Next Article in Journal
Acid Properties of GO and Reduced GO as Determined by Microcalorimetry, FTIR, and Kinetics of Cellulose Hydrolysis-Hydrogenolysis
Next Article in Special Issue
Bio-DEE Synthesis and Dehydrogenation Coupling of Bio-Ethanol to Bio-Butanol over Multicomponent Mixed Metal Oxide Catalysts
Previous Article in Journal
New Insight into the In Situ SO2 Poisoning Mechanism over Cu-SSZ-13 for the Selective Catalytic Reduction of NOx with NH3
Previous Article in Special Issue
Entrapping Immobilisation of Lipase on Biocomposite Hydrogels toward for Biodiesel Production from Waste Frying Acid Oil
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dehydration of Biomass-Derived Butanediols over Rare Earth Zirconate Catalysts

1
Graduate School of Engineering, Chiba University, Yayoi, Inage 263-8522, Japan
2
Graduate School of Science and Engineering, Ehime University, 2–5 Bunkyocho, Matsuyama, Ehime 790-8577, Japan
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(12), 1392; https://doi.org/10.3390/catal10121392
Submission received: 29 October 2020 / Revised: 24 November 2020 / Accepted: 27 November 2020 / Published: 29 November 2020
(This article belongs to the Special Issue Catalysis in Biomass Valorization)

Abstract

:
The aim of this work is to develop an effective catalyst for the conversion of butanediols, which is derivable from biomass, to valuable chemicals such as unsaturated alcohols. The dehydration of 1,4-, 1,3-, and 2,3-butanediol to form unsaturated alcohols such as 3-buten-1-ol, 2-buten-1-ol, and 3-buten-2-ol was studied in a vapor-phase flow reactor over sixteen rare earth zirconate catalysts at 325 °C. Rare earth zirconates with high crystallinity and high specific surface area were prepared in a hydrothermal treatment of co-precipitated hydroxide. Zirconates with heavy rare earth metals, especially Y2Zr2O7 with an oxygen-defected fluorite structure, showed high catalytic performance of selective dehydration of 1,4-butanediol to 3-buten-1-ol and also of 1,3-butanediol to form 3-buten-2-ol and 2-buten-1-ol, while the zirconate catalysts were less active in the dehydration of 2,3-butanediol. The calcination of Y2Zr2O7 significantly affected the catalytic activity of the dehydration of 1,4-butanediol: a calcination temperature of Y2Zr2O7 at 900 °C or higher was efficient for selective formation of unsaturated alcohols. Y2Zr2O7 with high crystallinity exhibits the highest productivity of 3-buten-1-ol from 1,4-butanediol at 325 °C.

Graphical Abstract

1. Introduction

Biomass has been regarded as an alternative resource to manufacture useful chemicals in the petrochemical industry. Butanediols (BDOs) such as 2,3-butanediol (2,3-BDO), 1,4-butanediol (1,4-BDO) and 1,3-butanediol (1,3-BDO) are derivable in the direct microbial conversion of renewable materials, and they are summarized in several reviews [1,2,3,4,5,6,7]. The microbial production of 2,3-BDO is most widely investigated among the BDO isomers and has a long history over a century [1]. 2,3-BDO is produced from corncob and kenaf core as well as sugars such as glucose, xylose, and sucrose. The production of 1,4-BDO from biomass has developed rapidly [8,9,10,11]: Yim et al. first reported the direct production of 1,4-BDO from glucose by using Escherichia coli strains [8]. 1,4-BDO can be also produced through the chemical process of hydrogenation of biomass-derived succinic acid [12]. 1,3-BDO can be produced from glucose using Escherichia coli [13,14]. 1,3-BDO can be also produced from 4-hydroxy-2-butanone, which is derived from biomass-derived levulinic acid, through microbial fermentation [15,16]. The biomass-derived BDOs can be utilized to produce chemicals such as 1,3-butadiene (BD), 3-buten-1-ol (3B1ol), tetrahydrofuran (THF), etc., which are summarized in recent reviews [17,18,19,20].
The direct production of BD from BDOs has been reported [21,22,23,24,25,26]. Phosphate catalysts are effective for the BD production from 2,3-BDO [21,22]. Acidic catalysts such as zeolites and mesoporous solid acids are active for the BD formation from 1,3-BDO [23,24,25,26] while propylene is produced as a major by-product. In our pioneering research, it has been found that 1,4-BDO is preferentially dehydrated to 3B1ol in the vapor phase over ZrO2 catalysts [27,28]. Pure rare earth oxides such as CeO2, Er2O3, and Yb2O3 are also active for the reaction and more selective to 3B1ol than ZrO2 catalysts [17,29,30]. Zhang et al. reported an active CaO-ZrO2 acid-base catalyst for the dehydration of 1,4-BDO [31]. ZrO2-supported Yb2O3 and CaO also show high selectivity to 3B1ol [32,33]. We recently investigated the catalytic dehydration of 1,4- and 1,3-BDO over yttria-stabilized tetragonal zirconia (YSZ) catalysts [34]. In the dehydration of 1,4-BDO, a crystalline YSZ with a Y2O3 content of 3.2 wt.% exhibits an excellent stable catalytic activity: the highest selectivity to 3B1ol of 75.3% at the 1,4-BDO conversion of 26.1% is obtained at 325 °C. The YSZ catalyst also shows an excellent catalytic activity in the dehydration of 1,3-BDO: the highest selectivity to unsaturated alcohols (UOLs) such as 3-buten-2-ol (3B2ol) and 2-buten-1-ol (2B1ol) over 98% is attained at a conversion of 66%.
We have reported an excellent Yb2Zr2O7 catalyst with high crystalline of oxygen-defected fluorite for the vapor-phase dehydration of 1,3-BDO to produce UOLs [35]. The sample prepared through a hydrothermal (HT) process in an ammonia media is confirmed to be oxygen-defected type cubic fluorite, Yb2Zr2O7, while the as-prepared co-precipitate sample is amorphous. The Yb2Zr2O7 maintains a high specific surface area (SA) as ca. 40 m2 g−1 even after being calcined at temperatures higher than 800 °C, in contrast to the fact that catalysts without HT are readily sintered at the temperatures. The Yb2Zr2O7 calcined at 900 °C shows the best catalytic performance: 1,3-BDO conversion of 82% is achieved with the total selectivity to UOLs higher than 95% at 325 °C. Also, Fang et al. have recently reported that Ln2Zr2O7 (Ln = La, Pr, and Sm) pyrochlores as well as Y2Zr2O7 fluorite show low-temperature activity for the oxidative coupling of methane [36]. Thus, we expected that rare earth zirconate (REZrO) samples including Yb2Zr2O7 would have a great possibility as a catalyst to activate glycol compounds and to produce UOLs through selective dehydration of BDOs.
In this study, we focused on the efficient production of UOLs from 1,4-BDO as well as BDO isomers such as 1,3- and 2,3-BDO (Figure 1), and we performed catalyst screening in 16 REZrO samples. In the preliminary screening, we have found out Y2Zr2O7 as a promising catalyst in the dehydration of 1,4- and 1,3-BDO to produce 3B1ol and a mixture of 2B1ol and 3B2ol, respectively. Detail reaction conditions that Y2Zr2O7 works efficiently in the catalytic dehydration of 1,4-BDO were estimated, and reaction mechanisms for the dehydration of 1,4-BDO over Y2Zr2O7 were discussed.

2. Results

2.1. Structural Property of REZrO Catalysts

Various REZrO samples were prepared by HT treatment using ammonia as a precipitant and a mineralizer. Table 1 lists SA of REZrO samples calcined at 900 °C. The SA values were in the range between 16 and 39 m2 g−1. To confirm that the composition of RE/Zr, XRF analyses of REZrOs were performed, and the results are also summarized in Table 1. A typical N2 adsorption-desorption isotherm of YZrO is shown in Figure S1. In most of the samples except LaZrO, PrZrO, CeZrO, and ScZrO, the ratios of RE/Zr were close to 1 or a little larger, and one of the reasons is that Zr reagent has Hf as an impurity. This indicates that the REZrO samples have a composition of RE/Zr = ca. 1. LaZrO, PrZrO, CeZrO, and ScZrO samples, however, had a RE/Zr ratio less than 1.
Figure 2 depicts continuous scanning XRD patterns of 16 REZrO samples prepared through HT process using ammonia followed by calcination at 900 °C, where RE = La, Ce, Pr, Nd, Sm, Eu, Gd, Tb, Dy, Ho, Y, Er, Tm, Yb, Lu, and Sc. In the zirconates of light rare earth such as La, Ce, Pr, Nd, Sm, and Eu together with GdZrO, the samples were composed of a zirconate and a separated rare earth oxide. For example, the XRD pattern of CeZrO fitted with the patterns of tetragonal Zr0.88Ce0.12O2 (PDF 01-082-1398, P42/nmc, No. 137) and cubic fluorite-type Ce0.91Zr0.0.9O2 (PDF 01-075-9496, Fm3m, No. 225) phases, and EuZrO had two phases such as cubic defect fluorite-type Eu2Zr2O7 (PDF 01-078-1292, Fm3m, No. 225) and cubic bixbyite Eu2O3 (PDF 01-073-6281, I213, No. 199). In addition, no monoclinic ZrO2 phase was observed in the REZrO samples, whereas pure ZrO2 prepared in the HT process without rare earth elements has a monoclinic structure [34]. Although La2O3-ZrO2, Nd2O3-ZrO2, and Sm2O3-ZrO2 prepared in HT process using KOH as a mineralizer are pyrochlore-type RE2Zr2O7 (Fd3m, No. 227), for example, La2Zr2O7 (PDF 01-070-5602) and Sm2Zr2O7 (PDF 00-024-1012) [35], LaZrO, NdZrO, and SmZrO were composed of different phases in the present paper using ammonia as a mineralizer. In contrast to the light rare earth zirconates, REZrOs of heavy rare earth such as TbZrO, DyZrO, HoZrO, YZrO, ErZrO, TmZrO, YbZrO, and LuZrO were composed of a single phase of zirconate, cubic fluorite-type RE2Zr2O7 such as Dy2Zr2O7 (PDF 01-078-1293), Ho2Zr2O7 (PDF 01-078-1294), Er2Zr2O7 (PDF 01-078-1299), and Yb2Zr2O7 (PDF 01-078-1300). However, YZrO was not assigned; both pyrochlore-type (PDF 01-074-9311) and fluorite-type Y2Zr2O7 (PDF 01-081-8080) were possible, as shown in Figure S2. To confirm the structure of YZrO, Rietveld analysis was performed in Section 2.3. Among the REZrOs, the ScZrO sample has a rhombohedral structure, Zr3Sc4O12 (PDF 01-071-1022, R-3, No. 148). The XRD results indicate that the ScZrO sample has Sc-rich composition.

2.2. Dehydration of Different BDOs over Various REZrO Catalysts

The dehydration of three BDOs such as 1,4-, 1,3-, and 2,3-BDO was performed over REZrO catalysts at a reaction temperature of 325 °C and a space time, W/F, of 0.31 h where W and F are the catalyst weight (g) and the flow rate of BDO (g h−1), respectively. Table 2 shows the results in the dehydration of 1,4-BDO over 16 REZrO catalysts calcined at 900 °C. Six REZrOs such as LaZrO, PrZrO, NdZrO, CeZrO, SmZrO, and ScZrO showed low conversions of 1,4-BDO below 40% with the selectivity to 3B1ol lower than 66%. Over LaZrO, NdZrO, and ScZrO, THF was the major product. In contrast, the other ten REZrO catalysts showed the 1,4-BDO conversions higher than 50% with the 3B1ol selectivity over 75%. Among the REZrOs, the selectivity to the total of 3B1ol, 2B1ol, and BD, which maintained a straight C4 carbon chain structure, exceeded 90% over DyZrO, YZrO, and ErZrO. In particular, YZrO showed the highest conversion of 77.6% as well as the highest 3B1ol selectivity of 87.7%. It was confirmed that the catalytic activity of YZrO was higher than the most active CaO/ZrO2 catalyst that had been ever reported [33].
Table 3 shows the comparison of reactivity of different BDOs in the dehydration of 1,4-, 1,3-, and 2,3-BDO over several REZrO catalysts calcined at 900 °C under the same reaction conditions as those in Table 2. In the dehydration of 1,3-BDO over 6 selected REZrO catalysts such as LaZrO, CeZrO, YZrO, YbZrO, LuZrO, and ScZrO, YZrO and YbZrO catalysts showed the 1,3-BDO conversions higher than 89% whereas LaZrO and ScZrO showed low conversions below 50%. The catalysts had high selectivity to UOLs over 80%. The sum of selectivity to UOLs and BD exceeded 94%, except LaZrO. In particular, over YZrO, YbZrO, and LuZrO, the sum of selectivity to UOLs and BD exceeded 97%. In the same way as 1,4-BDO, YZrO showed the highest 1,3-BDO conversion of 93.9% with high UOLs selectivity of 90.2%, while the selectivity to BD was the highest.
In the dehydration of 2,3-BDO over several REZrO catalysts calcined at 900 °C (Table 3), under the same reaction conditions, the REZrO catalysts showed 2,3-BDO conversions lower than 20% with the selectivity to 3B2ol at most 50%. Unfortunately, the catalytic activity of REZrO as well as YZrO is inferior to the catalysts such as ZrO2 and Sc2O3 previously reported in the references [38,39]. Thus, in the catalyst screening, it is concluded that the REZrO catalysts, especially YZrO, are active for the vapor-phase dehydration of 1,4- and 1,3-BDO, but they are not suitable for the activation of 2,3-BDO. In each REZrO catalyst, the highest UOLs yield was obtained in the dehydration of 1,3-BDO and the lowest one in 2,3-BDO. Therefore, the reactivity order of BDOs over REZrO catalysts was 1,3-BDO > 1,4-BDO > 2,3-BDO. Because YZrO showed the highest 1,4-BDO conversion and the highest 3B1ol selectivity, we investigated the YZrO catalyst in detail in the following sections.

2.3. Effect of Calcination Temperature on the Structure of YZrO Catalysts and Catalytic Activity in the Dehydration of 1,4-BDO

Table 4 lists SA of YZrO samples calcined at different temperatures. The SA value was decreased when the raising calcination temperature from 91 to 16 m2 g−1 at the temperature range between 600 and 1050 °C. The particle size calculated from the SA, DBET, is also summarized in Table 4. Figure 3 shows continuous scanning XRD analysis of YZrO samples calcined at different temperatures. The as-prepared sample was crystallized during the HT process. All the samples calcined at 600–1050 °C had a major diffraction peak at 29.9 degree. The crystallite size, DXRD, which was estimated from of the major diffraction peak using the Scherrer equation, is also summarized. The DXRD value was simply increased when raising the calcination temperature. There is a significant difference between the crystallite size and the particle size: DXRD was much smaller than DBET. To confirm the morphology of catalysts calcined at different temperatures.
For YZrO, we observed the FE-SEM image (Figure 4). The SEM image clearly shows that the YZrO sample calcined at 900 °C is composed of agglomerates of particles with the size of 30–50 nm, which is consistent with the DBET value of 31 nm rather than the DXRD of 6.6 nm. This indicates that the particles shown in the SEM image are non-porous primary particles.
Even at a high calcination temperature of 1050 °C, the result of Rietveld analysis was not satisfied within the fitting errors because of low diffraction intensity. Figure 5 shows the step-scanning XRD pattern of YZrO sample calcined at 1200 °C. We confirmed diffraction angles are the same between YZrO calcined at 1050 and 1200 °C. Thus, Rietveld analysis was performed in a YZrO sample calcined at 1200 °C to refine the structure parameter (Figure 5). We confirmed that the sample was composed of defect fluorite-type Y2Zr2O7, but not pyrochlore. Table 5 lists the refined structure parameters of Y2Zr2O7.
Black points and red lines represent observed and calculated patterns, respectively. Green vertical marks indicate the Bragg reflection positions of defected-fluorite Y2Zr2O7 phase. Blue line at the bottom shows the difference between the observed and calculated patterns (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article).
Figure 6 shows the effect of calcination temperature of YZrO on the conversion of 1,4-BDO, and the numerical data are located in Table S1. The 1,4-BDO conversion was maximized at 900 °C, and the selectivity to 3B1ol increased with raising the calcination temperature. The selectivity to THF showed an opposite trend to the selectivity to 3B1ol. The conversion, which increases with raising calcination temperature, is in reverse proportion to specific surface area at the range between 600 and 900 °C.

2.4. Catalytic Dehydration of 1,4-BDO over YZrO Calcined at 900 °C under Different Reaction Conditions

Figure 7 shows the change in conversion with different space time at 325 °C. The space time, W/F, was controlled by the variation of W at a constant F, where the numerical data are listed in Table S2. The 1,4-BDO conversion simply increased with increasing the space time. At a W/F = 0.94 h, 99.4% of the 1,4-BDO conversion was attained. The selectivity to 3B1ol gradually decreased from 87.7 to 77.9% with the increase in the selectivity to 2B1ol and BD, while the sum of selectivity to UOLs and BD seemed to be constant. This means that both 2B1ol and BD are regarded as the secondary products produced from 3B1ol. The selectivity to THF, however, was almost constant in the variation of W/F. Therefore, 3B1ol and THF are the primary products in the dehydration of 1,4-BDO. In addition, the UOLs formation rate at 325 °C was almost constant at conversions lower than 80% (Table S2): the rate is reduced at a W/F = 0.31 h by 10% by comparing at 0.19 h. At high conversions near 100%, 2B1ol and BD increased with decreasing 3B1ol (Figure 7 and Table 6).
Table 6 summarized the conversion data at different reaction temperatures in the dehydration of 1,4-BDO over YZrO calcined at 900 °C. With raising the reaction temperature, the conversion was increased, and the complete conversion was attained at 360 °C. The selectivity to 3B1ol was decreased through the maximum at 325 °C: the selectivity to 3B1ol and UOLs was 87.7 and 90.5%, respectively. At a temperature of 350 °C or higher, BD was produced together with the production of propylene, which was formed via the decomposition of 3B1ol [40,41]. The propylene produced at 375 °C was comparable to that of CO2: the selectivity to CO2 at 375 °C was 1.8%, which indicates that a molar ratio of propylene/CO2 is 1.07.

2.5. Effect of Carrier Gas on the Catalytic Dehydration of 1,4-BDO over YZrO Calcined at 900 °C

The effect of carrier gas in the dehydration of 1,4-BDO over YZrO catalyst was investigated to clarify the role of basic and acidic sites of YZrO catalyst: either CO2, NH3, or H2 was adapted as a carrier gas instead of N2 gas. Table 7 shows the effect of carrier gas on the conversion of 1,4-BDO over YZrO at 325 and 300 °C. In CO2 flow, the conversion, the 3B1ol selectivity, and the 3B1ol productivity were smaller than those in N2 flow. Thus, CO2 gas could work as a poison of basic sites. In NH3 flow at 325 °C, in contrast, the conversion level and the 3B1ol productivity were slightly higher than those in N2 flow: an NH3 seems to act as an accelerator, but not as a poison. It is much clearer, at 300 °C, that NH3 gas accelerates the reaction. An NH3 gas may act as a base catalyst together with catalyst surface for the abstraction of hydrogen from 1,4-BDO. It is possible that either weakly adsorbed NH3 or gas-phase one may act as a catalyst.
Then, we adopted pre-adsorption of NH3 before the reaction to confirm the above-mentioned hypothesis. In Figure S3, the conversion at TOS of the initial 2 h was higher than 25%, which is higher than that in N2, but it reduced to 13% after TOS of 2 h. This indicates either hydrogen-bonded or weakly adsorbed molecular NH3 acts as a base catalyst, but it removes from the surface during the reaction. Thus, NH3 adsorbed on YZrO at 300 °C acted as a poison of acidic sites in N2 flow (Table 7). In the formation of THF, however, the selectivity to THF was preferential in CO2 flow, and it was depressed in NH3 flow at both 325 and 300 °C. In H2 flow, the 1,4-BDO conversion was higher than that in N2 flow at 325 °C. H2 gas accelerates the formation of 3B1ol in the dehydration of 1,4-BDO over YZrO catalyst. The acceleration effect of H2 will be discussed in Section 3.3.
Figure 8 shows the changes in the conversion of 1,4-BDO with time on stream in long run tests. The catalytic activity, i.e., the 1,4-BDO conversion and the selectivity to each product, was stabilized at a conversion of ca. 87% in H2 flow for 30 h. In N2 flow, however, a gradual decrease in the 1,4-BDO conversion was observed; the conversion was decreased from 79 to 75% for 30 h. As shown in Table 7, the 1,4-BDO conversion in N2 was lower than that in H2 although the selectivity to 3B1ol was higher. The highest 3B1ol productivity of 27.0 mol h−1 kg−1was obtained in H2 flow, while it was 24.2 mol h−1 kg−1 in N2 flow.

2.6. Acid-base Properties of YZrO Catalyst Calcined at 600 and 900 °C

Figure 9 illustrates temperature-programmed desorption (TPD) profiles of adsorbed NH3 and CO2 on YZrO calcined at 600 and 900 °C. In Figure 9A, the amount of desorbed CO2 from YZrO calcined at 600 °C is more than that of NH3. The amounts of acidic and basic sites were calculated to be 0.193 and 0.313 mmol g−1 up to 600 °C, respectively. In contrast, the TPD profiles of NH3 and CO2 adsorbed on YZrO calcined at 900 °C were quite similar to each other (Figure 9B): the amounts of acidic and basic sites were calculated to be 0.123 and 0.127 mmol g−1 up to 500 °C, respectively.
Acid strength distributions obtained for both catalysts showed little variations in the low temperature range. In the high temperature range above 400 °C, the reduction of NH3 released during TPD for the sample calcined at 900 °C suggests a slight reduction of the stronger acid sites when rising the calcination temperature from 600 to 900 ºC. In addition, the total number of acidic sites was clearly reduced, when rising the calcination temperature. Base strength distribution showed the same trend as the acid strength distribution, when comparing both YZrO catalysts. Acidic and basic sites on which NH3 and CO2 adsorbed at temperatures higher than 325 °C are considered to be poisoned in the reaction at 325 °C. As long as strong acidic and strong basic sites measured by the desorption at >325 °C, the amounts of acidic and basic sites of YZrO calcined at 600 °C were calculated to be 0.041 and 0.074 mmol g−1, while those of YZrO calcined at 900 °C were 0.015 and 0.022 mmol g−1, respectively.

3. Discussion

3.1. Structure of REZrO as an Effective Catalyst for the Dehydration of BDOs

In the preparation of zirconate, KOH is frequently used as a mineralizer: La2Zr2O7 is synthesized through the HT process of double hydrous oxides of La and Zr using KOH as the mineralizer [42]. Liu et al. have synthesized REZrO ceramics such as Yb2Zr2O7 and Gd2Zr2O7 at calcination temperatures above 1550 °C [43]. An example of REZrO for catalyst use is rare; Fang et al. recently reported that oxygen-defected fluorite Y2Zr2O7 is prepared with a simple co-precipitation method using ammonia as a precipitant: the sample calcined at 800 °C has SA as small as 7 m2 g−1 and is used for methane oxidative coupling [36]. We previously prepared REZrO samples in HT conditions using KOH solution, but they had small SA values at most 10 m2 g−1 and showed low conversion in the dehydration of 1,3-BDO [35]. However, Yb2Zr2O7 prepared by HT aging using ammonia has a high SA of 49 m2 g−1 and shows an excellent catalytic activity in the dehydration of 1,3-BDO. Thus, HT aging with ammonia mineralizer would be promising for the preparation of catalyst samples with high crystallinity and high SA. In this paper, we obtained REZrO samples with high SA between 16 and 39 m2 g−1 (Table 1) in HT treatment using ammonia as a mineralizer.
In our previous paper [35], La2O3-ZrO2, Nd2O3-ZrO2, and Sm2O3-ZrO2 prepared in the HT process using KOH as a mineralizer were determined to be pyrochlore-type Ln2Zr2O7. The reason is that the ionic radius ratio of Ln/Zr determines the structure of Ln2Zr2O7 [44]: it is reported that the pyrochlore-type structure is favored at the ionic radius ratio of Ln/Zr between 1.46 and 1.78, while the defect fluorite-type structure is favorable at a ratio less than 1.46. At the ratio of 1.46, Gd2Zr2O7 could have both structures. In the continuous scanning XRD patterns (Figure 2), structures of the samples are assigned by comparing the PDF database. Most of the structural data of REZrOs of heavy rare earth were fitted with the database of cubic fluorite-type RE2Zr2O7, as described in Section 2.1. In the case of YZrO, however, the structure was fitted with both pyrochlore and fluorite phases. In Figure 5, Rietveld analysis of the YZrO calcined at 1200 °C clearly indicates the structure is fluorite. It is confirmed that the active YZrO is composed of defect fluorite-type Y2Zr2O7, but not pyrochlore. The structure is the same as that of oxygen-defected fluorite Yb2Zr2O7, which is reported previously [35].

3.2. Dehydration of Different BDOs over Various REZrO Catalysts

In the catalyst screening, it is concluded that the reactivity order of BDOs over each REZrO catalyst is 1,3-BDO > 1,4-BDO > 2,3-BDO. In this work, YZrO is the most active catalyst as a weight basis for the vapor-phase dehydration of 1,4- and 1,3-BDO. We have previously reported efficient rare earth oxide catalysts for the dehydration of 1,4-BDO to produce 3B1ol [17]: rare earth oxides such as Er2O3, Yb2O3, and Lu2O3 show the highest 3B1ol formation rates in the catalytic reaction. The formation rate varies with the ionic radius of the rare earth cation.
Here, we summarized the relationship between the formation rate and the ionic radius of the rare earth cation. Figure 10 shows the intrinsic activity of REZrO catalyst: the formation rate of UOLs per unit surface area with the variation of ionic radius of rare earth cation (Figure 10A) and average ionic radius between Zr4+ and rare earth cation, R(RE+Zr)/2 (Figure 10B). The formation rate of UOLs per unit surface area was calculated from the data in Table 2, Table 3, and Table S3 for the dehydration of 1,4-, 2,3-, and 1,3-BDO, respectively. In the present study, REZrO catalysts with heavy rare earth cations such as DyZrO, HoZrO, YZrO, ErZrO, YbZrO, and LuZrO show the comparable formation rate of UOLs, YbZrO shows the highest UOLs formation rates among the REZrOs in the dehydration of both 1,3- and 1,4-BDO (Figure 10A). In comparison with pure rare earth oxides, we adapted a mathematical mean of ionic radius between RE and Zr, which is listed in Table 1. The mathematical mean of ionic radius between Gd and Zr, R(Gd+Zr)/2, is calculated to be 0.0889 nm and that between Dy and Zr, R(Dy+Zr)/2 = 0.0876 nm, where RGd = 0.0938, RDy = 0.0912, and RZr = 0.0840 nm [37]. In Figure 10B, the formation rates over REZrO are larger than those over pure rare earth oxides reported in Ref. [45]. Pure CeO2 has exceptionally high activity in the dehydration of 1,3-BDO; it shows the highest formation rate of 3.85 mmol h−1 m−2 at the ionic radius of Ce4+, 0.0970 nm, as shown in Figure 10B with the dotted line. Active REZrO catalysts appear in a narrow range of ionic radius. The mean ionic radii of 0.0889 nm for GdZrO and 0.0854 nm for YbZrO are close to the ionic radii of Er3+ in Er2O3 and Lu3+ in Lu2O3, which are 0.0890 and 0.0861 nm, respectively [45]. The mean ionic radius of 0.0870 nm, R(Y+Zr)/2, for YZrO is close to the ionic radius of Yb3+ of 0.0868 in Yb2O3, which is reported to be an active catalyst for the dehydration of 1,4-BDO [46]. This suggests that crystallites with specific size of unit cell are efficient for the selective formation of 3B1ol. Besides, because yttrium is abundant and cheap among rare earth elements, we investigated the YZrO catalyst in detail.

3.3. Significant Parameters on the Catalytic Activity of YZrO

As shown in Figure 6, the calcination temperature of YZrO significantly affects the conversion of 1,4-BDO. The effect of calcination temperature on the 1,4-BDO conversion over YZrO is quite similar to those observed in the reaction over pure rare earth oxides such as Er2O3, Lu2O3 [46], and CeO2 [47]. The significant increase in the activity is probably attributed to the change in the ratio of exposed crystal surface with the particle sizes.
In Table 7, H2 carrier gas accelerates the formation of 3B1ol at 325 °C. It has been reported in the literature that the acceleration by H2 in the dehydration of BDOs: 1,4-BDO over CaO/ZrO2 [33] and 2,3-BDO over Sc2O3 [39]. In the dehydration of 1,3-BDO over Er2O3, however, the formation of UOLs is depressed in H2 flow [45]. The effect of H2 carrier gas depends on the types of reactants and catalysts. Anyway, H2 gas accelerates the formation of 3B1ol in the dehydration of 1,4-BDO over YZrO catalyst, Y2Zr2O7. Fluorite Y2Zr2O7 has essentially oxygen defect sites on the catalyst surface, as discussed in the following section. It is well known that ZrO2 as well as fluorite CeO2 has oxygen storage capacity [48]. In addition to the original oxygen defects, oxygen vacancies could be generated on the surface of YZrO due to the oxidation of H2. The generation of oxygen vacancies by the oxidation of H2 might be the reason for enhancing the catalytic performance of YZrO in H2 flow.
In Table S2, the variation of the UOLs formation rate with W/F was almost constant at conversions lower than 80%: the formation rate was 0.69 mmol m−2 h−1 at a conversion of 77.6% while that at 51.5% conversion was 0.75 mmol m−2 h−1. Then, we estimated the effect of external mass transfer using Carberry number (Ca) [49]. Ca was calculated to be 0.05 at a conversion of 73%. Because an external concentration gradient is absent at Ca < 0.05, the effect of external mass transfer would be small at a conversion of lower than 73%. In contrast, at high conversions over 73%, a catalyst efficiency becomes low due to mass transfer limitation in this study. Thus, in Figure 10, some of the formation rate of UOLs calculated from the conversion data over 73% could be underestimated. At high conversions near 100% without 1,4-BDO, subsequent reactions proceeded: the dehydration of 3B1ol to BD and the isomerization of 3B1ol proceed (Figure 7).
Besides, we exhibited that BD was produced as a by-product in the dehydration of 1,4-BDO over YZrO even at 300 °C (Table 6). We have previously reported that 1,4-BDO can be selectively converted to BD via an intermediate of 3B1ol over Yb2O3 at 360 °C [40]. To establish an efficient catalytic system for the production of BD from 1,4-BDO, we need selective dehydration catalysts for 3B1ol without decomposition to propylene. We could obtain several candidates for the stepwise dehydration of 1,4-BDO (Table 2). Thus, the BD formation in the dehydration of 1,4-BDO will be reported in the following paper in a near future.
The dehydration of alcohols readily proceeds over strong acidic catalysts, but it is not easy that single dehydration of glycols proceeds to produce selectively UOLs [50,51]: the reason is that step-wise reactions of the produced UOLs frequently proceed over acidic catalysts. Catalysts with acid-base properties are reported to be effective for the formation of 3B1ol in the dehydration of 1,4-BDO [33,46]. NH3 and CO2 gases work as weak poisons for CaO/ZrO2 and Er2O3 catalysts to reduce the yield of 3B1ol whereas the dehydration proceeds even in the NH3 atmosphere. In the dehydration of 2,3-BDO, both NH3 and CO2 also work as the poisons for the formation of 3B2ol over CaO/ZrO2 [52]. In Table 7, CO2 gas works as a poison of basic sites, but gaseous NH3 accelerates the dehydration. Although the reason for the acceleration by NH3 is not clear, an NH3 gas may act as a base catalyst together with catalyst surface for the abstraction of hydrogen from 1,4-BDO. In a similar manner, over pure ZrO2 with acidic property, it is observed that the formation of 3B1ol form 1,4-BDO is accelerated under NH3 flow conditions [33]. As shown in Figure S3, either hydrogen-bonded or weakly adsorbed molecular NH3 acts as a base catalyst. After it removes from the surface during the reaction, NH3 adsorbed on YZrO acted as a poison of acidic sites. In the formation of THF, however, the formation of THF is affected by the acidic sites because the selectivity is depressed by NH3. A similar trend in the THF formation is reported in the dehydration of 1,4-BDO [33,46].

3.4. Mechanistic Considerations on the Dehydration of 1,4-BDO over YZrO Catalyst

Although YZrO samples have both acidic and basic sites, the TPD results (Figure 9) cannot explain the results that YZrO samples calcined at high temperatures have high catalytic performance. Thus, it is reasonable that the change in the ratio of exposed crystal surface with large particle sizes could be the main reason for the catalytic performance enhanced in the YZrO calcined at 900 °C.
As shown in Section 3.2 (Table 5), YZrO samples are composed of Y2Zr2O7 with an oxygen-defected fluorite structure, which is the same structure of previously reported Yb2Zr2O7 [35]. We have discussed structures of active sites for the dehydration of 1,3-BDO using a crystal model of Yb2Zr2O7 and proposed probable model structures of active sites. As shown in the previous report [35], well-crystalized Yb2Zr2O7 inevitably has oxygen defect sites on the surface. The defect site exposes three cations such as Zr4+ and Yb3+, which act as an acidic site, and it is surrounded by six O2− anions, which act as a basic site. Because the crystal structure of Y2Zr2O7 is the same as that of Yb2Zr2O7, cations such as Zr4+ and Y3+ exposed on the O2− defect site could work as acidic sites, and an O2− anion surrounding the defect site would work as a basic site (adsorption site in Figure 11A). It is reasonable that the selective dehydration of BDOs to UOLs would proceed via an acid-base concerted mechanism over the basic site of O2− and the acidic sites of Zr4+ and Y3+.
In the dehydration of BDOs, UOLs formation could be initiated by tridentate coordination with two OH groups of BDO on the acidic cation sites and a hydrogen of BDO on the basic O2− site. We speculate possible tridentate coordination structures of BDO adsorbed on YZrO surface (Figure 11A). In Figure 11A, an acceptable adsorption structure of 1,4-BDO has a 6-membered ring, Y3+-O-C-C-H-O2−, a 7-membered ring, O2−-H-C-C-C-O-Zr4+, and an 8-membered ring, Y3+-O-C4-O-Zr4+. In the dehydration of 1,3- and 2,3-BDO, adsorption structures have 7- and 6-membered ring, Y3+-O-Cn-O-Zr4+ (n = 3 and 2), respectively. It is reasonable that the accessibility of hydrogen after the coordination of HO groups, Y3+-O-Cn-O-Zr4+ (n = 3 and 2), determines the reactivity because the anchoring OH group is important for the formation of UOLs [15,34]. The low reactivity of 2,3-BDO is probably caused by large strain in the tridentate coordination structure, which has two 6-membered rings such as Y3+-O-C2-O-Zr4+ and O2−-H-C-C-O-Y3+, and a 7-membered ring of O2−-H-C-C-C-O-Zr4, as shown in Figure 11A. The 6-membered ring for the fixation of OH groups as Y3+-O-C-C-O-Zr4+ will induce large strain in the tridentate coordination. On the other hand, in the dehydration of 1,3- and 1,4-BDO, 7- and 8-membered ring coordination, Y3+-O-Cn-O-Zr4+ (n = 3 and 4), respectively, which is looser than 6-membered ring of Y3+-O-C2-O-Zr4+ in 2,3-BDO, can allow a hydrogen of BDO to access a basic O2− site to form tridentate coordination structure. It is speculated that the 7-membered ring, Y3+-O-C3-O-Zr4+, in 1,3-BDO is more stable than the 8-membered ring, Y3+-O-C4-O-Zr4+, in 1,4-BDO. It is probable that this could be the reason for the difference in the reactivity of BDOs.
Figure 11B proposes a speculative reaction mechanism for the dehydration of 1,4-BDO to produce 3B1ol. The selective dehydration of 1,4-BDO to produce 3B1ol over the defect site is explained by the adsorption as the form of tridentate coordination of 1,4-BDO to the 2 cations and an O2− anion followed by sequential dehydration: the position-2 hydrogen of 1,4-BDO is firstly abstracted by a basic O2− anion and then the position-1 hydroxyl group is subsequently or simultaneously abstracted by an acidic Y3+ cation. Probably, the first and the second steps in Figure 11B would proceed simultaneously. Another OH group at position 4 plays an important role in anchoring 1,4-BDO to the catalyst surface. Thus, the selective dehydration of 1,4-BDO to 3B1ol could proceed via the speculative base-acid concerted mechanism, in a similar manner to the 1,4-BDO dehydration catalyzed by Er2O3 catalyst, which is supported by theoretical calculation [30].
In the formation of THF, a speculative model could be proposed (Figure 11C). The strength of the acidic site is not strong so that carbocation generated by the abstraction of the OH group could not be formed. The TPD results in Figure 9 could explain the result that the YZrO calcined at 600 °C, which has acidic sites stronger than that calcined at 900 °C, shows the higher selectivity to THF than that calcined at 900 °C (Figure 6). Thus, base-acid concerted mechanism can be considered. However, the adsorption structure in Figure 11C is highly strained because it has an 8-membered ring, Y3+-O-C4-O-Zr4+, and a 9-membered ring, Y3+-O-C4-O-H-O2−. The suppression of the major by-product, THF, is significant to improve the selectivity to 3B1ol.

4. Materials and Methods

4.1. Materials

ZrO(NO3)2·2H2O, La(NO3)3·6H2O, Pr(NO3)3·nH2O, and Sm(NO3)3·6H2O were purchased from Wako Pure Chemicals Industries, Ltd., Osaka, Japan, and Ce(NO3)3·6H2O was purchased from Nakalai Tesque, Inc., Kyoto, Japan. Other rare earth (RE) nitrates such as Y(NO3)3·6H2O were purchased from Sigma-Aldrich, St. Louis, Missouri, US. In the preparation of Y2O3-ZrO2 with a Y/Zr ratio of 1, for example, 3.24 g of Y(NO3)3·6H2O and 2.26 g of ZrO(NO3)2·2H2O were dissolved in 50 cm3 of distilled water. Then, the pH value of the solution was adjusted to 10 with 25 wt.% ammonia water under stirring. After the solution was transferred into 100 cm3 Teflon-lined autoclave, HT treatment was typically performed at 200 °C for 24 h. The recovered precipitate was centrifuged, washed with distilled water, dried at 110 °C for 18 h, and finally calcined at 900 °C for 3 h. Hereafter, the resulted Y2O3-ZrO2 is named as YZrO. In the case of REZrO, a sample was prepared in the same way as YZrO except using a rare earth nitrate instead of using Y(NO3)3·6H2O. The resulting zirconate with 50 mol% RE is named as REZrO. For example, the 50 mol% Dy-containing ZrO2 is named as DyZrO.
Reactant chemicals such as 1,4- and 1,3-BDO were purchased from Wako Pure Chemicals Industries, Ltd., Osaka, Japan. 2,3-BDO was purchased from Tokyo Chemical Industry, Co., Ltd. Tokyo, Japan. They were used for the catalytic reaction without any further purification.

4.2. Characterization of Catalysts

X-ray fluorescence (XRF) analysis for the estimation of RE content in REZrO samples was performed by EDX-900HS (Shimadzu, Kyoto, Japan). The continuous scanning X-ray diffraction (XRD) was recorded on New D8 ADVANCE (Bruker Japan, Yokohama, Japan) with Cu Kα radiation (λ = 0.154 nm) to identify the crystal phase of the catalyst. A crystallite size, DXRD, of YZrO sample was calculated from the full width at half maximum (FWHM) by Scherrer equation: DXRD = /(βcosθ), where K (=0.9), λ, β, and θ are the shape factor, X-ray wavelength, FWHM, and Bragg’s diffraction angle, respectively. The step-scanning XRD of YZrO calcined at 1200 °C was recorded on XRD-7000 (Shimadzu, Kyoto, Japan) with Cu Kα radiation in order to refine structure parameter by Rietveld analysis, which was operated by using RIETAN-FP program [53]. High-resolution field emission scanning electron microscope (FE-SEM) image was observed on JSM-7100FA microscope (JEOL, Tokyo, Japan) operated at 15 kV.
The N2 adsorption isotherm of catalyst was measured in a self-made gas adsorption apparatus at −196 °C, and the SA of catalyst was calculated from the N2 isotherm fitted with the Brunauer–Emmett–Teller (BET) equation. A particle size of Y2Zr2O7, DBET, was calculated by the following equation, DBET = 6/d SA, assuming that the primary particles are spherical or cubic where d is the density of Y2Zr2O7 (d = 5.36 g cm−3, which is an arithmetic average of 5.03 and 5.68 g cm−3 for Y2O3 and ZrO2, respectively [54]). The calculation of DBET was also performed to compare the morphology of catalyst [55,56,57].
TPD of adsorbed NH3 and CO2 was performed using BELCATII (Microtrac BEL Corp., Osaka, Japan) with thermal conductivity detector (TCD) to estimate the acidity and the basicity of catalysts, respectively. The measurement conditions are the following: preheating the sample in He flow at 500 °C for 1 h; adsorption of NH3 or CO2 at 100 °C for 1 h; desorption temperature controlled from 100 to 800 °C at an increment rate of 10 °C min−1. The TCD detector can detect gases such as water and oxygen desorbed from the catalyst. Thus, we also performed a blank TPD experiment for the sample preheated at 500 °C without adsorption of NH3. Figure S4 shows the blank TCD signal profile as a function of temperature together with the original TCD signals of adsorbed NH3 and CO2. A difference TCD signal between NH3 adsorption and the blank TPD profile was converted to a NH3-TPD profile using molar sensitivity of NH3. In a similar manner, the CO2-TPD profile was obtained from the difference TCD signal between CO2 adsorption and the blank TPD profile. The amounts of acidic and basic sites were estimated from the integral of the desorption profiles of NH3 and CO2, respectively.

4.3. Catalytic Reaction

The vapor-phase dehydration of a BDO was performed in a fixed-bed down-flow glass tube reactor under an atmospheric pressure of N2 at a temperature between 300 and 375 °C, typically at 325 °C. After 0.50 g of a catalyst with the granule size of 53–250 μm had been heated in N2 flow at the prescribed reaction temperature for 1 h, a reactant BDO was fed into the reactor at a liquid feed rate of 1.60 g h−1 together with an N2 carrier gas at 30 cm3 min−1. The reaction effluent at 325 °C was collected at −78 °C every hour, whereas it was collected at 0 °C in the case of the conversion over 80% and the reaction temperature over 325 °C. Differences in the temperature of cold trap were described in the SI file (Table S4). The collected liquid products were identified using a gas chromatograph (GC) equipped with mass-spectrometer (QP5050A, Shimadzu, Kyoto, Japan) and a capillary column (DB-WAX, a length of 30 m, JW Scientific, Osaka, Japan), and they were analyzed by a GC (GC-8A, Shimadzu, Kyoto, Japan) equipped with flame ionization detector and a capillary column of Inert Cap WAX-HT (an inner diameter of 0.53 mm and a length of 30 m, GL-Science, Tokyo, Japan). 1-Hexanol was used as an internal standard for the quantitative GC analysis. The gaseous products, such as BD, propylene, and CO2, collected using a gastight syringe with a volume of 0.5 cm3 were analyzed by another GC-8A equipped with TCD and a packed column (VZ-7, a length of 6 m, GL-Science, Tokyo, Japan) using the N2 carrier gas as an internal standard.
The catalytic reaction was performed during the initial TOS of 5 h to assure the stability of the catalytic activity. All the catalysts showed a little decay in the initial period of 1 h so that we evaluated the catalytic activity we evaluated by averaging the data of conversion and selectivity for 4 h from TOS of 1 to 5 h excluding the first period of 1 h. The mass recovered in the effluent liquid was more than 97% of the mass feed. The error of reproducibility was within 3%. The conversion and selectivity were calculated according to the following equations:
Conversion   ( % ) = ( 1 mole   of   the   recovered   reactant mole   of   the   fed   reactant ) × 100
Selectivity   ( mol % ) = mole   of   carbone   in   the   specific   product mole   of   carbon   in   the   fed   reactant   ×   Conversion / 100 × 100
The poisoning experiment of YZrO catalyst was performed in the catalytic dehydration of 1,4-BDO at 300 and 325 °C to confirm whether acidic and basic carrier gases affect the catalytic activity or not. Instead of using N2 gas, CO2, NH3, or H2 was used as a carrier gas at a flow rate of 30 cm3 min−1 in the catalytic reaction. The reaction effluent was collected at 0 °C in the poisoning experiment. In another poisoning experiment, the catalytic reaction was performed in N2 carrier gas flow at 300 °C, after NH3 gas had been flowed at a rate of 30 cm3 min−1 and 300 °C for 0.5 h followed by preheating at 300 °C for 1 h.

5. Conclusions

The vapor-phase catalytic dehydration of 1,4-BDO to form unsaturated alcohols such as 3B1ol was investigated over 16 REZrO samples together with the dehydration of 1,3- and 2,3-BDO. REZrO with heavy rare earth metals, especially oxygen-defected fluorite Y2Zr2O7, showed high conversion of 1,4-BDO to 3B1ol and also high conversion of 1,3-BDO to form 3B2ol and 2B1ol, while 2,3-BDO was less reactive than the other BDOs over REZrO catalysts. The high crystallinity of oxygen-defected fluorite Y2Zr2O7 with high surface area was synthesized in the HT treatment of do-precipitate hydroxide in ammonia media. Y2Zr2O7 has advantage in the industrial application because of the highest conversion based on weight. The calcination and crystallinity of Y2Zr2O7 affected the catalytic activity of the dehydration of BDOs: a calcination temperature of 900 °C or higher was efficient for selective formation of UOLs. In the dehydration of 1,4-BDO, 3B1ol, and THF were the primary products over Y2Zr2O7 catalyst, while 3B1ol was further reacted to form 2B1ol and BD. To the best of our knowledge, Y2Zr2O7 showed the highest 3B1ol productivity of 27.0 mol h−1 kg−1 in the dehydration of 1,4-BDO at 325 °C. In the poisoning experiment, both CO2 and NH3 behaved as poisons during the dehydration of 1,4-BDO. This suggests that the reaction proceeds via acid-base concerted mechanism.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/12/1392/s1, Figure S1, N2 adsorption-desorption isotherm of YZrO; Figure S2, XRD pattern of YZrO calcined at 900 °C together with PDF data; Figure S3, Changes in catalytic activity of YZrO calcined at 900 °C with TOS with and without NH3 pre-adsorption prior to the reaction; Figure S4, original TCD signal profiles in the TPD measurement of NH3 and CO2 desorbed from YZrO calcined at 600 and 900 °C; Table S1, Dehydration of 1,4-BDO over YZrO calcined at different temperatures; Table S2, Dehydration of 1,4-BDO over YZrO calcined at 900 °C with different space time; Table S3, Dehydration of 1,3-BDO over sixteen REZrO catalysts; Table S4, Dehydration of 1,4-BDO over YZrO calcined at 900 °C in different cold trap.

Author Contributions

A.M., Y.M., K.N., F.S. and S.S. summarized the data and prepared the original draft; A.M. and Y.M. prepared samples; A.M., Y.M. and K.N. performed catalytic tests in the dehydration of butanediols; A.M., Y.M., F.S. and R.T. characterized the samples; Y.Y. supervised the analytical work; S.S. designed the project, acquired funding, and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by JSPS KAKENHI KIBAN B, grant number JP18H01784.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ji, X.J.; Huang, H.; Ouyang, P.K. Microbial 2,3-butanediol production: A state-of-the-art review. Biotechnol. Adv. 2011, 29, 351–364. [Google Scholar] [CrossRef] [PubMed]
  2. Zeng, A.P.; Sabra, W. Microbial production of diols as platform chemicals: Recent progresses. Curr. Opin. Biotechnol. 2011, 22, 749–757. [Google Scholar] [CrossRef] [PubMed]
  3. 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]
  4. Jiang, T.; Liu, W.; Zou, H.; Cheng, T.; Tian, N.; Xian, M. Microbial production of short chain diols. Microb. Cell Fact. 2014, 13, 165. [Google Scholar] [CrossRef]
  5. Sabra, W.; Groeger, C.; Zeng, A.P. Microbial cell factories for diol production. Adv. Biochem. Eng. Biotechnol. 2016, 155, 165–197. [Google Scholar]
  6. Yang, T.; Rao, Z.; Zhang, X.; Xu, M.; Xu, Z.; Yang, S.T. Metabolic engineering strategies for acetoin and 2,3-butanediol production: Advances and prospects. Critic. Rev. Biotechnol. 2017, 37, 990–1005. [Google Scholar] [CrossRef]
  7. Biz, A.; Proulx, S.; Xu, Z.; Siddartha, K.; Indrayanti, A.M.; Mahadevan, R. Systems biology based metabolic engineering for non-natural chemicals. Biotechnol. Adv. 2019, 37, 107379. [Google Scholar] [CrossRef]
  8. Yim, H.; Haselbeck, R.; Niu, W.; Pujol-Baxley, C.; Burgard, A.; Boldt, J.; Khandurina, J.; Trawick, J.D.; Osterhout, R.E.; Stephen, R.; et al. Metabolic engineering of Escherichia coli for direct production of 1,4-butanediol. Nat. Chem. Biol. 2011, 7, 445–452. [Google Scholar] [CrossRef]
  9. Liu, H.; Lu, T. Autonomous production of 1,4-butanediol via a de novo biosynthesis pathway in engineered Escherichia coli. Metab. Eng. 2015, 29, 135–141. [Google Scholar] [CrossRef]
  10. Wang, J.; Jain, R.; Shen, X.; Sun, X.; Cheng, M.; Liao, J.C.; Yuan, Q.; Yan, Y. Rational engineering of diol dehydratase enables 1,4-butanediol biosynthesis from xylose. Metab. Eng. 2017, 40, 148–156. [Google Scholar] [CrossRef] [Green Version]
  11. Tai, C.; Shen, Y.; Guo, Y.; Tao, F. Biosynthesis of 1,4-butanediol from erythritol using whole-cell catalysis. Biocatal. Biotrans. 2019, 37, 92–96. [Google Scholar]
  12. Ly, B.K.; Minh, D.P.; Pinel, C.; Besson, M.; Tapin, B.; Epron, F.; Especel, C. Effect of addition mode of Re in bimetallic Pd-Re/TiO2 catalysts upon the selective aqueous-phase hydrogenation of succinic acid to 1,4-butanediol. Top. Catal. 2012, 55, 466–473. [Google Scholar] [CrossRef]
  13. Kataoka, N.; Vangnai, A.S.; Tajima, T.; Nakashimada, Y.; Kato, J. Improvement of (R)-1,3-butanediol production by engineered Escherichia coli. J. Biosci. Bioeng. 2013, 115, 475–480. [Google Scholar] [CrossRef] [PubMed]
  14. Nemr, K.; Müller, J.E.N.; Joo, J.C.; Gawand, P.; Choudhary, R.; Mendonca, B.; Lu, S.; Yu, X.; Yakunin, A.F.; Mahadevan, R. Engineering a short, aldolase-based pathway for (R)-1,3-butanediol production in Escherichia coli. Metab. Eng. 2018, 48, 13–24. [Google Scholar] [CrossRef]
  15. Matsuyama, A.; Yamamoto, H.; Kawada, N.; Kobayashi, Y. Industrial production of (R)-1,3-butanediol by new biocatalysts. J. Mol. Catal. B Enzym. 2001, 11, 513–521. [Google Scholar] [CrossRef]
  16. Itoh, N.; Nakamura, M.; Inoue, K.; Makino, Y. Continuous production of chiral 1,3-butanediol using immobilized biocatalysts in a packed bed reactor: Promising biocatalysis method with an asymmetric hydrogen-transfer bioreduction. Appl. Microbiol. Biotechnol. 2007, 75, 1249–1256. [Google Scholar] [CrossRef]
  17. Sato, S.; Sato, F.; Gotoh, H.; Yamada, Y. Selective dehydration of alkanediols into unsaturated alcohols over rare earth oxide catalysts. ACS Catal. 2013, 3, 721–734. [Google Scholar] [CrossRef]
  18. 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] [Green Version]
  19. Zeng, F.; Hohn, K. Catalytic conversion of biomass-derived compounds to C4 chemicals. Catalysis 2019, 31, 1–36. [Google Scholar]
  20. Sun, D.; Li, Y.; Yang, C.; Su, Y.; Yamada, Y.; Sato, S. Production of 1,3-butadiene from biomass-derived C4 alcohols. Fuel Proc. Technol. 2020, 197, 106193. [Google Scholar] [CrossRef]
  21. 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 sodiumphosphates. Appl. Catal. A Gen. 2016, 511, 156–167. [Google Scholar] [CrossRef]
  22. 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] [Green Version]
  23. 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]
  24. 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]
  25. 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]
  26. Lee, J.H.; Hong, S.B. Dehydration of 1,3-butanediol to butadiene over medium-pore zeolites: Another example of reaction intermediate shape selectivity. Appl. Catal. B Environ. 2021, 280, 119446. [Google Scholar] [CrossRef]
  27. Yamamoto, N.; Sato, S.; Takahashi, R.; Inui, K. Synthesis of homoallyl alcohol from 1,4-butanediol over ZrO2 catalyst. Catal. Commun. 2005, 6, 480–484. [Google Scholar] [CrossRef]
  28. Yamamoto, N.; Sato, F.; Takahashi, R.; Inui, K. Synthesis of 3-buten-1-ol from 1,4-butanediol over ZrO2 catalyst. J. Mol. Catal. A Chem. 2006, 243, 52–59. [Google Scholar] [CrossRef]
  29. Sato, F.; Yamada, Y.; Sato, S. Preparation of Er2O3 nanorod catalyst without using organic additive and its application to catalytic dehydration of 1,4-butanediol. Chem. Lett. 2012, 41, 593–594. [Google Scholar] [CrossRef] [Green Version]
  30. Sato, F.; Sato, S.; Yamada, Y.; Nakamura, M.; Shiga, A. Acid–base concerted mechanism in the dehydration of 1,4-butanediol over bixbyite rare earth oxide catalysts. Catal. Today 2014, 226, 124–133. [Google Scholar] [CrossRef]
  31. Zhang, Q.; Zhang, Y.; Li, H.; Gao, C.; Zhao, Y. Heterogeneous CaO-ZrO2 acid–base bifunctional catalysts for vapor-phase selective dehydration of 1,4-butanediol to 3-buten-1-ol. Appl. Catal. A Gen. 2013, 466, 233–239. [Google Scholar] [CrossRef]
  32. Inoue, H.; Sato, S.; Takahashi, R.; Izawa, Y.; Ohno, H.; Takahashi, K. Dehydration of 1,4-butanediol over supported rare earth oxide catalysts. Appl. Catal. A Gen. 2009, 352, 66–73. [Google Scholar] [CrossRef]
  33. Duan, H.; Hirota, T.; 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]
  34. Ohtsuka, S.; Nemoto, T.; Yotsumoto, R.; Yamada, Y.; Sato, F.; Takahashi, R.; Sato, S. Vapor-phase catalytic dehydration of butanediols to unsaturated alcohols over yttria-stabilized zirconia catalysts. Appl. Catal. A Gen. 2019, 575, 48–57. [Google Scholar] [CrossRef]
  35. Nemoto, T.; Yamada, Y.; Sato, F.; Takahashi, R.; Sato, S. Catalytic dehydration of 1,3-butanediol over oxygen-defected fluorite Yb2Zr2O7. Mol. Catal. 2019, 473, 110399. [Google Scholar] [CrossRef]
  36. Fang, X.; Xia, L.; Peng, L.; Luo, Y.; Xu, J.; Xu, L.; Xu, X.; Liu, W.; Zheng, R.; Wang, X. Ln2Zr2O7 compounds (Ln = La, Pr, Sm, Y) with varied rare earth A sites for low temperature oxidative coupling of methane. Chin. Chem. Lett. 2019, 30, 1141–1146. [Google Scholar] [CrossRef]
  37. Shanon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Cryst. 1976, A32, 751–767. [Google Scholar] [CrossRef]
  38. Duan, H.; Sun, D.; Yamada, Y.; Sato, S. Dehydration of 2,3-butanediol into 3-buten-2-ol catalyzed by ZrO2. Catal. Commun. 2014, 48, 1–4. [Google Scholar] [CrossRef]
  39. Duan, H.; Yamada, Y.; Sato, S. Vapor-phase catalytic dehydration of 2,3-butanediol into 3-buten-2-ol over Sc2O3. Chem. Lett. 2014, 43, 1773–1775. [Google Scholar] [CrossRef] [Green Version]
  40. 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] [Green Version]
  41. 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]
  42. Chen, D.; Xu, R. Hydrothermal synthesis and characterization of La2M2O7 (M = Ti, Zr) powders. Mater. Res. Bull. 1998, 33, 409–417. [Google Scholar] [CrossRef]
  43. Liu, Z.-G.; Ouyang, J.-H.; Zhou, Y.; Li, J.; Xia, X.-L. Densification, structure, and thermophysical properties of ytterbium-gadolinium zirconate ceramics. Int. J. Appl. Ceram. Technol. 2009, 6, 485–491. [Google Scholar] [CrossRef]
  44. Walker, J.D.S.; Hayes, J.R.; Gaultois, M.W.; Aluri, E.R.; Grosvenor, A.P. A case for oxygen deficiency in Gd2Ti2−xZrxO7 pyrochlore-type oxides. J. Alloys Compd. 2013, 565, 44–49. [Google Scholar] [CrossRef]
  45. Gotoh, H.; Yamada, Y.; Sato, S. Dehydration of 1,3-butanediol over rare earth oxides. Appl. Catal. A Gen. 2010, 377, 92–98. [Google Scholar] [CrossRef]
  46. Sato, S.; Takahashi, R.; Kobune, M.; Inoue, H.; Izawa, Y.; Ohno, H.; Takahashi, K. Dehydration of 1,4-butanediol over rare earth oxides. Appl. Catal. A Gen. 2009, 356, 64–71. [Google Scholar] [CrossRef]
  47. Kobune, M.; Sato, S.; Takahashi, R. Surface-structure sensitivity of CeO2 for several catalytic reactions. J. Mol. Catal. A Chem. 2008, 279, 10–19. [Google Scholar] [CrossRef]
  48. Sharma, S.; Hilaire, S.; Vohs, J.M.; Gorte, R.J.; Jen, H.-W. Evidence for oxidation of ceria by CO2. J. Catal. 2000, 190, 199–204. [Google Scholar] [CrossRef]
  49. Moulijn, J.A.; Tarfaoui, A.; Kapteijn, F. General aspects of catalyst testing. Catal. Today 1991, 11, 1–12. [Google Scholar] [CrossRef]
  50. Sato, S.; Takahashi, R.; Sodesawa, T.; Yamamoto, N. Dehydration of 1,4-butanediol into 3-buten-1-ol catalyzed by ceria. Catal. Commun. 2004, 5, 397–400. [Google Scholar] [CrossRef]
  51. Ichikawa, N.; Sato, S.; Takahashi, R.; Sodesawa, T. Catalytic reaction of 1,3-butanediol over solid acids. J. Mol. Catal. A Chem. 2006, 256, 106–112. [Google Scholar] [CrossRef]
  52. Duan, H.; Yamada, Y.; Sato, S. Selective dehydration of 2,3-butanediol to 3-buten-2-ol over ZrO2 modified with CaO. Appl. Catal. A Gen. 2014, 487, 226–233. [Google Scholar] [CrossRef] [Green Version]
  53. Izumi, F.; Momma, K. Three-dimensional visualization in powder diffraction. Solid State Phenom. 2007, 130, 15–20. [Google Scholar] [CrossRef]
  54. Lide, D.R. CRC Handbook of Chemistry and Physics, 82th ed.; CRC Press: London, UK, 2001; pp. 4-95–4-96. [Google Scholar]
  55. Takahashi, R.; Yamada, I.; Iwata, A.; Kurahashi, N.; Yoshida, S.; Sato, S. Synthesis of 3-buten-1-ol from 1,4-butanediol over indium oxide. Appl. Catal. A Gen. 2010, 383, 134–140. [Google Scholar] [CrossRef]
  56. Wattanasiriwech, D.; Wattanasiriwech, S. Effects of fuel contents and surface modification on the sol-gel combustion Ce0.9 Gd0.1O1.95 nanopowder. Energy Procedia 2013, 34, 524–533. [Google Scholar] [CrossRef] [Green Version]
  57. Yavetskiy, R.P.; Kosyanov, D.Y.; Baumer, V.N.; Doroshenko, A.G.; Fedorov, A.I.; Matveevskaya, N.A.; Tolmachev, A.V.; Vovk, O.M. Low-agglomerated yttria nanopowders via decomposition of sulfate-doped precursor with transient morphology. J. Rare Earth 2014, 32, 320–325. [Google Scholar] [CrossRef]
Figure 1. Dehydration of butanediols.
Figure 1. Dehydration of butanediols.
Catalysts 10 01392 g001
Figure 2. XRD patterns of REZrO samples calcined at 900 °C.
Figure 2. XRD patterns of REZrO samples calcined at 900 °C.
Catalysts 10 01392 g002
Figure 3. Continuous scanning XRD patterns of YZrO.
Figure 3. Continuous scanning XRD patterns of YZrO.
Catalysts 10 01392 g003
Figure 4. FE-SEM image of YZrO calcined at 900 °C. A scale bar in the photo indicates 100 nm.
Figure 4. FE-SEM image of YZrO calcined at 900 °C. A scale bar in the photo indicates 100 nm.
Catalysts 10 01392 g004
Figure 5. Step-scanning XRD pattern of YZrO calcined at 1200 °C with calculated XRD pattern.
Figure 5. Step-scanning XRD pattern of YZrO calcined at 1200 °C with calculated XRD pattern.
Catalysts 10 01392 g005
Figure 6. Effect of calcination temperature of YZrO catalyst on the dehydration of 1,4-BDO and SA.
Figure 6. Effect of calcination temperature of YZrO catalyst on the dehydration of 1,4-BDO and SA.
Catalysts 10 01392 g006
Figure 7. Change in the conversion of 1,4-BDO with space time over YZrO calcined at 900 °C.
Figure 7. Change in the conversion of 1,4-BDO with space time over YZrO calcined at 900 °C.
Catalysts 10 01392 g007
Figure 8. Long-run stability in the conversion of 1,4-BDO over YZrO calcined at 900 °C. Reaction temperature, 325 °C; carrier gas flow rate, 30 cm3 min−1; W/F = 0.31 h.
Figure 8. Long-run stability in the conversion of 1,4-BDO over YZrO calcined at 900 °C. Reaction temperature, 325 °C; carrier gas flow rate, 30 cm3 min−1; W/F = 0.31 h.
Catalysts 10 01392 g008
Figure 9. TPD profiles of adsorbed NH3 and CO2 on YZrO calcined at 600 (A) and 900 °C (B).
Figure 9. TPD profiles of adsorbed NH3 and CO2 on YZrO calcined at 600 (A) and 900 °C (B).
Catalysts 10 01392 g009
Figure 10. Changes in the formation rate of unsaturated alcohols (UOLs) in the dehydration of BDOs over REZrO catalysts at 325 °C with ionic radius of rare earth cation (A) and average ionic radius between Zr4+ and rare earth cation, R(RE+Zr)/2 (B). A dotted line in Figure 10B indicates the formation rate in the dehydration of 1,3-BDO over pure rare earth oxide with ionic radius of rare earth cation, reported in Ref. [45].
Figure 10. Changes in the formation rate of unsaturated alcohols (UOLs) in the dehydration of BDOs over REZrO catalysts at 325 °C with ionic radius of rare earth cation (A) and average ionic radius between Zr4+ and rare earth cation, R(RE+Zr)/2 (B). A dotted line in Figure 10B indicates the formation rate in the dehydration of 1,3-BDO over pure rare earth oxide with ionic radius of rare earth cation, reported in Ref. [45].
Catalysts 10 01392 g010
Figure 11. Possible coordination structures with adsorption site (A) and reaction mechanism for the dehydration of 1,4-BDO to form 3B1ol (B) and THF (C).
Figure 11. Possible coordination structures with adsorption site (A) and reaction mechanism for the dehydration of 1,4-BDO to form 3B1ol (B) and THF (C).
Catalysts 10 01392 g011
Table 1. Properties of REZrO catalysts calcined at 900 °C.
Table 1. Properties of REZrO catalysts calcined at 900 °C.
CatalystIonic Radius of Rare Earth Cation a (nm)Average Ionic
Radius, R b (nm)
SAc (m2 g−1)Molar Ratio of RE/Zr/Hf dRatio of RE/Zr d
LaZrO0.10320.09361643.3/56.1/0.60.77
PrZrO0.09900.09152945.1/54.0/0.90.84
NdZrO0.09830.091152552.8/46.6/0.61.13
CeZrO0.09700.09053047.1/52.3.0.60.90
SmZrO0.09580.08993252.9/46.7/0.41.13
EuZrO0.09470.089353551.0/48.4/0.61.05
GdZrO0.09380.08893552.2/47.1/0.71.11
TbZrO0.09230.088153451.8/47.6/0.61.09
DyZrO0.09120.08763253.2/46.2/0.61.15
HoZrO0.09010.087053053.5/45.9/0.61.17
YZrO0.09000.08703651.2/48.0/0.81.07
ErZrO0.08900.08653054.1/45.5/0.41.19
TmZrO0.08800.08603953.4/46.2/0.41.16
YbZrO0.08680.08542755.0/44.6/0.41.23
LuZrO0.08610.085052652.8/47.0/0.21.12
ScZrO0.07450.079252327.2/71.8/1.00.38
a Ionic radius of trivalent rare earth cation with 6 coordination except that of Ce4+ with 8 coordination. The data are cited from Ref. [37]. b Average ionic radius between Zr4+ and rare earth cation. c Specific surface area. d Estimated by XRF. An average value in five measurements.
Table 2. Dehydration of 1,4-BDO over REZrO catalysts calcined at 900 °C.
Table 2. Dehydration of 1,4-BDO over REZrO catalysts calcined at 900 °C.
CatalystConversionSelectivity (mol%)
(%)3B1ol2B1olUOLsBDTHFGBLOthers
LaZrO29.16.10.56.60.088.43.11.9
PrZrO25.261.71.563.20.032.72.51.6
NdZrO37.732.40.833.20.063.70.52.6
CeZrO24.760.85.666.40.024.95.63.1
SmZrO37.365.81.367.10.130.80.91.1
EuZrO52.882.62.885.40.810.82.20.8
GdZrO68.581.51.983.40.515.30.40.4
TbZrO52.476.11.177.20.621.00.90.3
DyZrO69.486.52.789.21.67.90.40.9
HoZrO51.282.61.784.30.813.30.70.9
YZrO77.687.72.890.51.27.50.60.2
ErZrO64.485.84.089.81.97.40.30.6
TmZrO55.982.81.384.11.113.30.31.2
YbZrO61.384.82.287.00.811.50.30.4
LuZrO51.284.32.086.32.69.00.31.8
ScZrO35.530.00.130.10.068.90.30.7
Conversion and selectivity are averaged at time on stream (TOS) between 1–5 h. Reaction conditions: temperature, 325 °C; W/F, 0.31 h; catalyst weight, 0.50 g; N2 carrier gas flow rate, 30 cm3 min−1. 3B1ol, 3-buten-1-ol; 2B1ol, 2-buten-1-ol; UOLs = 3B1ol + 2B1ol; BD, 1,3-butadiene; THF, tetrahydrofuran; GBL, γ-butyrolactone. Others include ethanol, 1-butanol, and some unidentified products.
Table 3. Comparison in the dehydration of 1,3-, 1,4- and 2,3-BDO over REZrO calcined at 900 °C.
Table 3. Comparison in the dehydration of 1,3-, 1,4- and 2,3-BDO over REZrO calcined at 900 °C.
CatalystReactantConv.Selectivity (mol%)
(%)3B2ol3B1ol2B1olUOLsBDMEKTHF3H2BOOthers
LaZrO1,3-BDO40.244.9 1.235.181.21.7 6.8--10.3 a
1,4-BDO29.1-6.1 0.56.60-88.4-5.0 b
2,3-BDO4.828.3--28.3016.1-37.118.5 c
CeZrO1,3-BDO64.755.8 1.035.191.92.8 0.7--4.6 a
1,4-BDO24.7-60.8 5.666.40-24.9-8.7 b
2,3-BDO6.615.4--15.4029.4-33.421.8 c
YZrO1,3-BDO93.952.1 1.336.890.27.2 0.2--2.4 a
1,4-BDO77.6-87.72.890.51.2-7.5-0.8 b
2,3-BDO17.030.3--30.3010.9-26.832.0 c
YbZrO1,3-BDO89.853.9 1.440.996.21.0 1.0--1.8 a
1,4-BDO61.3-84.8 2.287.00.8-11.5-0.7 b
2,3-BDO12.246.9--46.90 7.3-28.417.4 c
LuZrO1,3-BDO83.050.8 1.742.995.42.0 0.9--1.7 a
1,4-BDO51.2-84.3 2.086.32.6-9.0-2.1 b
2,3-BDO11.350.8--50.8010.1-24.214.9 c
ScZrO1,3-BDO20.643.112.739.795.50 1.4--3.1 a
1,4-BDO35.5-30.00.130.10-68.9-1.0 b
2,3-BDO18.447.4--47.4 018.7-16.017.9 c
Reaction conditions are the same as those in Table 2. 3B2ol, 3-buten-2-ol; 3B1ol, 3-buten-1-ol; 2B1ol, 2-buten-1-ol; UOLs = 3B2ol+3B1ol+2B1ol; BD, 1,3-butadiene; MEK, butanone; THF, tetrahydrofuran; 3H2BO, 3-hydroxy-2-butanone. a Others include ethanol, acetone, 1-butanol, 3-buten-2-one, etc. b Others include γ-butyrolactone, 1-butanol, ethanol, etc. c Others include 2-methylpropanal, 2-methyl-1-propanol, 2-butanol, ethanol, acetone, etc.
Table 4. Physical properties of YZrO calcined at different temperatures.
Table 4. Physical properties of YZrO calcined at different temperatures.
CalcinationSAFWHM at 2θ = 29.9°Particle Size,Crystallite Size,
(°C)(m2 g−1)(degree)DBET (nm)DXRD (nm)
600911.792124.6
800511.439225.7
900361.242316.6
1000190.6705812
1050160.5437015
Table 5. Refined structure parameters of Y2Zr2O7 and YZrO calcined at 1200 °C.
Table 5. Refined structure parameters of Y2Zr2O7 and YZrO calcined at 1200 °C.
SampleY2Zr2O7
Space GroupFm3ma
a (Å)5.19878B (Y)1.81615
α (degree)90.0000B (Zr)1.77582
β (degree)90.0000B (O)3.46843
γ (degree)90.0000Rwp (%)12.229%
g (Y) b0.5000RB (%)2.039%
g (Zr) b0.5000RF (%)1.593%
g (O) b0.8750S1.3891
g: occupancy; B: isotropic atomic displacement parameter. a Atomic sites: (Y, Zr) 4a (0, 0, 0) and O 8c (1/4, 1/4, 1/4). b Fixed during the refinement.
Table 6. Dehydration of 1,4-BDO over YZrO at different reaction temperatures.
Table 6. Dehydration of 1,4-BDO over YZrO at different reaction temperatures.
TemperatureConv.Selectivity (mol%)
(°C)(%)3B1ol2B1olUOLsBDPropyleneTHFGBLOthers
30025.984.01.185.10.40.012.40.61.5
32577.687.72.890.51.20.07.50.60.2
35099.870.27.177.311.50.09.90.21.1
360100.053.010.163.122.20.011.50.13.1
375100.023.84.428.258.85.83.70.13.4 a
The catalyst sample was calcined at 900 °C. Conversion and selectivity were averaged at TOS between 1–5 h. Reaction conditions: W/F, 0.31 h; catalyst weight, 0.50 g; N2 flow rate, 30 cm3 min−1. 3B1ol, 3-buten-1-ol; 2B1ol, 2-buten-1-ol; UOLs = 3B1ol + 2B1ol; BD, 1,3-butadiene; THF, tetrahydrofuran; GBL, γ-butyrolactone. Others include ethanol, 1-butanol, CO2, and some unidentified products. a includes 1.8 mol% of CO2.
Table 7. Effects of carrier gas on the dehydration of 1,4-BDO over YZrO calcined at 900 °C.
Table 7. Effects of carrier gas on the dehydration of 1,4-BDO over YZrO calcined at 900 °C.
Carrier GasConversionSelectivity (mol%)Productivity of 3B1ol
(%)3B1ol2B1olUOLsBDTHFGBLOthersmol h−1 kg−1
T = 325 °C
N277.687.72.890.51.27.50.60.224.2
CO274.664.61.566.10.332.30.80.517.1
NH381.985.56.592.00.06.80.11.124.9
H287.886.73.390.01.08.40.40.227.0
T = 300 °C
N225.984.01.185.10.412.40.61.57.7
CO219.260.81.562.30.034.01.62.14.1
NH332.387.61.589.10.09.90.20.710.1
N2 (NH3) a16.076.610.386.90.07.81.53.84.4
Conversion and selectivity are averaged at TOS between 1–5 h. Reaction conditions: T, reaction temperature; W/F, 0.31 h; carrier gas flow rate, 30 cm3 min−1. 3B1ol, 3-buten-1-ol; 2B1ol, 2-buten-1-ol; UOLs = 3B1ol + 2B1ol; BD, 1,3-butadiene; THF, tetrahydrofuran; GBL, γ-butyrolactone. Others include ethanol, 1-butanol, and some unidentified products. a Prior to the preheating in N2 at 300 °C for 1 h, NH3 was flowed at a rate of 30 cm3 min−1 and 300 °C for 0.5 h.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Matsuda, A.; Matsumura, Y.; Nakazono, K.; Sato, F.; Takahashi, R.; Yamada, Y.; Sato, S. Dehydration of Biomass-Derived Butanediols over Rare Earth Zirconate Catalysts. Catalysts 2020, 10, 1392. https://doi.org/10.3390/catal10121392

AMA Style

Matsuda A, Matsumura Y, Nakazono K, Sato F, Takahashi R, Yamada Y, Sato S. Dehydration of Biomass-Derived Butanediols over Rare Earth Zirconate Catalysts. Catalysts. 2020; 10(12):1392. https://doi.org/10.3390/catal10121392

Chicago/Turabian Style

Matsuda, Asami, Yoshitaka Matsumura, Kazuki Nakazono, Fumiya Sato, Ryoji Takahashi, Yasuhiro Yamada, and Satoshi Sato. 2020. "Dehydration of Biomass-Derived Butanediols over Rare Earth Zirconate Catalysts" Catalysts 10, no. 12: 1392. https://doi.org/10.3390/catal10121392

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