Sie sind auf Seite 1von 6

Molecular Catalysis 461 (2018) 80–85

Contents lists available at ScienceDirect

Molecular Catalysis
journal homepage: www.elsevier.com/locate/mcat

Catalytic properties of microporous zeolites in the catalytic cracking of m- T


diisopropylbenzene

Byung-Geon Parka, Kyong-Hwan Chungb,
a
Department of Food and Nutrition, Kwangju Women’s University, 165 Sanjung-dong, Gwangju, 62396, Republic of Korea
b
Department of Environmental Engineering, Sunchon National University, 255 Jungang-ro, Sunchon, Jeonnam, 57922, Republic of Korea

A R T I C LE I N FO A B S T R A C T

Keywords: The catalytic cracking of m-diisopropylbenzene, which is produced as a byproduct of the phenol production
Catalytic cracking process, was studied over various microporous zeolite catalysts. The zeolites had different acidity and pore
m-Diisopropylbenzene structures. The highest conversion of m-diisopropylbenzene in catalytic cracking was achieved on the MWW
MWW zeolite zeolite catalyst. BEA and FAU zeolite catalysts also showed considerably high conversion. The MWW zeolite
Cumene
exhibited a high selectivity for benzene and cumene compared to the other zeolites. The conversion of m-dii-
Benzene
sopropylbenzene on a MWW zeolite catalyst exceeded 90% at 400 °C. The yield of cumene plus benzene was
approximately70% in catalytic cracking over MWW zeolite. The deactivation of MWW zeolite catalysts by
carbon deposition on the catalyst could be neglected at high reaction temperatures, whereas it was quite serious
on the zeolite catalysts with the MOR structure. Benzene might be produced by a secondary reaction of inter-
mediates in the large pore channels during the retention time.

1. Introduction chloride catalysts [18]. In the process of cumene production, wasteful


byproducts are obtained, which consist of isomers of diisopro-
Aromatic compounds, such as benzene, toluene and xylene, are pylbenzene (DIPB), triisopropylbenzene (TIPB), butylbenzene, and
important raw materials in the petro-chemical and fine chemical in- heavy unknown aromatics [19].
dustry. On the other hand, their demand is not always met by the m-Diisopropylbenzene (m-DIPB) contains two isopropyl groups on
supply. The amount of benzene produced is insufficient to meet the benzene and is formed in higher quantities than other DIPB isomers and
demand, whereas the production of toluene exceeds its demand. To TIPB in the byproduct. DIPBs react with benzene and can be used as an
control the proportion between demand and supply, benzene from p- intermediate for the phenol production process. The byproduct can be
xylene is produced from selective toluene disproportionation [1–3] and used as a fuel oil in the industrial field without treatment, but it has a
the alkylation of aromatic compounds [4–6]. In other methods, benzene problem because aromatic compounds in the byproduct do not burn
is supplied from the catalytic cracking of aromatic byproducts formed completely.
in the naphtha cracking or phenol production process [7–9] The catalytic cracking behaviors of aromatic compounds depend on
Many processes have been developed as industrial syntheses for the the acidity and pore structure of the catalyst. The strength of acidity
production of phenol, but only the Hock process for cumene oxidation affects the cracking behavior and catalyst deactivation by the deposi-
[10,11] and the Dow process for toluene oxidation [12–14] are im- tion of polymerized high boiling hydrocarbons. The cracking behavior
portant industrially. The other processes have been rejected because of on zeolite catalysts, which have linear pore channels, exhibit high para-
their high cost. The cumene-phenol process is derived from the dis- selectivity, but serious catalyst deactivation can occur in the pore
covery of cumene hydroperoxide and its cleavage to acetone and phenol channels, which block the pores easily by high-boiling aromatic com-
[15,16]. Phenol is produced predominantly by this cumene-phenol pounds. On the other hand, MWW zeolite has a pillow-type structure
process in plants. The process consists of the cumene production pro- formed with two 10-membered rings, which are multidimensional
cess and phenol production process. Two byproducts were also pro- channels [20]. One of them is sinusoidal and the other containing large
duced during all processes, which reached approximately 10% against supercages is defined by the 12-membered rings with a 7.1 Å inner
the total amount of phenol production [17]. Cumene is produced from diameter and an 18.2 Å height [21]. In MWW zeolite, no high-boiling
benzene alkylation with propene on phosphoric acid or aluminum aromatic products were produced due to restrictions by the pore


Corresponding author.
E-mail address: chung-sea@hanmail.net (K.-H. Chung).

https://doi.org/10.1016/j.mcat.2018.10.008
Received 25 June 2018; Received in revised form 11 October 2018; Accepted 14 October 2018
2468-8231/ © 2018 Published by Elsevier B.V.
B.-G. Park, K.-H. Chung Molecular Catalysis 461 (2018) 80–85

entrance, so carbon deposition was difficult. Thus, MWW zeolite has compositions of the zeolites were analyzed by energy dispersive X-ray
been used in the reaction process using relatively large molecular re- (EDS; Norans, S-MAX 400) as an attachment to the scanning electron
actants, such as the alkylation of isopropy1naphtalene [22], skeletal microscope. The N2 isotherms were recorded using an automatic vo-
isomerization of 1-butene [23], and disproportionation of toluene [24]. lumetric adsorption apparatus (Miraei SI, Porosity-QX) at the liquid N2
In this study, the catalytic cracking behaviors of m-DIPB and the temperature. Evacuation of the samples was carried out at 130 °C for 1 h
byproduct were examined on various zeolite catalysts, which have before N2 adsorption.
different pore structures and acidity. Variation of the product dis- Gravimetric measurements of o-xylene adsorption-desorption were
tribution by the difference in pore structure of the zeolite catalysts was performed on a highly sensitive microbalance (quartz glass spring
also investigated based on the adsorption-desorption behavior of o-xy- balance). After the zeolite catalysts (0.1 g) were evacuated at 550 °C for
lene. The product distribution suggested a potential route for the re- 2 h under a pressure of less than 10−4 Torr, the measurements were
covery of profitable raw materials from the byproducts formed in the taken at 120 °C and at an o-xylene pressure of 10 Torr. The Brunauer-
isopropylation of benzene as a step in phenol production. Emmett-Teller (BET) surface area of the catalysts before and after the
cracking reaction was measured from the N2-adsorption isotherm at
2. Experimental liquid nitrogen temperature. NH3-temperature programmed desorption
(NH3-TPD) was performed according to the typical method. The cata-
2.1. Materials lysts were evacuated at 550 °C for 1 h and then exposed to NH3 gas flow
at room temperature for 30 min. The temperature was increased to
m-Diisopropylbenzene (m-DIPB, Aldrich, 96%) and byproducts were 550 °C at a constant heating rate of 10 °C/min under helium gas flow.
provided from Kum-Ho P&B Co. and used as reactants. MWW (MCM-22; The relative amount of ammonia desorbed from the catalysts was de-
Si/Al = 13) zeolite was prepared according to the methodology re- termined by TCD.
ported elsewhere [25]. A gel containing hexamethyleneimine was em-
ployed as the structure directing agent. The gel was crystallized in a
Teflon-lined autoclave at 150 °C for 9 days with stirring. Na-form zeo- 3. Results and discussions
lites were transformed to the H-form by ion exchange with 0.2 M
NH4NO3. Hydrothermally synthesized MFI (ZSM-5; Si/Al = 13) was 3.1. Physicochemical properties of the zeolites
applied after ion exchange into the H-forms. MOR (Si/Al = 10, Tosoh
Co.), FAU (Si/Al = 14, Tosoh Co.), and BEA (Si/Al = 13, PQ Co.) Fig. 1 presents XRD patterns of the zeolites. The characteristic peaks
zeolites were also introduced in the reaction. The Si/Al ratios of the of the zeolites were in accordance with those reported elsewhere [26].
catalysts were adjusted to 10 to 14 to exclude the influence of the Fig. 2 presents SEM images of the zeolites. The MWW zeolite was larger
different acidity. Table 1 lists the Si/Al molar ratios and physical than 10 μm, whereas that of BEA zeolite was less than 0.2 μm. The
properties of the zeolites. particle sizes of FAU and MFI were uniform at 0.5 μm. The type of N2-
isotherms of the zeolites is represented as a Langmuir isotherm, which
2.2. Catalytic cracking of m-DIPB was derived from a uniform micropore. The BET surface areas were
200–700 m2/g, as listed in Table 1.
The catalytic cracking reactions were performed in a fixed bed and Fig. 3 presents the NH3-TPD profiles of the zeolites. The profile of
continuous flow reactor at atmospheric pressure. After the catalyst was FAU zeolite revealed a single desorption profile. In contrast, the profiles
pretreated at 500 °C for 1 h, the reactant (m-DIPB) was fed from a mi- of the other zeolites were obtained as double profiles at 200 °C and
crofeeder with nitrogen gas as a carrier at the desired W (catalyst 300 °C–500 °C, respectively. The first profile appeared at approximately
weight) / F (flow rate of total gas) ratio and temperature. The eluted 200 °C due to the desorption of NH3 adsorbed physically on the surface
compounds were trapped by an ice bath and analyzed by gas chroma- of the zeolites. The second profile was assigned to the desorption of NH3
tography (GC, Donam Co., FID, column; DB-1, J&W Scientific Co.). on the acid sites of the zeolites [27]. The NH3-TPD profiles of MFI and
Conversion was defined as 1-(rate of m-DIPB recovery) / (rate of m- MOR zeolites indicated the presence of acid sites on the structure. In
DIPB feed). The yield was defined as the (rate of formation of the dis-
cussed compound) / (rate of m-DIPB feed).

2.3. Characterization of the zeolites


MWW
X-ray diffraction (XRD, D-MAXS-3000/PC, Rigaku) was carried out
using nickel filtered Cu Kα X-rays (40 kV, 40 mA) with a 2°/min scan
MFI
rate. The morphologies and particle sizes of the zeolites were estimated
Intensity (a. u.)

by scanning electron microscopy (SEM, Hitachi S-4800). The Si and Al

Table 1 MOR
Physical properties of the zeolite catalysts used in this work.
Zeolite Si/Al Pore BET Micropore Source
molar diameter (Å) Surface volumea (cm3/
ratio (-) area (m2/g) g)
FAU
MWW 13 5.6 × 5.6 420 0.15 synthesized
BEA 13 7.6 × 6.4, 690 0.19 PQ Co.
5.5 × 5.5
FAU 14 7.4 × 7.4 700 0.24 JRC-Z-HY5.5 BEA
MOR 10 6.5 × 7.0, 410 0.13 JRC-Z-HM10
2.6 × 5.7
5 10 15 20 25 30 35 40
MFI 14 5.3 × 5.6, 260 0.12 synthesized
5.1 × 5.5 2
a
determined from t-plot method. Fig. 1. XRD patterns of MWW, BEA, MOR, FAU, and MFI zeolites.

81
B.-G. Park, K.-H. Chung Molecular Catalysis 461 (2018) 80–85

Fig. 2. SEM images of MWW, BEA, MOR, FAU, and MFI zeolites.

contrast, the amount of NH3 desorbed at lower temperatures was larger Fig. 4 presents the adsorption behavior of o-xylene on MWW, BEA,
on the BEA and FAU zeolites. The acid strength can be evaluated by the MOR, FAU, and MFI zeolites. The behaviors indicate the diffusion of o-
maximum peak temperature (Tmax) of desorption peak relating to the xylene inside the pores of the zeolites. o-Xylene adsorbed rapidly on
activation energy for desorption of ammonia [28]. The acid strength of BEA and FAU, which have a wide pore space. The levels of o-xylene
MWW zeolite was similar to that of MFI zeolite. The order of the acid adsorption on MOR and MWW were smaller than those on the BEA and
strength was FAU≒BEA < MWW≒MFI < MOR. FAU zeolite catalysts. The adsorption of o-xylene on MFI zeolite, which

82
B.-G. Park, K.-H. Chung Molecular Catalysis 461 (2018) 80–85

BEA MWW
FAU MOR Tmax
MFI

TCD response (a. u.)


Tmax Tmax

0 200 400 600 0 200 400 600


Temperature (oC) Temperature (oC)
Fig. 3. NH3-TPD profiles from MWW, BEA, MOR, FAU, and MFI zeolites.

has a narrow pore entrance and bent pore structure, was slower than
250 BEA those of the other zeolites. This suggests that the stereo hindrance by
FAU different pore structures of the zeolites might be influenced in the re-
actions.
200

3.2. Influence of the pore structure of the zeolites in cracking


150
Conversion and selectivity are influenced by the acidity and pore
structures of the catalysts in the catalytic reaction using porous solid
100
-

acid catalysts. Therefore, the Si/Al molar ratios of the zeolite catalysts
MWW
were adjusted to10-14 to reduce the effect of acidity of the catalysts.
The catalytic properties of the zeolites might be influenced by the dif-
50 MFI ference in pore structure and acidities derived from different zeolite
MOR
structures. Fig. 5 shows the conversion and chemical composition of the
0 products from the catalytic cracking of m-DIPB on the zeolites. The
0 30 60 90 120 results differed remarkably according to the zeolite species. Conver-
sions on the MWW and BEA zeolite catalysts were as high as 80% at the
initial reaction. Deactivation of MWW and BEA zeolites occurred slowly
Fig. 4. Adsorption of o-xylene on MWW, BEA, MOR, FAU, and MFI zeolites.
as the reaction proceeded. In contrast, conversions on the MOR and
FAU zeolites were ca. 50%, and the conversions were reduced to below
20% after a 2 h process time. The conversion on MFI zeolite was ex-
cessively low.

m-DIPB p & o-DIPB Cumene Benzene Others

MWW
100
MFI FAU MOR BEA MWW

80 BEA
Catalysts

60 MOR

40
FAU

20
MFI

0
0 1 2 3 4 5 0 20 40 60 80 100

Time on stream (h) Composition (%)

Fig. 5. Changes in the conversion with time on stream (A) and chemical composition of outlet gas (B) in the dealkylation of m-DIPB on various zeolites at 350 °C W/
F = 5.90 g·h·mol−1 and loading amount of catalyst = 0.2 g. Product was collected for 60 min after 2 h of time on stream.

83
B.-G. Park, K.-H. Chung Molecular Catalysis 461 (2018) 80–85

Fig. 6. Changes in the conversion with time on stream (A) and chemical composition of outlet gas (B) in the dealkylation of m-DIPB on MWW zeolites. W/
F = 5.90 g·h·mol−1 and loading amount of catalyst = 0.2 g. Product was collected for 60 min after 2 h of time on stream.

m-DIPB p & o-DIPB Cumene Benzene Others zeolites because MOR zeolite has a linear type pore structure. Heavy
0.05 g
carbon deposition also occurred on the FAU zeolite because large mo-
lecular products might be generated in their supercage. Carbon de-
position on the catalysts led to the deactivation of the reaction.
Conversion on MFI zeolite was low due to the narrow pores inside the
MFI zeolite, which suppressed the diffusion of m-DIPB inside its pore. In
0.1 g contrast, the catalytic activities of BEA and MWW zeolite were main-
Loading amount

tained without deactivation because their pore structure hindered the


formation of an intermediate polymer. Their yield of the products was
also high. In particular, the yield of benzene increased on MWW zeolite.
0.2 g
Benzene might be formed due to cumene, which is generated by the
dealkylation of m-DIPB inside the pores.

3.3. Catalytic cracking of m-DIPB on MWW zeolite


0.3 g
Cumene, benzene, and p-DIPB were the main products in the m-
DIPB cracking reaction. Light hydrocarbons consisting of C3∼C5, tri-
isopropylbenzene (TIPB), and o-DIPB were produced in much smaller
0.4 g quantities. As shown in Fig. 5(A), high conversion was observed on
MWW, FAU, and BEA zeolite, which have a relatively large pore size
compared to MFI zeolite. The pore structure of MWW zeolite is com-
0 20 40 60 80 100 posed of a 10-oxygen member ring pore entrance and a 12-oxygen
Composition (%) member ring pore inside the channel [25]. Fig. 6 shows the variation of
conversion of m-DIPB and the chemical composition of the products in
Fig. 7. Chemical composition of products with the loading amount of MWW
the dealkylation of m-DIPB on MWW zeolite. The conversion of m-DIPB
zeolite in the dealkylation of m-DIPB at 350 °C. Product was collected for
increased and zeolite deactivation decreased with increasing reaction
30 min after 30 min of time on stream.
temperatures, as shown in Fig. 6(A). The intermediates of the aromatic
compounds are formed easily at high reaction temperatures. The pro-
The product distribution varies with the species of zeolite in the duct distribution also varied with the reaction temperature, as shown in
reaction. Fig. 5(B) shows the composition of products obtained in cat- Fig. 6(B). p-DIPB and o-DIPB were generated mainly by the iso-
alytic cracking. The product obtained from the isomerization and merization of m-DIPB at 250 °C. At 350 °C, the main product was cu-
cracking reaction was less than 10% on the FAU and MFI zeolite cat- mene formed by the dealkylation of m-DIPB. The production of benzene
alysts. o-DIPB and p-DIPB were generated mainly by the isomerization and light hydrocarbons increased with increasing reaction temperature.
of m-DIPB on MOR zeolite. In contrast, small amounts of benzene and Fig. 7 presents the chemical composition of products with the
cumene were produced by the dealkylation of m-DIPB. On the BEA and loading of MWW zeolite in the dealkylation of m-DIPB at 350 °C. The
MWW zeolites, the products of the dealkylation of m-DIPB were larger conversion of m-DIPB was improved and the yields of cumene plus
in quantity than those by the isomerization of m-DIPB. The yields of benzene were also enhanced as the catalyst loading was increased. In
benzene plus cumene were as high as ca. 50% on MWW zeolite com- particular, the conversion of m-DIPB exceeded 90%, and the yield of
pared to the other zeolites. benzene plus cumene was ca. 70% under a 0.4 g catalyst loading at
Carbon deposition inside the pores took place easily on the MOR 350 °C.

84
B.-G. Park, K.-H. Chung Molecular Catalysis 461 (2018) 80–85

4. Conclusion glassy carbon electrodes in aqueous solution and its interaction with ascorbic and
gallic acids, Electrochim. Acta 185 (2015) 1–5.
[12] S. Peta, T. Zhang, V. Dubovoy, K. Koh, T. Asefa, Facile synthesis of efficient and
The conversion of m-DIPB and product distribution were influenced selective Ti-containing mesoporous silica catalysts for tolueneoxidation, Mol. Catal.
by the pore structure of the zeolites in the catalytic cracking of m-DIPB. 444 (2018) 34–41.
MWW zeolite showed higher conversion than the other zeolites as well [13] H. Wang, Y. Lu, Y.X. Han, C. Lu, H. Wan, Z. Xu, S. Zheng, Enhanced catalytic to-
luene oxidation by interaction between copper oxide and manganese oxide in Cu-O-
as longer catalytic activity with slow deactivation. The selectivity of the Mn/γ-Al2O3 catalysts, Appl. Surf. Sci. 420 (2017) 260–266.
MWW zeolite for benzene and cumene was higher than those of the [14] Y. Zhang, H. Zhang, Y. Yan, Metal-organic chemical vapor deposition of Cu(acac)2
other zeolites. The pillow-type pore structure of MWW zeolite could for the synthesis of Cu/ZSM-5 catalysts for the oxidation of toluene, Microporous
Mesoporous Mater. 261 (2018) 244–251.
suppress secondary reactions, such as isomerization, which led to high [15] D.H.R. Barton, N.C. Delanghe, New catalysts for the conversion of cumene hydro-
yields of cumene and benzene. The catalytic behavior of MWW zeolite peroxide into phenol, Tetrahedron Lett. 38 (1997) 6351–6354.
in the catalytic cracking of m-DIPB was superior to that of the other [16] S.-H. Wu, Runaway reaction and thermal explosion evaluation of cumene hydro-
peroxide (CHP) in the oxidation process, Thermochim. Acta 559 (2013) 92–97.
zeolites.
[17] J.P. Fortuin, H.I. Waterman, Production of phenol from cumene, Chem. Eng. Sci. 2
(1953) 182–192.
References [18] J. Zhai, Y. Liu, L. Li, Y. Zhu, W. Zhong, L. Sun, Applications of dividing wall column
technology to industrial-scale cumene production, Chem. Eng. Res. Des. 102 (2015)
138–149.
[1] S. Suganuma, K. Nakamura, A. Okuda, N. Katada, Enhancement of catalytic activity [19] R.J. Schmidt, Industrial catalytic processes-phenol production, Appl. Catal. A: Gen.
for toluene disproportionation by loading Lewis acidic nickel species on ZSM-5 280 (2005) 89–103.
zeolite, Mol. Catal. 435 (2017) 110–117. [20] X. Du, Y. Yang, D. Shi, C. Lin, Y. Qiu, J. Sun, The construction of a series of hier-
[2] M. Kubů, N. Žilková, S.I. Zones, C.-Y. Chen, S. Al-Khattaf, J. Čejka, Three-dimen- archical MWW-type zeolites and their catalytic performances for bulky aldol con-
sional 10-ring zeolites: the activities in toluene alkylation and disproportionation, densation, Microporous Mesoporous Mater. 268 (2018) 117–124.
Catal. Today 259 (2016) 97–106. [21] Y. Shi, E. Xing, W. Xie, F. Zhang, X. Mu, X. Shu, Directing gel: an effective method
[3] D. Mitsuyoshi, K. Kuroiwa, Y. Kataoka, T. Nakagawa, M. Kosaka, K. Nakamura, tailoring morphology of MWW zeolites and their catalytic performance in liquid-
S. Suganuma, Y. Araki, N. Katada, Shape selectivity in toluene disproportionation phase alkylation of benzene with ethylene, Microporous Mesoporous Mater. 215
into para-xylene generated by chemical vapor deposition of tetramethoxysilane on (2015) 8–19.
MFI zeolite catalyst, Microporous Mesoporous Mater. 242 (2017) 118–126. [22] N.S. Nesterenko, S.E. Timoshin, A.S. Kuznetsov, V. Montouillout, F. Thibault-
[4] S. Peta, T. Zhang, V. Dubovoy, K. Koh, M. Hu, X. Wang, T. Asefa, Template-free Starzyck, C. Fernande, J.P. Gilson, F. Fajula, I.I. Ivanova, Transalkylation of 1,4-
synthesis of highly selective amorphous aluminosilicate catalyst for toluene alky- diisopropylbenzene with naphthalene over dealuminated mordenites, Stud. Surf.
lation, Appl. Catal. A: Gen. 556 (2018) 155–159. Sci. Catal. 154 (2004) 2163–2168.
[5] P.A. Gushchin, I.M. Kolesnikov, V.A. Vinokurov, E.V. Ivanov, V.A. Lyubimenko, [23] Yu.P. Khitev, I.I. Ivanova, Yu.G. Kolyagin, O.A. Ponomareva, Skeletal isomerization
V.N. Borshch, Alkylation of benzene with ethylene in the presence of dimethyldi- of 1-butene over micro/mesoporous materials based on FER zeolite, Appl. Catal. A:
chlorosilane, J. Catal. 352 (2017) 75–82. Gen. 441-442 (2012) 124–135.
[6] Z. Lei, L. Liu, C. Dai, Insight into the reaction mechanism and charge transfer [24] Q. Shi, J.C. Gonçalves, A.F.P. Ferreira, M.G. Plaza, A.E. Rodrigues, Xylene iso-
analysis for the alkylation of benzene with propylene over H-beta zeolite, Mol. merization side reactions over Beta zeolite: disproportionation and transalkylation
Catal. 454 (2018) 1–11. of C8 aromatics and toluene, Appl. Catal. A Gen. 562 (2018) 198–205.
[7] Y. Wei, J. Hong, W. Ji, Thermal characterization and pyrolysis of digestate for [25] M. Zhu, Y. Liu, Y. Yao, J. Jiang, F. Zhang, Z. Yang, Z. Lu, I. Kumakiri, X. Chen,
phenol production, Fuel 232 (2018) 141–146. H. Kita, Preparation and catalytic performance of Ti-MWW zeolite membrane for
[8] L.R. Cappellari, J. Chiappero, M.V. Santoro, W. Giordano, E. Banchio, Inducing phenol hydroxylation, Microporous Mesoporous Mater. 268 (2018) 84–87.
phenolic production and volatile organic compounds emission by inoculating [26] M.M.J. Treacy, J.B. Higgins, R. von Ballmoos, Collection of Simulated XRD Powder
Mentha piperita with plant growth-promoting rhizobacteria, Sci. Hortic. 220 (2017) Patterns for Zeolites, Elsevier, 1996, pp. 392–530.
193–198. [27] N. Katada, H. Igi, J.-H. Kim, M. Niwa, Determination of the acidic properties of
[9] B. Pang, S. Yang, W. Fang, T.-Q. Yuan, D.S. Argyropoulos, R.-C. Sun, Structure- zeolite by theoretical analysis of temperature-programmed desorption of ammonia
property relationships for technical lignins for the production of lignin-phenol- based on adsorption equilibrium, J. Phys. Chem. B 101 (1997) 5969–5977.
formaldehyde resins, Ind. Crops Prod. 108 (2017) 316–326. [28] Y. Miyamoto, N. Katada, M. Niwa, Acidity of β zeolite with different Si/Al2 ratio as
[10] C. Mu, Y. Cao, H. Wang, H. Yu, F. Peng, A kinetics study on cumeneoxidation measured by temperature programmed desorption of ammonia, Microporous
catalyzed by carbon nanotubes: effect of N-doping, Chem. Eng. Sci. 177 (2018) Mesoporous Mater. 40 (2000) 271–281.
391–398.
[11] R. Estévez, J.M.R. Mellado, M. Mayén, Oxidation of cumene hydroperoxide on

85

Das könnte Ihnen auch gefallen