Anda di halaman 1dari 8

DOI: 10.1002/slct.

201700869 Full Papers

z Sustainable Chemistry

Novel Acid Catalysts from Waste-Tire-Derived Carbon:


Application in Wasteto-Biofuel Conversion
Zachary D. Hood+,[a, b] Shiba P. Adhikari+,[c, d] Yunchao Li,[e, f] Amit K. Naskar,[f, g] Legna Figueroa-
Cosme,[a] Younan Xia,[a, h] Miaofang Chi,[b] Marcus W. Wright,[c] Abdou Lachgar,*[c, d] and M.
Parans Paranthaman*[e, f]

Many inexpensive biofuel feedstocks, including those contain- environmentally benign process that involved the sequential
ing free fatty acids (FFAs) in high concentrations, are typically treatment with L-cysteine, dithiothreitol, and H2O2. When
disposed of as waste due to our inability to efficiently convert benchmarked against the same waste-tire derived carbon
them into usable biofuels. Here we demonstrate that carbon material treated with concentrated sulfuric acid at 150 8C,
derived from waste tires could be functionalized with sulfonic similar catalytic activity was observed. Both catalysts could also
acid (-SO3H) to effectively catalyze the esterification of oleic effectively convert oleic acid or a mixture of fatty acids and
acid or a mixture of fatty acids to usable biofuels. Waste tires soybean oil to usable biofuels at 65 8C and 1 atm without
were converted to hard carbon, then functionalized with leaching of the catalytic sites.
catalytically active -SO3H groups on the surface through an

Introduction
This research addresses two major issues related to sustain-
[a] Z. D. Hood,+ L. Figueroa-Cosme, Prof. Y. Xia ability, namely, waste and energy production, by converting
School of Chemistry and Biochemistry waste tires to carbon-based solid acid catalysts through an
Georgia Institute of Technology environmentally benign process and by further demonstrating
Atlanta, GA, 30332 (USA)
the use of the catalysts for esterification of oleic acid and/or a
[b] Z. D. Hood,+ Dr. M. Chi
Center for Nanophase Materials Sciences mixture of fatty acids and soybean oil to produce biofuels. On a
Oak Ridge National Laboratory yearly basis, approximately 290 million tires are disposed of in
Oak Ridge, TN, 37831 (USA) the United States, of which 27 million end up in landfills[1].
[c] S. P. Adhikari,+ Prof. M. W. Wright, Prof. A. Lachgar
Since tire rubber is non-biodegradable, it inherently poses
Department of Chemistry
Wake Forest University major health and environmental concerns. Recently, pyrolysis of
Winston-Salem, NC, 27109 (USA) waste tires into high value-added hard carbon has received
E-mail: lachgar@wfu.edu increasing interest as it provides an inexpensive carbon source
[d] S. P. Adhikari,+ Prof. A. Lachgar
for a wide variety of applications, including the production of
Center for Energy, Environment, and Sustainability (CEES)
Wake Forest University supports for electrocatalysts, anodes for Li-/Na-ion and Li S
Winston-Salem, NC, 27109 (USA) batteries, and electrodes for supercapacitors[2]. The present
[e] Y. Li, Dr. M. P. Paranthaman demonstration adds another major application for carbon
Chemical Sciences Division
material derived from waste tires. More significantly, most used
Oak Ridge National Laboratory
Oak Ridge, TN, 37831 (USA) vegetable oils or animal fats contain large amounts of free fatty
E-mail: paranthamanm@ornl.gov acids (FFAs) and therefore cannot be used in homogeneous
[f] Y. Li, Dr. A. K. Naskar, Dr. M. P. Paranthaman base-catalyzed transesterification, a technology commonly
The Bredesen Center for Interdisciplinary Research and Graduate Educa-
used to produce biofuels from high-quality refined fats. These
tion
The University of Tennessee inexpensive feedstocks require a pretreatment step in which a
Knoxville, TN, 37996 (USA) strong acid such as H2SO4 is used for the esterification of FFAs
[g] Dr. A. K. Naskar to biofuel. The use of strong acids has drawbacks such as
Materials Science and Technology Division
additional neutralization steps and the necessity of the
Oak Ridge National Laboratory
Oak Ridge, TN, 37831 (USA) utilization of expensive corrosion-resistant equipment. The
[h] Prof. Y. Xia carbon-based solid acid catalyst described in this report offers a
The Wallace H. Coulter Department of Biomedical Engineering viable alternative for this process.
Georgia Institute of Technology
Herein, we report a systematic study of the synthesis,
Atlanta, GA, 30332 (USA)
[+] These authors contributed equally to this publication thermal stability, surface chemistry, and catalytic activity of
Supporting information for this article is available on the WWW under carbon-based solid acid catalysts derived from waste tires. The
https://doi.org/10.1002/slct.201700869 solid acid catalysts based on waste-tire derived carbon were

ChemistrySelect 2017, 2, 4975 4982 4975  2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers

Scheme 1. Schematic illustration showing the synthesis of sulfonated tire-derived carbon catalyst (STC-cys) through sequential treatment with L-cysteine, DTT,
and H2O2.

prepared using two different methodsthe conventional Likewise, the oxygen contents in STC-cys and STC were
method based upon the treatment with concentrated sulfuric increased by 60% and 110%, respectively, relative to the parent
acid at 150 8C and a newly developed method involving the carbon. The hydrogen content increased most significantly,
sequential treatment with L-cysteine, dithiothreitol, and H2O2 reaching 270% and 400% for STC-cys and STC, respectively.
under mild conditions. Both methods produced sulfonated It is critical for the carbon-based catalysts to be thermally
carbon with comparable catalytic activity toward the esterifica- stable so they will not decompose or oxidize during the
tion reaction for the effective conversion of waste oils catalytic process, which is typically performed at temperatures
containing a mixture of FFAs to advanced biofuels. The in the range of 65100 8C[5]. Thermogravimetric analysis (TGA)
catalysts retained their catalytic activity even after exhaustive and powder X-ray diffraction (PXRD) studies were conducted
methanol leaching. To the best of our knowledge, this work for STC-cys and STC to evaluate their thermal and structural
represents the first report on the demonstration of an environ- stabilities. The TGA data are shown in Figure S1. In contrast to
mentally benign method to produce solid acid catalysts from the tire-derived carbon, both STC-cys and STC carbon showed
waste tires. low weight loss upon heating to ~ 250 8C due to the loss of a
small amount of surface-adsorbed water, followed by a second
weight loss between 250600 8C for STC-cys and 250500 8C
Results and Discussion
for STC attributed to the elimination of sulfur as SO2 or SO3 [6].
Scheme 1 shows a schematic procedure for the production of a The most important result of this thermal stability study is the
solid acid catalyst from the carbon material derived from waste fact that STC-cys was thermally as stable as the tire-derived
tires. The process starts with the use of L-cysteine as a sulfur carbon (both started to decompose at ~ 650 8C), while STC
source to coat the surface of carbon. Most of the sulfur atoms undergoes thermal decomposition at ~ 550 8C. The lower
derived from cysteine molecules are covalently connected to decomposition temperature of STC is believed to result from
each other through the disulfide linkage while some of them the surface treatment with concentrated sulfuric acid[6]. This
may exist in the thiol form. The disulfide bonds can be result also suggests that the cysteine-based procedure for
efficiently reduced with dithiothreitol (DTT) under basic con- modifying the tire-derived carbon is a milder sulfonation
ditions (pH > 7) to generate thiols, which are then oxidized method that does not adversely affect the carbon support. As
with H2O2 to generate sulfonic acid groups. It is worth noting shown by the PXRD patterns in Figure S2, all samples exhibited
that this method is widely used by biochemists interested in broad diffraction peaks at 2q between 15 and 358, correspond-
the study of biochemical behavior of sulfur-containing pro- ing to non-graphitic carbon[7]. Upon heating the samples at
teins[3]. The final product is designated here as the STC-cys 800 8C under argon for 15 h, the PXRD patterns were retained
catalyst. For comparison, a batch of waste-tire derived carbon for all three samples, further confirming their good thermal
was sulfonated by soaking it in concentrated sulfuric acid at stability.
150 8C for 15 h with vigorous stirring and subsequently washing X-ray photoelectron spectroscopy (XPS) was used to
with copious amounts of DI water[4]. Hereafter, this sample is analyze the surface chemistry before and after sulfonation. The
referred to as the STC catalyst. XPS spectra for the S 2p peak are shown in Figure 2. The spin-
Table S1 shows elemental microanalysis data for the tire- orbital-split doublets (S 2p3/2 and S 2p1/2) at ~ 164 eV can be
derived carbon, STC-cys, and STC. For an easy comparison of assigned to the thiophene group that is retained from the
the elemental compositions, an empirical formula was derived parent carbon, while the spin-orbital-split doublets (S 2p3/2 and
for each sample, with nitrogen being excluded from the S 2p1/2) observed at 169 eV indicates the presence of sulfonic
calculation because of its low content. Both STC-cys and STC acid group (-SO3H)[8]. The observation of O 1 s peak further
showed a significant increase in hydrogen, sulfur, and oxygen supports the presence of sulfonic acid in both STC-cys and STC,
(per mole of carbon) in comparison with the starting tire- where deconvolution of this peak details the existence of SO3
derived carbon. Specifically, STC-cys and STC contained 40% groups (Figure S3). The XPS spectra for the C 1 s peak in
and 80%, respectively, more sulfur relative to the parent carbon. Figure S4 shows further evidence for the presence of oxygen in

ChemistrySelect 2017, 2, 4975 4982 4976  2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers

Table 1. Surface area, pore volume, pore size, and average particle size for the tire carbon, STC-cys catalyst, and STC catalyst.

BET Surface Area Pore Volume Pore Size Particle Size Range Average particle size (mm) based on
(m2 g 1) (cc g 1) (nm) (nm mm) SEM analysis

Pyrolyzed waste tire-derived carbon 224 0.10 3.06 54 74 ~8


Sulfonated tire carbon from L-cysteine 26 0.18 3.82 148 107 ~ 34
(STC-cys) catalyst
Sulfonated tire carbon (STC) catalyst 16 0.08 3.82 153 - 154 ~ 45

STC-cys and STC. Based on XPS data, two mechanisms are and STC catalysts corresponds to an increase in sulfonic acid
proposed to account for the formation of sulfonic acid groups concentration.
on STC-cys and STC (Figure S5): coupling to carboxylic func- Upon sulfonation, both STC-cys and STC exhibited morphol-
tional groups on the surface of the carbon and attachment to ogy similar to that of the parent tire-derived carbon. Figure 1
the 2- or 5- position on the thiophenic heterocycle, respectively. shows scanning electron microscopy (SEM) and transmission
electron microscopy (TEM) images of the three carbon materi-
als. Before sulfonation, the tire-derived carbon showed irregular
morphology, composed of both nano- and micro-scale particles
with an average particle size of ~ 8 mm. After surface
modification with cysteine or treatment with concentrated H2
SO4, the carbon maintained its granular morphology while the
overall particle size increased. The STC-cys had an average
particle size of ~ 34 mm. Representative high magnification SEM
images for these three samples are shown in Figure S6. The
increase in average particle size for STC-cys was confirmed by
Brunauer-Emmett-Teller (BET) surface area measurements,
where the surface areas were determined to be 224 and 26 m2
g 1 for the tire-derived carbon and STC-cys, respectively (shown
in Table 1). On the other hand, the STC catalyst had an average
particle size ~ 45 mm, together with a BET surface area of 16 m2
g 1. Nitrogen adsorption-desorption isotherms confirm that all
the three carbon samples were porous (Figure S7). Using BJH
analysis, the pore volume and pore size distributions were
found to be 0.10 cc g 1 and 3.06 nm for the tire-derived carbon,
0.18 cc g 1 and 3.82 nm for STC-cys, and 0.08 cc g 1 and
3.82 nm for STC. Both STC-cys and STC had similar pore volume
and pore size after sulfonation, although their surface areas
decreased significantly, consistent with the increase in average
particle size. The lower surface area and higher -SO3H
concentration observed for both STC-cys and STC compared to
that of the parent carbon, could indicate that functionalization
occurred both on the surface as well as within the pores of the
carbon precursor[10]. SEM images with corresponding EDS
mapping of sulfur for pyrolyzed waste tire, STC-cys and STC
catalysts are shown in Figure S8. EDS elemental mapping of the
original tire-derived carbon clearly shows the presence of sulfur
Figure 1. SEM (A C) and TEM (D F) images of the A, D) carbon derived from
waste tires by pyrolysis, B, E) STC-cys catalyst, and C, F) STC catalyst. G) XPS due to the presence of thiophene groups consistent with XPS
spectra of the S 2p region recorded from the three samples. data. After surface modification, significantly higher level of
sulfur was observed, in agreement with XPS data, which
confirmed the presence of both thiophene and sulfonic acid.
The catalytic activity of STC-cys and STC toward the
The density of acid functional groups was determined by esterification of FFAs to fatty acid methyl esters (FAMEs) was
titration to be 2.23  0.22 mmol/g for STC-cys and 2.73  studied using oleic acid (OA) as a model because it has a long
0.17 mmol/g for STC. This acid density is significantly higher fatty chain and is present in high concentration in typical FFA-
than that of other acid catalysts such as Amberlyst-15 and rich feedstocks. The activity of the catalysts was also tested
Nafion, which was reported to be 0.8 mmol/g[9]. Elemental using a mixture of fatty acids, or fatty acids and soybean oil[10
11]
analysis data suggest that the increase of sulfur in the STC-cys . The optimum STC-cys catalyst loading was determined by

ChemistrySelect 2017, 2, 4975 4982 4977  2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers

Figure 2. Evaluation of the STC-cys catalyst for oleic acid to biofuel conversion. A) Optimization of catalyst loading reveals 10 wt.% of STC-cys for > 98%
conversion from oleic acid to methyl oleate in 4 h in the presence of 10:1 methanol to oleic acid concentration. B) Time-dependent study using different wt.%
of STC-cys. C) Optimization of MeOH/FFA content using 10 wt.% of STC-cys shows that a 10:1 methanol to oleic acid content is necessary to achieve high
conversions at a reaction time of 4 h. D) Water effect study using 10 and 0.5 wt.% of STC-cys and a 10:1 MeOH to oleic acid ratio, where the H2O wt.% is in
terms of oleic acid content. For each of these studies, the weight percent of the STC-cys catalyst was measured in terms of the oleic acid mass and the
temperature of the reaction was maintained at 65 8C.

varying the catalyst to FFA (OA) weight % using methanol to for STC (Figure S12). When a 2.5 molar ratio of methanol to OA
FFA molar ratio of 10.0. When 0.5 wt.% of the STC-cys catalyst was used, the conversion was found to be 9295% for the STC-
was used, 7383% conversion of OA was obtained after 4 h cys.
(Figure 2 A). The highest conversion (~ 100 %) was achieved Homogeneous acid-based catalysis generally follows sec-
when 10 wt.% STC-cys was used (Figure 2B). The STC catalyst ond-order reaction kinetics, which is the case for sulfuric acid.[13]
showed a similar trend (Figure S9). Percent conversion of OA Heterogeneous catalysts, on the other hand, are typically more
vs. reaction time using the STC catalyst is shown in Figure S10. complex due to surface-adsorbate interactions. As such, the
Kinetic studies of both catalysts were performed, with 0.25 and reaction order of each reactant was determined by varying the
0.5 wt.% catalyst/OA ratios and the reaction performed under concentration of methanol and oleic acid while monitoring the
reflux with a methanol/OA molar ratio set at 6.0 (Figure S11). initial reaction rates when the reactions were performed at
With 0.5 wt.% of catalyst added the mass-normalized initial 65 8C using low catalyst loading (0.25 - 0.5 wt.%) (Table S2;
reaction rate constants were determined to be 3.8 and Table S3). The concentration of both methanol and OA had
6.4 mmol min 1 g 1 for STC-cys and STC catalysts, respectively. positive effects on the initial reaction rate. The initial reaction
These values are comparable to other catalysts currently used rate increased from 18.0 to 27.7 mM min 1 for the STC-cys
for esterification reactions, such as Amberlyst resins and other catalyst and 31.8 to 50.8 mM min 1 for the STC catalyst when
mesoporous silicas.[12] The effect of methanol concentration on the methanol to OA molar ratio was increased from 3 to 6. For
the catalytic activity of STC-cys was also evaluated. In the OA, the initial reaction rates increased from 13.1 to 18.0 mM
presence of a 5.0 molar ratio of methanol to OA with 10.0 wt.% min 1 for the STC-cys catalyst and 23.5 to 31.8 mM min 1 for
STC-cys catalyst added, the conversion of OA to methyl ester the STC catalyst when the OA to methanol molar ratio
was 9598% (Figure 2C), comparable to the conversion found increased from 0.167 to 0.333, suggesting that increased

ChemistrySelect 2017, 2, 4975 4982 4978  2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers
concentrations of OA relative to methanol also has a positive (0.5wt.%), it was apparent that the reaction kinetics were
effect on the reaction rate. The apparent reaction orders were suppressed by water, where the initial reaction rate is
determined using power law approximation, which were 0.62- decreased by 3234% when 20 wt.% H2O is added to the
0.68 order for methanol, 0.44-0.46 order for oleic acid, and 0.68- reaction. This decrease in catalytic activity can be a result of the
0.71 for the catalyst. ability of water to solvate free protons rather than methanol,
The chemical stability of both STC-cys and STC toward which is similar to previous reports.[4b, 13] Increasing the catalyst
water was investigated not only because water is a byproduct loading overcomes this issue, where our results indicate that
of the esterification reaction but also because generally, FFA- 10.0 wt.% of STC-cys and STC could effectively convert OA to
rich feedstocks contain high content of water, necessitating methyl ester even in the presence of a relatively large amount
catalysts that can function in aqueous environments.[13] The of water.
activity was evaluated using both 0.5 and 10 wt.% catalyst and Waste oils contain a broad range of FFA contents, so it is
a 10:1 methanol to FFA ratio in the presence of 5, 10, 15, and critical to determine the efficacy of a catalyst for feedstocks
20 wt.% of H2O (relative to FFA content). A small decrease in with different FFA contents by evaluating the FFA conversion
the % conversion was indeed observed as the wt.% of H2O was to methyl esters in a ternary system involving methanol, OA,
increased with a catalyst loading of 10.0 wt.%. A 9496% and triglycerides (soybean oil, SO). We investigated feedstocks
conversion was observed after the addition of 10 wt.% H2O with varying ratios of OA to SO and added 10.0 equivalents of
while a 8790% conversion was attained for samples contain- methanol in terms of FFA content (Figure S14). The results
ing 20 wt.% H2O (Figure 2D). With a decreased catalyst loading suggest that STC-cys could convert OA to methyl esters when
(0.5 wt.%), the oleic acid conversion decreased by ~ 12% when the feedstock contained an increasing amount of FFA. The
20 wt.% H2O was added to the reaction. The STC catalyst lower conversions observed for feedstocks with > 70% SO can
showed the same trend with comparable OA to ester be attributed to less favorable equilibrium position at lower
conversions (Figure S13). We further evaluated the effect of concentrations of OA in the FFA/triglyceride feedstock. To
water on acid-catalyzed esterification by evaluating the initial maintain a favorable equilibrium for the STC-cys catalyst, it is
reaction kinetics (Figure 3). With decreased catalyst loadings critical to have feedstocks with > 30% FFA, together with
enough methanol. To further explore the catalytic activity of
STC-cys and STC a mixture of FFAs (19.5% palmitic acid, 6.0%
stearic acid, 35.4% oleic acid, and 38.2% linoleic acid) and a 1:1
mixture of FFAs (in the same ratio) with soybean oil were
studied. The STC-cys and STC catalysts converted > 90% of the
FFAs to FAMEs, using 10.0 wt.% in terms of FFA content and
similar reaction conditions used in the initial catalyst studies
(Figure S15). Figure S16 presents images of the feedstock
before and after esterification, where the initial feedstock
consists of a solid, fatty mass and the product is composed of
liquid FAMEs at room temperature.
The active sites in many carbon-based solid acid catalysts
can potentially leach into the reaction mixture, which not only
compromises the ability of the catalyst to be used for
subsequent reactions but also leads to wrong conclusions with
respect to heterogeneous versus homogeneous catalysis. We
determined the chemical stability of STC-cys and STC toward
leaching by subjecting the catalysts to exhaustive methanol
leaching in a Soxhlet extractor. A total of 4 Soxhlet extractions
were performed for each catalyst, and the dried solids were
used to determine the catalysts activity (Figure 4, A and B). The
conversion was found to be 95, 94, 92 and 87% for STC-cys
after cycles 1, 2, 3, and 4, respectively. Similar results were
obtained for STC. This excellent retention in catalytic efficiency
suggests that both STC-cys and STC were able to retain the
-SO3H groups during exhaustive methanol leaching. This result
Figure 3. Evaluation of water sensitivity for the A) STC-cys and B) STC was confirmed by XPS studies in the S 2p region, which clearly
catalysts. The molar ratio of methanol to oleic acid was set at 10, and the show the presence of -SO3H groups (Figure 4, C and D).
water concentration is in terms of the oleic acid content. The catalyst loading Furthermore, when significant leaching of -SO3H active sites
was 0.5 wt.% (in terms of oleic acid content).
occurs in carbon-based materials, the leachate will typically
have good catalytic activity[4b, 14]. To determine the catalytic
activity of the leachate, 1.00 mL of the leachate was added to
the reaction mixture in lieu of STC-cys and the % conversion

ChemistrySelect 2017, 2, 4975 4982 4979  2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers

Figure 4. Catalytic activity of converting oleic acid to methyl oleate for A) the STC-cys catalyst and B) the STC catalyst after each exhaustive Soxhlet extraction
cycle. XPS spectra of the S2p region for C) the STC-cys catalyst and D) the STC catalyst after each Soxhlet extraction with methanol. The XPS results show
almost no change in the sulfur content, supporting the robustness for both the STC-cys and STC.

was determined. The conversion using leachates from cycles 1, and three drops of metallic mercury yielded similar results for
2, 3, and 4 for STC-cys catalyst was found to be 22, 12, 10, and the STC-cys and STC catalysts. Upon adding one drop of
10%, respectively. Except for the leachate from the first mercury and reacting for 4 h, the conversion of OA to methyl
methanol treatment, the conversions are comparable to control ester was retained at a level of 9599%. Similarly, when three
reactions performed with no catalyst added. The 22% con- drops of mercury were added, 9399% conversion was
version for the first leachate could be attributed to excess observed. A similar variation was found with decreased catalyst
sulfonic acid groups loosely bound to the surface of the tire- loadings. These results confirm the heterogeneous nature for
derived carbon. Moreover, the pH of the first leachate was both the STC-cys and STC catalysts.
slightly acidic (pH 5.5-6.0), while the pH values of all subse- In comparison to previously explored solid acid catalysts,
quent leachate solutions were close to 7, further supporting STC-cys and STC catalysts were found to have remarkable
the conclusion that the SO3H active sites were strongly bound ability in converting OA to FAMEs, even after continued use.
to the waste-tire-derived carbon. Elemental analysis of the Table 2 lists several reported carbon-based catalysts and their
leached carbon samples presented in Table S1 also supports ability to perform esterification of different FFAs to methyl ester
the retention of catalytically active sulfonic acid groups. in comparison with the STC-cys and STC catalysts. Sulfonated
To confirm the heterogeneous catalysis nature of both STC- hydrothermal carbon from glucose, sulfonated carbonized
cys and STC, mercury poisoning were performed by adding corncobs, and sulfonated lignin all showed good catalytic
metallic mercury into the reaction mixture[15]. In the presence of activity toward the esterification of different FFAs in the initial
mercury, the catalytic activity of a homogeneous catalyst will cycle, however, with increased use, their activity dropped
be quenched as a result of mercury poisoning[4b, 15]. In contrast, significantly[4b, 16]. These carbon materials tended to leach their
the retention of catalytic activity upon the addition of mercury active sites, which explain their loss in catalytic ability. In the
suggests heterogeneous catalysis. As shown in Figure S17, one case of sulfonated, ordered mesoporous carbon catalysts (i. e.,

ChemistrySelect 2017, 2, 4975 4982 4980  2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers

Table 2. Comparison of the state-of-the-art carbon catalysts for FFA to biofuel conversion.

Catalyst Reaction Conditions Catalyst Conversion TOF (s 1) Ea Recyclability Reference


Loading of FFA (%) (kcal/
mol)

Sulfonated tire carbon from Methanol to oleic acid molar 10 wt.% 98  2.0 0.07 10.2  Conversion of FFA (%): 91.8  This work
L-cysteine ratio: 10; 65 8C, 4 h (10 min) 1.5 0.9 after leaching 3x
(STC-cys)
Sulfonated tire carbon (STC) Methanol to oleic acid molar 10 wt.% 99  1.0 0.09 9.8  Conversion of FFA (%): 91.0  This work
ratio: 10; 65 8C, 4 h (10 min) 1.5 1.2 after leaching 3x
Sulfonated hydrothermal Methanol to oleic acid molar 10 wt.% 96 0.07 11  2 Conversion of FFA (%): 72 [4b]
carbons (from glucose) ratio: 10; 65 8C, 4 h (10 min) after leaching 3x
Sulfonated-carbonized corn- Methanol to oil molar ratio: 3 wt.% 98 N/A N/A Conversion of FFA (%): < 20 [16b]
cobs 32; 80 8C, 6 h after 2 cycles
Sulfonated lignin Methanol to oil molar ratio: 9; 7 wt.% 97 N/A N/A Conversion of FFA (%): < 80 [16c]
70 8C, 5 h after 4 cycles
PhSO3-H-OMCs Ethanol to oleic acid molar 3.5 wt.% 65-75 ~ 0.006 N/A Conversion of FFA (%): 6570 [17]
ratio: 10; 65 8C, 1.25 h (10 min) after 4 cycles
Amberlyst-15 Methanol to oleic acid molar 5.1 wt.% 95 ~ 0.004 17.46 N/A [18a, 18b]
ratio: 8; 120 8C, 1.25 h
Mesoporous silica (SBA-15- Methanol to palmitic acid 10 wt.% 85 0.0018 9.6  N/A [18c]
SO3H P123) molar ratio: 20; 85 8C, 3 h 0.2

PhSO3-H-OMCs), about 6575% of OA was converted to the carbon-derived from waste tires and environmentally benign
methyl ester in initial cycles, and after continued use, the protocols for surface modification. The obvious advantages of
percent conversion did not significantly decrease[17]. This STC-cys catalyst include relatively benign functionalization
retention in catalytic activity was attributed to strong binding process involving L-cysteine, very good catalytic performance,
of the aryl sulfonic acid group to the substrate through a and stability. The work shows promise for both green and
robust C C and C S bonds, which allows the catalyst to economic production of biofuels using inexpensive FFA-rich
preserve SO3H. In the case of the STC-cys and STC catalysts, feedstocks. The sulfonated, tire-derived carbon should also find
the carbon support was thermally stable up to temperatures use in acid-catalyzed chemical reactions beyond biofuel
> 200 8C and each catalyst exhibited excellent recyclability, applications.
indicating that the sulfonic acids were strongly attached to the
carbon support. Other heterogeneous catalysts, such as
Supporting Information Summary
Amberlyst-15 and mesoporous silica (SBA-15-SO3H P123) have
also been shown to convert free fatty acids to methyl esters Details of the experimental procedures, characterization data,
through esterification.[18] However, major drawbacks of sulfo- and evaluation of catalytic activity are described in the
nated silicas include decreased acid site concentration, diffu- Supporting Information.
sion limitations, and decreased turn-over frequency (TOFs).[19]
The tire-derived carbon catalysts have a number of advantages
Acknowledgements
over other previously reported catalysts. In the case of
sulfonated carbonized corncobs, the carbon originates from The evaluation of the new materials as novel carbon catalysts
biodegradable source, whereas in the newly developed cata- was performed equally at Wake Forest University (WFU),
lysts, the carbon originates from waste that has adverse Department of Chemistry and at Oak Ridge National Labora-
environmental impact and is the cause of major health tory. The work at WFU was funded by a National Academy of
issues.[16b] Additionally, mesoporous carbons were synthesized Sciences award NAS Sub-Grant 2000006099, and the work at
for the PhSO3-H-OMCs catalyst, which generally need longer ORNL (functionalize carbon to use as a catalyst) was sponsored
synthetic procedures and relatively expensive precursors to by the U.S. Department of Energy, Office of Science, Office of
create the carbon support. These catalysts were reported to Basic Energy Sciences, Materials Sciences and Engineering
achieve a maximum of ~ 74% conversion of OA to FAMEs, Division. The research on the conversion of recycled tires to
which, in comparison, is much lower than the conversions carbon powders was funded by Oak Ridge National Labora-
reported here for the both STC-cys and STC.[17] torys Technology Innovation Program. The authors gratefully
acknowledge financial support from the Wake Forest University
Center for Energy, Environment, and Sustainability. A portion of
Conclusion
this research was completed at the Center for Nanophase
To summarize, it is of critical importance to develop new Materials Sciences, which is a DOE Office of Science User
carbon-based solid acid catalysts for the conversion of FFAs to Facility. ZDH gratefully acknowledges a Graduate Research
usable biofuels. This report demonstrates a new direction for Fellowship from the National Science Foundation (DGE-
the design of solid acid catalysts by emphasizing the use of 1148903) and the Georgia Tech-ORNL Fellowship.

ChemistrySelect 2017, 2, 4975 4982 4981  2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim
Full Papers
Notice: This manuscript was coauthored by UT-Battelle, LLC [5] a) J. A. Melero, J. Iglesias, G. Morales, Green Chem. 2009, 11, 12851308;
under Contract No. DE-AC05-00OR22725 with the U.S. Depart- b) F. Su, Y. Guo, Green Chem. 2014, 16, 29342957.
[6] J. R. Kastner, J. Miller, D. P. Geller, J. Locklin, L. H. Keith, T. Johnson, Catal.
ment of Energy. The United States Government retains and the Today 2012, 190, 122132.
publisher, by accepting the article for publication, acknowl- [7] a) K. Nakajima, M. Hara, S. Hayashi, J. Am. Ceram. Soc. 2007, 90, 3725
edges that the United States Government retains a non- 3734; b) M.-H. Zong, Z.-Q. Duan, W.-Y. Lou, T. J. Smith, H. Wu, Green
exclusive, paid-up, irrevocable, world-wide license to publish or Chem. 2007, 9, 434437.
[8] a) D. G. Castner, K. Hinds, D. W. Grainger, Langmuir 1996, 12, 50835086;
reproduce the published form of this manuscript, or allow b) X.-L. Wei, M. Fahlman, A. Epstein, Macromolecules 1999, 32, 3114
others to do so, for United States Government purposes. The 3117; c) P. Siril, N. Shiju, D. Brown, K. Wilson, Appl. Catal., A 2009, 364,
Department of Energy will provide public access to these 95100; d) S.-A. Wohlgemuth, F. Vilela, M.-M. Titirici, M. Antonietti, Green
results of federally sponsored research in accordance with the Chem. 2012, 14, 741749; e) H. Okawa, T. Wada, H. Sasabe, K. Kajikawa,
K. Seki, Y. Ouchi, Jpn. J. Appl. Phys. 2000, 39, 252.
DOE Public Access Plan (http://energy.gov/downloads/doe- [9] L. Wu, Y. Wu, X. Liang, Kinetics and Catalysis 2013, 54, 378381.
public-access-plan). [10] Q. Shu, J. Gao, Z. Nawaz, Y. Liao, D. Wang, J. Wang, Appl. Energy 2010,
87, 25892596.
[11] B.-X. Peng, Q. Shu, J.-F. Wang, G.-R. Wang, D.-Z. Wang, M.-H. Han, Process
Conflict of Interest Saf. Environ. Prot. 2008, 86, 441447.
[12] A. F. Lee, J. A. Bennett, J. C. Manayil, K. Wilson, Chem. Soc. Rev. 2014, 43,
The authors declare no conflict of interest. 78877916.
[13] Y. Liu, E. Lotero, J. G. Goodwin, J. Catal. 2006, 242, 278286.
[14] D. E. Lopez, J. G. Goodwin, D. A. Bruce, E. Lotero, Appl. Catal., A 2005,
Keywords: biofuel conversion carbon catalysts 295, 97105.
esterification free fatty acids recycled tires [15] a) D. R. Anton, R. H. Crabtree, Organometallics 1983, 2, 855859; b) C. A.
Jaska, I. Manners, J. Am. Chem. Soc. 2004, 126, 97769785.
[16] a) J. Marchetti, V. Miguel, A. Errazu, Fuel Process. Technol. 2008, 89, 740
[1] M. Zhi, F. Yang, F. Meng, M. Li, A. Manivannan, N. Wu, ACS Sustainable 748; b) R. A. Arancon, H. R. Barros Jr, A. M. Balu, C. Vargas, R. Luque,
Chem. Eng. 2014, 2, 15921598. Green Chem. 2011, 13, 31623167; c) F. Guo, Z.-L. Xiu, Z.-X. Liang, Appl.
[2] a) M. Boota, M. P. Paranthaman, A. K. Naskar, Y. Li, K. Akato, Y. Gogotsi, Energy 2012, 98, 4752.
ChemSusChem 2015, 8, 35763581; b) A. K. Naskar, Z. Bi, Y. Li, S. K. Akato, [17] R. Liu, X. Wang, X. Zhao, P. Feng, Carbon 2008, 46, 16641669.
D. Saha, M. Chi, C. A. Bridges, M. P. Paranthaman, RSC Adv. 2014, 4, [18] a) N. zbay, N. Oktar, N. A. Tapan, Fuel 2008, 87, 17891798; b) I.
3821338221; c) Y. Li, M. P. Paranthaman, K. Akato, A. K. Naskar, A. M. Noshadi, R. K. Kumar, B. Kanjilal, R. Parnas, H. Liu, J. Li, F. Liu, Catal. Lett.
Levine, R. J. Lee, S.-O. Kim, J. Zhang, S. Dai, A. Manthiram, J. Power 2013, 143, 792797; c) I. K. Mbaraka, D. R. Radu, V. S.-Y. Lin, B. H. Shanks,
Sources 2016, 316, 232238; d) B.-C. Yu, J.-W. Jung, K. Park, J. B. J. Catal. 2003, 219, 329336.
Goodenough, Energy Environ. Sci. 2017, 10, 8690. [19] D. Y. Leung, X. Wu, M. Leung, Appl. Energy 2010, 87, 10831095.
[3] K. G. Reddie, K. S. Carroll, Curr. Opin. Chem. Biol. 2008, 12, 746754.
[4] a) M. Toda, A. Takagaki, M. Okamura, J. N. Kondo, S. Hayashi, K. Domen,
M. Hara, Nature 2005, 438, 178178; b) C. A. Deshmane, M. W. Wright, A. Submitted: April 24, 2017
Lachgar, M. Rohlfing, Z. Liu, J. Le, B. E. Hanson, Bioresour. Technol. 2013, Revised: June 6, 2017
147, 597604. Accepted: June 20, 2017

ChemistrySelect 2017, 2, 4975 4982 4982  2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

Anda mungkin juga menyukai