Next Article in Journal
Solvent-Free Microwave-Induced Oxidation of Alcohols Catalyzed by Ferrite Magnetic Nanoparticles
Next Article in Special Issue
Facile, One-Pot, Two-Step, Strategy for the Production of Potential Bio-Diesel Candidates from Fructose
Previous Article in Journal
Electroreduction of CO2 into Ethanol over an Active Catalyst: Copper Supported on Titania
Previous Article in Special Issue
Insights into the Metal Salt Catalyzed 5-Ethoxymethylfurfural Synthesis from Carbohydrates
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Conversion of Cellulose to Lactic Acid by Using ZrO2–Al2O3 Catalysts

1
Research Institute for Chemical Process Technology, National Institute of Advanced Industrial Science and Technology (AIST), 4-2-1 Nigatake, Miyagino, Sendai 983-8551, Japan
2
Department of Chemical Technology, Faculty of Science, Chulalongkorn University, Pathumwan, Bangkok 10330, Thailand
3
Center of Excellence on Petrochemical and Materials Technology, Chulalongkorn University Research Building, Bangkok 10330, Thailand
*
Author to whom correspondence should be addressed.
Catalysts 2017, 7(7), 221; https://doi.org/10.3390/catal7070221
Submission received: 26 June 2017 / Revised: 19 July 2017 / Accepted: 19 July 2017 / Published: 21 July 2017
(This article belongs to the Special Issue Catalysis of Biomass-Derived Molecules)

Abstract

:
Lactic acid has a wide range of applications in many industries, both as an ingredient and as an intermediate. Here, we investigated the catalytic conversion of cellulose to lactic acid by using heterogeneous mixed-oxide catalysts containing ZrO2. Although pure ZrO2 has catalytic activity for the conversion of cellulose to lactic acid, the yield of lactic acid obtained is not satisfactory. In contrast, a series of ZrO2–Al2O3 catalysts containing various percentages of ZrO2 provided higher yields of lactic acid. The ZrO2–Al2O3 catalysts had more Lewis acid sites and far fewer base sites than ZrO2. This suggests that the Lewis acid sites on ZrO2–Al2O3 catalysts are more important than the base sites for the conversion of cellulose to lactic acid.

1. Introduction

Biomass is a renewable organic resource that contains components that are essential for the production of many important chemicals and fuels [1,2,3]. Inedible lignocellulosic biomass is plant biomass that is mainly composed of cellulose, hemicellulose, and lignin. Currently, cellulose is the most promising component for biomass utilization because cellulose is the major component of lignocellulosic biomass [4,5,6,7].
Lactic acid is an organic compound that is used in large quantities as the food, cosmetic, pharmaceutical, and chemical productions [8,9]. Moreover, lactic acid is an important intermediate for conversion to other products such as propylene glycol, acrylic acid, and polylactic acid [8,10]. Conventionally, lactic acid is produced via the fermentation of sugars from starch [11]; however, the rate of production using this method is low, and the pH of the solution during fermentation must be kept in the range 5–7 to provide an environment suitable for lactic acid-fermenting bacteria. Thus, catalytic conversion of inedible cellulose is an attractive alternative to fermentation for the production of lactic acid.
Recently, several homogeneous catalysts (e.g., PbCl2 and ErCl3) have been investigated for their suitability for the conversion of cellulose to lactic acid [12,13]. These catalysts provide a high yield of lactic acid; however, their use in industries remains unrealized because of difficult separation of the products from the catalyst as well as the poor stability and recyclability of the catalysts. Therefore, heterogeneous catalysts for the conversion of cellulose to lactic acid are desired because products can be easily separated from this type of catalyst. Several heterogeneous catalysts have been reported to have catalytic activity for the conversion of cellulose to lactic acid, such as LaCoO3, NbF5–AlF3, and AlW, which provided yields of lactic acid of 24%, 27%, and 28%, respectively [14,15,16]. However, these solid catalysts have a disadvantage that the metal species leach into solution during the reaction—2.4% of the Co and 1.5% of the La from LaCoO3 [14] and 1.5% of the W from AlW [16] leached into solution. Additionally, these heterogeneous catalysts are relatively expensive, making their industrial use difficult. Previously, we reported a simple ZrO2 catalyst for the conversion of cellulose to lactic acid (yield, 21.2%) that did not leach metal species into the reaction solution [17]. In addition, the ZrO2 catalyst was stable in hot water. However, the yield of lactic acid obtained with this catalyst was less than those reported for the heterogeneous catalysts. Thus, the aim of the present study was to examine how to increase the yield of lactic acid obtainable with ZrO2-based catalysts. Dispersion of active species on supports such as metal oxides is a common means of increasing the catalytic activity of heterogeneous catalysts. Therefore, here we compared the catalytic activity of a series of mixed-oxide catalysts containing various amounts of ZrO2 (5%, 10%, or 20%) on an Al2O3 support with that of pure ZrO2. We found that the 10%ZrO2–Al2O3 catalyst showed activity for the production of lactic acid from cellulose with a product yield of 25.3%, which was cheaper than the reported heterogeneous catalysts.

2. Results and Discussion

2.1. Effect of Ball-Milling Treatment on Cellulose Conversion

In this paper, pulverization of cellulose by ball-milling has been applied for the increase of cellulose conversion [18,19,20]. The conversion of cellulose was only 25% in the case of un-milled cellulose (473 K, 6 h, without catalysts) and increased to 86% by the ball-milling treatment of cellulose. Figure 1 shows XRD patterns of un-milled and milled cellulose. The cellulose crystallinity was decreased by the ball-milling treatment as revealed by the peak broadening at 2θ = 22.6 degree (crystalline plane 002) in XRD patterns (Figure 1) [19,21], indicating that the cellulose conversion increased with decreasing cellulose crystallinity. In the reaction pathway, cellulose was firstly hydrolyzed to soluble glucose; thus, cellulose with high crystallinity was hydrolyzed very slowly at 473 K from the result of the low conversion of un-milled cellulose.

2.2. Catalytic Conversion of Cellulose to Lactic Acid by Using ZrO2–Al2O3

Previously, we investigated the catalytic conversion of cellulose to lactic acid by using various transition metal oxides (i.e., ZrO2, Al2O3, TiO2, Fe3O4, V2O5, CeO2, Y2O3, Tm2O3, HfO2, Ga2O3, MgO, La2O3, Nb2O5, and Ta2O5) [17]. Of these catalysts, ZrO2 had the highest catalytic activity (lactic acid yield, 21.2%; reaction temperature, 473 K). Therefore, in the present study, we examined the use of an Al2O3 support to increase the catalytic activity of ZrO2 for the conversion of cellulose to lactic acid. The yields of lactic acid obtained with ZrO2–Al2O3 mixed-oxide catalysts containing different percentages of ZrO2 at 473 K were higher than those obtained by using the individual metal oxides, indicating that the dispersion of active ZrO2 species on the Al2O3 support increased the catalytic activity of ZrO2 (Figure 2). The maximum yield of lactic acid (25.3%) was obtained when 10%ZrO2–Al2O3 was used as the catalyst. The yield of lactic acid increased slightly as the amount of ZrO2 on the Al2O3 support was increased from 5% to 10%; however, the yield decreased as the amount of ZrO2 was increased to 20%, indicating that the yield of lactic acid may decrease further as the percentage of ZrO2 on the support increases.

2.3. Reaction Conditions

Next, we optimized the reaction conditions to obtain the highest possible yields of lactic acid with 10%ZrO2–Al2O3 used as the catalyst. Figure 3 shows the product yields obtained as a function of reaction time. The yield of lactic acid increased for up to 6 h, after which it decreased due to decomposition or polymerization of lactic acid.
Figure 4 shows the product yields obtained as a function of reaction temperature. The maximum yield of lactic acid (25.3%) was obtained at 473 K. At the lowest temperature examined (i.e., 453 K), the cellulose conversion reaction proceeded very slowly. At the highest temperature examined (i.e., 493 K), the yield of lactic acid was less than that at the optimum temperature because the lactic acid was likely decomposed at this temperature.
Finally, we investigated the reusability of the 10%ZrO2–Al2O3 catalyst for repeated conversion of cellulose to lactic acid. After the reaction with the optimized conditions (473 K, 6 h), the recovered solid, which contained the used 10%ZrO2–Al2O3 catalyst and any unreacted cellulose, was heated at 673 K for 15 h under an air atmosphere to remove carbon-based material accumulated on the surface of the catalyst. The yield of lactic acid obtained by using the reused 10%ZrO2–Al2O3 catalyst at 473 K for 6 h was 31.7%, which was higher than the yield obtained from the first reaction. One possible reason is that the calcination of the 10%ZrO2–Al2O3 catalyst after the reaction would eliminate contamination on the surface, resulting in the higher lactic acid yield. This result suggests that the 10%ZrO2–Al2O3 catalyst is recyclable.

2.4. Characterization

Next, we characterized the ZrO2–Al2O3 catalysts to further understand their catalytic activity. The X-ray diffraction (XRD) patterns of the catalysts are shown in Figure 5. In the XRD patterns of the three ZrO2–Al2O3 catalysts, peaks at 46.0° and 66.5° were attributed to γ-Al2O3 [22]. In the XRD patterns of 10%ZrO2–Al2O3 and 20%ZrO2–Al2O3, peaks at 30.3° and 50.7° were attributed to tetragonal ZrO2 [23]; however, these peaks were not seen in the XRD pattern of 5%ZrO2–Al2O3, indicating that ZrO2 was either highly dispersed or in the amorphous phase in this catalyst. The XRD pattern for pure ZrO2 revealed that it was in the monoclinic phase (Figure 5d).
In our previous paper, we hypothesized that the properties of the acid and base sites on ZrO2 play an important role in the catalytic activity of this oxide for the conversion of cellulose to lactic acid [17]. Therefore, we examined the acid and base properties of the ZrO2–Al2O3 catalysts by means of temperature-programmed desorption of ammonia (NH3-TPD) and temperature-programmed desorption of carbon dioxide (CO2-TPD), respectively. Figure 6 shows NH3-TPD profiles of the ZrO2–Al2O3 and ZrO2 catalysts. The NH3 desorption temperatures of the three ZrO2–Al2O3 catalysts were all 470 K, which was similar to that of ZrO2, indicating that the acid sites on the ZrO2–Al2O3 and ZrO2 catalysts were weakly acidic Lewis acid sites [24] (Figure 6). The shoulder peaks at 550 K were observed in the case of ZrO2–Al2O3 catalysts, which were ascribed to slightly stronger acid sites than weakly acid sites (peak at 470 K). The slightly stronger acid sites might be located on the boundary between ZrO2 and Al2O3. Although the XRD analysis revealed that the crystal phase of ZrO2 differed depending on whether it was in pure or mixed-oxide form, the NH3 desorption temperatures revealed that the acid strength was almost the same irrespective of form.
The desorption temperatures of CO2 from the ZrO2–Al2O3 and ZrO2 catalysts were in the range 340–450 K (Figure 7), indicating that the base sites on the ZrO2 catalysts were weakly basic [25]. The amount of CO2 desorbed from the ZrO2–Al2O3 catalysts was smaller than that desorbed from the ZrO2 catalyst.
Table 1 shows the amounts of acid and base sites on the ZrO2–Al2O3 and ZrO2 catalysts. The ZrO2–Al2O3 catalysts had more acid sites but far fewer base sites than the ZrO2 catalyst. In our previous paper, we suggested that the retro-aldol reaction (the conversion of fructose to glyceraldehyde and dihydroxyacetone)—the key step in the conversion of cellulose to lactic acid—involved both the acid and base sites on the ZrO2 catalyst [17]. However, the ZrO2–Al2O3 catalysts had more acid sites and far fewer base sites than did ZrO2. We discuss the reaction pathway in the next section.

2.5. Reaction Pathway

The conversion of cellulose to lactic acid is a multi-step reaction (Scheme 1). First, cellulose is converted to glucose by hydrolysis, which can be catalyzed by acid catalysts [26,27]. Then, glucose is isomerized to fructose by a Lewis acid or basic catalysts [28]. The key step in the conversion pathway from cellulose to lactic acid is the conversion of fructose to glyceraldehyde and dihydroxyacetone via a retro-aldol reaction involving C–C bond cleavage, which can be enhanced by Lewis acids [29,30]. In a previous paper, we hypothesized that the combination of acid and base sites on ZrO2 enhanced the conversion of fructose to glyceraldehyde and dihydroxyacetone [17]. However, in the present study, the yield of lactic acid obtained from cellulose by using the ZrO2–Al2O3 catalysts did not appear to depend on the relative amount of acid and base sites. That is, the ZrO2–Al2O3 catalysts provided a higher yield of lactic acid than did the ZrO2 catalyst, even though they had more acid sites and far fewer base sites than did ZrO2 (Table 1). This indicates that the number of base sites is not important for the catalytic conversion of cellulose to lactic acid with ZrO2–Al2O3 catalysts. Furthermore, although the Lewis acid sites on the ZrO2–Al2O3 catalysts played an important role in the conversion of cellulose to lactic acid, the yield of lactic acid obtained was not proportional to the number of acid sites. This suggests that dispersion of ZrO2 on a Al2O3 support and accessibility to the acid sites on the catalyst are the most important factors for the conversion of cellulose to lactic acid when using ZrO2–Al2O3 catalysts. Further studies are required to elucidate the details of the relationship between the active sites and the product yield and also the reaction mechanism.
In the final step of the reaction pathway, glyceraldehyde and dihydroxyacetone can be converted to lactic acid at 463 K without a catalyst [17]. However, the conversion of pyruvaldehyde to lactic acid is reported to require a catalyst with Lewis acid sites [31,32], which a ZrO2-based catalyst could provide.
Previously, we reported that we used a ZrO2 catalyst to convert cellulose to lactic acid with a yield of 21.2% [17]. However, this yield of lactic acid was less than that obtained with other reported heterogeneous catalysts (24% yield of lactic acid using LaCoO3 [14], 28% using AlW [16], and 27% using NbF5–AlF3 [15]). In the present study, we found that by using the 10%ZrO2–Al2O3 catalyst we could obtain a yield of lactic acid of 25.3%, which was comparable with the yields using the reported catalysts. One advantage of the 10%ZrO2–Al2O3 catalyst is that it can be obtained in lower cost than the reported heterogeneous catalysts. Additionally, a few percentages of the metal species in the reported catalysts leached into solution during the reaction. Conversely, the Zr species was not leached out from ZrO2 into water during the reaction [17].

3. Materials and Methods

3.1. Materials

ZrO2 (ZRO-7; reference catalyst from Catalysis Society of Japan) and mixed oxides containing ZrO2 and Al2O3 (5%ZrO2–Al2O3, 10%ZrO2–Al2O3, and 20%ZrO2–Al2O3) were obtained from Daiichi Kigenso Kagaku Kogyo Co., Ltd. (Osaka, Japan). Aluminum oxide (Al2O3) was obtained from Sigma–Aldrich Co., LLC (St. Louis, MO, USA). These materials were used without pretreatment.

3.2. Catalytic Reaction

Cellulose (microcrystalline cellulose; Merck KGaA, Darmstadt, Germany) was pulverized with a ball mill at 60 rpm. The conversion of cellulose to lactic acid was carried out in a stainless steel batch reactor with an inner volume of 100 cm3 (MMJ-100; OM Lab-Tech, Tochigi, Japan), as described in a previous paper [17]. Briefly, the reactor was loaded with ball-milled cellulose (0.5 g), catalyst (1.0 g), and water (50 g); purged with nitrogen gas (0.1 MPa); and then heated to 453–493 K for 3–9 h with screw stirring. After the reaction was allowed to run, the resulting mixture was filtered to separate the liquid from the solid. The quantitative analyses of lactic acid, levulinic acid, 5-hydroxymethylfurfural, and furfural in the liquid fractions were carried out with a gas chromatograph (GC-2014; Shimadzu, Kyoto, Japan) equipped with a flame ionization detector and an InertCap capillary column (GL Sciences Inc., Tokyo, Japan). The chemicals in the liquid fraction (e.g., glucose) were analyzed by using a high-performance liquid chromatograph (Shimadzu, Kyoto, Japan) equipped with a refractive index detector (RID-10A; Shimadzu, Kyoto, Japan), an ultraviolet/ visible detector (SPD-20AV; Shimadzu, Kyoto, Japan), and a Rezex RPM-Monosaccharide Pb+2 column (Phenomenex Inc., Torrance, CA, USA). The product yields were calculated based on moles of carbon as follows:
Product yield   ( % ) = ( Moles of carbon atoms in each product ) ( Moles of carbon atoms in cellulose used ) × 100
Conversion   ( % ) = ( 1 ( Weight of solid residue ) ( Weight of solid catalyst ) ( weight of cellulose used ) ) × 100

3.3. Catalyst Characterization

XRD patterns of the catalysts were determined by using a Rigaku SmartLab XRD system (Rigaku, Tokyo, Japan) with Cu Kα radiation in the 2θ range of 5–90°.
NH3-TPD profiles were determined with a TPD-1-AT instrument (Bel Japan, Inc., Osaka, Japan) with an online quadrupole mass spectrometer. Samples (ca. 0.05 g) were pretreated at 773 K in flowing helium for 1 h, saturated in flowing 5% ammonia diluted with helium (0.5 cm3 s−1) at 373 K for 30 min, and then treated at 373 K in flowing helium (0.83 cm3 s−1) for 1 h. The samples were then heated at a constant rate of 10 K min−1 from 373 to 953 K while the amount of NH3 desorbed was detected.
CO2-TPD profiles were determined with a Micromeritics 3FLEX 3500 chemisorption analyzer (Micromeritics, Norcross, GA, USA) with an online quadrupole mass spectrometer. Samples (ca. 0.2 g) were pretreated at 773 K in flowing helium for 1 h, saturated in flowing CO2 (0.5 cm3 s−1) at 323 K for 30 min, and then treated at 323 K in flowing helium (0.83 cm3 s−1) for 1 h. The samples were then heated at a constant rate of 10 K min−1 from 323 to 953 K while the amount of CO2 desorbed was detected.

4. Conclusions

Compared with using pure ZrO2 as the catalyst (yield, 21.2%), a greater yield of lactic acid was obtained from cellulose by using a 10%ZrO2–Al2O3 catalyst (yield, 25.3%) and optimized reaction conditions (reaction temperature, 473 K; reaction time, 6 h). XRD analysis revealed that the 10%ZrO2–Al2O3 catalyst contained tetragonal ZrO2 and γ-Al2O3. NH3-TPD and CO2-TPD analyses revealed that all three ZrO2–Al2O3 catalysts had more Lewis acid sites and far fewer base sites than did ZrO2. This suggests that the number of Lewis acid sites on the ZrO2–Al2O3 catalysts was more important than the number of base sites for the conversion of cellulose to lactic acid.

Acknowledgments

We acknowledge Daiichi Kigenso Kagaku Kogyo Co., Ltd. (Osaka, Japan) for providing the ZrO2–Al2O3 samples. This study was partially supported by a JSPS KAKENHI grant (JP17H00803), the Thailand Research Fund (IRG5780001), and an NRCT-NSFC joint funding project (NRCT/2558-104).

Author Contributions

P.W. and A.Y. conceived and designed the experiments; P.W. performed the catalytic reaction experiments; O.S. and N.M. analyzed the data; K.S. performed the NH3-TPD study; P.W., P.R., and A.Y. wrote the paper; all the authors discussed the results and commented on the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Huber, G.W.; Iborra, S.; Corma, A. Synthesis of Transportation Fuels from Biomass: Chemistry, Catalysts, and Engineering. Chem. Rev. 2006, 106, 4044–4098. [Google Scholar] [CrossRef] [PubMed]
  2. Petrus, L.; Noordermeer, M.A. Biomass to biofuels, a chemical perspective. Green Chem. 2006, 8, 861–867. [Google Scholar] [CrossRef]
  3. Corma, A.; Iborra, S.; Velty, A. Chemical Routes for the Transformation of Biomass into Chemicals. Chem. Rev. 2007, 107, 2411–2502. [Google Scholar] [CrossRef] [PubMed]
  4. Kobayashi, H.; Komanoya, T.; Guha, S.K.; Hara, K.; Fukuoka, A. Conversion of cellulose into renewable chemicals by supported metal catalysis. Appl. Catal. A 2011, 409–410, 13–20. [Google Scholar] [CrossRef]
  5. Gallezot, P. Conversion of biomass to selected chemical products. Chem. Soc. Rev. 2012, 41, 1538–1558. [Google Scholar] [CrossRef] [PubMed]
  6. Yamaguchi, A. Biomass Valorization in High-temperature Liquid Water. J. Jpn. Petrol. Inst. 2014, 57, 155–163. [Google Scholar] [CrossRef]
  7. Yamaguchi, A.; Sato, O.; Mimura, N.; Shirai, M. One-pot conversion of cellulose to isosorbide using supported metal catalysts and ion-exchange resin. Catal. Commun. 2015, 67, 59–63. [Google Scholar] [CrossRef]
  8. Datta, R.; Henry, M. Lactic acid: recent advances in products, processes and technologies—A review. J. Chem. Technol. Biotechnol. 2006, 81, 1119–1129. [Google Scholar] [CrossRef]
  9. Mäki-Arvela, P.; Simakova, I.L.; Salmi, T.; Murzin, D.Y. Production of Lactic Acid/Lactates from Biomass and Their Catalytic Transformations to Commodities. Chem. Rev. 2014, 114, 1909–1971. [Google Scholar] [CrossRef] [PubMed]
  10. Garlotta, D. A Literature Review of Poly(Lactic Acid). J. Polym. Environ. 2001, 9, 63–84. [Google Scholar] [CrossRef]
  11. Martinez, F.A.C.; Balciunas, E.M.; Salgado, J.M.; González, J.M.D.; Converti, A.; Oliveira, R.P.D.S. Lactic acid properties, applications and production: A review. Trends Food Sci. Technol. 2013, 30, 70–83. [Google Scholar] [CrossRef]
  12. Wang, Y.; Deng, W.; Wang, B.; Zhang, Q.; Wan, X.; Tang, Z.; Wang, Y.; Zhu, C.; Cao, Z.; Wang, G.; et al. Chemical synthesis of lactic acid from cellulose catalysed by lead(II) ions in water. Nat. Commun. 2013, 4, 2141. [Google Scholar] [CrossRef] [PubMed]
  13. Lei, X.; Wang, F.-F.; Liu, C.-L.; Yang, R.-Z.; Dong, W.-S. One-pot catalytic conversion of carbohydrate biomass to lactic acid using an ErCl3 catalyst. Appl. Catal. A 2014, 482, 78–83. [Google Scholar] [CrossRef]
  14. Yang, X.; Yang, L.; Fan, W.; Lin, H. Effect of redox properties of LaCoO3 perovskite catalyst on production of lactic acid from cellulosic biomass. Catal. Today 2016, 269, 56–64. [Google Scholar] [CrossRef]
  15. Coman, S.M.; Verziu, M.; Tirsoaga, A.; Jurca, B.; Teodorescu, C.; Kuncser, V.; Parvulescu, V.I.; Scholz, G.; Kemnitz, E. NbF5–AlF3 Catalysts: Design, Synthesis, and Application in Lactic Acid Synthesis from Cellulose. ACS Catal. 2015, 5, 3013–3026. [Google Scholar] [CrossRef]
  16. Chambon, F.; Rataboul, F.; Pinel, C.; Cabiac, A.; Guillon, E.; Essayem, N. Cellulose hydrothermal conversion promoted by heterogeneous Brønsted and Lewis acids: Remarkable efficiency of solid Lewis acids to produce lactic acid. Appl. Catal. B 2011, 105, 171–181. [Google Scholar] [CrossRef]
  17. Wattanapaphawong, P.; Reubroycharoen, P.; Yamaguchi, A. Conversion of cellulose into lactic acid using zirconium oxide catalysts. RSC Adv. 2017, 7, 18561–18568. [Google Scholar] [CrossRef]
  18. Teramoto, Y.; Tanaka, N.; Lee, S.-H.; Endo, T. Pretreatment of eucalyptus wood chips for enzymatic saccharification using combined sulfuric acid-free ethanol cooking and ball milling. Biotechnol. Bioeng. 2008, 99, 75–85. [Google Scholar] [CrossRef] [PubMed]
  19. Yamaguchi, A.; Hiyoshi, N.; Sato, O.; Bando, K.K.; Shirai, M. Gaseous Fuel Production from Nonrecyclable Paper Wastes Using Supported Metal Catalysts in High-Temperature Liquid Water. ChemSusChem 2010, 3, 737–741. [Google Scholar] [CrossRef] [PubMed]
  20. Yamaguchi, A.; Sato, O.; Mimura, N.; Hirosaki, Y.; Kobayashi, H.; Fukuoka, A.; Shirai, M. Direct Production of Sugar Alcohols from Wood Chips using Supported Platinum Catalysts in Water. Catal. Commun. 2014, 54, 22–26. [Google Scholar] [CrossRef]
  21. Zhao, H.; Kwak, J.H.; Wang, Y.; Franz, J.A.; White, J.M.; Holladay, J.E. Effects of Crystallinity on Dilute Acid Hydrolysis of Cellulose by Cellulose Ball-Milling Study. Energy Fuels 2006, 20, 807–811. [Google Scholar] [CrossRef]
  22. Said, A.E.-A.A.; El-Wahab, M.M.M.A.; El-Aal, M.A. Effect of ZrO2 on the catalytic performance of nano γ-Al2O3 in dehydration of methanol to dimethyl ether at relatively low temperature. Res. Chem. Intermed. 2016, 42, 1537–1556. [Google Scholar] [CrossRef]
  23. Zhang, D.; Duan, A.; Zhao, Z.; Wan, G.; Gao, Z.; Jiang, G.; Chi, K.; Chuang, K.H. Preparation, characterization and hydrotreating performances of ZrO2–Al2O3-supported NiMo catalysts. Catal. Today 2010, 149, 62–68. [Google Scholar] [CrossRef]
  24. Manrı́quez, M.E.; López, T.; Gómez, R.; Navarrete, J. Preparation of TiO2–ZrO2 mixed oxides with controlled acid–basic properties. J. Mol. Catal. A 2004, 220, 229–237. [Google Scholar] [CrossRef]
  25. Ma, Z.-Y.; Yang, C.; Wei, W.; Li, W.-H.; Sun, Y.-H. Surface properties and CO adsorption on zirconia polymorphs. J. Mol. Catal. A 2005, 227, 119–124. [Google Scholar] [CrossRef]
  26. Rinaldi, R.; Schuth, F. Acid Hydrolysis of Cellulose as the Entry Point into Biorefinery Schemes. ChemSusChem 2009, 2, 1096–1107. [Google Scholar] [CrossRef] [PubMed]
  27. Wang, J.; Xi, J.; Wang, Y. Recent advances in the catalytic production of glucose from lignocellulosic biomass. Green Chem. 2015, 17, 737–751. [Google Scholar] [CrossRef]
  28. Delidovich, I.; Palkovits, R. Catalytic Isomerization of Biomass-Derived Aldoses: A Review. ChemSusChem 2016, 9, 547–561. [Google Scholar] [CrossRef] [PubMed]
  29. Yang, L.; Yang, X.; Tian, E.; Lin, H. Direct Conversion of Cellulose into Ethyl Lactate in Supercritical Ethanol–Water Solutions. ChemSusChem 2016, 9, 36–41. [Google Scholar] [CrossRef] [PubMed]
  30. Yang, L.; Yang, X.; Tian, E.; Vattipalli, V.; Fan, W.; Lin, H. Mechanistic insights into the production of methyl lactate by catalytic conversion of carbohydrates on mesoporous Zr-SBA-15. J. Catal. 2016, 333, 207–216. [Google Scholar] [CrossRef]
  31. Pescarmona, P.P.; Janssen, K.P.F.; Delaet, C.; Stroobants, C.; Houthoofd, K.; Philippaerts, A.; De Jonghe, C.; Paul, J.S.; Jacobs, P.A.; Sels, B.F. Zeolite-catalysed conversion of C3 sugars to alkyl lactates. Green Chem. 2010, 12, 1083–1089. [Google Scholar] [CrossRef]
  32. Nakajima, K.; Noma, R.; Kitano, M.; Hara, M. Titania as an Early Transition Metal Oxide with a High Density of Lewis Acid Sites Workable in Water. J. Phys. Chem. C 2013, 117, 16028–16033. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of (a) un-milled cellulose and (b) milled cellulose.
Figure 1. XRD patterns of (a) un-milled cellulose and (b) milled cellulose.
Catalysts 07 00221 g001
Figure 2. Product yields obtained by using Al2O3, ZrO2, or ZrO2–Al2O3 mixed-oxide catalysts for the conversion of cellulose to lactic acid. Reaction conditions: ball-milled cellulose, 0.5 g; catalyst, 1.0 g; water, 50 g; reaction temperature, 473 K; reaction time, 6 h. HMF, 5-hydroxymethylfurfural.
Figure 2. Product yields obtained by using Al2O3, ZrO2, or ZrO2–Al2O3 mixed-oxide catalysts for the conversion of cellulose to lactic acid. Reaction conditions: ball-milled cellulose, 0.5 g; catalyst, 1.0 g; water, 50 g; reaction temperature, 473 K; reaction time, 6 h. HMF, 5-hydroxymethylfurfural.
Catalysts 07 00221 g002
Figure 3. Product yields obtained when 10%ZrO2–Al2O3 was used as the catalyst as a function of reaction time. Reaction conditions: ball-milled cellulose, 0.5 g; catalyst, 1.0 g; water, 50 g; reaction temperature, 473 K. HMF, 5-hydroxymethylfurfural.
Figure 3. Product yields obtained when 10%ZrO2–Al2O3 was used as the catalyst as a function of reaction time. Reaction conditions: ball-milled cellulose, 0.5 g; catalyst, 1.0 g; water, 50 g; reaction temperature, 473 K. HMF, 5-hydroxymethylfurfural.
Catalysts 07 00221 g003
Figure 4. Product yields obtained when 10%ZrO2–Al2O3 was used as the catalyst as a function of reaction temperature. Reaction conditions: ball-milled cellulose, 0.5 g; catalyst, 1.0 g; water, 50 g; reaction time, 6 h. HMF, 5-hydroxymethylfurfural.
Figure 4. Product yields obtained when 10%ZrO2–Al2O3 was used as the catalyst as a function of reaction temperature. Reaction conditions: ball-milled cellulose, 0.5 g; catalyst, 1.0 g; water, 50 g; reaction time, 6 h. HMF, 5-hydroxymethylfurfural.
Catalysts 07 00221 g004
Figure 5. X-ray diffraction patterns of (a) 5%ZrO2–Al2O3, (b) 10%ZrO2–Al2O3, (c) 20%ZrO2–Al2O3, and (d) ZrO2.
Figure 5. X-ray diffraction patterns of (a) 5%ZrO2–Al2O3, (b) 10%ZrO2–Al2O3, (c) 20%ZrO2–Al2O3, and (d) ZrO2.
Catalysts 07 00221 g005
Figure 6. Temperature-programmed desorption of ammonia profiles of (a) 5%ZrO2–Al2O3, (b) 10%ZrO2–Al2O3, (c) 20%ZrO2–Al2O3, and (d) ZrO2.
Figure 6. Temperature-programmed desorption of ammonia profiles of (a) 5%ZrO2–Al2O3, (b) 10%ZrO2–Al2O3, (c) 20%ZrO2–Al2O3, and (d) ZrO2.
Catalysts 07 00221 g006
Figure 7. Temperature-programmed desorption of carbon dioxide profiles of (a) 5%ZrO2–Al2O3, (b) 10%ZrO2–Al2O3, (c) 20%ZrO2–Al2O3, and (d) ZrO2.
Figure 7. Temperature-programmed desorption of carbon dioxide profiles of (a) 5%ZrO2–Al2O3, (b) 10%ZrO2–Al2O3, (c) 20%ZrO2–Al2O3, and (d) ZrO2.
Catalysts 07 00221 g007
Scheme 1. Reaction pathway for the conversion of cellulose to lactic acid.
Scheme 1. Reaction pathway for the conversion of cellulose to lactic acid.
Catalysts 07 00221 sch001
Table 1. Properties of the ZrO2–Al2O3 and ZrO2 catalysts.
Table 1. Properties of the ZrO2–Al2O3 and ZrO2 catalysts.
CatalystSurface Area 1/m2 g−1Crystal Phase of ZrO2 2Amount of Acid Sites 3/mmol g−1Amount of Base Sites 4/mmol g−1Yield of Lactic Acid 5/%
5%ZrO2–Al2O3135-0.1090.09325.0
10%ZrO2–Al2O3126Tetragonal0.1310.05725.3
20%ZrO2–Al2O3130Tetragonal0.1800.09821.9
ZrO2 6101Monoclinic0.1180.48521.2
1 Reported by manufacturer; 2 Determined by XRD; 3 Determined by temperature-programmed desorption of ammonia (NH3-TPD); 4 Determined by temperature-programmed desorption of carbon dioxide (CO2-TPD); 5 Reaction conditions: ball-milled cellulose, 0.5 g; catalyst, 1.0 g; water, 50 g; reaction temperature, 473 K; reaction time, 6 h; 6 Presented in reference [17].

Share and Cite

MDPI and ACS Style

Wattanapaphawong, P.; Sato, O.; Sato, K.; Mimura, N.; Reubroycharoen, P.; Yamaguchi, A. Conversion of Cellulose to Lactic Acid by Using ZrO2–Al2O3 Catalysts. Catalysts 2017, 7, 221. https://doi.org/10.3390/catal7070221

AMA Style

Wattanapaphawong P, Sato O, Sato K, Mimura N, Reubroycharoen P, Yamaguchi A. Conversion of Cellulose to Lactic Acid by Using ZrO2–Al2O3 Catalysts. Catalysts. 2017; 7(7):221. https://doi.org/10.3390/catal7070221

Chicago/Turabian Style

Wattanapaphawong, Panya, Osamu Sato, Koichi Sato, Naoki Mimura, Prasert Reubroycharoen, and Aritomo Yamaguchi. 2017. "Conversion of Cellulose to Lactic Acid by Using ZrO2–Al2O3 Catalysts" Catalysts 7, no. 7: 221. https://doi.org/10.3390/catal7070221

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