Download as pdf or txt
Download as pdf or txt
You are on page 1of 40

Journal Pre-proof

Catalysts for glycerol hydrogenolysis to 1,3-propanediol: a review of


chemical routes and market

Alisson Dias da Silva Ruy (Conceptualization) (Investigation)


(Methodology) (Writing - original draft) (Visualization), Rita Maria de
Brito Alves (Supervision) (Writing - review and editing), Thiago
Lewis Reis Hewer (Investigation) (Methodology) (Writing - original
draft), Danilo de Aguiar Pontes (Writing - original draft) (Writing -
review and editing), Leonardo Sena Gomes Teixeira (Investigation)
(Supervision), Luiz Antônio Magalhães Pontes (Conceptualization)
(Methodology) (Writing - original draft) (Supervision) (Writing -
review and editing)

PII: S0920-5861(20)30421-1
DOI: https://doi.org/10.1016/j.cattod.2020.06.035
Reference: CATTOD 12953

To appear in: Catalysis Today

Received Date: 30 March 2020


Revised Date: 30 May 2020
Accepted Date: 10 June 2020

Please cite this article as: da Silva Ruy AD, de Brito Alves RM, Reis Hewer TL, de Aguiar
Pontes D, Gomes Teixeira LS, Magalhães Pontes LA, Catalysts for glycerol hydrogenolysis to
1,3-propanediol: a review of chemical routes and market, Catalysis Today (2020),
doi: https://doi.org/10.1016/j.cattod.2020.06.035
This is a PDF file of an article that has undergone enhancements after acceptance, such as
the addition of a cover page and metadata, and formatting for readability, but it is not yet the
definitive version of record. This version will undergo additional copyediting, typesetting and
review before it is published in its final form, but we are providing this version to give early
visibility of the article. Please note that, during the production process, errors may be
discovered which could affect the content, and all legal disclaimers that apply to the journal
pertain.

© 2020 Published by Elsevier.


Catalysts for glycerol hydrogenolysis to 1,3-propanediol: a review of chemical
routes and market

Alisson Dias da Silva Ruyab; Rita Maria de Brito Alvesb; Thiago Lewis Reis Hewerb;
Danilo de Aguiar Pontesa; Leonardo Sena Gomes Teixeiraa; Luiz Antônio Magalhães
Pontes1a.
a
Universidade Federal da Bahia. Escola Politécnica. Departamento de Engenharia
Química. Rua Prof. Aristides Novis, nº 02, 2nd floor, Federação, 40210-630, Salvador,
Bahia, Brazil.
b
Universidade de São Paulo. Escola Politécnica. Departamento de Engenharia Química.

of
Av. Prof. Luciano Gualberto, nº 380, Tv. 3, 05508-010, São Paulo – SP, Brazil.

ro
Graphical abstract

-p
re
lP
na
ur

Highlights:
Jo

• 1,3-propanediol as a sustainable product obtained from biodiesel-derived glycerol


hydrogenolysis;
• Market and industrial routes for the production of 1,3-propanediol;
• Analysis of metallic and acid sites for the production of 1,3-propanediol;

1
Corresponding author at: Federal University of Bahia. Aristides Novis, nº 02, 2nd
floor, Federação, 40210-630, Salvador, Bahia, Brazil
E-mail address: uolpontes@uol.com.br
2

• Influence of Lewis/Brønsted acid ratio in glycerol conversion and 1,3-propanediol


selectivity.

ABSTRACT

1,3-propanediol (1,3-PDO) is obtained from renewable and environmentally friendly


glycerol. The current industrial production uses high cost genetically modified
microorganisms. Researchers have studied heterogeneous catalysts for more efficient
processes leading to higher competitiveness in the 1,3-PDO market. In this context, a
review of studies involving chemical routes for its production was performed,

of
evaluating process variables and, in particular, the influence of active acid and metallic
phases on the activity and selectivity to the desired product. Platinum, iridium, and

ro
copper were verified to be the most promising metals. Brønsted sites are responsible for
the higher selectivity to 1,3-PDO, while the reaction rate strongly depends on Lewis
-p
sites since glycerol adsorption takes place in these sites. Moreover, in order to decrease
operating costs, important parameters such as temperature, glycerol concentration in the
re
feed stream and the reactor type must be optimized.
lP

Keywords: glycerol, hydrogenolysis, 1,3-propanediol, catalytic routes, market


na
ur
Jo
1

1. INTRODUCTION

Biodiesel is a renewable and environmentally friendly alternative to fossil fuels,


since it is biodegradable, non-toxic, and sulfur and aromatic compound free. Besides,
biodiesel emits fewer pollutants such as greenhouse gases into the air. Its production
takes place by the transesterification reaction of vegetable oils and animal fat and
generates glycerol as a byproduct [1,2].
Glycerol offers a wide range of opportunities based on its transformation for
application in high added-value molecules synthesis due to its availability in the market,
which increases the biodiesel industry competitiveness [3]. Several studies use glycerol
in the production of building block molecules, such as acrylic acid, lactic acid,

of
acrylonitrile, 1,3-propanediol, 1,2-propanediol, glyceric acid, among others [4–6].
1,3-propanediol (1,3-PDO) is a three-carbon dialcohol and an important commercial

ro
chemical used as intermediates or solvents in pharmaceutical, food and textile
industries. It is also used in several applications, for example, consumer products such
-p
as cleaning products, cosmetics and personal care, as well as a fluid in heat exchangers
and coating. Its market value sits around US$ 4.000/t [7]. Polytrimethylene terephthalate
re
(PTT) and polyurethane (PU) are the most dominant applications.
Researchers have thoroughly studied new heterogeneous catalysts [8,9] and
lP

genetically modified microorganisms to produce 1,3-PDO from glycerol [7,10,11]. The


motivation lies in the fact that incentives from several countries to the biodiesel added
to diesel have led to a glycerol surplus in the world market.
na

In 2024, glycerol production is expected to reach 29% more than the largest
commercialized volume in 2018, 3.6 million tons [12]. The product commercial
competitiveness and the survival of existing plants depend on the economic and
ur

technological optimization of existing processes. The high price of 1,3-PDO is also


explained by the high costs of its separation and purification processes. The high boiling
Jo

point and high hydrophilicity are the main factors for its expensiveness [13]. Several
studies have been performed to develop new efficient separation methods, such as
reactive extraction [14], extraction by aqueous two-phase systems (ATPS) [13],
distillation, pervaporation and their combination [15].
Considering the high availability of glycerol and great economic potential of its use
as feedstock in the production of 1,3-PDO, this paper evaluates the current 1,3-PDO
2

market and its growth perspectives within the next few years, the chemical production
routes, considering heterogeneous catalyst types, their functionalities and operating
conditions, aiming at designing a new economically competitive industrial process.

2. MARKET AND INDUSTRIAL ROUTES FOR 1,3-PDO PRODUCTION

Currently 1,3-PDO is used as a monomer in the formation of polymers and directly


in end-use products (Fig. 1) and presents a good growth perspective.

Polytrimethylene Terephthalate
6% 12% Polyuretane
30%

of
Cosmetics & Personal Care
9%
Household
8%

ro
Engine Coolants

14% 21% Heat Transfer Fluid

-p
Others

Fig. 1. Global market of 1,3-PDO by application [16]


re
Polymers represent more than 50% of 1,3-PDO applications, representing most of the
market [16,17]. PTT represents 30% of all the current applications of 1,3-PDO,
lP

followed by polyurethane [16]. PTT could be a substitute to petrochemical-based


polymers, such as polyethylene terephthalate (PET) and polybutylene terephthalate
(PBT), due to its advantageous physical-chemical properties, such as elasticity, easiness
na

of coloring, and resilience, which open a new market for obtaining products used in the
textile and engineering plastics industries [18]. PTT is applied in the production of
ur

textiles for clothing, carpets, and upholstery [19]. Polyurethane is applied in the
production of elastomers, foam, and plasterwork [20].
1,3-PDO is also used for manufacturing cosmetics & personal care, hygiene and
Jo

domestic cleaning, and heat transfer fluid [16]. In lower amounts, 1,3-PDO can be used
as an additive to foodstuffs, pharmaceuticals, fragrances, resins, and insect repellants
[19].
An optimistic market expectation for 1,3-PDO (Fig. 2) is of US$ 1,442.77 million in
2027, with a CAGR (Compound Annual Growth Rate) of 14.2% as of 2018, when
revenues reached US$ 437 million [21].
3

of
ro
Fig. 2. Global market of 1,3-PDO by region [22]

The market encompasses three main regions: the Americas, Asia-Pacific, Europe,
-p
Middle East and Africa. In 2018, the Americas represented the main 1,3-PDO market,
corresponding to 38%, followed by Asia-Pacific, with 33%, and Europe, Middle East
re
and Africa totalizing 29% (Fig. 2). The expectation is that the Americas reach 40% in
2027, consolidating their leadership [22]. North America dominated the regional market
lP

in terms of both demand and market . The United States (USA) leads the current
production and consumption of 1,3-PDO and its derivatives, with DuPont and Shell
having plants within their territory. Asia, China and India have developed new
na

technologies for 1,3-PDO synthesis, having a large consumer market. China has several
producing companies that use biological routes [23]. A great consumption growth is
expected for the whole region, particularly in applications for cosmetics, personal care,
ur

and cleaning products [16]. In Europe, countries such as France, United Kingdom (UK)
and Germany have increased the 1,3-PDO applications for the automotive industry,
Jo

cosmetics and personal care [22]. A new production unit is expected to start up in
France, in 2020, driving the European market forward.
Emerging markets such as Asia-Pacific and Central and South America are expected
to have the highest growth rates over the forecast period. Emerging economies, such as
China, India, Brazil, and Argentina, are also expected to lead the growth of the regional
market. Countries such as China, Brazil, Russia, and Argentina are taking important
4

measures to reduce their dependence on fossil fuels and to promote the production of
biodiesel [24].
The main 1,3-PDO producers are DuPont Tate & Lyle, and Shell. The DuPont
process uses genetically modified Escherichia coli (E. coli) and corn syrup as feedstock
[25,26]. The product is then commercialized under the trademarks Susterra™, and
Zemea™, with varying degrees of purity according to their application. Susterra™ is a
1,3-PDO industrial grade, which is copolymerized with terephthalic acid to produce
polytrimethylene terephthalate (PTT), under the trade name Sorona® [27]. Zemea™ is
high purity 1,3-PDO, used in cosmetics, personal care, food, flavors, laundry, cleaning
products, and pharmaceuticals [28]. Shell produces and commercializes 1,3-PDO and,

of
particularly, PTT, with the trade name Corterra®. The Shell Process uses a
heterogenous catalytic route [29], and its first plant is located in Geismar, Louisiana.
Several new companies have developed technologies mainly based on

ro
biotechnological processes. Metabolic-Explorer, METEX, is building a new plant in
France [30]. In Asia, there is a wide proliferation of new plants already in operation:
-p
Zhangjiagang Glory; Shangdong Mingxing; Chenneng; Henan Tianguan; and Shanghai
Demao [31].
re
In this scenario of wide spreading, its use as a polymer and market growth, 1,3-PDO
may make the transition from a traditional specialty chemical to a commodity chemical,
lP

with a significant reduction in production costs. However, its industrial-scale production


is still costly and complex. The Shell process requires high energy consumption, besides
the high catalyst costs. The biochemical route takes place in a bioreactor fermentation
na

step at mild conditions. Nevertheless, as disadvantages, this process presents high


enzyme costs, with more complex mechanisms due to the growth of microorganisms
and uses more expensive equipment than the petrochemical route.
ur

The biological conversion of glycerol into 1,3-PDO can be achieved through two
distinct routes. The first one is the DuPont Tate & Lyle BioProducts process, which
Jo

uses a recombinant strain of E. coli from glucose. The commercial process is protected
by company-owned patents [32,33]. The second route stems from glycerol using natural
microorganisms, mutagenesis of strains or genetically modified strains [7].
The biological process of 1,3-PDO production is commercialized by DuPont Tate &
Lyle BioProducts, which operates a plant in London, Tennessee (USA), with capacity of
63.5 thousand tons/year. The E. coli strains convert glycerol into 1,3-PDO, using
glucose as a primary fermentation substrate. First, glucose from corn syrup is used to
5

produce glycerol. Then, the same microorganism converts glycerol into 3-HPA, which
is then reduced to 1,3-PDO [27,29]. The formation of byproducts such as butyrate,
ethanol, and ethyl acetate during the glycolysis reaction, the concentration of glycerol,
and the product itself may inhibit this process, reducing 1,3-PDO yield [34–36].
The chemical synthesis of 1,3-PDO is performed by the Degussa and Shell
petrochemical processes in two steps and produces 3-hydroxypropionaldehyde (3-HPA)
as an intermediate. The process developed by Degussa was acquired by Dupont in 1998,
and discontinued afterwards [17]. The process uses acrolein as feedstock, and occurs in
two steps. First, acrolein is hydrated in an aqueous solution at temperature ranging
between 50-70 ºC, in order to form 3-HPA, which is hydrogenated over nickel catalysts

of
at pressure of 135 bar, and temperature of 75-145 ºC [29,37]. The Shell process uses
ethylene oxide and syngas, via the hydroformylation reaction, to form 3-HPA, which is
hydrogenated into 1,3-PDO in the presence of ruthenium, rhodium, or cobalt based

ro
catalysts at 100 atm and 100-150 ºC [29,38].
The literature presents several studies of new more active and selective
-p
heterogeneous catalysts for producing 1,3-PDO from glycerol hydrogenolysis. These
catalysts have a metallic function, responsible for hydrogenation, and an acid function,
re
which promotes selective dehydration.

3. HETEROGENOUS CATALYSTS FOR THE HYDROGENOLYSIS OF


lP

GLYCEROL INTO 1,3-PDO

The hydrogenolysis of glycerol is a catalytic chemical reaction that involves breaking


na

a carbon-oxygen bond with simultaneous addition of a hydrogen atom. Firstly, the


secondary hydroxyl from glycerol is dehydrated into 3-HPA on acid sites, which is then
hydrogenated on metallic sites into 1,3-PDO, according to Fig. 3 [39,40].
ur
Jo
6

Fig. 3. General mechanism for the glycerol hydrogenolysis reaction. Adapted from Samudrala,
Kandasamy and Bhattacharya [8]

of
The technological challenge for optimizing the glycerol hydrogenolysis reaction into
1,3-PDO lies in the study of selective and controlled cleavage of the C-O bond to

ro
remove the secondary hydroxyl, bonded to the central carbon [41]. The secondary
hydroxyl reactivity is sterically reduced due to the presence of the two primary
-p
hydroxyls, making it less accessible to the catalyst active sites. For this reason, the
removal of the secondary hydroxyl is kinetically and thermodynamically less favorable.
re
Besides, the activation energy for dehydrating hydroxyls is similar, which hinders
selectively hydrolyzing the secondary hydroxyl [42,43].
Usually, the hydrogenolysis of glycerol to form 1,3-PDO occurs on Brønsted active
lP

sites at high hydrogen pressures (Route A). Although high temperatures may increase
the conversion of glycerol into 3-HPA, they also increase the production of acrolein and
na

favor the hydrogenolysis of 1,3-PDO to monoalcohols. The reaction must be performed


under mild conditions, generally below 200 ºC, to inhibit the subsequent dehydration of
1,3-PDO [41]. However, the reaction can also occur on Lewis acid sites, leading to the
ur

formation of 1,2-propanediol (1,2-PDO) (Route B). The dehydration of glycerol


produces hydroxyacetone, which undergoes hydrogenation to produce 1,2-PDO.
Jo

Hydroxyacetone is thermodynamically stabler than 3-HPA, which could favor the


occurrence of Route B [44].
The search for a promising catalyst to convert glycerol into 1,3-PDO has led to the
development of a series of bifunctional catalysts based on the acid/metal combination,
with a reasonable selectivity to the desired product [44,45]. In order to obtain high 1,3-
PDO selectivity, an appropriate metallic site combination (Pt, Ir, and Rh) and Brønsted
acids (MoOx, ReOx, and WOx) aiming at selectively activating the secondary glycerol
7

hydroxyl is necessary. Among the metallic components, platinum is the most widely
used on several supports, such as SiO2, ZrO2, Al2O3 [46]. However, the total acid sites
and the presence of Lewis sites are important for higher glycerol adsorption and
conversion.
The metallic sites activate the hydrogen molecule, while the acid sites strongly
interact with glycerol, considering that metallic oxides have affinities for trialcohol
hydroxyls [42,45]. Among the Brønsted acids, ReOx and WOx oxides demonstrate to
exert the best effect on the selective conversion of glycerol into 1,3-PDO [43,47].
From the literature analysis, three tables containing catalyst groups with their
characteristics were obtained, taking into account the active metallic site: platinum

of
(Table 1), iridium and rhenium (Table 2), and other metals (Table 3). The information
regarding the process conditions presented in the tables is in the original units according
to the authors. Conversion, selectivity, and yield are presented in % molar. The defined

ro
catalyst was that of the best results in each publication.
Table 1 presents an evaluation of 43 articles from the literature considering studies
-p
with catalysts containing platinum as the main active metallic phase, supported on
several kinds of oxides, mainly SiO2, WO3, ZrO2, and Al2O3.
re
Table 1.
Hydrogenolysis of glycerol into 1,3-PDO. Active metallic phase Pt, Pt-WOx; Pt (Au, Nb, and others)
lP

X S R
Catalysts Conditions Reactor References
(%) (%) (%)
na

180°C, 8MPa, 18h, 4 g


1 Pt/ZrW Batch 10 31 3 [9]
glycerol / 36 g H2O
180°C, 5.5MPa, 12h,
2 Pt/m-WO3 Batch 18 39 7 [48]
10wt% glycerol
ur

Pt/5WO3/10TiO2/ 453K, 5.5MPa, 12h,


3 Batch 15 51 8 [46]
SiO2 10wt% glycerol
140°C, 0.5MPa, 12h, 10
Jo

4 Pt/WOx/AlOOH mmol glycerol / 30 mL Batch 37 21 8 [49]


H2O
Pt-
483K, 6MPa, 50h, 10wt%
5 20%WOx/SAPO- Batch 48 19 9 [50]
glycerol
34a
483K, 50h, 3 g glycerol /
6 Pt-10WO3/SiO2 Batch 43 26 11 [51]
27 g H2O

180°C, 5MPa, 0.09h-1,


7 Pt-HSiW/ZrO2 Fixed bed 24 48 12 [52]
10wt% glycerol
8

160°C, 1MPa, 12h, 10wt%


8 Pt/0.1%Nb‐ WOx Batch 45 28 12 [53]
glycerol

443K, 8MPa, 18h, 3 mmol


9 Pt/WO3/ZrO2 Batch 56 24 13 [54]
glycerol / 0.2 mL DMI

170°C, 5.5MPa, 12h,


10 Pt/WO3/ZrO2 Batch 46 29 13 [55]
10wt% glycerol
200°C, 40bar, 18h, 8.3wt%
11 Pt/Al2O3 Batch 49 28 14 [56]
glycerol
260°C, 0.14h-1, 20wt%
12 Pt/WO3/Al2O3 Fixed bed 99 14 14 [57]
glycerol
155°C, 5MPa, 7.5h, 5wt%
13 Pt/Au/WO3 Batch 31 54 17 [58]
glycerol
14 Pt/WOx 413K, 1MPa, 12h Batch 60 36 22 [59]

of
200°C, 3MPa, 6h, 1.74 M
15 Pt/WO3/ZrO2 Batch 65 36 22 [60]
glycerol / 5 mL H2O

ro
PtNPs- 473K, 4MPa, 15h, 0.1
16 Batch 61 33 23 [61]
HSiW/mAl2O3 M/15 mL H2O
180°C, 8MPa, 24h, 4 g
17 Pt/WO3/ZrO2 Batch 78 30 23 [62]
glycerol / 36 g H2O

18 Pt–LiSiW/ZrO2
180°C, 5MPa, 0.09h-1,
10wt% glycerol
-p
Fixed bed 44 54 23 [63]
re
2Pt/WO3-Al2O3-
19 160°C, 6MPa, 12h, 3wt% Batch 48 56 27 [64]
SiO2
lP

180°C, 1MPa, 16h, 47 mg


20 Pt/WOx/Al2O3 Batch 98 28 27 [65]
Glycerol / 2 mL H2O

220°C, 45bar, 24h, 5wt%


21 Pt/WOx/Al2O3 Batch 53 52 28 [66]
glycerol
na

180°C, 5MPa, 1h-1, 10wt%


22 5PtW/ZrSi Fixed bed 54 52 28 [67]
glycerol
180°C, 5MPa, 1h-1, 50wt%
23 6Pt/12.9W/Al2O3 Fixed bed 80 35 28 [68]
glycerol
ur

180°C, 8MPa, 18h, 4 g


24 Pt/ZrW38Mn3 Batch 56 51 28 [42]
glycerol / 36 g H2O
200°C, 90bar, 4h, 20wt%
Jo

25 9Pt/8WOx/Al2O3 Batch 80 39 31 [69]


glycerol
200°C, 6MPa, 0.045h-1,
26 Pt-15HSiW/SiO2 Fixed bed 81 39 31 [70]
10wt% glycerol
130°C, 4MPa, 60wt%
27 3Pt/WO3/10ZrO2 Fixed bed 70 46 32 [71]
glycerol
2Pt/20(W+Al)- 160°C, 6MPa, 12h, 3wt%
28 Batch 66 50 33 [72]
SBA-15b glycerol
9

9Pt-8WO3/γ- 200°C, 25bar, 16h, 5wt%


29 Batch 59 56 33 [73]
Al2O3 glycerol

453K, 5MPa, 15h, 3 mmol


30 Pt/WO3/Al2O3 Batch 54 59 33 [74]
Glycerol / 9 mL H2O

260°C, 0.1MPa, 2.09h-1,


31 2Pt/AlPO4 Fixed bed 100 35 35 [75]
10wt% glycerol
260°C, 0.1MPa, 1.02h-1,
32 2Pt/AlPO4 Fixed bed 100 35 35 [76]
10wt% glycerol

2Pt-10WO3/SBA- 210°C, 0.1MPa, 1.02h-1,


33 Fixed bed 86 42 36 [77]
15b 10wt% glycerol

453K, 3MPa, 10h, 1 mmol


34 PtAlOx/WO3 Batch 90 44 40 [78]

of
glycerol / 3 mL H2O

413K, 1MPa, 12h, 5wt%


35 0,10 AuPt/WOx Batch 81 52 42 [79]
glycerol

ro
160°C, 5MPa, 0.09h-1,
36 Pt–10WOx/Al2O3 Fixed bed 64 66 42 [80]
10wt% glycerol
Pt–
150°C, 4MPa, 0.2h-1, -p
37 Li2B4O7/WOx/Zr Fixed bed 91 50 45 [81]
40wt% glycerol
O2

225°C, 0.1MPa, 1.02h-1,


2Pt/H−HMc
re
38 Fixed bed 95 49 46 [82]
10wt% glycerol

413K, 8MPa, 24h, 30wt%


39 Pt–WOx/t-ZrO2 Batch 78 65 49 [83]
glycerol
lP

215°C, 0.1MPa, 1.02h-1,


40 Pt-5Cu/ H−HMc Fixed bed 90 59 53 [84]
10wt% glycerol
na

41 Pt-Sulfated ZrO2 170°C, 7.3MPa, 24h Batch 67 84 56 [85]

200°C, 0.1MPa, 1.02h-1,


42 Pt/S-MMTd Fixed bed 94 62 58 [8]
10wt% glycerol
ur

453K, 5MPa, 12h, 1 mmol


43 Pt/WOx/AlOOH Batch 100 66 66 [86]
Glycerol / 3 mL H2O
Jo

a
SAPO: silicoaluminophosphate; bSBA: mesoporous silica; cHM: modernite; MMT: montmorillonite;
d
NPs: Nanoparticles
Considering the process conditions and yield to 1,3-PDO over different catalysts, in
Table 1, important points of interest may be analyzed aiming at choosing a catalytic
process that can be developed for industrial applications. Most laboratory-scale
experiments are carried out using glycerol diluted in water solutions, at temperatures in
the range 273 K - 473 K, and high hydrogen pressure. High temperatures (> 453 K)
10

lead to consecutive reactions and low selectivity to 1,3-PDO. Glycerol dilution is


important to avoid its degradation with the formation of oligomers at high temperatures.
High hydrogen pressures keep the system in liquid phase and favors the glycerol
hydrogenolysis reaction, with the preferential formation of 1,3-PDO. Molecular
hydrogen is dissociated on platinum and undergoes spillover onto the support surface
containing Lewis acid sites where glycerol is adsorbed. The aqueous environment leads
to the formation of protons, which act as Brønsted acid sites on the active support
surface.

One must have in mind that the physical-chemical characterizations of the catalysts
listed in Table 1, such as metallic dispersion, interaction among metals, surface area,

of
crystallinity, strength and nature of the active support acid sites impact glycerol
conversion and yield to 1,3-PDO. However, many characterizations are performed ex-

ro
situ in dissimilar environments to reactional conditions. Particularly in aqueous phase
high-pressure glycerol conversion studies, attention must be paid to changes that may
-p
occur to the support crystalline form and catalyst acidic activity. Studies about glycerol
conversion on platinum catalysts supported on alumina or niobia report the role of the
Lewis and Brønsted acid sites using transmission FTIR spectroscopy with the aid of
re
density functional theory (DFT) calculations. It is worth noting that, in aqueous media,
Brønsted sites are formed in substitution to Lewis sites, and that the aluminum oxide
lP

crystalline phase undergoes transformations with the formation of boehmite according


to the process conditions [87,88].
na

The studies listed in Table 1 show that platinum was introduced together with
tungsten oxide (WOx), being present in most bifunctional catalysts used in the selective
glycerol hydrogenolysis. The Brønsted acid sites, present in WOx, play a key role in the
ur

selective glycerol hydrogenolysis. The number of active sites must be in a defined


proportion with the number of active metallic sites aiming at maximizing the selectivity
Jo

to 1,3-PDO. Acid strength is another important factor, because strong acid sites can lead
to consecutive reactions with the formation of alcohols, which decreases the product
selectivity [46,54].
WOx has been used as an active phase and as a support, despite usually having a
small surface area. According to Table 1, among the studies using WO3, only
PtAlOx/WO3 and AuPt/WOx catalysts present reasonable yields to 1,3-PDO, with 40%
and 42%, respectively [78,79].
11

Aluminum, primarily in the form of Al2O3, has a larger surface area than WO3, and
acts as an “anchor” for fixating glycerol, due to the presence of hydroxyl groups on its
surface, being one of the best materials for catalyzing the hydrogenolysis reaction, and
leading to good yields to 1,3-PDO, reaching 40% with Pt-AlOx/WO3 [78], 42.4% with
Pt-WOx/Al2O3 [80], and 66% with Pt/WOx/AlOOH [86]. In a review study [44],
catalysts based on the combination of Pt, WO3, and Al2O3 were found to be the most
effective for glycerol hydrogenolysis into 1,3-PDO. These catalysts remain the most
promising in the evaluation performed in this review. It has also been verified that the
good catalyst stability is attributed to strong interactions between the boehmite surface
and active Pt and WOx species. However, according to Table 1, the best glycerol yield

of
results were obtained for concentrations around 10 wt% glycerol in water, using fixed
bed reactors [8,84]. Testing these catalysts at higher glycerol concentrations for scale-
up purposes is a challenge.

ro
Moreover, other studies were performed considering iridium as active metallic phase
associated with rhenium metallic or its oxide. SiO2 was the most widely used support.
-p
The catalyst were listed in Table 2 in ascending order of yield .

Table 2.
re
Hydrogenolysis of glycerol into 1,3-PDO. Active metallic phase Ir, Ir-ReOx.
lP

X S R
Catalysts Conditions Reactor References
(%) (%) (%)
393K,
8MPa, 4h,
na

44 Ru-Ir-ReOx/SiO2 4g Batch 23 54 12 [89]


glycerol / 2
g H2O
393K,
ur

8MPa, 12h,
45 Ir–ReOx/SiO2 Batch 23 58 13 [90]
67wt%
glycerol
Jo

453K,
8MPa, 12h,
46 IrOx/H-ZSM-5 Batch 22 70 15 [45]
10wt%
glycerol
403K,
8MPa,
Fixed
47 Ir–ReOx/SiO2 0.5h-1, 61 31 19 [91]
bed
80wt%
glycerol
12

120°C,
8MPa, 12h,
48 Ir–Re/D-ASA-2.0a Batch 55 39 21 [92]
20wt%
glycerol
120°C,
8MPa, 12h,
49 Ir-Re/KIT-6-CRb Batch 63 35 22 [93]
20wt%
glycerol
393K,
8MPa, 24h,
50 Ir-ReOx/SiO2 Batch 69 47 32 [43]
67wt%
glycerol
393K,
Ir-ReOx/SiO2 + 8MPa, 36h,
51 Batch 75 44 33 [94]
H-ZSM-5 80wt%

of
glycerol
413K,
1MPa, 1h- Fixed
52 0.1Re/Pt/WOx/Al2O3 1 68 49 33 [95]

ro
, 5wt% bed
glycerol
393K,
8MPa, 36h, -p
53 Ir–ReOx/SiO2 Batch 81 47 38 [96]
80wt%
glycerol
a
D-ASA: De-aluminated Amorphous Silica-Alumina; bKIT: ordered mesoporous silica
re
The interest in new bifunctional catalysts for the hydrogenolysis reaction involving
lP

the association of a platinum group metal with oxophilic oxides includes Ir and Re
oxide, which represent an alternative to Pt and WOx. Just as platinum, iridium
decomposes hydrogen atoms, and acts as a metallic site. ReOx plays a similar role to
na

WOx, which anchors the primary glycerol hydroxyl, and forms an alkoxyde strongly
bonded to the surface, and protonating the secondary hydroxyl. However, the
ur

combination of Pt and W is still more effective than Ir and Re, according to the
comparison of yields to 1,3-PDO as shown in Tables 1 and 2.
The catalysts containing iridium and rhenium were reported for the first time by
Jo

Nakagawa et al. [96]. Ir-ReOx/SiO2 was the most effective for the hydrogenolysis of
glycerol. One may notice that, on average, they present lower conversion and equivalent
selectivity (Table 2) when compared to Pt, tungsten oxide, and aluminum-based
catalysts (Table 1). Wan et al. [45] obtained a selectivity of 70% to 1,3-PDO using IrOx
as acid site supported on an acidified zeolite (H-ZSM-5) as an alternative to ReOx. Such
high selectivity was reached due to the presence of Brønsted acid sites induced by
13

iridium on the catalyst interfaces at 8 MPa, 453 K and 12 h. The catalytic activity
strongly correlated to the total acidity regarding the Brønsted acid, which, for this
catalyst, was 46:26 mmol/g. This demonstrates that the concentration of total acids
when compared to Brønsted acids may be a factor that interferes with the selectivity to
1,3-PDO.
In Table 2, note that the best yields to 1,3-PDO occur in batch reactors at high
hydrogen pressures and high glycerol concentration in water (80 wt%) at mild
temperature (393 K) to avoid reactant degradation [94,96]. An outlier from these
references is the high glycerol concentration, which could have technical and economic
advantages in a future scale-up. Although glycerol conversion is relevant (> 75%), the

of
selectivity to 1,3-PDO is still low, with considerable formation of 1-propanol and 1,2-
PDO. The catalytic test in continuous reactors could be a solution for increasing the
selectivity.

ro
Table 3 displays studies that considered other active metals (Cu, Ni, and Zr)
associated with an acid phase on different supports.

Table 3.
-p
Hydrogenolysis of glycerol into 1,3-PDO. Other active metallic phases.
re
X S R
Catalysts Conditions Reactors References
lP

(%) (%) (%)


453K,
3.5MPa,
54 Cu-WOx-TiO2 10h, Batch 13 32 4 [97]
na

10wt%
glycerol
200°C,
600psi,
ur

55 Ni-Zr/H-beta 10h, Batch 77 14 11 [98]


20wt%
glycerol
Jo

210°C,
10Cu- Fixed
56 0.54MPa, 83 32 27 [99]
20H4SiW12O40/SiO2 bed
0.1h-1
393K,
57 NiRe(7)/MSa Batch 98 47 46 [100]
5MPa, 4h
a
MS: Mesoporous Saponite
Some studies tried to replace noble metals (Pt, Ir, and Re) with others of lower cost,
such as Cu, Ni, and Zr. Li et al. [97] used copper as a metal for activating the hydrogen
molecule. However, the selectivity to 1,3-PDO was low, producing considerable
14

amounts of 1,2-PDO. Huang et al. [99] used copper with a heteropolyacid containing
tungsten, with a high conversion rate, but low yield to 1,3-PDO.
Gebretsadik et al. [100] produced modified mono and bimetallic Ni catalysts, with
various types of oxides MOx (M = Mo, V, W, and Re), which were tested as an
alternative path to glycerol hydrogenolysis to obtain 1,3-PDO. Catalysts containing Ni-
Cu, Ni-V, and Ni-W with or without the presence of Re were observed to be more
selective to form 1,2-PDO. When Ni was added to rhenium and supported on
mesoporous saponite (Ni-Re/MS), high conversion and moderate selectivity to 1,3-PDO
were obtained. The addition of a new step, with the formation of glycidol, leads to high
conversions, although it is economically disadvantageous due to the increase in the

of
process cost. The Ni-Re/MS catalyst presented the highest activity and high selectivity
to 1,3-PDO, with 98% conversion and 46.1% yield, at 393 K, 5 MPa H2 and 4 h in a
batch reactor.

ro
Considering the evaluation of Tables 1 to 3, it can be verified that the best process
conditions for scale-up take into account aqueous liquid phase systems, with high
-p
hydrogen pressure (5 to 8 MPa), medium-range temperatures (393 K to 473 K), and
preferably in continuous flow reactors. A challenge for scaling up from laboratory
re
bench to process design is the development of more studies at relevant glycerol
concentrations (over 80 wt%), which would especially decrease recycling and energy
lP

costs. A more detailed analysis of reactor type and conditions used in laboratory studies
considering future process design is presented in item 4 herein.
Another important point to be analyzed is the possible leaching of one of the active
na

phases when working with liquid phase systems. Supported heteropolyacids catalysts
were reported by Zhu et al., who used lithium in order to stabilize the support [52].
Other authors performed thermal treatment on Pt/W [65] and Ir/Re [43,94] catalyst
ur

precursors, solving or mitigating the problem in laboratory works. However, leaching


can be an extra challenge for designing industrial catalysts and should be studied.
Jo

Regarding the catalyst, an intense search should be carried out for selecting the best
active acid phase for the dehydration reaction. The attack of the glycerol secondary
hydroxyl and the simultaneous addition of hydrogen by the metallic site to preferably
form 1,3-PDO, may be evaluated. WO3 seems to be well established as an acidity
function provider, many times associated with other larger surface area oxide (Al2O3,
ZrO2, SiO2 among others), zeolite (mordenite) or clay (montmorillonite). Platinum is
the best metal for dissociating the hydrogen molecule, even though other metals (Ir
15

associated with ReOx among others) have been a target of research aiming at mitigating
consecutive hydrogenolysis reactions of 1,3-PDO to propanol, and reducing the catalyst
preparation costs (Cu, Ni).
Although comparing catalyst activity at different experimental conditions is
difficult, an approximate correlation analysis between the nature and quantity of
Brønsted and Lewis acid sites in the catalysts could be a way to evaluate the influence
of the acid site on glycerol conversion and selectivity to 1,3-PDO.

3.1. NATURE OF THE ACID SITES

In order to avoid undesirable products in the conversion of glycerol and orientate the

of
reactions pathway to increase selectivity to the 1,3-PDO, understanding the specific
active sites of the catalysts involved in each reaction step is fundamental. The strength

ro
and chemical nature (Lewis or Brønsted acid) of the acid site will determine which
glycerol hydroxyl groups will be attached onto the catalyst surface and consequently
affect the product formation. Fig. 4 presents total acidity, measured using TPD-NH3, as
-p
a function of glycerol conversion for several catalysts containing platinum on different
supports.
re
lP
na
ur
Jo

Fig. 4. Total acidity (mmol g-1) as a function of conversion on different catalysts. Reaction conditions:
reaction temperature: 160-200 °C, pressure: 4.0-5.0 MPa; 10wt% glycerol aqueous solution; H2/glycerol=
137:1 (molar ratio); WHSV= 0.075-0.1 h-1, fixed bed [52,63,67,70,80].

All the results shown in Fig. 4 were obtained under similar experimental conditions,
a continuous flow fixed bed reactor system with glycerol aqueous solution and the same
16

H2/glycerol molar ratio. Platinum was the active metallic phase responsible for the
dissociation of hydrogen molecules for all catalysts. The support type was the main
difference among the catalysts. It is observed that it is not possible to establish a direct
correlation between the materials total acidity and its performance in the glycerol
conversion. The largest number of acid sites, 1.6 mmol NH3 g.cat.-1, determined by
TPD-NH3, was observed for the Pt-H4SiW12O40/SiO2 catalyst. The best glycerol
conversion, 64%, was obtained for the Pt–10WOx/Al2O3 catalyst and its total acidity
was only 0.41 mmol NH3 gcat.-1.
Therefore, the discussion about the effect of catalyst acidity on glycerol conversion
without considering quantity, distribution, and kind of noble metal element present is

of
unreasonable. The difference in glycerol conversion is probably influenced by the
highest Pt dispersion. The Pt–10WOx/Al2O3 catalyst has 78% Pt dispersion, while Pt-
H4SiW12O40/SiO2 has only 19%. Furthermore, the conversion is greatly dependent on

ro
reaction temperature, space velocity, H2 pressure, and initial water content.
Tungsten oxides are traditionally used as co-catalysts for glycerol hydrogenolysis
-p
into 1,3-PDO due to the presence of the Brønsted sites on their surface. Zhu et al.
[52,63,70] prepared a series of catalysts with H4SiW12O40 Kegging-type heteropoly acid
re
as co-catalysts. The results of 1,3-PDO selectivity for the catalysts are summarized in
Fig. 5.
lP
na
ur
Jo

Fig. 5. Lewis/Brønsted acid site ratio as a function of 1,3-PDO selectivity on different catalysts
[52,63,70]
17

The increase in the Lewis/Brønsted acid sites ratio is accompanied by higher


selectivity for 1,3-PDO production. The Pt-H4SiW12O40/ZrO2 structure was doped with
lithium metal by the ion-exchange method, and the final material Pt-
Li2H2SiW12O40/ZrO2 showed better selectivity than unmodified catalysts, attaining 53%
1,3-PDO selectivity. The modification effect with Lithium probably includes the
adjustment of the acidity and the proportion between Lewis and Brønsted acid sites,
which optimized the glycerol hydrogenolysis to 1,3-PDO formation.
Besides the solid catalysts acid strength, the nature of acid sites (Lewis or Brønsted)
may play a key role in determining its selectivity for glycerol hydrogenolysis to 1,3-
PDO. The reaction sequence proposed in Fig. 3 and the results presented in Fig. 5 show

of
that the reaction of glycerol would be initiated by the dehydration involving either the
central or terminal -OH, which results in two different enol intermediates:
hydroxyacetone and 3-HPA. Hydroxyacetone formation proceeds via the dehydration of

ro
the terminal -OH group to form the tautomers 2-propene-1,2-diol and hydroxyacetone
and their hydrogenation then leads to 1,2-PDO. The key intermediate for forming 1,3-
-p
PDO is 3-HPA, which can easily rise to further hydrogenation for producing the
desirable 1,3-PDO [87].
re
Given the intermediaries importance in the selectivity of the glycerol hydrogenolysis
reaction, different studies have focused on the interaction of glycerol with the catalyst
lP

surface, specifically looking for a correlation between the role of Lewis and Brønsted
acid sites in the dehydration of glycerol using different spectroscopy techniques and
theoretical models [87,101–103].
na

Foo et al. [87] investigated the glycerol dehydration over niobia catalysts using FTIR
spectroscopy and DFT calculations to determine the role of Lewis and Brønsted sites in
the reaction pathway The authors applied the ex-situ impregnation method to attach the
ur

glycerol molecules onto the catalysts surface and the Brønsted sites were blocked by
Na+ to ensure the elucidation of the role of different types of acid sites. When
Jo

adsorption occurs on the Lewis site, glycerol forms a multidentate alkoxy species
through the terminal -OH, and favors dehydration to form hydroxyacetone and, finally,
1,2-PDO. Yet, the secondary alcohol group of glycerol is preferentially attached to the
Brønsted site and, in this position, the dehydration favors 3-HPA formation, and could
subsequently lead to 1,3-PDO by hydrogenation on the metallic particles [87,103]
The role of each type of acid site on the catalyst selectivity is reasonably understood
by the scientific community. Most of these studies use well-established techniques that
18

have been developed to characterize acid-base surface properties in the gas phase, such
as adsorption-desorption, calorimetry, and IR spectroscopy with probe molecules [104].
However, the acid-base properties of the catalyst surface are strongly affected by the
liquid medium in which it is immersed. Specifically, for reactions in polar solvents, e.g.
water in glycerol reactions, the superficial acid site could be strongly affected by the
chemisorption of water. These include the reaction between Lewis site with hydroxyl
ions from the water molecule (Lewis base) and the Brønsted site formation on the
catalyst surface [105]. Some authors reported the formation of Brønsted site generated
from Lewis acid sites in the presence of water vapor on niobium oxide catalysts
[87,106]. Gould et al. [104,107] reported the influence of the liquid phase - water,

of
ethanol, and acetonitrile - on the quantification of Brønsted acidity for four different
zeolites. The measurements were conducted by determining the pyridine content in the
liquid phase with a FTIR flow cell in an attenuated total reflection (ATR) configuration.

ro
The comparison with the gas phase measurements confirms that the solvent choice
impacts the Brønsted acidity of zeolites quantification [107].
-p
Dehydration is crucial for 1,3-PDO production and this step is claimed to occur on
the Brønsted acid sites of the catalysts in the aqueous system. Thus, improved
re
understanding of the liquid-solid interfaces could provide more accurate development
insights to design catalysts with better performance in glycerol conversion to 1,3-PDO.
lP

4. 1,3-PDO PRODUCTION PROCESSES EVALUATION

Process condition analysis is important for choosing the most promising catalysts to
na

be used in an industrial process. Some operating conditions need to be evaluated


individually due to their influence on yield and process economics. Reaction severity,
determined by the system temperature and pressure, and glycerol concentration at the
ur

reactor inlet are variables of great relevance in process optimization.


Temperatures over 180 ºC reduce the selectivity to 1,3-PDO due to the
Jo

hydrogenolysis of 1,3-PDO towards the formation of 1-propanol and 2-propanol.


Moreover, they may promote the breaking of the C-C bonds, leading to undesirable
products such as ethanol, methanol, ethane, and methane. Nevertheless, considering the
low cost of glycerol, the economic feasibility of the process can be reached at
temperatures higher than 180 ºC. As a result, there will be higher conversion and lower
selectivity to 1,3-PDO with formation of the afore mentioned byproducts, which are
tradeable at low price. Several studies have been carried out regarding the effect of
19

temperature on 1,3-PDO yields. It was verified that, in liquid phase batch reactors, the
highest conversion and selectivity (over 80% and 60%, respectively) were reached
within the temperature range from 120 ºC to 160 ºC [83,85,86]. The reaction pressure
varies along a wide range in the compiled experimental studies. Pressures over 5 MPa
may lead to better selectivity since they increase the solubility and hydrogen dispersion
in the aqueous medium, favoring the hydrogenation of 3-HPA at the moment it is
formed. As a disadvantage, high pressures increase initial investment costs due to more
resilient equipment specification requirements. Fan et al. [83] and Oh, Dash, Lee [85]
obtained yields over 48%, at pressures of 7 MPa or higher.
Glycerol at high concentrations is unstable at high temperatures (over 180 ºC),

of
forming oligomers. As can be verified in Tables 1 to 3, many bench scale studies use
diluted concentrations, most around 10% w/w glycerol in aqueous solutions. This
implies large energy expenditure in the process involved due to water recycling (or

ro
steam) at high temperatures. However, works such as those by Zhu et al. [81], and
Nakagawa et al. [96] achieved good conversion and selectivity using concentrations of
40% and 80%, respectively.
-p
Considering the studies listed in Tables 1 to 3 and aiming at the economics of the
re
process to be designed, the catalysts with best yield or medium yields with high
selectivity both in continuous and batch reactors were analyzed. Therefore, studies were
lP

selected according to the reactor type used, taking into account that continuous reactors
lead to lower initial investment for the same production.
na

4.1. PROCESS IN BATCH REACTORS

Glycerol conversion into 1,3-PDO in laboratory was performed in a liquid phase


batch reactor (generally aqueous). The studies that presented yields over 45% with
ur

considerable selectivity were evaluated in greater detail. Arundhathi et al. [86]


developed an efficient catalyst for glycerol selective hydrogenolysis to 1,3-PDO. The
Jo

authors dispersed platinum nanoporous particles over tungsten oxide supported on


boehmite (Pt/WOx/AlOOH). Even though a high yield was achieved in a batch reactor
at 453 K and 5 MPa H2 for 12 h, the initial glycerol concentration was too low (3%
w/w) and extrapolation for developing process design is difficult.
Fan et al. [83] studied the effects of the crystalline phases of the ZrO2 support, the
monoclinic (m-ZrO2) and the tetragonal (t-ZrO2) on the catalyst activity (Pt-
WOx/ZrO2), for the hydrogenolysis of glycerol at 1,3-PDO. Among the two catalysts,
20

the Pt-WOx/t-ZrO2 catalyst exhibited the highest yield (49.4%) and selectivity of 65%,
as well good stability after several reaction cycles. The superiority of the catalytic
performance of Pt-WOx/t-ZrO2, when compared to Pt-WOx/m-ZrO2 was due to higher
dispersion, interaction of Pt and WOx and generation of stronger Brønsted acid sites in
the presence of WOx, resulting in higher activity and selectivity to the desired product.
Both catalysts were tested in the same reaction conditions, 413 K and 8 MPa H2 for 24h.
The authors reaffirm the importance of using an appropriate crystallographic mold of
the catalyst support in a reaction that should be highly selective to the secondary
hydroxyl of glycerol.
Oh, Dash and Lee [85] studied the performance of a Pt-sulfated/ZrO2 catalyst for

of
selectively converting glycerol into 1,3-PDO. Sulfated zirconia is obtained after
treatment with sulfuric acid, aiming at obtaining Brønsted acid sites. The sulfate and
platinum ions were stabler in the tetragonal zirconia phase. The reaction conditions of

ro
170 °C and 7.3 MPa H2 for 24 h resulted in a yield of 55.6%, which was the highest
reported value for zirconia-supported catalysts, using DMI (1,3-dimethyl-2-
-p
imidazolidinone) as a solvent, since the results were poor in aqueous solution.
Note that the best results (yield > 45%) for batch reactors were those of Fan et al.
re
[83], Oh, Dash and Lee [85] and Arundhathi et al. [86]. However, these reactors are less
competitive for their use in industrial scale since, albeit versatile, they present
lP

disadvantages, such as the need for replacing the catalyst at every batch, as well as
loading and unloading procedures, decreasing the annual operation time and higher
initial investment due to the size of the equipment designed considering the same
na

production.

4.2. PROCESS IN FIXED BED REACTORS


ur

The tested catalysts in fixed bed reactors were analyzed according to two criteria:
good yields (over 45%) and moderate yields (30-45%), yet with high selectivity to 1,3-
Jo

PDO (over 70%). The second option considers that unreacted glycerol recycle
compensates the lower activity obtained by these catalysts. Priya et al. [82,84]
developed catalysts tested in continuous processes, containing mordenite, which is a
zeolite with many industrial applications, possesses acidic properties and can be used as
support, due to its thermal stability and good regeneration. The strong Brønsted acidity
of mordenite is responsible for the good performance in the hydrogenolysis of glycerol.
Priya et al. [82] impregnated Pt on mordenite (Pt/HM), while Priya et al. [84] developed
21

a bimetallic catalyst on mordenite Pt-Cu/HM. Both catalyts promoted the breaking of


the C-O bond over the C-C bond, although the Pt-Cu combination demonstrated more
efficiency for 1,3-PDO formation and presented higher catalytic activity than the
monometallic catalyst developed by Priya et al. [82]. The authors used similar reaction
conditions: 0.1 MPa H2; WHSV = 1.02 h-1, vapor phase and 10% w/w glycerol
concentration in water.
Zhu and Chen [81] studied the hydrogenolysis of glycerol over catalysts containing
platinum, passivated with lithium and boron composites, and tungsten and zirconia
oxides. The results showed that Li2B4O7/Wox/ZrO2 obtained 90.7% conversion at 150
°C and 4 MPa, with 1,3-PDO yield of 45% mol and excellent stability after 200 hours in

of
operation. The reaction occurred in a fixed bed reactor with 40% w/w glycerol in water.
The reaction products were 1,3-PDO, 1,2-PDO, n-propanol and iso-propanol. A low
formation of 1,2-PDO (1.2% mol) and high formation of monoalcohols (44% mol) were

ro
observed. The greater mass percentage of glycerol in the reactor feed is a major
advantage of this study. The low reaction temperature is also a highlight for scaling up
-p
the process. The high production of monoalcohols as byproducts is a challenge to be
faced, aiming at producing a catalyst that decreases the undesirable consecutive reaction
re
of 1,3-PDO hydrogenolysis.
Samudrala, Kandasamy and Bhattacharya [8] studied the use of montmorillonite clay
lP

activated with sulfuric acid, of low cost, as a support for platinum nanoparticles (Pt/S-
MMT). Montmorillonite is a clay mineral constituted of silica and alumina structures
and are adsorbents due to the existence of several active sites, such as surface exchange
na

sites and ion exchange sites. The catalyst performance was investigated in a fixed bed
reactor, varying reaction parameters, such as temperature, hydrogen flow rate, glycerol
concentration, WHSV, and contact time, to verify the optimal reaction conditions. The
ur

best yield was obtained for atmospheric pressure, 200 ºC; 10% w/w glycerol in water,
0.5 g of catalyst, H2 flow rate of 70 mL/min and WHSV of 1.02 h-1, obtaining 62%
Jo

selectivity to 1,3-PDO and 94% glycerol conversion. The resulting reaction products
were 1,3-PDO, 1,2-PDO, acrolein, hydroxyacetone, among others (1-propanol, 2-
propanol, acetone, and acetaldehyde). Although the dilution of glycerol in water is high
for an economically feasible process, the high selectivity to 1,3-PDO is one of the
greatest strengths of this study.
22

5. CONCLUSIONS

1,3-PDO can be industrially obtained from the hydrogenolysis of glycerol and is


considered one of the most promising platform molecules to replace fossil-derived
products. The increase in biodiesel production, and consequent high availability of
glycerol, is the primary factor for increasing its competitiveness in a market that is
expected to reach a CAGR of 14.2% over the next ten years. PTT, obtained from 1,3-
PDO, is the polymer of highest potential to compete with polymers of petrochemical
origin due to its unique physical-chemical characteristics, which ensure its better
quality. A new design of an industrial process must optimize the temperature and
concentration of glycerol in the reactor feed. The bifunctional catalyst preferably uses

of
Pt, Re, and Cu as an active metallic phase for the hydrogenation reaction, while the total
amount of acid sites and the balanced ratio of Brønsted and Lewis sites are the key rule

ro
for greater selectivity to 1,3-PDO in the dehydration reactions.

Credit Author Statement -p


re
Alisson Dias da Silva Ruy. Conceptualization; Investigation; Methodology;
Roles/Writing - original draft; Visualization
Rita Maria de Brito Alves. Supervision; Writing - review & editing
lP

Thiago Lewis Reis Hewer. Investigation; Methodology; Roles/Writing - original draft;


Danilo de Aguiar Pontes. Roles/Writing - original draft; Writing - review & editing;
na

Leonardo Sena Gomes Teixeira. Investigation; Supervision;


Luiz Antônio Magalhães Pontes. Conceptualization; Methodology; Roles/Writing -
original draft; Supervision; Writing - review & editing
ur

6. ACKNOWLEDGMENTS
Jo

The authors gratefully acknowledge the support of the RCGI – Research Centre for
Gas Innovation, hosted by the University of São Paulo (USP) and sponsored by
FAPESP – the State of São Paulo Research Foundation (2014/50279-4), and
Shell Brasil. The authors also acknowledge FAPESB – the State of Bahia Research
Foundation. This study was financed in part by the Coordenação de Aperfeiçoamento de
23

Pessoal de Nível Superior - Brasil (CAPES – Coordination for the Improvement of


Higher Education Personnel) - Finance Code 001.

of
ro
-p
re
lP
na
ur
Jo
24

REFERENCES

[1] S.N. Gebremariam, J.M. Marchetti, Economics of biodiesel production: Review,


Energy Convers. Manag. 168 (2018) 74–84.
https://doi.org/10.1016/j.enconman.2018.05.002.

[2] Y. Wang, Z. Fang, F. Zhang, Esterification of oleic acid to biodiesel catalyzed by


a highly acidic carbonaceous catalyst, Catal. Today. 319 (2019) 172–181.
https://doi.org/10.1016/j.cattod.2018.06.041.

[3] A. Hejna, P. Kosmela, K. Formela, Ł. Piszczyk, J.T. Haponiuk, Potential


applications of crude glycerol in polymer technology – Current state and

of
perspectives, Renew. Sustain. Energy Rev. 66 (2016) 449–475.
https://doi.org/10.1016/j.rser.2016.08.020.

ro
[4] S. Bagheri, N. Muhd, W.A. Yehye, Catalytic conversion of biodiesel derived raw
glycerol to value added products, Renew. Sustain. Energy Rev. 41 (2015) 113–

[5]
-p
127. https://doi.org/10.1016/j.rser.2014.08.031.

J.E. Castanheiro, J. Vital, I.M. Fonseca, A.M. Ramos, Glycerol conversion into
re
biofuel additives by acetalization with pentanal over heteropolyacids
immobilized on zeolites, Catal. Today. (2019) 0–1.
lP

https://doi.org/10.1016/j.cattod.2019.04.048.

[6] H. Zhao, L. Zheng, X. Li, P. Chen, Z. Hou, Hydrogenolysis of glycerol to 1,2-


propanediol over Cu-based catalysts: A short review, Catal. Today. (2019).
na

https://doi.org/10.1016/j.cattod.2019.03.011.

[7] M. Yang, J. Yun, H. Zhang, T.A. Magocha, H. Zabed, Y. Xue, E. Fokum, W.


ur

Sun, X. Qi, Genetically engineered strains: Application and advances for 1,3-
Propanediol production from glycerol, Food Technol. Biotechnol. 56 (2018) 3–
15. https://doi.org/10.17113/ftb.56.01.18.5444.
Jo

[8] S.P. Samudrala, S. Kandasamy, S. Bhattacharya, Turning Biodiesel Waste


Glycerol into 1,3-Propanediol: Catalytic Performance of Sulphuric acid-
Activated Montmorillonite Supported Platinum Catalysts in Glycerol
Hydrogenolysis, Sci. Rep. 8 (2018) 1–12. https://doi.org/10.1038/s41598-018-
25787-w.

[9] W. Zhou, Y. Zhao, S. Wang, X. Ma, The effect of metal properties on the
25

reaction routes of glycerol hydrogenolysis over platinum and ruthenium catalysts,


Catal. Today. 298 (2017) 2–8. https://doi.org/10.1016/j.cattod.2017.07.021.

[10] N.Y. Kalisa, X. Gao, Current Trends in Efficient Production of 1,3-Propanediol,


J. Acad. Ind. Res. 6 (2018) 137–148.

[11] G.P. Da Silva, C.J.B. De Lima, J. Contiero, Production and productivity of 1,3-
propanediol from glycerol by Klebsiella pneumoniae GLC29, Catal. Today. 257
(2015) 259–266. https://doi.org/10.1016/j.cattod.2014.05.016.

[12] Mordor Intelligence, Global Glycerin Market - Growth, Trends, and Forecast
(2019 - 2024), (2019). https://research.mordorintelligence.com/reports/glycerin-
market (accessed October 9, 2019).

of
[13] Z. Li, L. Yan, J. Zhou, X. Wang, Y. Sun, Z.L. Xiu, Two-step salting-out

ro
extraction of 1,3-propanediol, butyric acid and acetic acid from fermentation
broths, Sep. Purif. Technol. 209 (2019) 246–253.
https://doi.org/10.1016/j.seppur.2018.07.021.
-p
[14] M. Matsumoto, K. Nagai, K. Kondo, Reactive extraction of 1,3-propanediol with
re
aldehydes in the presence of a hydrophobic acidic ionic liquid as a catalyst,
Solvent Extr. Res. Dev. 22 (2015) 209–213.
https://doi.org/10.15261/serdj.22.209.
lP

[15] S. Wang, L. Qiu, H. Dai, X. Zeng, B. Fang, Highly pure 1,3-propanediol:


Separation and purification from crude glycerol-based fermentation, Eng. Life
na

Sci. 15 (2015) 788–796. https://doi.org/10.1002/elsc.201500004.

[16] Market Research Future, 1,3-Propanediol Market Research Report - Forecast to


2023, (2020). https://www.marketresearchfuture.com/reports/1-3-propanediol-
ur

market-5722 (accessed September 20, 2019).

[17] MILLION INSIGHTS, 1,3-Propanediol (PDO) Market to be Driven by Rising


Jo

Infiltration of Polyurethane & High Demand for Polyesters Till 2022 | Million
Insights, (2020). https://www.millioninsights.com/industry-reports/1-3-
propanediol-pdo-
market?utm_source=prnewswire&utm_medium=referral&utm_campaign=prn_0
9Jan2020_pdo_rd2 (accessed May 8, 2020).

[18] N. Vivek, A. Pandey, P. Binod, An efficient aqueous two phase systems using
26

dual inorganic electrolytes to separate 1,3-propanediol from the fermented broth,


Bioresour. Technol. 254 (2018) 239–246.
https://doi.org/10.1016/j.biortech.2018.01.076.

[19] N. Vivek, A. Pandey, P. Binod, Production and Applications of 1,3-Propanediol,


in: Curr. Dev. Biotechnol. Bioeng. Prod. Isol. Purif. Ind. Prod., Elsevier B.V.,
2017: pp. 719–738. https://doi.org/10.1016/B978-0-444-63662-1.00031-2.

[20] D. Tjahjasari, T. Kaeding, A.P. Zeng, 1,3-Propanediol and


Polytrimethyleneterephthalate, in: Compr. Biotechnol. Second Ed., Second Edi,
Elsevier B.V., Hamburg, 2011: pp. 229–242. https://doi.org/10.1016/B978-0-08-
088504-9.00160-4.

of
[21] Business Wire, Global 1,3-Propanediol (PDO) Market Outlook Report 2018-
2027, (2019).

ro
https://www.businesswire.com/news/home/20191001005864/en/Global-13-
Propanediol-PDO-Market-Outlook-Report-2018-2027 (accessed October 9,
2019).
-p
[22] RESEARCH AND MARKETS, 1,3-Propanediol (PDO) Market by Application
re
(Polytrimethylene terephthalate (PTT), Cosmetics, Personal Care & Cleaning
Products, Polyurethane (PU)) and Region (Americas, APAC, Europe, Middle
lP

East & Africa (EMEA)) - Global Forecast to 2024, (2019).


https://www.researchandmarkets.com/reports/4832576/13-propanediol-pdo-
market-by-
na

application?utm_source=dynamic&utm_medium=GNOM&utm_code=v2zwv4&
utm_campaign=1330673+-+1%2C3-
Propanediol+(PDO)+Markets%3A+Global+Analysis%2C+Insights%2C+Trends
ur

+and+Opportunitie (accessed May 10, 2020).


Jo

[23] MARKETS AND MARKETS, 1,3-Propanediol (PDO) Market by Application


(Polytrimethylene terephthalate (PTT), Cosmetics, Personal Care & Cleaning
Products, Polyurethane (PU)) and Region (Americas, APAC, Europe, Middle
East & Africa (EMEA)) - Global Forecast to 2024, (2020).
https://www.marketsandmarkets.com/Market-Reports/1-3-propanediol-pdo-
market-760.html (accessed May 1, 2020).

[24] Grand View Research, 1,3 Propanediol (PDO) Market Size, Share & Trends
27

Analysis Report By Application (Polytrimethylene Terephthalate (PTT),


Polyurethane, Personal Care & Detergents) And Segment Forecasts To 2022,
(2015). https://www.grandviewresearch.com/industry-analysis/1-3-propanediol-
pdo-market/methodology (accessed May 20, 2020).

[25] Y.Q. Sun, J.T. Shen, L. Yan, J.J. Zhou, L.L. Jiang, Y. Chen, J.L. Yuan, E.M.
Feng, Z.L. Xiu, Advances in bioconversion of glycerol to 1,3-propanediol:
Prospects and challenges, Process Biochem. 71 (2018) 134–146.
https://doi.org/10.1016/j.procbio.2018.05.009.

[26] Y. Ma, Y. Wang, P.J. Morgan, R.E. Jackson, X. Liu, G.C. Saunders, F.
Lorenzini, A.C. Marr, Designing effective homogeneous catalysis for glycerol

of
valorisation: selective synthesis of a value-added aldehyde from 1,3-propanediol
via hydrogen transfer catalysed by a highly recyclable, fluorinated Cp*Ir(NHC)

ro
catalyst, Catal. Today. 307 (2018) 248–259.
https://doi.org/10.1016/j.cattod.2017.09.036.
-p
[27] F. Cavani, S. Albonetti, F. Basile, A. Gandini, Chemicals and Fuels from Bio-
Based Building Blocks, John Wiley & Sons, Weinheim, 2016.
re
[28] Biotechnology Innovation Organization, Renewable Chemical Platforms
Building the Biobased Economy, Ind. Biotechnol. 14 (2018) 109–111.
lP

https://doi.org/https://doi.org/10.1089/ind.2018.29135.bio.

[29] S.P. Samudrala, Glycerol Transformation to Value-Added 1,3-Propanediol


Production: A Paradigm for a Sustainable Biorefinery Process, in: Glycerine
na

Prod. Transform. Innov. Platf. Sustain. Biorefinery Energy, IntechOpen, 2019: p.


24. https://doi.org/10.5772/intechopen.83694.
ur

[30] Metabolic Explorer, Metabolic EXplorer in the 1st half of 2019, (2019).
https://www.metabolic-explorer.com/2019/09/27/first-half-2019/ (accessed
Jo

October 10, 2019).

[31] Pro Market Research, Global 1,3-Propanediol (PDO) Market 2019 - by


Manufacturers, Regions, Type, Application, Sales, Revenue, and Forecast to
2025, (2019). https://www.promarketresearch.com/global-1-3-propanediol-pdo-
market-2018-by-
30533.html?utm_source=ourcryptojournal.com&utm_medium=Krishna
(accessed October 9, 2019).
28

[32] A.K. Hilaly, T.P. Binder, Method of recovering 1,3-propanediol from


fermentation broth, U.S. Patent No. 6,479,716, 2002.

[33] C.E. Nakamura, A.A. Gatenby, A.K.-H. Hsu, R.D. La Reau, S.L. Haynie, M.
Diaz-Torres, D.E. Trimbur, G.M. Whited, V. Nagarajan, M.S. Payne, S.K.
Picataggio, R. V. Nair, Method for the production of 1,3-propanediol by
recombinant microorganisms, U.S. Patent No. 6,013,494, 2000.

[34] S.K. Brar, S.J. Sarma, K. PaKShiraJan, Platform Chemical Biorefinery: Future
Green Chemistry, Elsevier Inc., The Netherlands, 2016.

[35] A. Pandey, S. Negi, C.R. Soccol, Current Developments in Biotechnology and


Bioengineering: Production, Isolation and Purification of Industrial Products,

of
Elsevier, Netherlands, 2017.

ro
[36] J.J. Varghese, L. Cao, C. Robertson, Y. Yang, L.F. Gladden, A.A. Lapkin, S.H.
Mushrif, Synergistic Contribution of the Acidic Metal Oxide-Metal Couple and
Solvent Environment in the Selective Hydrogenolysis of Glycerol: A Combined
-p
Experimental and Computational Study Using ReOx -Ir as the Catalyst, ACS
Catal. 9 (2019) 485–503. https://doi.org/10.1021/acscatal.8b03079.
re
[37] J. Feng, B. Xu, D.R. Liu, W. Xiong, J.B. Wang, Production of 1,3-Propanediol
by Catalytic Hydrogenolysis of Glycerol, Adv. Mater. Res. 791–793 (2013) 16–
lP

19. https://doi.org/10.4028/www.scientific.net/amr.791-793.16.

[38] K. Tomishige, Y. Nakagawa, M. Tamura, Production of Diols from Biomass, in:


na

Prod. Platf. Chem. from Sustain. Resour., 2017: pp. 343–372.


https://doi.org/10.1007/978-981-10-4172-3.

[39] C.S. Lee, M.K. Aroua, W.M.A.W. Daud, P. Cognet, Y. Pérès-Lucchese, P.-L.
ur

Fabre, O. Reynes, L. Latapie, A review: Conversion of bioglycerol into 1,3-


propanediol via biological and chemical method, Renew. Sustain. Energy Rev.
Jo

42 (2015) 963–972. https://doi.org/10.1016/j.rser.2014.10.033.

[40] Y. Jiang, X. Wang, Q. Cao, L. Dong, J. Guan, X. Mu, Chemical Conversion of


Biomass to Green Chemicals, in: M. Xian (Ed.), Sustain. Prod. Bulk Chem.,
Springer, 2016: p. 32. https://doi.org/10.1007/978-94-017-7475-8.

[41] D. Sun, Y. Yamada, S. Sato, W. Ueda, Glycerol hydrogenolysis into useful C3


chemicals, Appl. Catal. B Environ. 193 (2016) 75–92.
29

https://doi.org/10.1016/j.apcatb.2016.04.013.

[42] W. Zhou, J. Luo, Y. Wang, J. Liu, Y. Zhao, S. Wang, X. Ma, WOx domain size,
acid properties and mechanistic aspects of glycerol hydrogenolysis over
Pt/WOx/ZrO2, Appl. Catal. B Environ. 242 (2019) 410–421.
https://doi.org/10.1016/j.apcatb.2018.10.006.

[43] L. Liu, S. Kawakami, Y. Nakagawa, M. Tamura, K. Tomishige, Highly active


iridium–rhenium catalyst condensed on silica support for hydrogenolysis of
glycerol to 1,3-propanediol, Appl. Catal. B Environ. 256 (2019) 117775.
https://doi.org/10.1016/j.apcatb.2019.117775.

[44] Y. Nakagawa, M. Tamura, K. Tomishige, Catalytic materials for the

of
hydrogenolysis of glycerol to 1,3-propanediol, J. Mater. Chem. A. 2 (2014)
6688–6702. https://doi.org/10.1039/c3ta15384c.

ro
[45] X. Wan, Q. Zhang, M. Zhu, Y. Zhao, Y. Liu, C. Zhou, Y. Yang, Interface
synergy between IrOx and H-ZSM-5 in selective C–O hydrogenolysis of glycerol
-p
toward 1,3-propanediol, J. Catal. 375 (2019) 339–350.
https://doi.org/10.1016/j.jcat.2019.06.025.
re
[46] L. Gong, Y. Lu, Y. Ding, R. Lin, J. Li, W. Dong, T. Wang, W. Chen, Selective
hydrogenolysis of glycerol to 1,3-propanediol over a Pt/WO3/TiO2/SiO2 catalyst
lP

in aqueous media, Appl. Catal. A Gen. 390 (2010) 119–126.


https://doi.org/10.1016/j.apcata.2010.10.002.
na

[47] R. Rinaldi, Catalytic Hydrogenation for Biomass Valorization, The Royal


Society of Chemistry, Cambridge, 2015.

[48] L. Liu, Y. Zhang, A. Wang, T. Zhang, Mesoporous WO3 supported Pt catalyst


ur

for hydrogenolysis of glycerol to 1,3-propanediol, Chinese J. Catal. 33 (2012)


1257–1261. https://doi.org/10.1016/S1872-2067(11)60425-7.
Jo

[49] P. Uttraporn, P. Praserthdam, Effect of Calcination Temperature and Support


Type of Pt/WOx/boehmite Catalyst on 1,3-propanediol Production from
Hydrogenolysis of Glycerol, in: IOP Conf. Ser. Mater. Sci. Eng., 2019: p.
012012. https://doi.org/10.1088/1757-899X/559/1/012012.

[50] G. Shi, J. Xu, Z. Song, Z. Cao, K. Jin, S. Xu, X. Yan, Selective hydrogenolysis of
glycerol to 1,3-propanediol over Pt-WOx/SAPO-34 catalysts, Mol. Catal. 456
30

(2018) 22–30. https://doi.org/10.1016/j.mcat.2018.06.018.

[51] G. Shi, Z. Cao, J. Xu, K. Jin, Y. Bao, S. Xu, Effect of WOx Doping into Pt/SiO2
Catalysts for Glycerol Hydrogenolysis to 1,3-Propanediol in Liquid Phase, Catal.
Letters. 148 (2018) 2304–2314. https://doi.org/10.1007/s10562-018-2464-7.

[52] S. Zhu, X. Gao, Y. Zhu, Y. Zhu, X. Xiang, C. Hu, Y. Li, Alkaline metals
modified Pt–H4SiW12O40/ZrO2 catalysts for the selective hydrogenolysis of
glycerol to 1,3-propanediol, Appl. Catal. B Environ. 448 (2013) 60–67.
https://doi.org/10.1016/j.mcat.2018.01.029.

[53] M. Yang, X. Zhao, Y. Ren, J. Wang, N. Lei, A. Wang, T. Zhang, Pt/Nb-WOx for
the chemoselective hydrogenolysis of glycerol to 1,3-propanediol: Nb dopant

of
pacifying the over-reduction of WOx supports, Cuihua Xuebao/Chinese J. Catal.
39 (2018) 1027–1037. https://doi.org/10.1016/S1872-2067(18)63074-8.

ro
[54] T. Kurosaka, H. Maruyama, I. Naribayashi, Y. Sasaki, Production of 1,3-
propanediol by hydrogenolysis of glycerol catalyzed by Pt/WO3/ZrO2, Catal.
-p
Commun. 9 (2008) 1360–1363. https://doi.org/10.1016/j.catcom.2007.11.034.
re
[55] L. Gong, Y. Lü, Y. Ding, R. Lin, J. Li, W. Dong, T. Wang, W. Chen, Solvent
Effect on Selective Dehydroxylation of Glycerol to 1,3-Propanediol over a
Pt/WO3/ZrO2 Catalyst, Chinese J. Catal. 30 (2009) 1189–1191.
lP

https://doi.org/10.1016/s1872-2067(08)60142-4.

[56] J. Dam, K. Djanashvili, F. Kapteijn, U. Hanefeld, Pt/Al2O3 Catalyzed 1,3-


na

Propanediol Formation from Glycerol using Tungsten Additives, ChemCatChem.


5 (2013) 497–505. https://doi.org/10.1002/cctc.201200469.

[57] M. Edake, M. Dalil, M.J. Darabi Mahboub, J.L. Dubois, G.S. Patience, Catalytic
ur

glycerol hydrogenolysis to 1,3-propanediol in a gas-solid fluidized bed, RSC


Adv. 7 (2017) 3853–3860. https://doi.org/10.1039/c6ra27248g.
Jo

[58] C. Yang, F. Zhang, N. Lei, M. Yang, F. Liu, Z. Miao, Y. Sun, X. Zhao, A. Wang,
Understanding the promotional effect of Au on Pt/WO3 in hydrogenolysis of
glycerol to 1,3-propanediol, Cuihua Xuebao/Chinese J. Catal. 39 (2018) 1366–
1372. https://doi.org/10.1016/S1872-2067(18)63103-1.

[59] J. Wang, X. Zhao, N. Lei, L. Li, L. Zhang, S. Xu, S. Miao, X. Pan, A. Wang, T.
Zhang, Hydrogenolysis of Glycerol to 1,3-propanediol under Low Hydrogen
31

Pressure over WOx-Supported Single/Pseudo-Single Atom Pt Catalyst,


ChemSusChem. 9 (2016) 784–790. https://doi.org/10.1002/cssc.201501506.

[60] Z. Wu, M. Zhang, Y. Yao, J. Wang, D. Wang, M. Zhang, Y. Li, One-pot catalytic
production of 1,3-propanediol and γ-valerolactone from glycerol and levulinic
acid, Catal. Today. 302 (2018) 217–226.
https://doi.org/10.1016/j.cattod.2017.02.035.

[61] M. Gu, Z. Shen, L. Yang, B. Peng, W. Dong, W. Zhang, Y. Zhang, The Effect of
Catalytic Structure Modification on Hydrogenolysis of Glycerol into 1,3-
Propanediol over Platinum Nanoparticles and Ordered Mesoporous Alumina
Assembled Catalysts, Ind. Eng. Chem. Res. 56 (2017) 13572–13581.

of
https://doi.org/10.1021/acs.iecr.7b02899.

[62] W. Zhou, Y. Zhao, Y. Wang, S. Wang, X. Ma, Glycerol Hydrogenolysis to 1,3-

ro
Propanediol on Pt/WO3/ZrO2: Hydrogen Spillover Facilitated by Pt (111)
Formation, ChemCatChem. 8 (2016) 3663–3671.
https://doi.org/10.1002/cctc.201600981.
-p
[63] S. Zhu, Y. Qiu, Y. Zhu, S. Hao, H. Zheng, Y. Li, Hydrogenolysis of glycerol to
re
1,3-propanediol over bifunctional catalysts containing Pt and heteropolyacids,
Catal. Today. 212 (2013) 120–126. https://doi.org/10.1016/j.cattod.2012.09.011.
lP

[64] S. Feng, B. Zhao, L. Liu, J. Dong, Platinum Supported on WO3 ‑ Doped


Aluminosilicate: A Highly E fficient Catalyst for Selective Hydrogenolysis of
Glycerol to 1,3-Propanediol, Ind. Eng. Chem. Res. 56 (2017) 11065–11074.
na

https://doi.org/10.1021/acs.iecr.7b02951.

[65] O. Kirichenko, V. Nissenbaum, G. Kapustin, E. Redina, K. Vikanova, N.


ur

Davshan, L. Kustov, Thermal analysis of intermediates formed during


preparation of a Pt/WOx/Al2O3 catalyst for 1,3-propanediol synthesis from
Jo

glycerol, J. Therm. Anal. Calorim. 5 (2019) 1–14.


https://doi.org/10.1007/s10973-019-08322-5.

[66] S. García-Fernández, I. Gandarias, J. Requies, M.B. Güemez, S. Bennici, A.


Auroux, P.L. Arias, New approaches to the Pt/WOx/Al2O3 catalytic system
behavior for the selective glycerol hydrogenolysis to 1,3-propanediol, J. Catal.
323 (2015) 65–75. https://doi.org/10.1016/j.jcat.2014.12.028.
32

[67] S. Zhu, X. Gao, Y. Zhu, J. Cui, H. Zheng, Y. Li, SiO2 promoted Pt/WOx/ZrO2
catalysts for the selective hydrogenolysis of glycerol to 1,3-propanediol, Appl.
Catal. B Environ. 158–159 (2014) 391–399.
https://doi.org/10.1016/j.apcatb.2014.04.049.

[68] N. Lei, X. Zhao, B. Hou, M. Yang, M. Zhou, F. Liu, Effective Hydrogenolysis of


Glycerol to 1,3-Propanediol over Metal-Acid Concerted Pt/WOx/Al2O3
Catalysts, ChemCatChem. (2019) 1–11. https://doi.org/10.1002/cctc.201900689.

[69] S. García-Fernández, I. Gandarias, J. Requies, F. Soulimani, P.L. Arias, B.M.


Weckhuysen, The role of tungsten oxide in the selective hydrogenolysis of
glycerol to 1,3-propanediol over Pt/WOx/Al2O3, Appl. Catal. B Environ. 204

of
(2017) 260–272. https://doi.org/10.1016/j.apcatb.2016.11.016.

[70] S. Zhu, Y. Zhu, S. Hao, L. Chen, B. Zhang, Y. Li, Aqueous-phase

ro
hydrogenolysis of glycerol to 1,3-propanediol over Pt-H4SiW12O40/SiO2, Catal.
Letters. 142 (2012) 267–274. https://doi.org/10.1007/s10562-011-0757-1.
-p
[71] L.Z. Qin, M.J. Song, C.L. Chen, Aqueous-phase deoxygenation of glycerol to
1,3-propanediol over Pt/WO3/ZrO2 catalysts in a fixed-bed reactor, Green Chem.
re
12 (2010) 1466–1472. https://doi.org/10.1039/c0gc00005a.

[72] S. Feng, B. Zhao, Y. Liang, L. Liu, J. Dong, Improving Selectivity to 1,3-


lP

Propanediol for Glycerol Hydrogenolysis Using W- and Al-Incorporated SBA-15


as Support for Pt Nanoparticles, Ind. Eng. Chem. Res. 58 (2019) 2661–2671.
https://doi.org/10.1021/acs.iecr.8b03982.
na

[73] S. García-Fernández, I. Gandarias, Y. Tejido-Núñez, J. Requies, P.L. Arias,


Influence of the support of bimetallic Pt-WOx catalysts on the 1,3-propanediol
ur

formation from glycerol, ChemCatChem. 9 (2017) 4508–4519.


https://doi.org/10.1002/cctc.201701067.
Jo

[74] T. Aihara, H. Kobayashi, S. Feng, H. Miura, T. Shishido, Effect of WO3 loading


on the activity of Pt/WO3/Al2O3 catalysts in selective hydrogenolysis of
glycerol to 1,3-propanediol, Chem. Lett. 46 (2017) 1497–1500.
https://doi.org/10.1246/cl.170601.

[75] S.S. Priya, V.P. Kumar, M.L. Kantam, S.K. Bhargava, K.V.R. Chary, Vapour-
phase hydrogenolysis of glycerol to 1,3-propanediol over supported pt catalysts:
33

The effect of supports on the catalytic functionalities, Catal. Letters. 144 (2014)
2129–2143. https://doi.org/10.1007/s10562-014-1395-1.

[76] S.S. Priya, P. Kumar, L. Kantam, S.K. Bhargava, K.V.R. Chary, Catalytic
performance of Pt/AlPO4 catalysts for selective hydrogenolysis of glycerol to
1,3- propanediol in the vapour phase, RSC Adv. 4 (2014) 51893–51903.
https://doi.org/10.1039/C4RA09357G.

[77] S.S. Priya, V.P. Kumar, M.L. Kantam, S.K. Bhargava, A. Srikanth, K.V.R.
Chary, High Efficiency Conversion of Glycerol to 1,3-Propanediol Using a
Novel Platinum-Tungsten Catalyst Supported on SBA-15, Ind. Eng. Chem. Res.
54 (2015) 9104–9115. https://doi.org/10.1021/acs.iecr.5b01814.

of
[78] T. Mizugaki, T. Yamakawa, R. Arundhathi, T. Mitsudome, K. Jitsukawa, K.
Kaneda, Selective Hydrogenolysis of Glycerol to 1,3-Propanediol Catalyzed by

ro
Pt Nanoparticles–AlOx/WO3, Chem. Lett. 41 (2012) 1720–1722.
https://doi.org/10.1246/cl.2012.1720.
-p
[79] X. Zhao, J. Wang, M. Yang, N. Lei, L. Li, B. Hou, S. Miao, X. Pan, A. Wang, T.
Zhang, Selective Hydrogenolysis of Glycerol to 1,3-Propanediol: Manipulating
re
the Frustrated Lewis Pairs by Introducing Gold to Pt/WOx, ChemSusChem. 10
(2017) 819–824. https://doi.org/10.1002/cssc.201601503.
lP

[80] S. Zhu, X. Gao, Y. Zhu, Y. Li, Promoting effect of WOx on selective


hydrogenolysis of glycerol to 1,3-propanediol over bifunctional Pt-WOx/Al2O3
catalysts, J. Mol. Catal. A Chem. 398 (2015) 391–398.
na

https://doi.org/10.1016/j.molcata.2014.12.021.

[81] M. Zhu, C. Chen, Hydrogenolysis of glycerol to 1,3-propanediol over Li2B4O7-


ur

modified tungsten–zirconium composite oxides supported platinum catalyst,


React. Kinet. Mech. Catal. 124 (2018) 683–699. https://doi.org/10.1007/s11144-
Jo

018-1379-z.

[82] S.S. Priya, P. Bhanuchander, V.P. Kumar, D.K. Dumbre, S.R. Periasamy, S.K.
Bhargava, M.L. Kantam, K.V.R. Chary, Platinum Supported on H‑ Mordenite: A
Highly Efficient Catalyst for Selective Hydrogenolysis of Glycerol to 1,3-
Propanediol, ACS Sustain. Chem. Eng. 4 (2016) 1212–1222.
https://doi.org/10.1021/acssuschemeng.5b01272.
34

[83] Y. Fan, S. Cheng, H. Wang, J. Tian, S. Xie, Y. Pei, M. Qiao, B. Zong, Pt–WOx
on monoclinic or tetrahedral ZrO2: Crystal phase effect of zirconia on glycerol
hydrogenolysis to 1,3-propanediol, Appl. Catal. B Environ. 217 (2017) 331–341.
https://doi.org/10.1016/j.apcatb.2017.06.011.

[84] S.S. Priya, P. Bhanuchander, V.P. Kumar, S.K. Bhargava, K.V.R. Chary,
Activity and Selectivity of Platinum − Copper Bimetallic Catalysts Supported on
Mordenite for Glycerol Hydrogenolysis to 1,3-Propanediol, Ind. Eng. Chem. Res.
55 (2016) 4461–4472. https://doi.org/10.1021/acs.iecr.6b00161.

[85] J. Oh, S. Dash, H. Lee, Selective conversion of glycerol to 1,3-propanediol using


Pt-sulfated zirconia, Green Chem. 13 (2011) 2004–2007.

of
https://doi.org/10.1039/c1gc15263g.

[86] R. Arundhathi, T. Mizugaki, T. Mitsudome, K. Jitsukawa, K. Kaneda, Highly

ro
selective hydrogenolysis of glycerol to 1,3-propanediol over a boehmite-
supported platinum/tungsten catalyst, ChemSusChem. 6 (2013) 1345–1347.
https://doi.org/10.1002/cssc.201300196.
-p
[87] G.S. Foo, D. Wei, D.S. Sholl, C. Sievers, Role of Lewis and Brønsted acid sites
re
in the dehydration of glycerol over niobia, ACS Catal. 4 (2014) 3180–3192.
https://doi.org/10.1021/cs5006376.
lP

[88] J.R. Copeland, G.S. Foo, L.A. Harrison, C. Sievers, In situ ATR-IR study on
aqueous phase reforming reactions of glycerol over a Pt/γ-Al2O3 catalyst, Catal.
Today. 205 (2013) 49–59. https://doi.org/10.1016/j.cattod.2012.08.002.
na

[89] M. Tamura, Y. Amada, S. Liu, Z. Yuan, Y. Nakagawa, K. Tomishige, Promoting


effect of Ru on Ir-ReOx/SiO2 catalyst in hydrogenolysis of glycerol, J. Mol.
ur

Catal. A Chem. 389 (2014) 177–187.


https://doi.org/10.1016/j.molcata.2013.09.015.
Jo

[90] Y. Amada, Y. Shinmi, S. Koso, T. Kubota, Y. Nakagawa, K. Tomishige,


Reaction mechanism of the glycerol hydrogenolysis to 1,3-propanediol over Ir-
ReOx/SiO2 catalyst, Appl. Catal. B Environ. 105 (2011) 117–127.
https://doi.org/10.1016/j.apcatb.2011.04.001.

[91] W. Luo, Y. Lyu, L. Gong, H. Du, Selective hydrogenolysis of glycerol to 1,3-


propanediol over egg-shell type Ir–ReOx catalysts, RSC Adv. 6 (2016) 13600–
35

13608. https://doi.org/10.1039/c5ra24808f.

[92] C. Deng, L. Leng, X. Duan, J. Zhou, X. Zhou, W. Yuan, Support effect on the
bimetallic structure of Ir-Re catalysts and their performances in glycerol
hydrogenolysis, J. Mol. Catal. A Chem. 410 (2015) 81–88.
https://doi.org/10.1016/j.molcata.2015.09.009.

[93] C. Deng, X. Duan, J. Zhou, X. Zhou, W. Yuan, S.L. Scott, Ir-Re alloy as a highly
active catalyst for the hydrogenolysis of glycerol to 1,3-propanediol, Catal. Sci.
Technol. 5 (2015) 1540–1547. https://doi.org/10.1039/c4cy01285b.

[94] Y. Nakagawa, X. Ning, Y. Amada, K. Tomishige, Solid acid co-catalyst for the
hydrogenolysis of glycerol to 1,3-propanediol over Ir-ReOx/SiO2, Appl. Catal. A

of
Gen. 434 (2012) 128–134. https://doi.org/10.1016/j.apcata.2012.05.009.

ro
[95] J. Wang, N. Lei, C. Yang, Y. Su, X. Zhao, A. Wang, Effect of promoters on the
selective hydrogenolysis of glycerol over Pt/W‐ containing catalysts, Chinese J.
Catal. 37 (2016) 1513–1519. https://doi.org/10.1016/S1872.
-p
[96] Y. Nakagawa, Y. Shinmi, S. Koso, K. Tomishige, Direct hydrogenolysis of
re
glycerol into 1,3-propanediol over rhenium-modified iridium catalyst, J. Catal.
272 (2010) 191–194. https://doi.org/10.1016/j.jcat.2010.04.009.
lP

[97] D. Li, Z. Zhou, J. Qin, Y. Li, Z. Liu, W. Wu, Cu-WOx-TiO2 Catalysts by


Modified Evaporation-Induced Self-Assembly Method for Glycerol
Hydrogenolysis to 1,3-Propanediol, ChemistrySelect. 3 (2018) 2479–2486.
na

https://doi.org/10.1002/slct.201703090.

[98] A. Kant, Y. He, A. Jawad, X. Li, F. Rezaei, J.D. Smith, A.A. Rownaghi,
Hydrogenolysis of glycerol over Ni, Cu, Zn, and Zr supported on H-beta, Chem.
ur

Eng. J. 317 (2017) 1–8. https://doi.org/10.1016/j.cej.2017.02.064.

[99] L. Huang, Y. Zhu, H. Zheng, G. Ding, Y. Li, Direct conversion of glycerol into
Jo

1,3-propanediol over Cu-H 4SiW12O40/SiO2in vapor phase, Catal. Letters. 131


(2009) 312–320. https://doi.org/10.1007/s10562-009-9914-1.

[100] F.B. Gebretsadik, J. Llorca, Y. Cesteros, M.P. Salagre Carneiro, Hydrogenolysis


of glycidol as alternative route to obtain selectively 1,3-propanediol using MOx
modified Ni-Cu catalysts supported on acid mesoporous saponite,
ChemCatChem. 9 (2017) 3670–3680. https://doi.org/10.1002/cctc.201700536.
36

[101] E. Yoda, A. Ootawa, Dehydration of glycerol on H-MFI zeolite investigated by


FT-IR, Appl. Catal. A Gen. 360 (2009) 66–70.
https://doi.org/10.1016/j.apcata.2009.03.009.

[102] A. Alhanash, E.F. Kozhevnikova, I. V. Kozhevnikov, Gas-phase dehydration of


glycerol to acrolein catalysed by caesium heteropoly salt, Appl. Catal. A Gen.
378 (2010) 11–18. https://doi.org/10.1016/j.apcata.2010.01.043.

[103] S.H. Chai, H.P. Wang, Y. Liang, B.Q. Xu, Sustainable production of acrolein:
Investigation of solid acid-base catalysts for gas-phase dehydration of glycerol,
Green Chem. 9 (2007) 1130–1136. https://doi.org/10.1039/b702200j.

[104] N.S. Gould, B. Xu, Catalyst characterization in the presence of solvent:

of
development of liquid phase structure–activity relationships, Chem. Sci. 9 (2018)
281–287. https://doi.org/10.1039/C7SC03728G.

ro
[105] B. Kasprzyk-Hordern, Chemistry of alumina, reactions in aqueous solution and
its application in water treatment, Adv. Colloid Interface Sci. 110 (2004) 19–48.
-p
https://doi.org/10.1016/j.cis.2004.02.002.
re
[106] K.M.A. Santos, E.M. Albuquerque, G. Innocenti, L.E.P. Borges, C. Sievers,
M.A. Fraga, The Role of Brønsted and Water‐ Tolerant Lewis Acid Sites in the
Cascade Aqueous‐ Phase Reaction of Triose to Lactic Acid, ChemCatChem. 11
lP

(2019) 3054–3063. https://doi.org/10.1002/cctc.201900519.

[107] N.S. Gould, B. Xu, Quantification of acid site densities on zeolites in the
na

presence of solvents via determination of extinction coefficients of adsorbed


pyridine, J. Catal. 358 (2018) 80–88.
https://doi.org/https://doi.org/10.1016/j.jcat.2017.11.016.
ur
Jo

You might also like