [115] it received much attention [116,117].
As displayed in Figure 8, vinyl bromide and magnesium are used to prepare the THF solution of vinylmagnesium bromide, followed by the addition of acetophenone (Figure 8a). The mixture must be maintained under gentle reflux and then maintained with stirring for 1 h. A saturated aqueous solution of ammonium chloride is applied for the hydrolysis process to obtain the crude product. After the distillation and removal of the solvent, methylphenylvinylcarbinol (Figure 8b) is afforded with a yield of 75%. Methylphenylvinylcarbinol, aniline hydrobromide, and hydroquinone are charged into a round-bottomed flask equipped with a Vigreux column. The mixture is heated at 100 °C in an oil bath, followed by gentle distillation with the temperature increasing to 150 °C. The fraction with a b.p. of 57–63 °C is collected and dried by anhydrous calcium chloride to obtain 2-PB (Figure 8c). Recently, Yao et al. [118] used acetophenone as the starting material to prepare 2-PB via elaborate manipulations, which are performed under a dried and oxygen-free atmosphere, and the solvents involved in the reaction must be purified. In the typical procedure, acetophenone is first converted into 2-phenylbut-3-en-2-ol in the presence of vinylmagnesium bromide at 0 °C, followed by a dehydration reaction to afford 2-PB in the presence of pyridinium p-toluenesulfonate at 80 °C; THF and toluene are employed as solvents, respectively. Similar to 2-PB, 1-phenyl-1,3-butadiene (1-PB) and 1-(4-methypenyl)-1,3-butadiene (1-MPB) have also gained much attention [120,121,122,123,124,125].
Moreover, Yao et al. synthesized 2-(4-methoxyphenyl)-1,3-butadiene (2-MOPB) through the dehydration of 2-(4-methoxyphenyl) but-3-en-2-ol in the presence of pyridinium p-toluenesulfonate which is used as a dehydrating agent. The manipulations were conducted under mild conditions, with a moderate yield of 46% [119]. Phenyl-1,3-butadiene derivatives provide a promising platform for the synthesis of polar-group functionalized 1,3-butadienes by introducing substituents into phenyl [126,127,128]. In our previous work, 4-methyl-4-phenyl-1,3-dioxane (MPD) was synthesized via Prins condensation of alpha-methylstyrene with formaldehyde and MPD could be easily converted into 2-PB over solid acid catalysts at above 250 °C [129]. Notably, 2-PB can hardly be modified with functional groups in the presence of a conjugated bond. However, it is possible to introduce functional groups into MPD, and then the functionalized MPD can be easily converted into functionalized 2-PB and 1-PB (Figure 9). Therefore, this is a promising candidate route for the synthesis of 2-PB with functional groups on a large scale.

5. Polar-Group Functionalized 1,3-butadienes

A large number of functionalized dienes with polar groups containing O, N, and Si, such as vinyloxy, ethoxymethyl, cyanomethyl, amine derivatives, alkyloxysilyl, alkylsilyl, alkyloxymethyl, N,N-dialkylamino dimethylsilyl, etc., have been reported.
2-Vinyloxy-1,3-butadiene is a potentially valuable building block and monomer, as it can be taken as either a functionalized 1,3-butadiene or vinyl ether (CH2=CHOR). It has been synthesized by a hydrative trimerization reaction using acetylene and water as raw materials at 80–115 °C in the KOH·H2O/DMSO system (Figure 10) [130,131]. It is suggested that one molecule of acetylene reacts with one molecule of water to form acetaldehyde (Figure 10c), which undergoes ethynylation in the presence of another molecule of acetylene to give acetylenic alcohol (Figure 10d), followed by the vinylation to afford (Figure 10e). After isomerization of acetylene-allene-diene, 2-vinyloxy-1,3-butadiene is obtained [132]. Recently, Vitkovskaya et al. studied the mechanism of this reaction by using the MP2/6-311++G**//B3LYP/6-31+G* quantum chemical framework [133,134]. They clarify that the formation of acetaldehyde and its ethynylation occur in complexes of 5DMSO·KOH, and the addition of the hydroxide ion to the acetylene molecule is the rate-determining step of this reaction.
2-Ethoxymethyl-1,3-butadiene is synthesized as shown in Figure 11 [135,136,137]. The prepared 2-bromomethyl-1,3-butadiene (Figure 11a) is added dropwise to sodium ethoxide solution at 0 °C, followed by stirring overnight at 40 °C; 2-ethoxymethyl-1,3-butadiene (Figure 11b) is obtained after distillation under reduced pressure at a yield of 63%. Dry THF is employed as the solvent. The used 2-bromomethyl-1,3-butadiene can be synthesized as follows: firstly, bromine is added dropwise to isoprene (Figure 11h) in a dry ice/acetone bath (maintained below –20 °C) to obtain 1,4-dibromo-2-methyl-2-butene (Figure 11i) with a yield of 100%. Following the addition of 1,3-dimethyl-3,4,5,6-tetrahydro-2(1H)-pyrimidinone (DMPU), 1,4-dibromo-2-methyl-2-butene is converted into 2-bromomethyl-1,3-butadiene which is collected by applying a vacuum.
2-Cyanomethyl-1,3-butadiene can be used for the production of synthetic rubbers that show excellent oil and solvent resistance, as well as adhesive properties owing to the introduction of cyano into butadiene [138]. As is known, a higher acrylonitrile content in NBR can give rise to a higher resistance to hydrocarbons and impermeability to gases, but it also leads to lower resilience and temperature flexibility [145,146], which indicates that oil resistance and low-temperature flexibility are mutually incompatible. Cyano-substituted monomers could provide an ideal solution to this problem [139]. Jing et al. [138,139] synthesized 2-cyanomethyl-1,3-butadiene in three steps, with the synthesis of 2-bromomethyl-1,3-butadiene from isoprene and Br2 via the first two steps (Figure 11). In the third step, sodium cyanide, acetonitrile, and tetrabutylammonium chloride are added into the reaction mixture. After stirring at room temperature for 48 h, the mixture is quenched, extracted, washed, and distillated to afford 2-cyanomethyl-1,3-butadiene (Figure 11c). The total yield is about 35%.
Quite different from 2-cyanomethyl-1,3-butadiene, the introduction of amine groups into dienes could remarkably change the properties of their final polymers, such as adhesion and solubility. It has been verified that the solubility of this kind of polymer in polar solvents is better than that of a polymer without polar-side chains, such as polybutadiene and polyisoprene. Moreover, the polymer containing the amine always presents chemical and structural versatility. Therefore, a lot of effort has been put into the synthesis of 1,3-butadiene containing amines (Figure 11) [140,141]. 2-((N,N-Dimethylamino)methyl)-1,3-butadiene (Figure 11d) is synthesized by adding 2-bromomethyl-1,3-butadiene dropwise into the mixture of diethyl amine and ether at 0 °C, followed by stirring overnight at room temperature. In this process, an aqueous solution of dimethyl amine (40%) is used. After distillation, a colorless liquid product is obtained with a 65% yield [142,143,144]. By a similar method, 2-((N,N-diethylamino) methyl)-1,3-butadiene (Figure 11e) and 2-((N,N-di-n-propylamino)methyl)-1,3-butadiene (Figure 11g) are synthesized using 2-bromomethyl-1,3-butadiene and the corresponding amines, and both of their yields can reach 80%. An alternative route is displayed in Figure 12.
The pre-prepared 2-bromomethyl-1,3-butadiene can also be synthesized as shown in Figure 13 [144]. Prior to charging condensed SO2 as a liquid into a stainless-steel reactor, an acetone/dry ice bath is employed to maintain the mixture temperature below –10 °C. After the introduction of isoprene, methanol, and hydroquinone, the reactor is sealed and heated to 85 °C for 4 h with stirring. After cooling the mixture to room temperature, deionized water is added to allow for recrystallization of the product in Figure 13a, which is obtained with a yield of 90%. Then, the product (Figure 13a), together with N-bromosuccinimide, benzyl peroxide, and chloroform, is charged into the round-bottom flask equipped with a condenser. After refluxing for 20 h, evaporation of the solvent and recrystallization of the product are carried out to afford the product in Figure 13b with a yield of 25%. A trace amount of hydroquinone and this product (Figure 13b) are charged into a preheated flask (170 °C) using an oil bath. After melting the solid, a vacuum is applied and the greenish-brown liquid product (Figure 13c) is gained with a yield of 80% (2-bromomethyl-1,3-butadiene).
Sato et al. [147] reported a synthetic route of 2-(triisopropoxysilyl)-l,3-butadiene in 1984; since then, this kind of functional diene has attracted wide interest. In this route, 2-(triisopropoxysilyl)-1,3-butadiene is prepared using 1,4-dichloro-2-(trichlorosilyl)-2-butene and isopropyl alcohol as starting materials (Figure 14) [148,149]. Firstly, the solution of isopropyl alcohol in triethylamine is added dropwise to a THF solution of l,4-dichloro-2-(trichlorosilyl)-2-butene (Figure 14a) at 0 °C. After stirring overnight at 20 °C, the mixture is heated to 60 °C and kept for several hours, followed by cooling to 0 °C. Hexane is applied to precipitate the inorganic salt, which is removed using Hyflo-Super-Cel” (Johns-Manville Co.). Then, l,4-dichloro-2-(triisopropoxysilyl)-2-butene (Figure 14b) is obtained by distillation. In the second step, the THF solution of l,4-dichloro-2-(triisopropoxysilyl)-2-butene (Figure 14b) is dropwise added to a mixture of zinc powder in THF under refluxing condition. After refluxing for 2 h, the reaction mixture is cooled down to 0 °C. In this stage, pentane is used to precipitate the inorganic salt. 2-(Triisopropoxysilyl)-l,3-butadiene (Figure 14c) is afforded with a yield of 64% after removing the salt and solvent. By a similar method, 2-triethoxysilyl-1,3-butadiene (Figure 14d) is synthesized from 1,4-dichloro-2-(trichlorosilyl)-2-butene and ethanol, and 2-(diisopropoxymethylsilyl)-l,3-butadiene (Figure 14e) and 2-(dimethylisopropoxysilyl)-l,3-butadiene (Figure 14f) are synthesized via a reaction of isopropyl alcohol with 1,4-dichloro-2- (dichloromethylsilyl)-2-butene or l,4-dichloro-2-(dimethylchlorosilyl)-2-butene, respectively.
2-Triethoxymethyl-1,3-butadiene is synthesized by a coupling reaction of tetraethyl orthocarbonate and 2-(1,3-butadienyl) magnesium chloride (Figure 14) [150]. A THF solution of tetraethyl orthocarbonate is charged into a flask equipped with a reflux condenser. A THF solution of 2-(1,3-butadienyl) magnesium chloride is added dropwise under a N2 atmosphere at 55–65 °C. Then the reaction mixture is stirred for 48 h. After the reaction, the mixture is concentrated and poured into a NH4Cl solution containing ice, then extracted three times with ether. After removing the solvent, distillation is employed to attain 2-triethoxymethyl-1,3-butadiene (Figure 14j).
2-(Trimethylsilyl)-1,3-butadiene is synthesized as given in Figure 14. Trimethylsilyl chloride is added to a solution of pre-prepared 2-(1,3-butadienyl) magnesium chloride in THF at room temperature. Then the mixture is heated and refluxed for 3 h. After the reaction, the mixture is poured into 2 N HCl and extracted with pentane. After removing the solvent, 2-(trimethylsilyl)-1,3-butadiene is obtained (Figure 14k). The synthesis of pre-prepared 2-(1,3-butadienyl) magnesium chloride has been described elsewhere [151].
2-(N,N-Dialkylamino)dimethylsilyl-1,3-butadiens are synthesized by the reactions of 1,3-butadien-2-ylmagnesium chloride with the corresponding (N,N-dialkylamino)dimethylsilyl chloride (Figure 14) [135]. In typical procedures, 2-((N,N-diethylamino) dimethylsilyl)-1,3-butadiene (Figure 14n) is synthesized by dropwise adding a solution of (N,N-diethylamino) dimethylsilyl chloride (Figure 14m) to 1,3-butadien-2-ylmagnesium chloride (Figure 14h) under an atmosphere of nitrogen at 0 °C. THF is used as the solvent. Followed by refluxing for 12 h, the mixture is cooled, filtered, and washed with a solution of dry pentane/THF under a nitrogen atmosphere. After evaporation and distillation over LiAlH4, 2-((N,N-diethylamino) dimethylsilyl)-1,3-butadiene is afforded as a colorless liquid with a yield of 40%. According to similar procedures, (N,N-dibutylamino) dimethylsilyl chloride (Figure 14o), (1-pyrrolidinyl)-dimethylsilyl chloride (Figure 14q), and (N-(2′-(N′,N′-dimethylamino) ethyl)-N-methylamino)dimethylsilyl chloride (Figure 14s) are used to synthesize 2-((N,N-dibutylamino) dimethylsilyl)-1,3-butadiene (Figure 14p) at a 13% yield, 2-((1-Pyrrolidinyl) dimethylsilyl)-1,3-butadiene (Figure 14r) at a 68% yield, and 2-((N-(2′-(N′,N′-Dimethylamino) ethyl)-N-methylamino)-dimethylsilyl)-1,3-butadiene (Figure 14t) at a 71% yield. Moreover, 1,3-butadienes with abundant functional groups at the 2-position have been reported (Figure 15) [152,153].

6. Conclusions and Outlook

1,3-Butadienes (especially IB and IP) are the basic building blocks for the synthetic rubbers used in the tire fabrication industry, and it is needless to mention that their production still depends on petrochemicals in industry. Taking into consideration the depletion of oil reserves, the volatile price of petrochemicals, and the issues related to the environment, it is urgent to explore renewable resources for the sustainable production of 1,3-butadienes, especially for those in large demand in the industry [88,154,155]. The production of 1,3-butadienes from biomass-derived feedstocks is a prospective route, which is increasingly competitive in polymer production [156]. With regards to sustainability, great achievements have been obtained in the production of IB from ethanol in both academia and the industry. Unfortunately, it is still a great challenge to improve the BD yield and to develop robust catalysts because the aldol coupling and the dehydrogenation steps involved in this technology cannot be efficiently performed on simple metal oxide catalysts [157].
Bio-derived C4 alcohols and especially unsaturated C4 alcohols have been declared promising feedstocks for the production of BD via the dehydration process; even the two-step synthesis of BD from C4 alcohols via UOLs displays a total BD yield of 96.6% over Yb2O3 at 360 °C. A great deal of work still needs to be done, especially for the improving catalyst life for industry usage. For the sustainable production of IP, a lot of efforts must be made. Although some biomass-derived chemicals, such as itaconic acid, have been explored in the last 2 years, it is still difficult to meet the requirements for industrial-scale production. Extensive research studies on IP synthesis are urgently required, and the exploration of novel platform chemicals faces challenges but will be highly rewarding. Fortunately, the US Department of Energy (DOE) “Top 10” report (the DOE outlined research needs for bio-based products in 2004) and its revision provide guidance to choose potential candidate chemicals for IP synthesis, as well as BD [158,159,160].
In contrast, functionalized 1,3-butadienes, without any research on their synthesis from biomass-derived chemicals, are always synthesized at the lab scale by complicated operations under rigorous conditions, e.g., without water or oxygen, such as the synthesis of these monomers with phenyl, silicon, cyano, and amino groups. However, they play an essential role in enhancing the existing properties of synthetic rubber, particularly for certain special applications. Therefore, exploration of a simple catalytic route to synthesize functionalized-1,3-butadienes at a large scale is still needed. In China, research work is ongoing in this direction at the Key Laboratory of High-Performance Synthetic Rubber and its Composite Materials, Changchun Institute of Applied Chemistry, CAS and is expected to continue in the next few years.

Author Contributions

Y.Q. performed the literature search and drafted the manuscript; Z.L. and C.B. made great efforts to revise the manuscript; S.L., L.C., Q.D., J.H., and W.D. participated in the checking of this work.

Funding

This work was supported by the National Natural Science Foundation of China (No. 51873203); National Natural Science Foundation of China (No. 51473156); Key Projects of Jilin Province Science and Technology Development Plan (No. 20160204028GX); and Key Projects of Jilin Province Science and Technology Development Plan (No. 20180201087GX).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

1,3-BDO1,3-Butanediol
1,4-BDO1,4-Butanediol
1-MPB1-(4-Methypenyl)-1,3-butadiene
1-PB1-Phenyl-1,3-butadiene
2,3-BDO2,3-Butanediol
2B1OL2-Buten-1-ol
2-MOPB2-(4-Methoxyphenyl)-1,3-butadiene
2-MTHF2-Methyltetrahydrofuran
2-PB2-Phenyl-1,3-butadiene
3B1OL3-Buten-1-ol
3B2OL3-Buten-2-ol
3-MTHF3-Methyltetrahydrofuran
BD1,3-Butadiene
BDOsButanediols
BRButadiene rubber
CRChloroprene rubber
DMD4,4-Dimethyldioxane-1,3
IPIsoprene
IRIsoprene rubber
MPD4-Methyl-4-phenyl-1,3-dioxane
MTBEMethyl tertiarybutyl ether
MVKMethyl ethyl ketone
NBRNitrile butadiene rubber
NBNatural rubber
SBRStyrene butadiene rubber
TgGlass transition temperature
THFTetrahydrofuran
UOLsUnsaturated C4 alcohols

References

  1. Sarkar, P.; Bhowmick, A.K. Sustainable rubbers and rubber additives. J. Appl. Polym. Sci. 2018, 135, 45701. [Google Scholar] [CrossRef]
  2. Zubov, A.; Pokorny, J.; Kosek, J. Styrene-butadiene rubber (SBR) production by emulsion polymerization: Dynamic modeling and intensification of the process. Chem. Eng. J. 2012, 207–208, 414–420. [Google Scholar] [CrossRef]
  3. Thiele, S.K.H.; Wilson, D.R. Alternate Transition Metal Complex Based Diene Polymerization. J. Macromol. Sci. Part C 2003, 43, 581–628. [Google Scholar] [CrossRef]
  4. Porri, L.; Giarrusso, A. Conjugated diene polymerization, Pergamon Press plc. Compr. Polym. Sci. 1989, 4, 53–108. [Google Scholar]
  5. Baugh, L.S.; Canich, J.A.M. Stereoselective Polymerization with Single-Site Catalysts; CRC Press: Boca Raton, FL, USA, 2007. [Google Scholar]
  6. Park, S.; Takeuchi, D.; Osakada, K. Pd Complex-Promoted Cyclopolymerization of Functionalized α,ω-Dienes and Copolymerization with Ethylene to Afford Polymers with Cyclic Repeating Units. J. Am. Chem. Soc. 2006, 128, 3510–3511. [Google Scholar] [CrossRef] [PubMed]
  7. Dai, Q.; Zhang, X.; Hu, Y.; He, J.; Shi, C.; Li, Y.; Bai, C. Regulation of the cis-1,4- and trans-1,4-Polybutadiene Multiblock Copolymers via Chain Shuttling Polymerization Using a Ternary Neodymium Organic Sulfonate Catalyst. Macromolecules 2017, 50, 7887–7894. [Google Scholar] [CrossRef]
  8. Wang, B.; Bi, J.; Zhang, C.; Dai, Q.; Bai, C.; Zhang, X.; Hu, Y.; Jiang, L. Highly active and trans-1,4 specific polymerization of 1,3-butadiene catalyzed by 2-pyrazolyl substituted 1,10-phenanthroline ligated iron (II) complexes. Polymer 2013, 54, 5174–5181. [Google Scholar] [CrossRef]
  9. Hirao, A.; Goseki, R.; Ishizone, T. Advances in Living Anionic Polymerization: From Functional Monomers, Polymerization Systems, to Macromolecular Architectures. Macromolecules 2014, 47, 1883–1905. [Google Scholar] [CrossRef]
  10. Marconi, W.; Mazzei, A.; Lugli, G.; Bruzzone, M. Slereospecific copolymers 1, 3-butadiene-2-phenylbutadiene. J. Polym. Sci. Part C Polym. Symp. 1967, 16, 805–819. [Google Scholar] [CrossRef]
  11. Buonerba, A.; Cuomo, C.; Speranza, V.; Grassi, A. Crystalline Syndiotactic Polystyrene as Reinforcing Agent of cis-1,4-Polybutadiene Rubber. Macromolecules 2010, 43, 367–374. [Google Scholar] [CrossRef]
  12. Schneiderman, D.K.; Hillmyer, M.A. 50th Anniversary Perspective: There Is a Great Future in Sustainable Polymers. Macromolecules 2017, 50, 3733–3749. [Google Scholar] [CrossRef]
  13. Ezinkwo, G.O.; Tretyakov, V.P.; Aliyu, A.; Ilolov, A.M. Fundamental Issues of Catalytic Conversion of Bio-Ethanol into Butadiene. ChemBioEng Rev. 2014, 1, 194–203. [Google Scholar] [CrossRef]
  14. Gallo, J.M.; Bueno, J.; Schuchardt, U. Catalytic transformations of ethanol for biorefineries. J. Braz. Chem. Soc. 2014, 25, 2229–2243. [Google Scholar] [CrossRef]
  15. Bruijnincx, P.C.A.; Weckhuysen, B.M. Shale Gas Revolution: An Opportunity for the Production of Biobased Chemicals? Angew. Chem. Int. Ed. 2013, 52, 11980–11987. [Google Scholar] [CrossRef] [PubMed]
  16. DeRosa, S.E.; Allen, D.T. Impact of Natural Gas and Natural Gas Liquids Supplies on the United States Chemical Manufacturing Industry: Production Cost Effects and Identification of Bottleneck Intermediates. ACS Sustain. Chem. Eng. 2015, 3, 451–459. [Google Scholar] [CrossRef]
  17. Duan, H.; Yamada, Y.; Sato, S. Future Prospect of the Production of 1,3-Butadiene from Butanediols. Chem. Lett. 2016, 45, 1036–1047. [Google Scholar] [CrossRef]
  18. Ipatieff, V. To the question of the decomposition of ethyl alcohol due to various catalysts. J. Russ. Phys. Chem. Soc. 1903, 35, 449–452. [Google Scholar]
  19. Gorin, Y.A. On The Catalytic Conversion of Alcohols Into Hydrocarbons of The Divinyl Series. 2. A Study of The Process of Formation of Divinyl From Ethyl Alcohol. Zhurnal Obshchei Khimii 1946, 16, 283–294. [Google Scholar]
  20. Goldstein, R.F.; Waddams, A.L. The Petroleum Chemicals Industry; Spon: Hamburg, Germany, 1967. [Google Scholar]
  21. Sushkevich, V.L.; Ivanova, I.I.; Taarning, E. Ethanol conversion into butadiene over Zr-containing molecular sieves doped with silver. Green Chem. 2015, 17, 2552–2559. [Google Scholar] [CrossRef]
  22. Sushkevich, V.L.; Palagin, D.; Ivanova, I.I. With Open Arms: Open Sites of ZrBEA Zeolite Facilitate Selective Synthesis of Butadiene from Ethanol. ACS Catal. 2015, 5, 4833–4836. [Google Scholar] [CrossRef]
  23. Tret’yakov, V.F.; Talyshinskii, R.M.; Ilolov, A.M.; Maksimov, A.L.; Khadzhiev, S.N. Initiated conversion of ethanol to divinyl by the Lebedev reaction. Pet. Chem. 2014, 54, 195–206. [Google Scholar] [CrossRef]
  24. Gruver, V.; Sun, A.; Fripiat, J.J. Catalytic properties of aluminated sepiolite in ethanol conversion. Catal. Lett. 1995, 34, 359–364. [Google Scholar] [CrossRef]
  25. León, M.; Díaz, E.; Ordóñez, S. Ethanol catalytic condensation over Mg–Al mixed oxides derived from hydrotalcites. Catal. Today 2011, 164, 436–442. [Google Scholar] [CrossRef]
  26. Larina, O.V.; Kyriienko, P.I.; Soloviev, S.O. Effect of the Addition of Zirconium Dioxide on the Catalytic Properties of ZnO/MgO-SiO2 Compositions in the Production of 1,3-Butadiene from Ethanol. Theor. Exp. Chem. 2015, 51, 252–258. [Google Scholar] [CrossRef]
  27. Ochoa, J.V.; Bandinelli, C.; Vozniuk, O.; Chieregato, A.; Malmusi, A.; Recchi, C.; Cavani, F. An analysis of the chemical, physical and reactivity features of MgO–SiO2 catalysts for butadiene synthesis with the Lebedev process. Green Chem. 2016, 18, 1653–1663. [Google Scholar] [CrossRef]
  28. Angelici, C.; Meirer, F.; van der Eerden, A.M.J.; Schaink, H.L.; Goryachev, A.; Hofmann, J.P.; Hensen, E.J.M.; Weckhuysen, B.M.; Bruijnincx, P.C.A. Ex Situ and Operando Studies on the Role of Copper in Cu-Promoted SiO2–MgO Catalysts for the Lebedev Ethanol-to-Butadiene Process. ACS Catal. 2015, 5, 6005–6015. [Google Scholar] [CrossRef]
  29. Tsuchida, T.; Kubo, J.; Yoshioka, T.; Sakuma, S.; Takeguchi, T.; Ueda, W. Reaction of ethanol over hydroxyapatite affected by Ca/P ratio of catalyst. J. Catal. 2008, 259, 183–189. [Google Scholar] [CrossRef]
  30. Quattlebaum, W.; Toussaint, W.; Dunn, J. Deoxygenation of certain aldehydes and ketones: Preparation of butadiene and styrene1. J. Am. Chem. Soc. 1947, 69, 593–599. [Google Scholar] [CrossRef]
  31. Lebedev, S. Preparation of bivinyl directly from alcohol. I. Zh Obshch Khim 1933, 3, 698–717. [Google Scholar]
  32. Delacaillerie, J.B.D.; Gruver, V.; Fripiat, J.J. Modification of the Surface Properties of Natural Phyllosilicate Sepiolite by Secondary Isomorphic Substitution. J. Catal. 1995, 151, 420–430. [Google Scholar] [CrossRef]
  33. Arundale, E.; Mikeska, L. The Olefin-Aldehyde Condensation. The Prins Reaction. Chem. Rev. 1952, 51, 505–555. [Google Scholar] [CrossRef]
  34. Kvisle, S.; Aguero, A.; Sneeden, R.P.A. Transformation of ethanol into 1,3-butadiene over magnesium oxide/silica catalysts. Appl. Catal. 1988, 43, 117–131. [Google Scholar] [CrossRef]
  35. Bhattacharyya, S.K.; Avasthi, B.N. Catalytic conversion of ethanol to butadiene by two-step process in fluidised bed. J. Appl. Chem. 1966, 16, 239–244. [Google Scholar] [CrossRef]
  36. da Ros, S.; Jones, M.D.; Mattia, D.; Schwaab, M.; Noronha, F.B.; Pinto, J.C. Modelling the effects of reaction temperature and flow rate on the conversion of ethanol to 1, 3-butadiene. Appl. Catal. A Gen. 2017, 530, 37–47. [Google Scholar] [CrossRef]
  37. Baylon, R.A.; Sun, J.; Wang, Y. Conversion of ethanol to 1, 3-butadiene over Na doped ZnxZryOz mixed metal oxides. Catal. Today 2016, 259, 446–452. [Google Scholar] [CrossRef]
  38. Cheong, J.L.; Shao, Y.; Tan, S.J.R.; Li, X.; Zhang, Y.; Lee, S.S. Highly Active and Selective Zr/MCF Catalyst for Production of 1,3-Butadiene from Ethanol in a Dual Fixed Bed Reactor System. ACS Sustain. Chem. Eng. 2016, 4, 4887–4894. [Google Scholar] [CrossRef]
  39. Zhu, Q.; Wang, B.; Tan, T. Conversion of Ethanol and Acetaldehyde to Butadiene over MgO–SiO2 Catalysts: Effect of Reaction Parameters and Interaction between MgO and SiO2 on Catalytic Performance. ACS Sustain. Chem. Eng. 2017, 5, 722–733. [Google Scholar] [CrossRef]
  40. Hayashi, Y.; Akiyama, S.; Miyaji, A.; Sekiguchi, Y.; Sakamoto, Y.; Shiga, A.; Koyama, T.; Motokura, K.; Baba, T. Experimental and computational studies of the roles of MgO and Zn in talc for the selective formation of 1,3-butadiene in the conversion of ethanol. Phys. Chem. Chem. Phys. 2016, 18, 25191–25209. [Google Scholar] [CrossRef]
  41. Huang, X.; Men, Y.; Wang, J.; An, W.; Wang, Y. Highly active and selective binary MgO–SiO2 catalysts for the production of 1,3-butadiene from ethanol. Catal. Sci. Technol. 2017, 7, 168–180. [Google Scholar] [CrossRef]
  42. Shylesh, S.; Gokhale, A.A.; Scown, C.D.; Kim, D.; Ho, C.R.; Bell, A.T. From Sugars to Wheels: The Conversion of Ethanol to 1,3-Butadiene over Metal-Promoted Magnesia-Silicate Catalysts. ChemSusChem 2016, 9, 1462–1472. [Google Scholar] [CrossRef]
  43. Sushkevich, V.L.; Ivanova, I.I. Ag-Promoted ZrBEA Zeolites Obtained by Post-Synthetic Modification for Conversion of Ethanol to Butadiene. ChemSusChem 2016, 9, 2216–2225. [Google Scholar] [CrossRef] [PubMed]
  44. Klein, A.; Keisers, K.; Palkovits, R. Formation of 1,3-butadiene from ethanol in a two-step process using modified zeolite-β catalysts. Appl. Catal. A Gen. 2016, 514, 192–202. [Google Scholar] [CrossRef]
  45. Kyriienko, P.I.; Larina, O.V.; Soloviev, S.O.; Orlyk, S.M.; Dzwigaj, S. High selectivity of TaSiBEA zeolite catalysts in 1, 3-butadiene production from ethanol and acetaldehyde mixture. Catal. Commun. 2016, 77, 123–126. [Google Scholar] [CrossRef]
  46. Gao, M.; Jiang, H.; Zhang, M. The influence of calcination temperatures on the acid-based properties and catalytic activity for the 1, 3-butadiene synthesis from ethanol/acetaldehyde mixture. Appl. Surf. Sci. 2018, 439, 1072–1078. [Google Scholar] [CrossRef]
  47. Tripathi, A.; Faungnawakij, K.; Laobuthee, A.; Assabumrungrat, S.; Laosiripojna, N. Catalytic Activity of Bimetallic Cu-Ag/MgO-SiO2 Toward the Conversion of Ethanol to 1,3-Butadiene. Int. J. Chem. React. Eng. 2016, 14, 945–954. [Google Scholar] [CrossRef]
  48. Wang, Y.; Liu, S. Butadiene Production from Ethanol. J. Bioprocess Eng. Biorefinery 2012, 1, 33–43. [Google Scholar] [CrossRef]
  49. Makshina, E.V.; Dusselier, M.; Janssens, W.; Degrève, J.; Jacobs, P.A.; Sels, B.F. Review of old chemistry and new catalytic advances in the on-purpose synthesis of butadiene. Chem. Soc. Rev. 2014, 43, 7917–7953. [Google Scholar] [CrossRef] [Green Version]
  50. Angelici, C.; Weckhuysen, B.M.; Bruijnincx, P.C.A. Chemocatalytic Conversion of Ethanol into Butadiene and Other Bulk Chemicals. ChemSusChem 2013, 6, 1595–1614. [Google Scholar] [CrossRef]
  51. Lanzafame, P.; Centi, G.; Perathoner, S. Catalysis for biomass and CO2 use through solar energy: Opening new scenarios for a sustainable and low-carbon chemical production. Chem. Soc. Rev. 2014, 43, 7562–7580. [Google Scholar] [CrossRef]
  52. Pomalaza, G.; Capron, M.; Ordomsky, V.; Dumeignil, F. Recent breakthroughs in the conversion of ethanol to butadiene. Catalysts 2016, 6, 203. [Google Scholar] [CrossRef]
  53. Miyaji, A.; Hiza, M.; Sekiguchi, Y.; Akiyama, S.; Shiga, A.; Baba, T. Catalysis by MgO and the Role of Zn2+ in Talc Catalysts for the Selective Production of 1,3-Butadiene from Ethanol. J. Jpn. Pet. Inst. 2018, 61, 171–181. [Google Scholar] [CrossRef]
  54. Ostromislenskiy, J. Production of butadiene. J. Russ. Phys. Chem. Soc. 1915, 47, 1472–1506. [Google Scholar]
  55. Jing, F.; Katryniok, B.; Araque, M.; Wojcieszak, R.; Capron, M.; Paul, S.; Daturi, M.; Clacens, J.M.; de Campo, F.; Liebens, A.; et al. Direct dehydration of 1,3-butanediol into butadiene over aluminosilicate catalysts. Catal. Sci. Technol. 2016, 6, 5830–5840. [Google Scholar] [CrossRef] [Green Version]
  56. Jing, F.; Katryniok, B.; Paul, S.; Fang, L.; Liebens, A.; Shen, M.; Hu, B.; Dumeignil, F.; Pera-Titus, M. Al-doped SBA-15 Catalysts for Low-temperature Dehydration of 1,3-Butanediol into Butadiene. ChemCatChem 2017, 9, 258–262. [Google Scholar] [CrossRef]
  57. Zeng, F.; Tenn, W.J.; Aki, S.N.V.K.; Xu, J.; Liu, B.; Hohn, K.L. Influence of basicity on 1,3-butadiene formation from catalytic 2,3-butanediol dehydration over γ-alumina. J. Catal. 2016, 344, 77–89. [Google Scholar] [CrossRef]
  58. Fang, L.; Jing, F.; Lu, J.; Hu, B.; Pera-Titus, M. Nano-flowered Ce@MOR hybrids with modulated acid properties for the vapor-phase dehydration of 1,3-butanediol into butadiene. Green Chem. 2017, 19, 4610–4621. [Google Scholar] [CrossRef]
  59. Nguyen, N.T.T.; Matei-Rutkovska, F.; Huchede, M.; Jaillardon, K.; Qingyi, G.; Michel, C.; Millet, J.M.M. Production of 1,3-butadiene in one step catalytic dehydration of 2,3-butanediol. Catal. Today 2019, 323, 62–68. [Google Scholar] [CrossRef]
  60. Kim, W.; Shin, W.; Lee, K.J.; Song, H.; Kim, H.S.; Seung, D.; Filimonov, I.N. 2,3-Butanediol dehydration catalyzed by silica-supported sodium phosphates. Appl. Catal. A Gen. 2016, 511, 156–167. [Google Scholar] [CrossRef]
  61. Tsukamoto, D.; Sakami, S.; Ito, M.; Yamada, K.; Yonehara, T. Production of Bio-based 1,3-Butadiene by Highly Selective Dehydration of 2,3-Butanediol over SiO2-supported Cesium Dihydrogen Phosphate Catalyst. Chem. Lett. 2016, 45, 831–833. [Google Scholar] [CrossRef]
  62. Duan, H.; Yamada, Y.; Kubo, S.; Sato, S. Vapor-phase catalytic dehydration of 2,3-butanediol to 3-buten-2-ol over ZrO2 modified with alkaline earth metal oxides. Appl. Catal. A Gen. 2017, 530, 66–74. [Google Scholar] [CrossRef]
  63. Duan, H.; Hirota, T.; Ohtsuka, S.; Yamada, Y.; Sato, S. Vapor-phase catalytic dehydration of 1,4-butanediol to 3-buten-1-ol over modified ZrO2 catalysts. Appl. Catal. A Gen. 2017, 535, 9–16. [Google Scholar] [CrossRef]
  64. Sun, D.; Arai, S.; Duan, H.; Yamada, Y.; Sato, S. Vapor-phase dehydration of C4 unsaturated alcohols to 1,3-butadiene. Appl. Catal. A Gen. 2017, 531, 21–28. [Google Scholar] [CrossRef]
  65. Wang, Y.; Sun, D.; Yamada, Y.; Sato, S. Selective production of 1,3-butadiene in the dehydration of 1,4-butanediol over rare earth oxides. Appl. Catal. A Gen. 2018, 562, 11–18. [Google Scholar] [CrossRef]
  66. Choudhary, V.; Pinar, A.B.; Sandler, S.I.; Vlachos, D.G.; Lobo, R.F. Xylose Isomerization to Xylulose and its Dehydration to Furfural in Aqueous Media. ACS Catal. 2011, 1, 1724–1728. [Google Scholar] [CrossRef]
  67. Wang, S.; Vorotnikov, V.; Vlachos, D.G. Coverage-Induced Conformational Effects on Activity and Selectivity: Hydrogenation and Decarbonylation of Furfural on Pd(111). ACS Catal. 2015, 5, 104–112. [Google Scholar] [CrossRef]
  68. Li, X.; Jia, P.; Wang, T. Furfural: A Promising Platform Compound for Sustainable Production of C4 and C5 Chemicals. ACS Catal. 2016, 6, 7621–7640. [Google Scholar] [CrossRef]
  69. Choudhary, V.; Sandler, S.I.; Vlachos, D.G. Conversion of Xylose to Furfural Using Lewis and Brønsted Acid Catalysts in Aqueous Media. ACS Catal. 2012, 2, 2022–2028. [Google Scholar] [CrossRef]
  70. Abdelrahman, O.A.; Park, D.S.; Vinter, K.P.; Spanjers, C.S.; Ren, L.; Cho, H.J.; Vlachos, D.G.; Fan, W.; Tsapatsis, M.; Dauenhauer, P.J. Biomass-Derived Butadiene by Dehydra-Decyclization of Tetrahydrofuran. ACS Sustain. Chem. Eng. 2017, 5, 3732–3736. [Google Scholar] [CrossRef]
  71. Ezinkwo, G.; Tretjakov, V.; Talyshinky, R.; Ilolov, A.; Mutombo, T. Overview of the Catalytic Production of Isoprene from different raw materials; Prospects of Isoprene production from bio-ethanol. Catal. Sustain. Energy 2013, 1, 100–111. [Google Scholar] [CrossRef]
  72. Songsiri, N.; Rempel, G.L.; Prasassarakich, P. Liquid-phase synthesis of isoprene from MTBE and formalin using cesium salts of silicotungstic acid. Mol. Catal. 2017, 439, 41–49. [Google Scholar] [CrossRef]
  73. Songsiri, N.; Rempel, G.L.; Prasassarakich, P. Liquid-Phase Synthesis of Isoprene from Methyl tert-Butyl Ether and Formalin Using Keggin-Type Heteropolyacids. Ind. Eng. Chem. Res. 2016, 55, 8933–8940. [Google Scholar] [CrossRef]
  74. Abdelrahman, O.A.; Park, D.S.; Vinter, K.P.; Spanjers, C.S.; Ren, L.; Cho, H.J.; Zhang, K.; Fan, W.; Tsapatsis, M.; Dauenhauer, P.J. Renewable Isoprene by Sequential Hydrogenation of Itaconic Acid and Dehydra-Decyclization of 3-Methyl-Tetrahydrofuran. ACS Catal. 2017, 7, 1428–1431. [Google Scholar] [CrossRef]
  75. Dumitriu, E.; On, D.T.; Kaliaguine, S. Isoprene by Prins Condensation over Acidic Molecular Sieves. J. Catal. 1997, 170, 150–160. [Google Scholar] [CrossRef]
  76. Dumitriu, E.; Hulea, V.; Fechete, I.; Catrinescu, C.; Auroux, A.; Lacaze, J.-F.; Guimon, C. Prins condensation of isobutylene and formaldehyde over Fe-silicates of MFI structure. Appl. Catal. A Gen. 1999, 181, 15–28. [Google Scholar] [CrossRef]
  77. An, L.-D.; Jiang, Z.-C.; Yin, Y.-G. Deactivation of AgxSbyOz/SiO2 Catalyst for The Condensation of Isobutene and Formaldehyde to Isopropene. Stud. Surf. Sci. Catal. 1987, 34, 159–171. [Google Scholar]
  78. Dang, Z.; Gu, J.; Yu, L. X-ray photoelectron spectroscopic study of CuSO4—MgO/SiO2 catalysts for isoprene synthesis. Appl. Catal. 1990, 63, 259–266. [Google Scholar] [CrossRef]
  79. Dang, Z.; Gu, J.; Yu, L.; Zhang, C. Vapor-phase synthesis of isoprene from formaldehyde and isobutylene over CuSO4−MOx/SiO2 catalysts. React. Kinet. Catal. Lett. 1991, 43, 495–500. [Google Scholar] [CrossRef]
  80. Krzywicki, A.; Wilanowicz, T.; Malinowski, S. Catalytic and physico-chemical properties of the Al2O3−H3PO4 system, I. Vapor phase condensation of isobutylene and formaldehyde—The Prins reaction. React. Kinet. Catal. Lett. 1979, 11, 399–403. [Google Scholar] [CrossRef]
  81. Ai, M. The formation of isoprene by means of a vapor-phase prins reaction between formaldehyde and isobutene. J. Catal. 1987, 106, 280–286. [Google Scholar] [CrossRef]
  82. Sushkevich, V.L.; Ordomsky, V.V.; Ivanova, I.I. Synthesis of isoprene from formaldehyde and isobutene over phosphate catalysts. Appl. Catal. A Gen. 2012, 441–442, 21–29. [Google Scholar] [CrossRef]
  83. Sushkevich, V.L.; Ordomsky, V.V.; Ivanova, I.I. Isoprene synthesis from formaldehyde and isobutene over Keggin-type heteropolyacids supported on silica. Catal. Sci. Technol. 2016, 6, 6354–6364. [Google Scholar] [CrossRef]
  84. Yu, X.; Zhu, W.; Zhai, S.; Bao, Q.; Cheng, D.; Xia, Y.; Wang, Z.; Zhang, W. Prins condensation for the synthesis of isoprene from isobutylene and formaldehyde over sillica-supported H3SiW12O40 catalysts. React. Kinet. Mech. Catal. 2016, 117, 761–771. [Google Scholar] [CrossRef]
  85. Ivanova, I.; Sushkevich, V.L.; Kolyagin, Y.G.; Ordomsky, V.V. Catalysis by Coke Deposits: Synthesis of Isoprene over Solid Catalysts. Angew. Chem. 2013, 125, 13199–13202. [Google Scholar] [CrossRef]
  86. Qi, Y.; Cui, L.; Dai, Q.; Li, Y.; Bai, C. Assembly line synthesis of isoprene from formaldehyde and isobutene over SiO2-supported MoP catalysts with active deposited carbon. RSC Adv. 2017, 7, 37392–37401. [Google Scholar] [CrossRef]
  87. Qi, Y.; Cui, L.; Li, Y.; Dai, Q.; Bai, C. Development a facile way to restore reactivity of deactivated phosphate catalysts for Prins reaction with the assistance of carbon deposition. Catal. Commun. 2018, 106, 11–15. [Google Scholar] [CrossRef]
  88. Canter, N. Manufacturing tires from renewable resources. Tribol. Lubr. Technol. 2017, 73, 20–22. [Google Scholar]
  89. Willke, T.; Vorlop, K.-D. Biotechnological production of itaconic acid. Appl. Microbiol. Biotechnol. 2001, 56, 289–295. [Google Scholar] [CrossRef]
  90. Rozentsvet, V.A.; Kozlov, V.G.; Korovina, N.A.; Ivanova, V.P.; Kostjuk, S.V. Cationic polymerization of 1,3-pentadiene coinitiated by zinc halides. J. Appl. Polym. Sci. 2013, 128, 1771–1778. [Google Scholar] [CrossRef]
  91. Kostjuk, S.V. Recent progress in the Lewis acid co-initiated cationic polymerization of isobutylene and 1,3-dienes. RSC Adv. 2015, 5, 13125–13144. [Google Scholar] [CrossRef]
  92. Ren, L.; Liu, K.; He, Q.; Ou, E.; Lu, Y.; Xu, W. Anionic polymerization of 1,3-pentadiene in toluene: Homopolymer, alternating and block copolymers. RSC Adv. 2016, 6, 51533–51543. [Google Scholar] [CrossRef]
  93. Kumar, A.; Hackenberg, J.D.; Zhuo, G.; Steffens, A.M.; Mironov, O.; Saxton, R.J.; Goldman, A.S. High yields of piperylene in the transfer dehydrogenation of pentane catalyzed by pincer-ligated iridium complexes. J. Mol. Catal. A Chem. 2017, 426, 368–375. [Google Scholar] [CrossRef]
  94. Kundu, S.; Lyons, T.W.; Brookhart, M. Synthesis of Piperylene and Toluene via Transfer Dehydrogenation of Pentane and Pentene. ACS Catal. 2013, 3, 1768–1773. [Google Scholar] [CrossRef]
  95. Behr, A.; Neubert, P. Piperylene—A Versatile Basic Chemical in Catalysis. ChemCatChem 2014, 6, 412–428. [Google Scholar] [CrossRef]
  96. Burnette, L.W. The production of rubber from furfural. Rubber Chem. Technol. 1945, 18, 284–285. [Google Scholar] [CrossRef]
  97. Schniepp, L.; Geller, H. The Preparation of 1,3- and 1,4-Pentadienes from Furfural. J. Am. Chem. Soc. 1945, 67, 54–56. [Google Scholar] [CrossRef]
  98. Casas, E.; de Ancos, B.; Valderrama, M.J.; Cano, P.; Peinado, J.M. Pentadiene production from potassium sorbate by osmotolerant yeasts. Int. J. Food Microbiol. 2004, 94, 93–96. [Google Scholar] [CrossRef] [Green Version]
  99. Bui, P.; Cecilia, J.A.; Oyama, S.T.; Takagaki, A.; Infantes-Molina, A.; Zhao, H.; Li, D.; Rodríguez-Castellón, E.; López, A.J. Studies of the synthesis of transition metal phosphides and their activity in the hydrodeoxygenation of a biofuel model compound. J. Catal. 2012, 294, 184–198. [Google Scholar] [CrossRef]
  100. Sun, R.; Zheng, M.; Li, X.; Pang, J.; Wang, A.; Wang, X.; Zhang, T. Production of renewable 1,3-pentadiene from xylitol via formic acid-mediated deoxydehydration and palladium-catalyzed deoxygenation reactions. Green Chem. 2017, 19, 638–642. [Google Scholar] [CrossRef]
  101. Kumbhalkar, M.D.; Buchanan, J.S.; Huber, G.W.; Dumesic, J.A. Ring Opening of Biomass-Derived Cyclic Ethers to Dienes over Silica/Alumina. ACS Catal. 2017, 7, 5248–5256. [Google Scholar] [CrossRef]
  102. Chern, Y.-T.; Shiue, H.-C. Low Dielectric Constants of Soluble Polyimides Based on Adamantane. Macromolecules 1997, 30, 4646–4651. [Google Scholar] [CrossRef]
  103. Fukukawa, K.-I.; Shibasaki, Y.; Ueda, M. A Photosensitive Semi-Alicyclic Poly(benzoxazole) with High Transparency and Low Dielectric Constant. Macromolecules 2004, 37, 8256–8261. [Google Scholar] [CrossRef]
  104. Feng, F.; Mitsuishi, M.; Miyashita, T.; Okura, I.; Asai, K.; Amao, Y. Preparation of Polymer Langmuir−Blodgett Films Containing Porphyrin Chromophore. Langmuir 1999, 15, 8673–8677. [Google Scholar] [CrossRef]
  105. van Reenen, A.J.; Mathias, L.J.; Coetzee, L. Polymerization of olefins with bulky substituents. 1. Homo- and copolymerization of 3-(1-adamantyl)propene. Polymer 2004, 45, 799–804. [Google Scholar] [CrossRef]
  106. Acar, H.Y.; Jensen, J.J.; Thigpen, K.; McGowen, J.A.; Mathias, L.J. Evaluation of the Spacer Effect on Adamantane-Containing Vinyl Polymer Tg’s. Macromolecules 2000, 33, 3855–3859. [Google Scholar] [CrossRef]
  107. Ishizone, T.; Tajima, H.; Torimae, H.; Nakahama, S. Anionic Polymerizations of 1-Adamantyl Methacrylate and 3-Methacryloyloxy-1,1′-biadamantane. Macromol. Chem. Phys. 2002, 203, 2375–2384. [Google Scholar] [CrossRef]
  108. Hashimoto, T.; Makino, Y.; Urushisaki, M.; Sakaguchi, T. Living cationic polymerization of 2-adamantyl vinyl ether. J. Polym. Sci. Part A Polym. Chem. 2008, 46, 1629–1637. [Google Scholar] [CrossRef]
  109. Matsumoto, A.; Tanaka, S.; Otsu, T. Synthesis and characterization of poly (1-adamantyl methacrylate): Effects of the adamantyl group on radical polymerization kinetics and thermal properties of the polymer. Macromolecules 1991, 24, 4017–4024. [Google Scholar] [CrossRef]
  110. Kobayashi, S.; Matsuzawa, T.; Matsuoka, S.-I.; Tajima, H.; Ishizone, T. Living Anionic Polymerizations of 4-(1-Adamantyl)styrene and 3-(4-Vinylphenyl)-1,1‘-biadamantane. Macromolecules 2006, 39, 5979–5986. [Google Scholar] [CrossRef]
  111. Mathias, L.J.; Lewis, C.M.; Wiegel, K.N. Poly(ether ether ketone)s and Poly(ether sulfones) with Pendent Adamantyl Groups. Macromolecules 1997, 30, 5970–5975. [Google Scholar] [CrossRef]
  112. Kobayashi, S.; Kataoka, H.; Ishizone, T.; Kato, T.; Ono, T.; Kobukata, S.; Arimoto, K.; Ogi, H. Synthesis of well-defined random and block copolymers of 2-(1-adamantyl)-1,3-butadiene with isoprene via anionic polymerization. React. Funct. Polym. 2009, 69, 409–415. [Google Scholar] [CrossRef]
  113. Kobayashi, S.; Kataoka, H.; Ishizone, T. Synthesis of Well-Defined Poly(ethylene-alt-1-vinyladamantane) via Living Anionic Polymerization of 2-(1-Adamantyl)-1,3-butadiene, Followed by Hydrogenation. Macromolecules 2009, 42, 5017–5026. [Google Scholar] [CrossRef]
  114. Cai, Y.; Lu, J.; Zuo, D.; Li, S.; Cui, D.; Han, B.; Yang, W. Extremely High Glass Transition Temperature Hydrocarbon Polymers Prepared through Cationic Cyclization of Highly 3,4-Regulated Poly(Phenyl-1,3-Butadiene). Macromol. Rapid Commun. 2018, 39, 1800298. [Google Scholar] [CrossRef] [PubMed]
  115. Marvel, C.; Woolford, R. 2-Phenyl-1,3-butadiene and Related Compounds1. J. Org. Chem. 1958, 23, 1658–1660. [Google Scholar] [CrossRef]
  116. Pragliola, S.; Cipriano, M.; Boccia, A.C.; Longo, P. Polymerization of Phenyl-1,3-butadienes in the Presence of Ziegler-Natta Catalysts. Macromol. Rapid Commun. 2002, 23, 356–361. [Google Scholar] [CrossRef]
  117. Suzuki, T.; Tsuji, Y.; Takegami, Y.; Harwood, H.J. Microstructure of poly (2-phenylbutadiene) prepared by anionic initiators. Macromolecules 1979, 12, 234–239. [Google Scholar] [CrossRef]
  118. Yao, C.; Xie, H.; Cui, D. Highly 3,4-selective living polymerization of 2-phenyl-1,3-butadiene with amidino N-heterocyclic carbene ligated rare-earth metal bis(alkyl) complexes. RSC Adv. 2015, 5, 93507–93512. [Google Scholar] [CrossRef]
  119. Yao, C.; Liu, N.; Long, S.; Wu, C.; Cui, D. Highly cis-1,4-selective coordination polymerization of polar 2-(4-methoxyphenyl)-1,3-butadiene and copolymerization with isoprene using a β-diketiminato yttrium bis(alkyl) complex. Polym. Chem. 2016, 7, 1264–1270. [Google Scholar] [CrossRef]
  120. Cai, Y.; Lu, J.; Jing, G.; Yang, W.; Han, B. High-Glass-Transition-Temperature Hydrocarbon Polymers Produced through Cationic Cyclization of Diene Polymers with Various Microstructures. Macromolecules 2017, 50, 7498–7508. [Google Scholar] [CrossRef]
  121. Yao, C.; Lin, F.; Wang, M.; Liu, D.; Liu, B.; Liu, N.; Wang, Z.; Long, S.; Wu, C.; Cui, D. Highly Syndioselective 3,4-Trans Polymerization of (E)-1-(4-Methylphenyl)-1,3-butadiene by Fluorenyl N-Heterocyclic Carbene Ligated Lutetium Bis(alkyl) Precursor. Macromolecules 2015, 48, 1999–2005. [Google Scholar] [CrossRef]
  122. Masuda, T.; Otsuki, M.; Higashimura, T. Structure and reactivity in cationic polymerization of butadiene derivatives. II. 1-phenylbutadiene. J. Polym. Sci. Polym. Chem. Ed. 1974, 12, 1385–1394. [Google Scholar] [CrossRef]
  123. Suzuki, T.; Tsuji, Y.; Takegami, Y. Microstructure of poly (1-phenylbutadiene) prepared by anionic initiators. Macromolecules 1978, 11, 639–644. [Google Scholar] [CrossRef]
  124. Asami, R.; Hasegawa, K.-I.; Onoe, T. Cationic Polymerization of Phenylbutadienes. I. Cationic Polymerization of trans-1-Phenyl-1,3-butadieiie. Polym. J. 1976, 8, 43–52. [Google Scholar] [CrossRef] [Green Version]
  125. Ikeda, Y.; Ukai, J.; Ikeda, N.; Yamamoto, H. Stereoselective synthesis of (z)- and (e)-1,3-alkadienes from aldehydes using organotitanium and lithium reagents. Tetrahedron 1987, 43, 723–730. [Google Scholar] [CrossRef]
  126. Wang, Z.; Liu, D.; Cui, D. Statistically Syndioselective Coordination (Co)polymerization of 4-Methylthiostyrene. Macromolecules 2016, 49, 781–787. [Google Scholar] [CrossRef]
  127. Li, S.; Liu, D.; Wang, Z.; Cui, D. Development of Group 3 Catalysts for Alternating Copolymerization of Ethylene and Styrene Derivatives. ACS Catal. 2018, 8, 6086–6093. [Google Scholar] [CrossRef]
  128. Liu, D.; Wang, M.; Wang, Z.; Wu, C.; Pan, Y.; Cui, D. Stereoselective Copolymerization of Unprotected Polar and Nonpolar Styrenes by an Yttrium Precursor: Control of Polar-Group Distribution and Mechanism. Angew. Chem. Int. Ed. 2017, 56, 2714–2719. [Google Scholar] [CrossRef] [PubMed]
  129. Qi, Y.; Cui, F.; He, J.; Cui, L.; Li, Y.; Dai, Q.; Bai, C. Insight into performance of lactam-based Brønsted-acidic catalysts for Prins condensation and their self-separation in water. Mol. Catal. 2018, 445, 80–86. [Google Scholar] [CrossRef]
  130. Trofimov, B.A. Acetylene and its Derivatives in Reactions with Nucleophiles: Recent Advances and Current Trends. Curr. Org. Chem. 2002, 6, 1121–1162. [Google Scholar] [CrossRef]
  131. Boris, A.T.; Nina, K.G. Acetylene: New prospects of classical reactions. Russ. Chem. Rev. 2007, 76, 507–527. [Google Scholar]
  132. Trofimov, B. New Intermediates for Organic Synthesis based on Acetylene. Zeitschrift für Chemie 1986, 26, 41–49. [Google Scholar] [CrossRef]
  133. Vitkovskaya, N.M.; Larionova, E.Y.; Skitnevskaya, A.D.; Kobychev, V.B.; Trofimov, B.A. Nucleophilic addition of methanol and methanethiol to acetylene in the superbasic system KOH-DMSO: A quantum chemical model. Russ. Chem. Bull. 2013, 62, 26–32. [Google Scholar] [CrossRef]
  134. Vitkovskaya, N.M.; Larionova, E.Y.; Skitnevskaya, A.D.; Trofimov, B.A. Hydrative trimerization of acetylene into 2-vinyloxy-1,3-butadiene in the KOH/DMSO system: A quantum chemical insight. Tetrahedron Lett. 2015, 56, 1063–1066. [Google Scholar] [CrossRef]
  135. Hirao, A.; Hiraishi, Y.; Nakahama, S.; Takenaka, K. Polymerization of Monomers Containing Functional Silyl Groups. 13. Anionic Polymerization of 2-[(N,N-Dialkylamino)dimethylsilyl]-1,3-butadiene Derivatives. Macromolecules 1998, 31, 281–287. [Google Scholar] [CrossRef]
  136. Takenaka, K.; Akagawa, Y.; Takeshita, H.; Miya, M.; Shiomi, T. Polymerization of 1, 3-Dienes Containing Functional Groups 6: Unexpected Collapse of Monomer Structure in the Anionic Polymerization of 2-Ethoxymethyl-l, 3-butadiene. Polym. J. 2009, 41, 106–107. [Google Scholar] [CrossRef]
  137. Petzhold, C.; Morschhaeuser, R.; Kolshorn, H.; Stadler, R. On the Anionic Polymerization of (Dialkylamino) isoprenes. 2. Influence of the Tertiary Amino Group on the Polymer Microstructure. Macromolecules 1994, 27, 3707–3713. [Google Scholar] [CrossRef]
  138. Jing, Y.; Sheares, V.V. Polar Functionalized Diene-Based Materials. 1. Bulk, Solution, and Emulsion Free Radical Polymerization of 2-Cyanomethyl-1,3-butadiene. Macromolecules 2000, 33, 6255–6261. [Google Scholar] [CrossRef]
  139. Jing, Y.; Sheares, V.V. Polar, Functionalized Diene-Based Materials. 2. Free Radical Copolymerization Studies of 2-Cyanomethyl-1,3-butadiene with Styrene and Acrylonitrile. Macromolecules 2000, 33, 6262–6268. [Google Scholar] [CrossRef]
  140. Yang, Y.; Sheares, V.V. Synthesis of disubstituted amine-functionalized diene-based polymers. Polymer 2007, 48, 105–109. [Google Scholar] [CrossRef]
  141. Wu, L.; Sheares, V.V. Polar, functionalized diene-based materials. V. Free-radical polymerization of 2-[(N-benzyl-n-methylamino)methyl]-1,3-butadiene and copolymerization with styrene. J. Polym. Sci. Part A Polym. Chem. 2001, 39, 3227–3238. [Google Scholar] [CrossRef]
  142. Mannebach, G.; Morschhäuser, R.; Stadler, R.; Petzhold, C. On the anionic polymerization of dialkylaminoisoprenes, 4. Experimental observation of the unusual polymerization kinetics. Macromol. Chem. Phys. 1998, 199, 909–912. [Google Scholar] [CrossRef]
  143. Sheares, V.; Li, Y.; Emmick, T.; Martin, C.; Jing, Y.; Beery, M. Polar functionalized materials via free radical polymerization of substituted dienes. In Abstracts of Papers of the American Chemical Society; American Chemical Society: Washington, DC, USA, 1999; Volume 80, p. 79. [Google Scholar]
  144. Sheares, V.V.; Wu, L.; Li, Y.; Emmick, T.K. Polar, functionalized diene-based material. III. Free-radical polymerization of 2-[(N,N-dialkylamino)methyl]-1,3-butadienes. J. Polym. Sci. Part A Polym. Chem. 2000, 38, 4070–4080. [Google Scholar] [CrossRef]
  145. Blackley, D.C. Monomers for Synthetic Rubber Production, Synthetic Rubbers: Their Chemistry and Technology; Springer: Dordrecht, The Netherlands, 1983; pp. 32–58. [Google Scholar]
  146. Blackley, D.C. Synthetic Rubbers: Their Chemistry and Technology; Applied Science Publishers: Barking, UK, 1983; pp. 126–141. [Google Scholar]
  147. Sato, F.; Uchiyama, H.; Samaddar, A.K. 1984. Available online: https://pubs.acs.org/servlet/linkout?suffix=ol070089eb00011/ol070089eb00011_1&dbid=32&doi=10.1021%2Fol070089e&key=1%3ACAS%3A528%3ADyaL2MXisVOjtw%253D%253D (accessed on 17 November 2018).
  148. Takenaka, K.; Kawamoto, S.; Miya, M.; Takeshita, H.; Shiomi, T. Polymerization of 1,3-dienes containing functional groups: 8. Free-radical polymerization of 2-triethoxysilyl1,3-butadiene. Polym. Int. 2010, 59, 891–895. [Google Scholar] [CrossRef]
  149. Takenaka, K.; Hattori, T.; Hirao, A.; Nakahama, S. Polymerization of monomers containing functional silyl groups. 6. Anionic polymerization of 2-(trialkoxysilyl)-1,3-butadiene. Macromolecules 1989, 22, 1563–1567. [Google Scholar] [CrossRef]
  150. Takenaka, K.; Hanada, K.; Shiomi, T. Polymerization of 1,3-Dienes with Functional Groups. 1. Free-Radical Polymerization of 2-Triethoxymethyl-1,3-butadiene. Macromolecules 1999, 32, 3875–3877. [Google Scholar] [CrossRef]
  151. Nunomoto, S.; Yamashita, Y. Reaction of 2-(1,3-butadienyl) magnesium chloride with carbonyl compounds and epoxides. A regioselectivity study. J. Org. Chem. 1979, 44, 4788–4791. [Google Scholar] [CrossRef]
  152. Leicht, H.; Göttker-Schnetmann, I.; Mecking, S. Stereoselective Copolymerization of Butadiene and Functionalized 1,3-Dienes. ACS Macro Lett. 2016, 5, 777–780. [Google Scholar] [CrossRef]
  153. Beery, M.D.; Rath, M.K.; Sheares, V.V. Polar, Functionalized Diene-Based Materials. 4. Polymerization Studies of 2,3-Bis(4-ethoxy-4-oxobutyl)-1,3-butadiene and Copolymerization with Styrene. Macromolecules 2001, 34, 2469–2475. [Google Scholar] [CrossRef]
  154. Marshall, J.D.; Toffel, M.W. Framing the Elusive Concept of Sustainability: A Sustainability Hierarchy. Environ. Sci. Technol. 2005, 39, 673–682. [Google Scholar] [CrossRef] [Green Version]
  155. Sikdar, S.K. Sustainability Perspective and Chemistry-Based Technologies. Ind. Eng. Chem. Res. 2007, 46, 4727–4733. [Google Scholar] [CrossRef]
  156. Hillmyer, M.A. The promise of plastics from plants. Science 2017, 358, 868–870. [Google Scholar] [CrossRef]
  157. Sushkevich, V.L.; Ivanova, I.I.; Ordomsky, V.V.; Taarning, E. Design of a Metal-Promoted Oxide Catalyst for the Selective Synthesis of Butadiene from Ethanol. ChemSusChem 2014, 7, 2527–2536. [Google Scholar] [CrossRef] [PubMed]
  158. Becker, J.; Lange, A.; Fabarius, J.; Wittmann, C. Top value platform chemicals: Bio-based production of organic acids. Curr. Opin. Biotechnol. 2015, 36, 168–175. [Google Scholar] [CrossRef] [PubMed]
  159. Jang, Y.-S.; Kim, B.; Shin, J.H.; Choi, Y.J.; Choi, S.; Song, C.W.; Lee, J.; Park, H.G.; Lee, S.Y. Bio-based production of C2–C6 platform chemicals. Biotechnol. Bioeng. 2012, 109, 2437–2459. [Google Scholar] [CrossRef] [PubMed]
  160. Bozell, J.J.; Petersen, G.R. Technology development for the production of biobased products from biorefinery carbohydrates—the US Department of Energy’s “Top 10” revisited. Green Chem. 2010, 12, 539–554. [Google Scholar] [CrossRef]
Figure 1. The general reaction mechanism and major by-products of 1,3-butadiene (BD) production from ethanol [19,23,24,25,26,27,28,29,30,31,32,33,34,35].
Figure 1. The general reaction mechanism and major by-products of 1,3-butadiene (BD) production from ethanol [19,23,24,25,26,27,28,29,30,31,32,33,34,35].
Catalysts 09 00097 g001
Figure 2. Synthesis of BD from ethanol. Route (a) was proposed by Ipatiev at al. [18]; Route (b) was proposed by Ostromyslensky at al. [54].
Figure 2. Synthesis of BD from ethanol. Route (a) was proposed by Ipatiev at al. [18]; Route (b) was proposed by Ostromyslensky at al. [54].
Catalysts 09 00097 g002
Figure 3. Synthesis of BD from butanediols (BDOs) or C4 unsaturated alcohols (UOLs) [17,62,63,64,65]. Evolution from reagents and intermediates to the target BD is labeled by arrows, and the by-products are produced during Dehydration 1.
Figure 3. Synthesis of BD from butanediols (BDOs) or C4 unsaturated alcohols (UOLs) [17,62,63,64,65]. Evolution from reagents and intermediates to the target BD is labeled by arrows, and the by-products are produced during Dehydration 1.
Catalysts 09 00097 g003
Figure 4. Synthesis of BD from THF through the intermediate 3-buten-1-ol (3B1OL), and the by-products in the decyclization of THF [70].
Figure 4. Synthesis of BD from THF through the intermediate 3-buten-1-ol (3B1OL), and the by-products in the decyclization of THF [70].
Catalysts 09 00097 g004
Figure 5. Overview of isoprene (IP) production routes [71,72,73,74].
Figure 5. Overview of isoprene (IP) production routes [71,72,73,74].
Catalysts 09 00097 g005
Figure 6. Synthesis of 1,3-pentadiene from 2-methyltetrahydrofuran (2-MTHF) (a) and Xylitol (b) [100,101].
Figure 6. Synthesis of 1,3-pentadiene from 2-methyltetrahydrofuran (2-MTHF) (a) and Xylitol (b) [100,101].
Catalysts 09 00097 g006
Figure 7. Synthesis of 2-adamantyl-1,3-butadiene [112,113].
Figure 7. Synthesis of 2-adamantyl-1,3-butadiene [112,113].
Catalysts 09 00097 g007
Figure 8. Synthesis of 2-phenyl-1,3-butadiene [115,116,117,118,119].
Figure 8. Synthesis of 2-phenyl-1,3-butadiene [115,116,117,118,119].
Catalysts 09 00097 g008
Figure 9. A straightforward way to synthesize 2-phenyl-1,3-butadiene [129].
Figure 9. A straightforward way to synthesize 2-phenyl-1,3-butadiene [129].
Catalysts 09 00097 g009
Figure 10. The pathway of synthesis of 2-vinyloxy-1,3-butadiene [130,131,132].
Figure 10. The pathway of synthesis of 2-vinyloxy-1,3-butadiene [130,131,132].
Catalysts 09 00097 g010
Figure 11. Synthesis of functionalized 1,3-butadiene using 2-bromomethyl-1,3-butadiene [135,136,137,138,139,140,141,142,143,144].
Figure 11. Synthesis of functionalized 1,3-butadiene using 2-bromomethyl-1,3-butadiene [135,136,137,138,139,140,141,142,143,144].
Catalysts 09 00097 g011
Figure 12. Alternative route for the synthesis of amine groups functionalized 1,3-butadiene [144].
Figure 12. Alternative route for the synthesis of amine groups functionalized 1,3-butadiene [144].
Catalysts 09 00097 g012
Figure 13. Synthesis of 2-bromomethyl-1,3-butadiene [144].
Figure 13. Synthesis of 2-bromomethyl-1,3-butadiene [144].
Catalysts 09 00097 g013
Figure 14. Synthesis of 2-substituted 1,3-butadienes [135,147,148,149,150,151].
Figure 14. Synthesis of 2-substituted 1,3-butadienes [135,147,148,149,150,151].
Catalysts 09 00097 g014
Figure 15. Reported 2-substituted 1,3-butadienes [152,153].
Figure 15. Reported 2-substituted 1,3-butadienes [152,153].
Catalysts 09 00097 g015
Table 1. Synthesis of BD from ethanol over various catalyst systems recently.
Table 1. Synthesis of BD from ethanol over various catalyst systems recently.
CatalystReaction Temperature °CConversion%BD Selectivity %Ref.BD Formation rate/mmol (gcat.h)−1
ZnO–MgO/H–β280 35043.663.4[36]0.92
NaZn1Zr10Oz–H35054.428[37]9.07
Cu/MCF–Zr/MCF235, 4009637[38]25.92
MgO–SiO2–500 a500Total 29.780.7[39]0.96
Talc/Zn40048.461.0[40]8.53
MgO–SiO2 (65:35)4509577[41]25
3%Au/MgO–SiO2 300>45~60[42]2.4
Ag/Zr(3.3)BEA(38)32015~60[43]10.3
Cu/SiO2 MgO/H–β280100, 300-33% yield[44]-
Ta3.0SiBEA–EtOH/AA=3.235058.973.1[45]-
2%ZrO2/NanoSiO2–500 a32058.5293.18[46]-
5 % Ag/MgO–SiO2 275>50>28 yield[47]-
2.5% Cu-2.5%wt Ag/MgO–SiO2 30060>40 yield[47]-
a Acetaldehyde was co-fed with ethanol.

Share and Cite

MDPI and ACS Style

Qi, Y.; Liu, Z.; Liu, S.; Cui, L.; Dai, Q.; He, J.; Dong, W.; Bai, C. Synthesis of 1,3-Butadiene and Its 2-Substituted Monomers for Synthetic Rubbers. Catalysts 2019, 9, 97. https://doi.org/10.3390/catal9010097

AMA Style

Qi Y, Liu Z, Liu S, Cui L, Dai Q, He J, Dong W, Bai C. Synthesis of 1,3-Butadiene and Its 2-Substituted Monomers for Synthetic Rubbers. Catalysts. 2019; 9(1):97. https://doi.org/10.3390/catal9010097

Chicago/Turabian Style

Qi, Yanlong, Zaizhi Liu, Shijun Liu, Long Cui, Quanquan Dai, Jianyun He, Wei Dong, and Chenxi Bai. 2019. "Synthesis of 1,3-Butadiene and Its 2-Substituted Monomers for Synthetic Rubbers" Catalysts 9, no. 1: 97. https://doi.org/10.3390/catal9010097

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop