Elsevier

Catalysis Communications

Volume 46, 10 February 2014, Pages 66-70
Catalysis Communications

Short Communication
Thermodynamics and reaction pathways of furfuryl alcohol oligomer formation

https://doi.org/10.1016/j.catcom.2013.11.030Get rights and content

Highlights

  • Furfuryl alcohol monomer was converted into a broad range of oligomers.

  • Methylene-bridged, ether-bridged, terminal CH2OH, and internal OH dimers were produced.

  • The formation of terminal CH2OH is both kinetically and thermodynamically favored.

Abstract

The acid-catalyzed transformation of furfuryl alcohol (FA) monomer to oligomers has been studied in the liquid phase to investigate the reaction mechanisms and intermediate species by using a combination of quantitative reaction product measurements and density functional theory calculations. FA monomer was converted into oligomers with a broad range of carbon number: C9–C10, C14–C15, C19–C29, > C29. Based on the calculations, terminal CH2OH dimer formation is both kinetically and thermodynamically favored, consistent with the experimental results. The order for dimer production in the C9–C10 range follows terminal CH2OH > ether bridged–methylene bridged dimer > OH-carbon bridge.

Introduction

The conversion of lignocellulosic biomass into fuels and chemicals has been rigorously investigated as a response to the depletion of petroleum resources, increasing demand for oil, environmental concern over greenhouse gas emissions, and energy security [1], [2], [3]. Among the biomass derived chemicals, furfuryl alcohol (FA) is an attractive intermediate chemical for the production of levulinic acid [4], alkyl levulinate [5], and various useful polymer products [6], [7], [8]. FA can be obtained through the hydrogenation reaction of furfural and further converted into FA oligomers by the acid-catalyzed condensation reaction [9], [10], [11]. FA oligomers have been studied due to their application in fuel blending components [12] (Scheme 1). Buijtenen et al. have shown that after hydrogenation of FA oligomers with Ni/Al2O3, the fuel blend containing 10 vol.% of these C9–C20 carbon coupled FA oligomers fulfills the European diesel specification [12]. The acid-catalyzed condensation of FA is the major source for the FA oligomerization [13], [14]. Depending on the experimental conditions and method of analysis, different oligomer/polymer structures such as linear [9], [15], [16], cross-linked [17], [18], and open ring [19] have been proposed from the results of XRD [20], NMR [15], [21], [22], [23], UV–Vis [16], [21], [24], [25], IR [16], [20], [21], Raman spectroscopy [8], [9], [10], [13], [20], and DFT calculations [9]. We have recently shown that the FA monomer can dimerize in acidic solution and two possible structures can be produced through dehydration or deprotonation reactions (Eqs. (1), (2), the changes in free energy are given in kcal/mol) [9].

From the free energies formation of 2-hydroxymethyl-5(5-furfuryl)furan (HFF) is thermodynamically more favorable than 2,2′-difurfuryl ether (DFE). Even though, the formation of dimer from two FA monomers is thermodynamically favorable, because of the higher carbocation reactivity, dimer formation via the carbocation may be preferred for kinetic reasons. In addition to these two dimers, Khusnutdiv et al. reported that 2,2′-difurfurylmethane (DFM) can be also produced in the presence of the Rh(PPh3)3Cl catalyst using H2O and CCl4 as solvents [26]. The following mechanism for formation of DFM (Eq. (3)) was proposed by Sunjic et al. [27] and Korneeva et al. [28]. DFM and formaldehyde are produced through decomposition of DFE derived from the condensation of two FA monomers.

Using the 13C NMR spectroscopy, Choura et al. proposed a similar reaction mechanism (Eq. (4)) for the conversion of 5-methylfurfuryl alcohol (MFA) to the bis[2-(5-methylfuryl)]methane (BMFM) [15]. 2,2′-dimethylfurfuryl ether is formed first, which then decomposes into formaldehyde and BMFM.

Although the oligomerization of FA appears in a number of literature citations, the detailed understanding of reaction mechanisms and thermodynamics for the oligomerization processes (multiple pathways) is not available. Such an understanding will substantially enhance the utilization of FA as a platform chemical for the formation of longer chain hydrocarbons used in liquid transportation fuels. The aim of the presented study is to understand the reaction mechanism of oligomer formation and provide an accurate analysis of oligomer products. Here we have determined the product distribution of dimers (C9–C10) in detail and employed density functional theory to provide a qualitative confirmation of the experimentally observed dimer distribution trends.

Section snippets

Catalytic reaction

FA (98%) and sulfuric acid (99.999%) were obtained from Sigma-Aldrich and used without further purification. FA was mixed with dilute sulfuric acid (1 M) to catalyze the production of oligomers.

Initial GC/MS studies were undertaken to establish the identity of the major oligomeric species formed during the reaction (see Supporting information for the details of GC/MS experiment). Gas chromatographic analyses were performed using an Agilent 5890 Series II equipped with a flame ionization detector

Acid-catalyzed oligomerization of FA

Oligomerization of FA was performed with an aqueous solution containing 3.00 g FA and 0.03 g 1 M H2SO4 at 25 °C and ambient pressure, as shown in Fig. 1. During the reaction, the conversion values (mass basis) monotonically increase with increasing reaction time, and asymptotically reach ~ 32 wt.% in 24 h (Fig. 1a). Heavy oligomer (> C18) formation monotonically increases, while FA monomer (C5) wt.% decreases as a function of reaction time (Fig. 1b). It was found that up to 4 h, the wt% of > C18 was

Conclusions

In this work, we investigated the products and reaction mechanisms for the liquid-phase oligomerization of FA monomer using gas chromatography (experimental) and density functional methods (theoretical). The selectivity toward high molecular weight oligomers (> hexamer) increases with increasing reaction time. Four different kinds of dimers were observed: methylene bridged (D1), ether bridged (D2), terminal CH2OH (D3), and internal OH (D4) dimers. Theoretical studies were carried out to

Acknowledgments

This work was supported as part of the Institute for Atom-efficient Chemical Transformations (IACT), an Energy Frontier Research Center funded by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences. Argonne is managed by UChicago Argonne, LLC, for the U.S. Department of Energy under contract DE-AC02-06CH11357. Use of the computational resources from Center for Nanoscale Materials was supported by the U.S. Department of Energy, Office of Science, Office of Basic

References (32)

  • T. Kim et al.

    Catal. Today

    (2013)
  • P.S. Batista et al.

    Polymer

    (2000)
  • A. Gandini et al.

    Prog. Polym. Sci.

    (1997)
  • C. Mealares et al.

    Polymer

    (1996)
  • C.R. Mendonca et al.

    Chem. Phys. Lett.

    (2001)
  • P.S. Batista et al.

    Synth. Met.

    (1999)
  • R.S. Assary et al.

    Chem. Phys. Lett.

    (2010)
  • Y.R. Leshkov et al.

    Science

    (2006)
  • J.Q. Bond et al.

    Science

    (2010)
  • Y.J. Pagan-Torres et al.

    ACS Catal.

    (2010)
  • B. Girisuta et al.

    Green Chem.

    (2006)
  • W.D. van de Graaf, J.P. Lange, U.S. Patent 7,265,239,...
  • A. Gandini

    Small

    (2007)
  • N. Guigo et al.

    Phys. Chem. Chem. Phys.

    (2007)
  • B. Yi et al.

    J. Am. Chem. Soc.

    (2006)
  • T. Kim et al.

    ChemCatChem

    (2001)
  • Cited by (39)

    • New salicylic acid derivatives, double inhibitors of glycolate oxidase and lactate dehydrogenase, as effective agents decreasing oxalate production

      2022, European Journal of Medicinal Chemistry
      Citation Excerpt :

      Acid-catalyzed polymerization of 2-furylcarbinols (5, Scheme 2) to di-(2-furyl)methanes is a well-known process [34–36]. It has been explained via etherification [37] between two molecules of 5 (steps a-c) and further decomposition of the resulting ether (6) into formaldehyde (7) and the final difurylmethane (step d) [38]. One reference was found regarding polymerization of 2-furylmethanamines to di-(2-furyl)methanes [39].

    • Solid catalysts for furfuryl alcohol conversion to drop-in chemicals

      2021, Advanced Catalysis for Drop-in Chemicals
    • Esterification of furfuryl alcohol to butyl levulinate over ion-exchange resins

      2019, Fuel
      Citation Excerpt :

      Another feasible synthesis route is from furfuryl alcohol (FA) via acid-catalyzed alcoholysis. The main drawback of starting from furfuryl alcohol is that it easily polymerizes under acidic conditions at temperatures as low as 25 °C [19], decreasing the yield to BL. Recently, some authors have studied this reaction using several solid acid catalysts [20–26].

    View all citing articles on Scopus
    1

    Present address: Chemical and Molecular Engineering Department, Stony Brook University, NY 11787, USA.

    View full text