66
Formic and Levulinic Acid from Cellulose via Heterogeneous Catalysis Johan Ahlkvist

Formic and Levulinic Acid from Cellulose via Heterogeneous

  • Upload
    others

  • View
    6

  • Download
    1

Embed Size (px)

Citation preview

Page 1: Formic and Levulinic Acid from Cellulose via Heterogeneous

Formic and Levulinic Acid from Cellulose via Heterogeneous Catalysis

Johan Ahlkvist

Page 2: Formic and Levulinic Acid from Cellulose via Heterogeneous

Copyright © Johan Ahlkvist 2014

This work is protected by the Swedish Copyright Legislation (Act 1960:729)

ISBN: 978-91-7459-798-1

Digital version available at http://umu.diva-portal.org/

Printed by: VMC-KBC Umeå

Umeå, Sweden 2014

Page 3: Formic and Levulinic Acid from Cellulose via Heterogeneous

i

Abstract

The chemical industry of today is under increased pressure to develop novel

green materials, bio-fuels as well as sustainable chemicals for the chemical

industry. Indeed, the endeavour is to move towards more eco-friendly cost

efficient production processes and technologies and chemical transformation

of renewables has a central role considering the future sustainable supply of

chemicals and energy needed for societies. In the Nordic countries, the

importance of pulping and paper industry has been particularly pronounced

and the declining European demand on these products as a result of our

digitalizing world has forced the industry to look at alternative sources of

revenue and profitability. In this thesis, the production of levulinic and

formic acid from biomass and macromolecules has been studied. Further,

the optimum reaction conditions as well as the influence of the catalyst and

biomass type were also discussed.

Nordic sulphite and sulphate (Kraft) cellulose originating from two Nordic

pulp mills were used as raw materials in the catalytic synthesis of green

platform chemicals, levulinic and formic acids, respectively. The catalyst of

choice used in this study was a macro-porous, cationic ion-exchange resin,

Amberlyst 70, for which the optimal reaction conditions leading to best

yields were determined. Cellulose from Nordic pulp mills were used as raw

materials in the catalytic one-pot synthesis of ‘green’ levulinic and formic

acid. The kinetic experiments were performed in a temperature range of

150–200 °C and an initial substrate concentration regime ranging from 0.7

to 6.0 wt %. It was concluded that the most important parameters in the one-

pot hydrolysis of biomass were the reaction temperature, initial reactant

concentration, acid type as well as the raw material applied. The reaction

route includes dehydration of glucose to hydroxymethylfurfural as well as its

further rehydration to formic and levulinic acids. The theoretical maximum

yield can hardly be obtained due to formation of humins. For this system,

maximum yields of 59 mol % and 68 mol % were obtained for formic and

levulinic acid, respectively. The maximum yields were separately obtained in

a straight-forward conversion system only containing cellulose, water and

the heterogeneous catalyst. These yields were achieved at a reaction

temperature of 180 °C and an initial cellulose intake of 0.7 wt % and belong

to the upper range for solid catalysts so far presented in the literature.

The reaction network of the various chemical species involved was

investigated and a simple mechanistic approach involving first order

reaction kinetics was developed. The concept introduces a one-pot procedure

providing a feasible route to green platform chemicals obtained via

Page 4: Formic and Levulinic Acid from Cellulose via Heterogeneous

ii

conversion of coniferous soft wood pulp to levulinic and formic acids,

respectively. The model was able to describe the behaviour of the system in a

satisfactory manner (degree of explanation 97.8 %). Since the solid catalyst

proved to exhibit good mechanical strength under the experimental

conditions applied here and a one-pot procedure providing a route to green

platform chemicals was developed. A simplified reaction network of the

various chemical species involved was investigated and a mechanistic

approach involving first order reaction kinetics was developed.

Page 5: Formic and Levulinic Acid from Cellulose via Heterogeneous

iii

Table of Contents

Abstract i Table of Contents iii List of Papers v Abbreviations vi Introduction 1

Levulinic Acid 3 Reaction Network upon Production of Levulinic Acid from Biomass 3 The Influence of Reaction Parameters to the Yield of Levulinic Acid 4 The Influence of Raw Material Varieties 4 The Influence of Reaction Temperature 5 The Influence of Substrate Concentration 6 The Influence of Acid Type 7 The Influence of Acid 8 The Influence of Reaction Time 10 The Influence of Pre-treatment 11 The influence of Additives 13 The Influence of Agitation Speed 15

Materials and methods 17 Chemicals 17 Raw materials 17 Acidic polymer catalyst 18 Reactor system 18 Substrate and catalyst characterization 19 Nitrogen physisorption 20 Scanning electron microscopy 20 Acid-base titration 21 Product analysis 21 Liquid chromatography 22 Gas chromatography 23 Mass spectrometry 23 Derivatization for GC/MS 24 Definitions for the yield calculations of levulinic acid (LA) and formic acid (FA) 24 Modelling software and techniques 25

Results and Discussion 26 Catalyst Characterization 26 Substrate characterization 27 Auto-catalysis 29 Reusability of the catalyst, catalyst deactivation 29 Reaction network 30 The influence of reaction conditions and raw material on product yields 32

Page 6: Formic and Levulinic Acid from Cellulose via Heterogeneous

iv

Comparison between different wood pulps 32 Reaction with D-glucose 33 The influence of reaction temperature and catalyst to substrate dosage 33 The influence of reaction gas and pressure 36 Selectivity 37 Kinetic modelling 38 Fit of the model to experimental data 41

Conclusions 46 Acknowledgements 48 References 49

Page 7: Formic and Levulinic Acid from Cellulose via Heterogeneous

v

List of Papers

I Johan Ahlkvist, Samikannu Ajaikumar, William Larsson, Jyri-

Pekka Mikkola (2013). One-pot catalytic conversion of Nordic pulp

media into green platform chemicals. Applied Catalysis A, 454: 21-

29.

II Johan Ahlkvist, Samikannu Ajaikumar, William Larsson, Johan

Wärnå, Tapio Salmi, Jyri-Pekka Mikkola (2013). Reaction Network

upon One-pot Catalytic Conversion of Pulp. Chemical Engineering

Transactions, 32: 649-654.

III Johan Ahlkvist, Päivi Mäki-Arvela, Jyri-Pekka Mikkola. Macro-

molecules as a Source of Levulinic Acid (2013). The International

Review of Chemical Engineering. (submitted)

IV Johan Ahlkvist, Johan Wärnå, Tapio Salmi, Jyri-Pekka Mikkola

(2013). Experimental and Kinetic Modelling Studies upon

Conversion of Nordic Pulp into Levulinic Acid. Industrial &

Engineering Chemistry Research. (manuscript)

Authors’ contributions

I Planning, kinetic experiments, data analysis, most of the practical

work except some parts of the catalyst characterization, GC and

GC/MS measurements, writing article

II Planning, most of the practical work and data analysis except

mathematical modelling, major writing

III Major writing

IV Planning, kinetic experiments, most of the practical work and data

analysis except mathematical modelling, major writing

Page 8: Formic and Levulinic Acid from Cellulose via Heterogeneous

vi

Abbreviations

5-HMF 5-hydroxymethylfurfural

AD Air dried

AHP Alkaline hydrogen peroxide pretreatment

APPO Aqueous phase partial oxidation

BET Stephen Brunauer, Paul Hugh Emmett and Edward Teller

BJH Barrett, Joyner and Halenda

CE-MS Capillary electrophoresis-mass spectrometry

CPD Critical point dried

D Domsjö DP Degree of polymerization

FA Formic acid

GC Gas chromatography

GC-MS Gas chromatography-mass spectrometry

GVL Gamma-valerolactone

HAS High amylose corn starch

HMDS Hexametyldisilazane

HMF Hydroxymethylfurfural HPLC High-performance liquid chromatography

IUPAC International Union of Pure and Applied Chemistry

LA Levulinic acid

LC-MS Liquid chromatography-mass spectrometry

MCMC Markov Chain Monte Carlo method

MG Mechanical grinding

MilliQ Deionized water

MGRS Mechanical grinding rice straw

MGSERS Mechanical grinding steam explosion rice straw MS Mass spectrometry

M-S Metsä-Serla

MW Microwave

NS Normal starch

PAH Pressurized acid hydrolysis

RH Rice husk

RI Refractive index RID Refractive Index Detector

RS Rice straw

SE Steam explosion SEM Scanning electron microscopy

SERS Steam explosion rice straw SG Superfine grinding

SGRS Superfine grinding rice straw

Page 9: Formic and Levulinic Acid from Cellulose via Heterogeneous

vii

SGSERS Superfine grinding steam explosion rice straw

TMCS Chloromethylsilane (Trimethylchlorosilane) UV-vis Ultraviolet-visible spectroscopy

Page 10: Formic and Levulinic Acid from Cellulose via Heterogeneous

1

Introduction

One of the main reasons for developing biorefinery processes for our

bioresources is to reduce the emission of carbon dioxide and other

greenhouse gases. Also, fossil fuels are a limited source of energy and will be

depleted in the future. However, today the cost of processing renewables to

chemicals and fuels is often too high, partly due to the fact that the

traditional synthesis routes developed and optimized for hydrocarbons are

not well adaptable for renewables [1]. Development of new processes for

woody biomass or other lignocellulosic materials has attracted increasing

attention for some time now. In fact, commercial operations and pilot units

are either integrated as biorefineries combined with adjacent pulp and paper

mills (utilizing e.g. black liquor and other secondary streams as sources of

new products) or built as stand-alone processing units. In order to combine

and re-invent the material streams of the old industry, together with other

processes and practices of chemical industry, a lot of new R&D efforts are

required thus utilizing the existing and new technology solutions to produce

a range of fossil-free alternatives of future.

Lignocellulosic biomass such as wood is a superb raw material for converting

renewables into fuels or fuel additives and chemicals, in view of the fact that

the utilization of this kind of biomass does not compete with feed or food

production. Chemical products produced from food-crops (e.g. corn) are

questionable in a world where there is a lack of adequate nourishment for all

its inhabitants. In fact, sugar crops like e.g. sugarcane is often grown for non-

food purposes (primarily transportation fuel blends) on farmland that

competes with human related food production. After the initial boom of first

generation biochemicals and fuels made from edible biomass, an ever

increasing number of the so-called second generation biofuels produced

from non-edible biomass (e.g. agricultural residues, lignocellulosics and

other waste materials) have been introduced.

Wood, the most common source of biomass in the world, is typically

composed of 40-50% cellulose, 23-33% lignin and 25-35% of hemicellulose

[2]. Cellulose is considered as a promising alternative to non-renewable

natural resources for the sustainable supply of fuels and chemicals in the

future. Currently, extensive research is being carried out worldwide to

identify and study chemical or biological transformation pathways to convert

cellulose into biofuels and feedstock chemicals [3, 4]. Furthermore, the

hydrolysis process of cellulose has been cited as the main gateway to a

biorefinery scheme based on the transformation of carbohydrates [5]. The

Scholler process, which was developed in Germany in the 1920’s, was the

first hydrolysis process for acidic conversion of lignocellulosic biomass [6].

Page 11: Formic and Levulinic Acid from Cellulose via Heterogeneous

2

To date, several studies have reported the conversion of biomass into

levulinic acid (LA) using a dilute mineral acid, such as HCl and H2SO4 as

catalysts [7-10]. Although these hydrolysis reactions proceed smoothly, the

use of the mineral acids causes serious pollution and promotes equipment

corrosion. Moreover, any homogeneous catalysts are difficult to recover from

the reaction products for recycling [11]. Very limited attention has been given

to the use of solid acid catalysis for the depolymerization of cellulose in water

under realistic conditions for the possible exploitation, i.e. temperatures

below about 200 °C (to avoid secondary transformation of the produced

sugars). However, Rackemann et al. [12] summarizes the studies conducted

on the synthesis of levulinic acid using solid acid catalysts from various feed-

stocks. The catalyst used in this experimental study is a macro-porous,

cationic ion-exchange resin, Amberlyst 70 [13]. Today Amberlyst 70 is

reasonably well known and has been used with good results as a catalyst in

many types of reactions. Rinaldi et al. [14], Agirrezabal-Telleria et al. [15] as

well as Hu and Li [16] have recently used Amberlyst 70 in similar type of

reactions as our case.

Several kinetic studies using a various range of cellulosic materials can be

found in the literature. Saeman [17] was one of the pioneers when he in 1945

presented a modelling study based on hydrolysis reaction of Douglas fir into

glucose performed in batch mode. Also, several later studies are based on

Saemans model. In papers by Fagan et al. [18], Thompson et al. [19], and

Franzidis et al. [20], a plug-flow reactor was used to study the kinetics of the

hydrolysis reaction of Kraft paper slurries, Solka Floc as well as filter paper.

Also, Malester and co-workers [21, 22] were using Saemans model when the

kinetics of dilute hydrolysis of cellulose originating from municipal solid

wastes was investigated in a batch reactor. Nevertheless, only a few kinetic

reports [7, 23, 24] are available for the acid-catalyzed hydrolysis of cellulose

to levulinic acid before Girisuta et al. [9] who presented a more detailed

reaction mechanism of levulinic acid (LA) formation from cellulose through

glucose and 5-hydroxymethylfurfural (HMF), including the humin formation

reactions. Jing and Lu [25] also published a model describing LA production

based on a similar reaction mechanism performed in a batch autoclave using

glucose as the reactant. Further, Shen and Wyman [26] presented a

sophisticated model describing LA production in tubular batch reactors from

microcrystalline cellulose and their model accounted for the major

compounds as well as the side-reactions leading to by-products.

Hereupon we describe an experimental and modelling study of the acid-

catalyzed hydrolysis of soft wood dissolving pulp into levulinic acid through

glucose and HMF as well as xylose, furfural and formic acid from the

transformation of xylans in the pulp.

Page 12: Formic and Levulinic Acid from Cellulose via Heterogeneous

3

Our aim was to develop a procedure whereupon a reasonable cellulose-to-

catalyst ratio was applied. Also, avoiding the use of energy-intensive

pretreatments methods for cellulose, such as long-time ball-milling typically

used in most of the literature examples was the criteria set by ourselves [27].

Herein we report the direct catalytic conversion of wood pulp to levulinic

acid and formic acid by sequential hydrolysis, dehydration–hydration

reactions under relatively mild conditions.

Levulinic Acid

The first synthesis of LA by heating sucrose with mineral acids at high

temperature was reported in the 1840’s by the Dutch professor G. J. Mulder

[28]. This versatile renewable platform molecule has also been identified by

the US Department of Energy as one of the `12 top value-added biochemicals

[29].

Levulinic acid is a commodity chemical that finds applications for several

purposes, such as source of polymer resins, animal feed, food as well as

components of flavoring and fragrance industry, textile dyes, additives,

extenders for fuels, antifreeze products, antimicrobial agents, herbicides and

also plasticizers [12, 23]. Traditionally, the most convenient industrial route

to levulinic acid involves production from hexose sugars [23]. More recently,

the utilization of biomass and agricultural wastes for production of levulinic

acid has been subject to increased interest as a research topic, since it would

render the process cheaper due to the high availability and low cost of raw

materials. In fact, even ‘negative’ cost of the raw materials can be achieved

when wastes such as paper mill sludge are applied [30]. Importantly, a novel

process based on the use of waste products would not compete with food

supply. One-pot biomass hydrolysis to sugars followed by their dehydration

can be tuned to produce many different derivatives, including 5-

hydroxymethylfurfural and furfural. The former one can then consequently

be rehydrated to levulinic and formic acids. There are a few reviews available

that concern levulinic acid [12, 30, 31]. In a paper by Timokhin [31], mainly

the properties of levulinic acid and its utilization in organic synthesis were

reviewed, whereas transformation of lignocellulosics to levulinic acid is

emphasized in ref. [12].

Reaction Network upon Production of Levulinic Acid from

Biomass

All in all, one-pot synthesis of levulinic acid from biomass is a demanding

reaction since the whole process most conveniently starts with acid

Page 13: Formic and Levulinic Acid from Cellulose via Heterogeneous

4

hydrolysis of biomass thus giving rise to various issues in terms of waste

produced as well as the occurrence of multiple side reactions. The desired

raw material, lignocellulosic biomass, is mainly composed of cellulose,

hemicelluloses and lignin of which the first two can yield the desired

hexoses, the key intermediates in the production of levulinic acid.

Thereafter, sugars need to be dehydrated and rearranged to 5-

hydroxymethylfurfural and furfural depending on whether hexoses or

pentoses, respectively, are to be dehydrated. Consequently, hexoses can then

further rehydrate to formic and levulinic acids also requiring two equivalents

(moles) of water. Since two different organic carboxylic acids are formed, the

theoretical yield of levulinic acid from hexoses and cellulose has been

reported to reach 64.5 wt % and 71.5 wt % only, respectively [32]. It should,

however, be stated that the real yields of levulinic acid are typically lower

than the theoretical yields due to the harsh reaction conditions required for

the hydrolysis of macromolecules. Thus, the formation of humins is hardly to

be avoided.

The Influence of Reaction Parameters to the Yield of Levulinic

Acid

The influence of various reaction parameters during transformation of

biomass to levulinic acid, such as biomass type, reaction temperature, initial

concentration of feedstock, liquid-to-solid ratio, agitation speed in the

reactor, potential adjacent pretreatments of biomass as well as the influence

of additives is elucidated in this Section. The main emphasis was directed

towards selection of optimal reaction conditions to the yield of levulinic acid

and how it actually can be maximized.

The Influence of Raw Material Varieties

The influence of raw material upon production of levulinic acid is important

since, for example, the presence of lignin results in changes in the optimum

reaction conditions and usually more drastic conditions are needed for

lignocellulosic raw material compared to cellulose. It should be pointed out

here that it is difficult to directly compare the results on the basis of levulinic

acid yield, since in some cases the yield has been given on the basis of initial

raw material, cellulose content (as well as C6 content) in the raw material

(mass) or based on the theoretical yield of levulinic acid. The highest

theoretical yield of levulinic acid reported amounted to 92 % corresponding

to 65.9 % yield, based on the cellulose content in the raw material (paper

sludge) and achieved by a two-stage process at a temperature of 200–215 °C

[33]. Comparative theoretical yields of levulinic acid were achieved from rice

husk (82.9 %) [34], tabacco chops (82.5 %) [30], paddy straw (79.6 %) [10],

Page 14: Formic and Levulinic Acid from Cellulose via Heterogeneous

5

bagasse (78.8 %) [10] and wheat straw (77.4 %) [30]. All produced with a

strong acid, hydrochloric acid (HCl) and from a waste material. When

comparing different raw materials and the yield of levulinic acid, it was

observed that - as also stated by Kang et al. [35] - marine biomass can be

grown rapidly and is easily cultivated without the need for expensive

equipment, compared to land biomass. In addition, annual CO2 absorption

by marine biomass can be five to seven times higher than that of wood-

biomass and the carbohydrate content is higher and can easily be converted

to chemicals through proper chemical processes.

The Influence of Reaction Temperature

The influence of reaction temperature on the levulinic acid production has

been intensively investigated using different feedstock’s, such as cellulose [9,

36-38] wheat straw [8, 30], kernel grain sorghum [23], water hyacinth [39],

poplar sawdust [30], paper sludge [30], olive tree pruning [30] bagasse and

paddy straw [10]. The effect of reaction temperature on the cellulose

conversion into LA was investigated by Peng et al. [38] and the experiments

were conducted at 180–220 °C, respectively. The temperature has a

significant effect on LA yield. When the reaction temperature was increased

above 200 °C, the yield of LA began to fall, and intermediates such as

glucose and HMF couldn’t be detected any more. Higher temperatures thus

could accelerate the rate of chemical reactions, but undesired side reactions

like formation of humins also appeared at the same time. Therefore, higher

temperatures are unfavorable for the extent of reaction. Analogously to the

study of Peng et al. [38], an optimum temperature giving rise to high yield of

levulinic acid has been reported for the transformation of wheat straw at

about 210 °C [40]. Similar results with an optimum temperature was also

obtained by Szabolcs et al. (cellobiose) [41] and Shen et al. (cellulose) [26].

However, by contrast to cellulose transformation by Girisuta et al. [42], the

yield of levulinic acid was greater in the study by Fang et al. [23] at higher

temperatures (200 °C) than that obtained at lower ones (160 °C) using

kernel grain sorghum as the raw material. Very high yields of levulinic acid

were also obtained from bagasse and paddy straw at 220 °C [10]. These

results indicate that the presence of lignin requires a higher temperature in

order to maximize the yield of levulinic acid compared to the case when

cellulose is the raw material of choice although it is known that the humins

originate from glucose and HMF [36]. Furthermore, it has been confirmed

that levulinic acid is stable at 180 °C, even after 24 h [36]. However, it was

also observed that at very high temperatures, above 230 °C, the amount of

levulinic acid decreased in water solution due to dehydration to α- and β-

angelica lactone [10]. Generally, around 170 °C is considered the optimal

reaction temperature above which the humins are readily forming due to

Page 15: Formic and Levulinic Acid from Cellulose via Heterogeneous

6

higher equilibrium concentration of protonated pyranose forms of the sugars

[41].

The Influence of Substrate Concentration

Theoretically, a higher liquid-to-solid ratio should result in a higher yield of

levulinic acid. However, it has further been found that high carbohydrate

loading had an adverse effect on levulinic acid yield in the reaction mixture.

Therefore, one method to maximize the yield is to slowly add the

carbohydrate material to the reaction mixture in order to reduce the

concentration of the carbohydrate in the mixture at any given time and

increase the yield of levulinic acid. In this manner the final concentration of

levulinic acid will be gradually built up, resulting in a decreased cost of

materials and handling [43]. Generally, lower concentration of cellulose

resulted in higher yields of levulinic acid. For each levulinic acid molecule

generated, two molecules of water are needed; otherwise 5-

hydroxymethylfurfural (HMF) may be converted into humins [42]. These

observations are in line with other studies on the acid-catalyzed

decompositions of HMF and glucose, where the low substrate concentrations

favored high LA yields. The highest levulinic acid yield (60 mol %), at full

cellulose conversion, could be obtained at a temperature of 150 °C,

whereupon an initial cellulose addition of 1.7 wt % and a sulfuric acid

concentration of 1 M were applied [9]. The same phenomena have also been

noted in transformation of sorghum flour [23], corn starch [44], cotton straw

[45], and cellulose [26, 46].

Chang et al. [8] observed that the yield of levulinic acid in fact deteriorated

after a certain optimum ratio was reached as noted by Carlson [47]. Also,

Peng et al. [38] observed that the optimal concentration of substrate is very

important in terms of the efficient use of cellulose and the final levulinic acid

(LA) concentration reached. Higher LA yield was achieved at lower substrate

concentrations, but a higher LA concentration in the product is extremely

favorable since it results in lower energy consumption upon purification of

LA and also reduces the amount of waste water produced. When catalyst

dosage was fixed, increasing substrate concentration resulted in more

cellulose available for conversion which, in turn, meant that higher LA

concentration was achieved in the reaction product. However, the yield of LA

was found to decrease significantly by increasing substrate concentration

from 10 wt % to 15 wt %. The plausible reason for this is the product

feedback inhibition, reactivity diminution or insufficient catalyst dosage for

the additional cellulose [38]. In accordance to Yan et al. [10] also too much

or too little water inhibits the hydrolysis process. The primary hydrolysis of

the reaction system is significantly depressed if the water-to-solid ratio is too

Page 16: Formic and Levulinic Acid from Cellulose via Heterogeneous

7

low, at least in the case of bagasse. The reaction with paddy straw shows a

much less sharp change in terms of the yield of levulinic acid reached. The

highest practical yield was achieved when the water-to-solid wt-ratio was 8.5

for both raw materials.

The Influence of Acid Type

In terms of the mineral acid used, it seems that the use of hydrochloric acid

results in better selectivities than that of sulphuric acid. This was observed

e.g. by Galetti et al. [30] when levulinic acid was produced from paper

sludge. However, the highest theoretical yield ever, 92 %, is produced from

paper sludge by sulfuric acid [33]. When water insoluble biomass is

processed, homogeneous acid catalysts generally appear more suitable in

comparison to heterogeneous ones, considering their ability to deliver the

active species into the solid or swelled biomass [48]. So far H2SO4 has been

the most commonly used mineral acid, even if HCl, HBr, and H3PO4 also

have been frequently reported as applicable homogeneous catalysts [49].

However, HBr and H2SO4 have been attempted, albeit with poor

performance compared to HCl at low temperatures [50]. In all cases, diluted

HCl always demonstrated superior behavior in comparison to H2SO4 as the

catalyst. Also, significant deposition of humin solid by-products on the

reactor walls was observed and when the raw biomass had a significant

content of calcium salts, like paper-mill cellulose wastes or tobacco chops,

significant deposition of calcium sulphate was observed, even at lower H2SO4

concentrations. When higher H2SO4 concentrations were adopted, reactor

clogging took place. On the contrary, when HCl was the catalyst of choice

under optimized reaction conditions and at limited exposure times, no

precipitation of salts and humins deposition was observed. Consequently,

HCl always afforded higher LA yields with respect to H2SO4. However, when

Chang et al. [8] reported the optimization of levulinic acid production from

wheat straw using H2SO4 under optimal conditions the yield reached, based

on the weight of raw material, 19.86 %, which is comparable with the 19.3 %

value obtained upon use of HCl. The use of this volatile acid catalyst enables

furthermore the recovery of the produced LA by atmospheric/vacuum

distillation and steam stripping whereupon about 95 % of the acid catalyst

and water can be recycled [51]. These results are important in terms of the

economics of the whole process and considering the possible exploitation of

the solid residue, mainly due to lignin component. In fact, the residual lignin

has great potential for many applications such as a substrate for soil

conditioning, as a free radical scavenger for antioxidation, as a low-density

filler taking place of inorganic fillers, as a component for biodegradable

formulations or to be incinerated to energy. Again, if water-soluble

carbohydrates are employed as substrates, heterogeneous acid catalysts

Page 17: Formic and Levulinic Acid from Cellulose via Heterogeneous

8

appear attractive. A key property of heterogeneous catalysts is their very easy

separation from the reaction products. Furthermore, heterogeneous catalysts

are easily recycled. Also, they are environmentally friendly with respect to

safety and corrosiveness. Up until now, very few heterogeneous acid catalysts

have been tested for the conversion of carbohydrates to yield LA. Examples

thereof include ion-exchange resins [52], zeolites [53], pillared clays [54],

nafion [55], and solid super acid S2O82-/ZrO2

-SiO2-Sm2O3 [20]. NbP was

observed to be active in the conversion of inulin to yield LA under traditional

heating. Yields in the range of 20 wt % were achieved in 1 hour without

significant formation of solid by-products. LA was the predominant

monomer product, while a soluble monomer by-product identified in very

low concentrations was the intermediate 5-HMF. Interestingly, the use of

microwave (MW) irradiation allowed for significant time savings and, at the

same time, enhanced the yield from 20 up to about 30 wt %, corresponding

to 40 % of the theoretical yield [30]. This investigation, carried out in a

slurry batch reactor, appears very promising. Many heterogeneous catalysts

up to now investigated in the synthesis of LA generally required very long

reaction times (15-25 h) to reach comparable yields as e.g. in refs. [52-55].

The results of the study presented above can well be compared with those

obtained with homogeneous HCl under similar thermal treatment

conditions. Nevertheless, the heterogeneous NbP has one significant

drawback which is that NbP remains encapsulated in the solid lignin. The

encapsulated catalyst can be recovered and recycled only after suitable

thermal treatments [30]. In terms of Lewis or Brönsted acids, the former

types are assumed more efficient concerning levulinic acid production [56].

As a general observation, hydrochloric acid appears to be slightly better

catalyst than sulfuric acid although the potential release of chlorine into the

biosphere speaks against the use of HCl [41].

The Influence of Acid

Fang and Hanna [23] observed that higher sulfuric acid concentration also

significantly increased the levulinic acid yield. Yield of levulinic acid was

positively correlated with mineral acid concentration, which greatly boosted

the yield of levulinic acid by the higher rate of hydrolysis of carbohydrates to

organic acids. As observed by many authors, also Peng et al. [38] noticed

that the yield of LA increased rapidly with increasing catalyst dosage until an

optimum level was reached. After this only a minor increase of the LA yield

was achieved. Consequently, the addition of excess catalyst has little or even

negative effect on the LA yield. Accordingly, when the requirements of the

reaction are satisfied, no more should be added. Similar observation was

noted upon LA production catalyzed by other catalysts such as H2SO4 [39]

and HCl [8]. When working with CrCl3 (0.02 M), less catalyst was needed

Page 18: Formic and Levulinic Acid from Cellulose via Heterogeneous

9

than in the case of H2SO4 (0.1–1 M) [8, 9, 39] and HCl (~1.2 M) [10]. In this

case, the most beneficial concentration of CrCl3 was found to be 0.02 M

when considering the economic aspects and the efficiency. Too high

concentration and long reaction time were in favor of causing side-reactions

that may negatively affect the rate of hydrolysis of the cellulose to levulinic

acid [37]. The plausible reason for this is the existence of parallel reactions

upon transformations occurring [45]. A higher catalyst concentration was

not only unnecessary but undesirable in that it was found to complicate the

purification of the final product [57]. If higher acid concentrations are used,

difficulties are encountered in terms of the operation of the process and, at

the same time, no beneficial yield gains could be observed [44]. Chang et al.

[8] also observed that mineral acid with higher concentration also caused

severe corrosion in the reactor equipment – a well known fact. Mineral acid

concentration in combination with increasing reaction temperature,

determines how corrosive the reactants are to the metal parts of the reactor.

At a given temperature there is a maximum allowable concentration of a

mineral acid for a specific metal or alloy [23]. The disposal of the catalysts

will not only waste resources, but also cause environmental pollution, so it is

extremely important that particularly metal catalysts like chromium are

recycled [38].

In order to verify the optimum acid concentration, Szabolcs et al. [41]

performed dehydration of cellulose and cellobiose in a concentration range

of 0.1–2.5 M H2SO4. The optimum conditions for cellobiose were found to be

2 M H2SO4 at 170 °C for 10 min; and for cellulose 2 M H2SO4 at 170 °C for 30

min. In fact, Yan et al. [10] also found that low concentration of

hydrochloride acid resulted in too low concentration of H+ ions thus being

unable to facilitate cellulose hydrolysis and, consequently, a low reaction rate

was observed. In the case of raw materials containing primarily C5 sugars,

low acid concentrations and high temperatures were beneficial when aiming

at high furfural yields. Typical streams of this kind are e.g. pre-hydrolysis

liquors or pulping/papermaking waste streams. On the other hand, as

expected, low acid concentrations and short reaction times lead to low

levulinic acid yields. However, using low acid concentrations is problematic,

because the lignin present in the biomass consumes a fraction of the acid

during the reaction. Due to the low initial concentration of acid, the amount

of acid lost is significant, and the LA yield decreased from 57 % to 27 % upon

third batch of corn stover being added to the reaction mixture [58]. In case

of starch, higher sulfuric acid concentrations favor levulinic acid formation

and a large weight ratio of 0.3 (sulfuric acid-to-starch) was found optimal

[59]. In case of sulfuric acid, the yield of levulinic acid falls at higher

temperatures due to oxidation of raw material by the acid demonstrated by

the presence of hydrogen sulfide formed [50]. Shen and Wyman observed

Page 19: Formic and Levulinic Acid from Cellulose via Heterogeneous

10

that an increase in HCl acid concentration caused a slight increase in the

maximum LA yield from 54.8 % to 59.7 %, respectively. This outcome is

consistent with the role of hydrochloric acid as a catalyst not changing

chemical equilibrium under ideal conditions even though the chemical

reaction rate would increase with increasing acid concentration. As a result,

the maximum LA concentration and yield changed little for the three acid

concentrations applied. However, increasing acid concentration reduced the

time needed to attain the maximum LA yield from about 20 min with an acid

concentration of 0.309 M to only 6 min at 0.926 M. This result confirmed

that acids accelerated the reactions. Moreover, LA concentrations and yields

dropped if the solution was held beyond the optimum time due to by-

product formation [26].

The Influence of Reaction Time

The normal trend in the most reactions carried out at moderate reaction

temperatures is that increasing reaction time also results in significantly

increased levulinic acid yield. However, Chang et al. [8] also observed that

time had a relative significant correlation with other variables. Temperature,

acid concentration and liquid-to-solid ratio influenced significantly the time

required for the over-all reaction to proceed. Moreover, an optimum

temperature existed in the conversion process. It was suggested that if the

hydrolysis was continued for too long a time, the yield decreased, again as

the result of side reactions occurring that consumed the levulinic acid

already formed. Peng et al. [38] noted that the amount of LA increased very

fast during the first 140 min. After this it was observed that only a slight rise

occurred after 140 min of reaction time. Finally, it was found that 180 min

was sufficient for completion of the hydrolysis reaction. Yan et al. [10]

observed that bagasse was hydrolyzed rapidly during the first 10 min after

which the hydrolysis rate slowed down and the yield rose gently during the

last 20 min. whereas the rate of the hydrolysis reaction for paddy straw

reached its highest point after 25 min. The reaction proceeded well for 10

min, and then the reaction progressed much more sluggishly. The reason for

these differences in kinetics may lie in the different cellulose content,

different cellulose type and the intrinsic composition structures. The

reaction was interrupted after 45 min because side reactions started to occur

when the degradation products were exposed to high temperatures for a

prolonged period of time. Thomas [57] also noticed that continuing the

digestion of levulose, dextrose or starch with hydrochloric acid for periods

longer than one hour actually resulted in decreased yields of levulinic acid,

the effect decreasing in this order. Alonso et al. [58] observed that short

reaction times and low acid concentrations prevent furfural degradation

reactions and leave the cellulose mostly unconverted, leading to low LA

Page 20: Formic and Levulinic Acid from Cellulose via Heterogeneous

11

yields (less than 11 %). On the other hand, Cha et al. [59] were using a

reactive extrusion process to hydrolyze starch in the presence of 5 wt %

sulfuric acid. They noticed that the levulinic acid yield increased as reaction

time increased from 20 to 60 min. However, the yields in the samples

containing 25 wt % extruded starch and further reacted at 40 and 60 min of

reaction time were not significantly different. Wang et al. [37] also

investigated the effect of the reaction temperature on the levulinic acid yield.

It was concluded that the reaction rate is very sensitive to the temperature

and the rate was reduced dramatically at lower temperatures. As already

stated, higher temperatures favor the production of more LA, also observed

by Kang et al. [35].

The Influence of Pre-treatment

First of all it should be pointed out here that it is extremely difficult to

directly compare the values of levulinic acid yield reported in the literature.

It is very important to note that some of the raw materials were not really

‘pure’ raw materials in its proper sense. In fact, several starting materials

had already been exposed to some type of pre-treatment or purification

processes. Because of these reasons the results obtained may display various

‘unfair’ advantages in comparison to studies where the starting materials

were not pre-treated in any form. When Galletti et al. [30] were using poplar

sawdust as a substrate, the performance could be improved not only by

adopting MW irradiation, but also by carrying out traditional heating in two

steps: a pre-hydrolysis at lower temperature and a successive conversion

step at 200 °C. The pretreatment at a lower temperature, such as 120 °C for

2 h, dissolves part of the hemicelluloses thus yielding cellulose more easily

accessible and, consequently, increasing the yield of levulinic acid during the

second hydrolysis step. Upon a preliminary screening, the adoption of the

pre-hydrolysis step appears as a more immediate approach in order to

improve the LA yields from raw biomass in comparison to MW irradiation

[30].

Chen et al. [60] found out that steam explosion (SE) combined with

superfine grinding (SG) of rice straw (RS) could effectively increase levulinic

acid (LA) yield due to the reduction of particle size and the increase of

specific surface area which would enhancing the accessibility of cellulose

[61]. It was found that LA yield reached upon a SGSERS sequence amounted

to 22.8 %, accounting for 70 % of the theoretical value, which was 61 %

higher than obtained when applying the sequence mechanical grinding (MG)

SERS. When comparing MGRS and SGRS sequences, LA yield of MGSERS

and SGSERS were increased by 95 % and 68 %, respectively. Pretreatment

lowers the polymerization grade and the crystallinity of the cellulose, which

Page 21: Formic and Levulinic Acid from Cellulose via Heterogeneous

12

can exert a significant influence on the hydrolysis yield. Additionally, the

chemical pretreatment stage of the raw material aims to partially remove the

lignin bound to the hemicelluloses without degrading the cellulosic chain

[62]. For this reason Bevilaqua et al. [34] conducted in their study some

pretreatments focusing on the delignification of the biomass and the

elimination of non-contributing substances, like extractives and metals. To

establish the best precondition for pressurized acid hydrolysis (PAH), rice

husk (RH) was submitted to different pretreatments. From an analysis of the

pretreatment procedures, it can be seen that the strategy of applying aqueous

extraction via Soxhlet extraction resulted in a higher LA yield (18.2 g L-1),

probably due to the elimination of interfering metals and an improved

performance PAH on RH. There was no apparent advantage in pretreatment

with oxalic acid aiming at removing metals or with hydrogen peroxide upon

an alkaline medium (delignification). It is evident, in this case, that a part of

the glucose was released during the pretreatment. Soxhlet extraction of RH

with an organic solvent, with the aim of eliminating extractives and

pretreatment with chlorite (RH delignification and facilitation of cellulose

hydrolysis), did not exert any significant influence on the LA yield. Also,

Ghorpade and Hanna [63] patented a process for producing levulinic acid

using an extrusion process, which is advantageous in that it is continuous,

requires fewer steps and reduces reaction time compared with a batch

reactor process. However, the extrusion process is limited due to the loss of

evaporated gases containing HMF. In Cha and Hanna’s [59] study extrusion

has been used as pre-treatment in a ‘two-stage’ process. After extrusion, the

extrudates were reacted with additional water and sulfuric acid in a

pressurized reactor. Extrusion processing increased levulinic acid yield

because the extrusion processing increased the hydrolysis of starch in the

presence of an acid. The yields of glucose and levulinic acid increased at the

higher extrusion temperature and slower screw speed (longer residence

time). High amylose corn starch (HAS) was also converted more readily to

glucose and levulinic acid than normal starch (NS). Extrusion of both starch

varieties resulted in even higher yields.

The cellulose hydrolysis study of Stijn Van de Vyver et al. [64] indicates that

the reaction rate is strongly dependent on the degree of crystallinity, well in

agreement with previous studies. Since the cleavage of the b-1,4-glycosidic

bonds is considered to be rate-determining, hydrolysis of ball-milled

cellulose gave the highest conversion, carbon efficiency and selectivity

among the five cellulose fractions, with a LA yield higher than 27 %.

Interestingly, the observed LA yield corresponds well to the results for

cellobiose, cellohexaose and even starch. The latter observation should be

explained based on the low crystallinity index of 4.1 % for ball-milled

cellulose when compared to the substantially higher crystallinities for

Page 22: Formic and Levulinic Acid from Cellulose via Heterogeneous

13

commercial cellulose substrates. Generally, a limited accessibility of the b-

1,4-glycosidic linkages of cellulose is likely responsible for slower hydrolysis

rates at higher crystallinity. Alonso et al. [58] have been using corn stover

which has been subjected to an alkaline hydrogen peroxide pretreatment

(AHP), which partially removes the lignin [65, 66]. By using AHP corn

stover, the yield of LA increases slightly to 69 % in the first batch, and

remained high (overall 53 %) after the third batch of stover. Removing the

lignin was not necessary for the process; however the sulfuric acid was

retained in the reactor, and not consumed by the lignin. The use of the AHP

corn stover allows for lower acid concentrations and shorter reaction times

to be employed, resulting in furfural yields up to 96 %, with low conversion

of cellulose.

The influence of Additives

Already in the 1940’s Thompson [43] noticed when converting hydrol

(mother liquor from industrial, crystalline corn sugar) that the presence of

sodium chloride in the hydrolysis mixture will not only increase the speed of

the reaction but will also increase the yield of levulinic acid. The

phenomenon was at that time not fully understood, but it was speculated

that the presence of the sodium chloride in the reaction mixture increases

the activity of the mineral acid. The best results were obtained when the salt-

to-water wt-ratio was 6. Galletti et al. [30] also observed that the addition of

a cheap electrolyte to the reaction medium, such as NaCl, increased the LA

yield for some substrates such as wheat straw and tobacco chops. When

water slurry of tobacco chops powder containing 25 wt % of cellulose was

employed as the starting material, the actual yield of LA reached about 83 %

of the theoretical yield. The reason given was the combined effect of both the

addition of the electrolyte and the pre-hydrolysis step at 120 °C. This yield is

one of the highest ever reported using a waste biomass in a batch process.

Also, more generally, the yields reached with the different substrates are

promising if compared with those that have been reported until now [51]. In

essence, it is known that the addition of an electrolyte can increase the rate

of the heterogeneous acid hydrolysis of cellulose to convert to glucose and

then LA, while enhancing the accessibility of the rest portion of the cellulose

[67]. Liu et al. [68] also reported that inorganic salts could increase the

hydrolysis rate of hemicellulose components during dilute acid

pretreatment. This salt-effect has been explained by applying the Donnan’s

theory of membrane equilibria to the heterogeneous hydrolysis of the

cellulose in dilute acid. The difference of concentration of H3O+ ions in the

two phases can be minimized by the addition of an inert, diffusible salt. The

addition of an electrolyte should result in an increase in the concentration of

H3O+ ions in the cellulose phase, thus increasing its swelling and the rate of

Page 23: Formic and Levulinic Acid from Cellulose via Heterogeneous

14

protonation of reaction sites [30]. The effect of added salts on Nafion as a

solid acid catalyst for the hydrolysis of cellulose was also investigated by

Potvin et al. [69]. The result of the investigation was that the yield of

levulinic acid can be increased 5-fold from 14 % to 72 % upon incorporation

of a 25 wt % NaCl solution. The use of a different salt, such as potassium

chloride, was also investigated to determine if the salt effect was general.

Reactions using 10 wt % KCl was less powerful than reactions with

comparable concentrations of NaCl. On the contrary, 5 wt % KCl solution

induced higher product yields than the use of 5 wt % NaCl solution. The

yields were not significantly different between the sodium and potassium

chloride trials indicating that cation size may not have a large influence on

the product yield. They hypothesize that the greater concentration of NaCl

allows for a greater concentration of ions in solution. These charged particles

interact with the hydrogen bonding network of the cellulose structure. The

dramatic increase in yield indicates that sodium chloride is effective in

interrupting the hydrogen bonding network at high temperatures and

pressures. Wettstein et al. [70] have also been using a solute additive for the

conversion of cellulose. They have demonstrated a biphasic process to make

GVL (gamma-valerolactone), in which cellulose is deconstructed to produce

LA and FA using a dilute aqueous solution of hydrochloric acid (HCl) with a

sufficient amount of a dissolved solute in the solution to yield a biphasic

system with GVL. The top phase is an organic phase rich in GVL and LA,

while the bottom phase is an aqueous phase rich in mineral acid. The

maximum yield of LA achieved was 72 % after the first batch of cellulose and

using 1.25 M HCl saturated with 35 wt % NaCl and a 1:1 aqueous-to-organic

wt-ratio. During the reaction, the GVL continuously extracts the LA and FA

from the aqueous phase, and solubilizes the humins that form. The aqueous

phase can be used to carry out additional cellulose deconstructions. The

organic phase is then transferred to a hydrogenation reactor, where the LA is

converted into GVL (using a Ru–Sn/C catalyst) that can be used directly or

further upgraded to chemicals [71-73] or transportation fuels [74, 75]. A

biphasic reaction system using an aqueous phase containing a phase

modifier (e.g., salt and sugars) and GVL as a solvent offers significant

advantages in the production of GVL from biomass derived cellulose. This

system solves the following three important issues that limit current

processing approaches: eliminating the need for a filtration step after

cellulose deconstruction, eliminating the need for a step to separate the

product from the solvent (the product is the solvent), and using a renewable

solvent. Alonso et al. [58] have taken the use of GVL a step further in the

conversion of lignocellulosic biomass (corn stover). By using GVL as a

solvent broadens the optimal processing conditions allowing the conversion

of hemicelluloses and cellulose simultaneously in a single reactor, thus

eliminating pre-treatment steps to fractionate biomass and simplifying

Page 24: Formic and Levulinic Acid from Cellulose via Heterogeneous

15

product separation since GVL is a product of the process. Other important

advantages are elimination of reactor clogging and other solids handling

problems because GVL effectively solubilizes the biomass (and humins). The

GVL solvent also increases the rate of cellulose conversion and decreases the

rate of furfural degradation and is completely miscible with water, allowing

wet biomass to be used in the process.

Lin et al. [46] was converting cellulose over a ZrO2 catalyst by a one-pot

catalytic aqueous phase partial oxidation (APPO) process. Under the APPO

conditions, the molar ratio of formic acid to LA was only ~0.35 with a ZrO2

catalyst. The low formic/levulinic acid molar ratio implies that, unlike acid

hydrolysis, APPO follows a pathway in which formic acid is no longer a

major by-product. Nevertheless, the LA yield is sensitive to the oxygen

partial pressure and the maximum yield of LA was obtained with 2.8 % O2.

Redox, rather than acid-base, properties of the ZrO2 catalyst play the central

role in the oxidative deconstruction of cellulose in the APPO reactions.

Higher oxygen partial pressures resulted in lower LA yields as LA is further

degraded to acetic acid and CO2 by strong oxidation. The results also show

that there is a decrease of 5 % in LA yield when not feeding O2 compared to

the case with 2.8 % O2 loading. Another unique feature of the APPO process

is the low yield of water insoluble humins. It is proposed that HMF is the

precursor of humins during the acid hydrolysis of cellulose [76]. However,

the proposed APPO reaction mechanism indicates that gluconic acid and not

HMF, is the key intermediate to produce LA. This is very likely the reason

why the yield of humins was lower than in a pure acid catalyzed hydrolysis

reaction.

The Influence of Agitation Speed

It is well known that agitation speed is an important factor in solid-liquid

phase catalytic reaction systems. If the agitation speed is too low the reaction

may suffer from low reaction rate, low conversion and low product yields etc.

This is a result of high interfacial mass transfer resistances between the two

phases. Increasing the agitation speed might increase the contacting area

and relative velocity between the two phases, and hence diminish any

external diffusion limitations [77]. The effect of agitation speed on the yields

of reaction products was, in fact, investigated by Peng et al. [38]. The

hydrolytic reaction was carried out by varying the agitation speed from 0–

600 rpm using cellulose as a feedstock and chromium chloride as a catalyst.

As can be seen, the interfacial mass transfer resistances between the solid

surface and water phase appear as negligible when the agitation speed is

higher than 200 rpm considering levulinic acid yield. However, one could

not observe any clear influence of the agitation speed in terms of the yields of

Page 25: Formic and Levulinic Acid from Cellulose via Heterogeneous

16

glucose and 5-hydroximethylfurfural (HMF) - most probably because they

are properly dissolved in the liquid and thus can be well contacted with the

catalyst. Nevertheless, the reader should be reminded that the design of the

stirrer as well as the baffling of the reaction vessel have a significant

influence in terms of the mixing performance.

Page 26: Formic and Levulinic Acid from Cellulose via Heterogeneous

17

Materials and methods

Chemicals

The commercially acquired reagents and chemicals used were of analytical

grade and used without further purifications or other treatment procedures.

D-glucose (VWR International, >99 %), 5-hydroxymethylfurfural (Aldrich,

99 %), levulinic acid (Janssen Chimica, 98+ %) and formic acid (Sigma-

Aldrich, 98-100 %) were used to facilitate the calibration curves needed to

follow the reaction kinetics. HPLC grade sulfuric acid (Fluka, 50 %) was used

for eluent preparation. Pyridine (Fluka, ≥99 %), hexamethyldisilazane

(Fluka, ≥99 %) and chloromethylsilane (Fluka ≥99 %) were used for the

silylation. Phenolphthalein (Fluka) and Sodium chloride (Sigma-Aldrich,

≥99 %) were used for the acid titration of catalyst. Whenever water was used,

deionized water (MilliQ) was used to prepare all the solutions.

Raw materials

Most of the focus has been given to investigate the hydrolytic conversion of

Nordic, bleached, industrial pulp obtained as a gift from the Aditya Birla

Domsjö pulp mill in Sweden (a mixture of softwoods Norway Spruce and

Scots Pine), however Metsä-Serlas kraft pulp from Finland (hardwood birch,

Betula pendula) were compared as a competing raw material. The softwood

pulp utilized has a typical degree of polymerization (DP) in the range of 950

as an average whereas the DP of hardwood approaches 1900 on the basis of

viscosity measurements (Scandinavian Standard SCAN-CM 15:99) [78].

Kraft (sulphate) cellulose typically contains significant amounts of residual

hemicelluloses (16 wt % xylose pentose) [79] whereas sulphite pulp is lean on

hemicelluloses (sum of xylose and mannose pentoses comprise 4.2 wt %)

[80]. Native Birch wood (Betula pendula) has been reported to contain 23 %

of xylose, bound in the form of the hemicellulose xylanes [81].

Monosaccharide’s originating from hemicelluloses of four different bleached

birch pulps from Finnish paper mills has been studied earlier and it was

found that the average amount of xylose found in the pulps resides in the

range of 14.3 wt % (13.6–15.3 wt %) [82]. The pulps were utilized without

any prior chemical treatments but both the sulphite and sulphate pulp were

initially disintegrated by a centrifugal mill (ZM 200, Retch, Germany) with a

cutting sieve to an average particle size of 0.35 mm in order to facilitate a

realistic comparison. Since the degradation and transformations of the

cellulose polymers initially proceeds via hydrolysis to glucose, also

transformation of pure D-glucose was investigated in order to obtain insight

into the general reaction behaviour.

Page 27: Formic and Levulinic Acid from Cellulose via Heterogeneous

18

Acidic polymer catalyst

The acidic polymer based catalyst, Amberlyst 70, was obtained as a gift from

the Rohm and Haas Company [13]. The catalyst is a macro-porous,

polymeric, cationic ion exchange resin material applicable in high-

temperature applications and is designed to be used in various processes

such as olefin hydration, esterification and aromatic alkylation. In the

aforementioned applications, Amberlyst 70 offers good performance over

conventional resins due to its high thermal stability (up to 190 °C), catalytic

activity and good mechanical strength. Importantly, very low sulphonate

leaching has also been confirmed. The catalyst has an appearance as dark

brown and spherical particles with an effective particle size of around 0.50

mm. The average pore diameter corresponds to 220 Å (22 nm) and the

nitrogen physisorption (BET area) to 36 m2 g−1 [13, 83, 84].

Reactor system

All kinetic experiments were carried out batch wise in a standard 300 ml

laboratory scale, three-phase stainless-steel autoclave (Parr Inc, USA)

designed for high pressures and temperatures (see Fig. 1.). The temperature

was measured with a thermocouple and automatically adjusted with the

inbuilt temperature regulator (Parr 4848 Reactor Controller). The autoclave

was equipped with a sampling outlet fitted with a 7 μm metal sinter frit

designed to prevent the catalyst particles as well as pulp pieces to leave the

reaction vessel upon sampling of the liquid phase. The hydrolytic

degradation of wood pulp took place in the presence of catalysts in an

aqueous media and high stirring rates (1000 rpm) were applied in order to

eliminate any external mass transfer limitations. The reaction parameters

studied, such as the reaction time, temperature, pressure as well as the

nature of the catalyst was of crucial importance in terms of optimal

conversion of the biopolymers to platform chemicals. Each and every

experiment was commenced by charging the reaction vessel with a slurry

containing 1–9 g of pulp dispersed in 141–149 ml of deionized water along

with 1–9 g of catalyst. Before heating the reaction mixture to the intended

temperature (150–200 °C), the system was degassed (<0.01 atm) with a high

power vacuum pump to remove any residual air in the solution. After

reaching the preset temperature the agitation was engaged, gas atmosphere

(CO2, Ar, H2) was applied at set pressure (3–50 bar excluding vapour

pressure from water) and this was considered as the initial start of the

reaction. A blank reaction (without catalyst) has also been performed at 200

°C with a higher CO2 pressure than normal reaction conditions (90 bar

instead of 50 bar) to exclude that carbonic acid catalyzes the reaction.

Throughout the experimental run, samples were taken at suitable intervals

Page 28: Formic and Levulinic Acid from Cellulose via Heterogeneous

19

and the sampling line was flushed with small volumes of the reaction

solution in between the samples to avoid any contamination between

consecutive samples. For the sake of precaution, the samples were once more

filtered and the composition was determined by means of high-performance

liquid chromatography (HPLC).

TI

PI

RPMI

Reactor Controller

RPMIC

TIC

PI

PC

Reactor

Sinter

Thermocouple

Heating jacket

Agitator

Coil

Sampling

Gas inlet

Water in/outlet

PI Pressure indicator

TI Temperature indicator

TIC Temperature indicator & controller

RPMI RPM indicator

RPMIC RPM indicator & controller

Figure 1. Schematic picture of the experimental set-up used for the catalytic transformation

of wood pulp in batch mode.

Substrate and catalyst characterization

The substrates and catalyst were characterized by means of the following

methods: scanning electron microscopy, nitrogen physisorption and acid-

base titration. Also the wood pulps were characterized by scanning Electron

Microscopy (SEM): The morphology of the two wood pulps, kraft pulp from

Metsä-Serla and the sulfite pulp from Aditya Birla (Domsjö) was investigated

by SEM (Cambridge Leica Stereoscan 360). The samples were prepared with

a thin layer (50 Å) of gold and carbon before they were introduced to the

measurement chamber. Some of the samples were also critical-point-dried

for a comparison with air-dried samples. Also the morphology of the

Amberlyst 70 catalyst was investigated by SEM (Cambridge Stereoscan 360).

Further nitrogen adsorption–desorption measurements were conducted for

the fresh and spent catalyst using Micromeritics TriStar 3000 porosimeter.

Adsorption–desorption isotherms were recorded at −196 °C after degassed

the samples at 120 °C. Surface areas were calculated by B.E.T method and

pore volumes were calculated from desorption isotherm. Pore size

Page 29: Formic and Levulinic Acid from Cellulose via Heterogeneous

20

distribution was estimated using the Barrett, Joyner and Halenda (BJH)

algorithm (ASAP-2010) available as built-in software from Micromeritics.

The acid site concentration of the catalyst was measured by means of a

conventional titration method [85]. 1 g of the resin was added to about 50 ml

of NaCl solution (200 g L-1). The cation exchange between H+ and Na+ was

allowed to take place for 24 h upon stirring of the system. The mixture was

then titrated with a 0.1 N NaOH solution. The measurement was carried out

on fresh catalyst and on spent catalysts exposed to the reaction temperatures

of 190 and 200 °C, respectively.

Nitrogen physisorption

Surface area, pore size distribution and pore volume is an attribute that is

used by catalyst manufacturers and users to monitor the activity and stability

of catalysts as well as a quality control tool during catalyst manufacture.

There are different methods used to measure surface area and each method

can yield different results. In physisorption, a monolayer formation of inert

gas, mostly nitrogen, is adsorbed by weak Van der Waals attractions on the

surface of a solid material or, in case of porous materials, also on the surface

of pores. Most widely known is the determination of the B.E.T (Stephen

Brunauer, Paul Hugh Emmett, and Edward Teller) surface area by gas

adsorption. Adsorption of nitrogen at a temperature of 77 K leads to a so-

called adsorption isotherm, sometimes referred to as the B.E.T. isotherm,

which is the most common way to express measured porosity of materials.

The principle of capillary condensation is usually applied to assess the

presence of pores, pore volume and pore size distribution. Pre-treatment of

the sample is performed at elevated temperature in vacuum or either by a

flowing inert gas in order to remove any contaminants.

Scanning electron microscopy

A scanning electron microscope (SEM) is a type of electron microscope that

produces micrographs of a solid specimen by scanning it with a focused

beam of high-energy electrons. The electrons interact with atoms in the

sample, producing signals that can be detected and that contain information

about the surface topography, chemical composition, crystalline structure

and orientation of materials making up the sample. In most applications,

information is collected over a selected area of the surface of the sample, and

a micrograph is generated that displays spatial variations in these properties.

Generally the electron beam is scanned in a raster scan pattern, and the

beam's position is combined with the detected signal. A two-dimensional

micrograph with a resolution better than 1 nanometer can be produced. The

most common mode of detection is by secondary electrons emitted by atoms

Page 30: Formic and Levulinic Acid from Cellulose via Heterogeneous

21

excited by the electron beam. The number of secondary electrons is a

function of the angle between the surface and the beam. On a flat surface, the

plume of secondary electrons is mostly contained by the sample. However on

a tilted surface, the plume is partially exposed and more electrons are

emitted. By scanning the sample and detecting the secondary electrons, a

micrograph displaying the tilt of the surface is generated.

Acid-base titration

The acid-base titration process consists of the addition of a solution with

known concentration, from a buret, to a measured volume of a solution of

the other reactant or to a weighed sample dissolved in water, until the same

number of equivalents of each substance has been used. The end point in the

titration is detected by means of a suitable indicator. Phenolphthalein, which

is red in basic solution and colorless in acid solution, is often used for acid-

base titrations. The end point, where the permanent color change just occurs

after thorough mixing, is usually very sharp. Often a drop or a fraction of a

drop will bring about a color change.

Product analysis

The quantitative composition of the samples in the liquid-phase was

analyzed off-line with high-performance liquid chromatography (HPLC,

Agilent Technologies 1200 Series) equipped with a refractive index (RI)

detector, a degasser, a double-channel binary pump and an auto sampler.

The analytical column of choice (Bio-Rad Aminex HPX-87H) is a 300 x 7.8

mm, a pre-packed HPLC column suited for carbohydrate and acidic

compounds analysis supplied in the proton form containing 9 μm particle

size with 8 % cross linkage and is designed to operate at a pH range of 1–3. 5

mM H2SO4 was utilized as the mobile phase with a flow rate of 0.5 cm3 min-1

and temperature of 60 °C. The samples were injected into the HPLCs directly

after the sample was taken without any pre-treatment other than filtering (2

µm) to prevent column damage. The analysis for a sample was completed

within 1 h. The concentrations of each compound in the liquid-phase mixture

were determined using calibration curves obtained by analyzing standard

solutions with known concentrations. For the sake of the clarity, instead of

more commonly applied UV-vis detector, refractive index (RI) detection is a

must upon analysis of carbohydrates and their derivatives since UV detection

is not capable of resolving most of the species. A Bio-Rad Aminex HPX-87C

carbohydrate column was also tested, however the acidic main compounds,

formic- and levulinic acid, peaks overlapped each other (same retention

times).

Page 31: Formic and Levulinic Acid from Cellulose via Heterogeneous

22

Other possible by-products of the acid-catalyzed hydrolysis of cellulose are

gas-phase components from thermal degradation (eventually catalyzed by

the autoclave walls) reactions of reactants, intermediates and/or products.

To gain insights into the extent of these reactions, a GC study has been

carried out to analyze the gas phase composition during an experiment

performed at 180 ◦C without any gas addition. A sample of non-condensable

gases were taken with a gas syringe and analyzed with a Varian 490-GC

Micro-GC. To identify the compounds in the liquid phase gas

chromatography-mass spectrometry (GC-MS) analysis have been performed.

The samples were derivatized to make the compounds volatile and stable,

suitable for gas chromatography analysis. About three drops (150 µl) of the

liquid samples were transferred to a little test tube and evaporated in a water

bath operating at 40 °C under nitrogen flow. As a reference a blank sample

containing only distilled water was dried and used. The dried samples were

silylated by addition of 150-200 µl of pyridine, 200 µl of HMDS silylation

reagent (hexametyldisilazane) and, finally, 100 µl of TMCS reagent

(chloromethylsilane). After half an hour of stirring at room temperature the

samples were ready for analysis. The samples were transferred into small

ampoules and analyzed with GC-MS. The GC-MS was equipped with an HP-1

column (25 m x 0.2 mm x 0.11 µm) and a flow of 0.8 ml/min (helium) was

used and the split was 15 ml/min. The heating program started at 60 °C with

a heating rate of 8 °C/min up to 325 °C.

Liquid chromatography

Although gas chromatography is widely used, it is limited to samples that are

thermally stable and easily volatilized. Nonvolatile samples, such as peptides

and carbohydrates, can be analyzed by GC, but only after they have been

made more volatile by a suitable chemical derivatization. For this reason, the

various techniques included within the general scope of liquid

chromatography are among the most commonly used separation techniques.

In high-performance liquid chromatography (HPLC), a liquid sample, or a

solid sample dissolved in a suitable solvent, is carried through a

chromatographic column by a liquid mobile phase. Separation is determined

by solute/stationary-phase interactions, including liquid–solid adsorption,

liquid–liquid partitioning, ion exchange and size exclusion, and by

solute/mobile-phase interactions. In each case, however, the basic

instrumentation is essentially the same [86]

Over the past three to four decades, high-performance liquid

chromatography (HPLC) has been the method of choice for the

determination of carbohydrate species, and as a result a large number of

HPLC methods have been developed to determine a wide variety of

Page 32: Formic and Levulinic Acid from Cellulose via Heterogeneous

23

carbohydrate samples. One of the most challenging areas of analytical

chemistry is the development of selective and sensitive methods for analyses

of carbohydrates. The main difficulties in the analysis of carbohydrates arise

from their considerable number of isomeric forms due to the various

possible configurations of the monosaccharides. Oligo- and polysaccharides

are composed of different combinations of them, forming linear or branched

polymeric species, many of which may differ only in the position of

attachment and anomeric configuration of the glycosidic linkages. RI

detection is the most popular detection technique for carbohydrates in HPLC

but a serious drawback is that it can only be used to monitor isocratic

separations [87].

Gas chromatography

In gas chromatography (GC) the sample, which may be a gas or liquid, is

injected into a stream of an inert gaseous mobile phase (carrier gas). The

sample is carried through a packed or capillary column where the sample’s

components separate based on their ability to distribute themselves between

the mobile and stationary phases. The most common mobile phases for GC

are He, Ar, and N2, which have the advantage of being chemically inert

toward both the sample and the stationary phase. The choice of which carrier

gas to use is often determined by the instrument’s detector [86].

Mass spectrometry

A mass spectrometer consists of three major parts: the ion source, the mass

analyzer, and the detector. The first mass spectrometer was used in England

by JJ Thompson in 1912, and by F. W. Aston in 1919, but the instrument that

resembles today's mass spectrometers was first constructed in 1932 [88].

Over the last two decades mass spectrometry has become one of the central

techniques in analytical chemistry. Its importance is now comparable to that

of the more traditional electrophoresis, liquid and gas separations

techniques, and it is often used in conjunction with them as so-called

“hyphenated” techniques, such as CE-MS, LC-MS and GC-MS. Mass

spectrometry is used for both quantitative and qualitative analysis and is

based on production of unstable ions from neutral species, that is, atoms,

molecules or clusters and the study of the decomposition (fragmentation) of

these. The separation should also be independent of the chemical

conformation of the species. When a compound is ionized in vacuum the

compound decays whereby new ions are formed. The apparatus must be kept

under vacuum (typically <10-6 mmHg) to prevent reactions between the

unstable fragment ions and gases, resulting in mass spectra that cannot be

interpreted. The ions are separated and deposited in a chart relative to the

Page 33: Formic and Levulinic Acid from Cellulose via Heterogeneous

24

proportion of ions towards the ion mass, resulting in a so-called mass

spectrum. Each compound decomposes in a manner which is unique

precisely for this compound and thus a mass spectrum is obtained that is

specific for that molecule. The decomposition gives us information about the

association structure, which allows the mass spectrum can be interpreted

and compound can be identified. All mass analyzers presently in use are

based on electromagnetism so ions are required to obtain separation. The

different types of mass spectrometers differ in regard to the mechanism by

which ions are separated. There are several types of mass analyzers used in

mass spectrometric research and they can be divided into different

categories, such as magnetic or pure electric, scanning/non-scanning (pulse

based), and trapping/non-trapping analyzers [89].

Derivatization for GC/MS

Silylation is the most widely used derivatization procedure for gas

chromatography and mass spectrometry analysis and have been used since

the late fifties. A wide variety of products (and functional groups) like

alcohols, carboxylic acids, amines, thiols, and phosphates etc. can be

derivatized by silylation. The process involves the replacement of an active

proton with a trialkylsilyl group, typically trimethylsilyl (-SiMe3). Compared

to their parent compounds, a silyl derivative generally are more volatile, less

polar, and more thermally stable and enables the GC-MS analysis of many

involatile and too unstable compounds [90].

Definitions for the yield calculations of levulinic acid (LA) and

formic acid (FA)

The yield of levulinic acid (LA) on molar basis (𝒀𝑳𝑨) was defined as the ratio

between the LA concentration in the reaction product (𝑪𝑳𝑨) and the concentration of the available C6-sugars (𝑪𝑪𝟔) in cellulose and hemicelluloses

fractions in the pulp:

𝑌𝐿𝐴(𝑚𝑜𝑙 %) =𝐶𝐿𝐴

𝐶𝐶6× 100 % (1)

It is also possible to define the yield of LA on a weight base (𝒀𝑳𝑨,𝒘𝒕),

corresponding to the mass ratio between the LA and the available C6-sugars

in the pulp:

Page 34: Formic and Levulinic Acid from Cellulose via Heterogeneous

25

𝑌𝐿𝐴,𝑤𝑡(𝑤𝑡 %) =𝐶𝐿𝐴 ×𝑀𝐿𝐴𝐶𝐶6 ×𝑀𝐶6

× 100 % (2)

In eq. (2), the terms and represent the molecular weight of LA (116 g mol–1)

and C6-sugars (180 g mol–1), respectively.

The yield of LA can also be defined as the ratio between the mass of LA and

the total mass of the oven-dried pulp:

𝑌𝐿𝐴,𝑡𝑜𝑡𝑎𝑙(𝑤𝑡 %) =𝐶𝐿𝐴×𝑀𝐿𝐴

𝑚𝑝𝑢𝑙𝑝× 100 % (3)

where 𝒎𝒑𝒖𝒍𝒑 represent the mass of the oven-dried pulp.

The yield of formic acid (FA) on molar basis (𝒀𝑭𝑨 ) was calculated in the

same manner except that the sugar concentration was based on the available

C6- and C5-sugars in cellulose and hemicelluloses fractions in the pulp. The

theoretical yield of LA from C6-sugars is 100 mol %, or 64.5 wt % due to the

co-production of formic acid.

Modelling software and techniques

For the modeling of the reaction kinetics, the software Modest® was used

[91]. The software solves the reactor model equations (system of ordinary

differential equations) with the backward difference method and optimizes

the parameter values using a hybrid method involving Simflex and

Levenberg-Marquardt methods.

The software minimizes the following objective function

𝑄 =∑

𝑡

∑(𝑐𝑖,𝑡,𝑒𝑥𝑝𝑖

− 𝑐𝑖,𝑡,𝑚𝑜𝑑𝑒𝑙)2𝑤𝑖,𝑡 (4)

Page 35: Formic and Levulinic Acid from Cellulose via Heterogeneous

26

Results and Discussion

Catalyst Characterization

In Fig. 2 we find the N2 adsorption–desorption isotherms of the fresh and

spent Amberlyst 70 catalysts. The results exhibit that the materials have a

sharp inflection at higher relative pressures in the range of 0.7–1.0p/p0.

This corresponds to type-V curves according to the IUPAC classification,

characteristic to large pore materials. The hysteresis loop observed at the

inflection, reflects the irregularity of the pore sizes over the surface. The

results indicate that the surface areas of fresh and spent catalysts are 0.165

and 35.72 m2 g−1, respectively; at the same time, the pore volumes were

measured to reside at 0.0015 and 0.168 cm3 g−1; and the average pore sizes at

37.5 and 18.16 nm, respectively. The results of the spent catalysts revealed

that the morphological properties are clearly altered during the reaction. The

surface area and pore volume of the spent catalyst were increased

enormously; this might be due to swelling nature of the catalyst in aqueous

reaction conditions.

Fig. 2. Nitrogen adsorption–desorption curves for (a) fresh Amberlyst 70 and (b) spent

Amberlyst 70.

The changes in the surface properties were further evidenced with SEM

images of the catalysts. The scanning electron micrographs (SEM) show

sponge like porous surface over the spent catalyst (see Fig. 3).

The exchange capacity of the fresh resin determined by the dry weight

titration method resulted in a capacity of 2.50 mequiv. g-1. This value was

rather well in line with the manufacturer’s data (2.55 mequiv. g-1). However,

when the catalyst was exposed to the reaction temperature of 190 °C, a

reduction in the number of acid sites at around 70 % of the original value

was obtained whereas 200 °C reaction temperature resulted in a reduction in

the number of acid sites of around 60 %, in comparison to the fresh catalyst.

Page 36: Formic and Levulinic Acid from Cellulose via Heterogeneous

27

Figure 3. Scanning electron micrographs of fresh (top) and spent (below) Amberlyst 70

catalyst (scales correspond to 500, 20, and 2 respective 0.5 µm).

Substrate characterization

Typical scanning electron micrographs of the wood pulps from Metsä-Serla

(M-S) and Domsjö (D), respectively, are depicted in Fig. 4. When looking at

the sample M-S 0.35 AD with fibres from Metsä-Serla, one can see that the

fibre surface is much smoother than the fibre surface of the sample from

Domsjö, D 0.35 AD. A probable reason for this is that all the hemicellulose

contained within the pulp was not dissolved from the kraft pulp from Metsä-

Serla; however, the sulfite pulp from Domsjö turned out to be almost

hemicellulose-free, resulting in a less glossy surface appearance. Upon a

comparison between air-dried and freeze-dried pulp, it was a bit more

difficult to draw any further conclusions. In theory, freeze drying should be

less destructive to the fibers; however, in this case it was difficult to see any

differences between the micrographs obtained for the wood fibre samples.

Page 37: Formic and Levulinic Acid from Cellulose via Heterogeneous

28

M-S 0.35 AD

D 0.35 AD

D 0.35 CPD

D CPD

D AD

Figure 4. Scanning electron micrographs of wood pulps. Scales correspond to 200, 20, and

10 µm. M-S (Metsä-Serla), D (Domsjö), AD (air-dried), CPD (critical-point-dried), 0.35

(grinded to; mm).

Page 38: Formic and Levulinic Acid from Cellulose via Heterogeneous

29

Auto-catalysis

Initially, reactions were also carried out without catalyst until to the reaction

temperature of 200 °C containing 10, 30, 50 and 80 mM formic and levulinic

acid, which corresponds to the concentration range we work within.

Consequently, we could observe that we have very little or negligible

autocatalysis occurring during our experiments. What could be observed was

that the hemicellulose and a very small part of the cellulose were converted

to intermediates and final products before the reaction subsided. It appeared

as if the solution was not sufficiently acidic to react the crystalline part of the

cellulose, but enough acidic (LA pKa = 4.59, FA pKa = 3.75 [92, 93]) to react

some of the amorphous part. However, the reaction rates were very slow

(particularly at low concentrations) and approximately around 5 wt% of the

total mass of pulp was converted before reaction died. In an experiment with

catalyst (heterogeneous catalysis), the contribution of autocatalysis in the

system is likely to be very negligible. The reason is that all the hemicellulose

and probably also the most part of the amorphous cellulose is already

converted before the concentration of formic acid and levulinic acid becomes

so high that it could contribute to autocatalysis. A blank reaction without the

catalyst was also performed at 200 °C and under higher CO2 pressure than at

normal reaction conditions (90 bars instead of 50 bars). The goal was to

determine whether the carbonic acid formed catalyzes the reaction. No pulp

conversion was observed; i.e. no. reaction products or intermediates (or

humins visually) were observed upon analysis.

Reusability of the catalyst, catalyst deactivation

The catalyst reusability was studied at a reaction temperature of 190 °C and

200 °C, respectively. After the first hydrolysis reaction, the resin was

separated from the reaction solution and washed with distilled water and

dried at 100 °C for 24 h. At 190 °C, the yield of LA an FA was approximately

similar as upon the first batch, but the reaction time was significantly longer.

At 200 °C reaction temperature, the catalyst activity reflects the results of

the acidic capacity measurements, i.e. lower product yields, lower reaction

rate and a higher formation of humins. The manufacturer, Rohm and Haas,

has suggested a maximum operating temperature of 190 °C, but some

investigators have reported that the resin could be considered as quite stable

even above this temperature [83]. Consequently, the loss of sulfonic groups

as a result of extended reaction times leads to the lower activity of the

catalyst. Nevertheless, an acid treatment should at least partially restore the

catalytic performance, but it has never been examined in practice.

Page 39: Formic and Levulinic Acid from Cellulose via Heterogeneous

30

Reaction network

A simplified reaction network for the acid-catalyzed hydrolysis of wood pulp to levulinic and formic acids is introduced in Fig. 5.

Figure 5. The reaction network upon production of the platform chemicals levulinic and formic acid from wood pulp.

In fact, the one-pot synthesis of levulinic acid from wood is a demanding

reaction since the whole process most conveniently starts with acid

hydrolysis of pulp thus giving rise to various issues in terms of waste

produced as well as the occurrence of multiple side reactions. Two reaction

pathways are involved: degradation of cellulose results into formation of

glucose, while the degradation of hemicelluloses generates xylose. The

glucose formed is under the experimental conditions further transformed to

5-hydroxymethylfurfural (HMF) which, in turn, is further converted in a

consecutive hydration-dehydration reaction sequence to levulinic (LA) and

formic acids (FA). The xylose pentoses originating from the hemicellulose

fractions transform to furfural which is finally decomposed to formic acid.

Furthermore, HMF can be decarbonylated into furfural [78]. The desired

raw material, lignocellulosic biomass, is mainly composed of cellulose,

hemicelluloses and lignin of which the first two can yield the desired

hexoses, the key intermediates in the production of levulinic acid.

O

OH O

O

O

OOH

OHOH

OH OH

O

OHCH3

O

+ OHO

Levulinic acid Formic acid

Wood pulp

O OH

OHOH

OH OH

Glucose Xylose

HMFFurfural

H2O H2O

2 H2O

3 H2O 3 H2O

Humins

(or Tar)

decarbonylation

OHO

Formic acid

Decomp. prod.

Page 40: Formic and Levulinic Acid from Cellulose via Heterogeneous

31

Thereafter, sugars need to be dehydrated and rearranged to 5-

hydroxymethylfurfural and furfural depending on whether hexoses or

pentoses, respectively, are to be dehydrated. Consequently, hexoses can then

further rehydrate to formic and levulinic acids also requiring two equivalents

(moles) of water.

Since two different organic carboxylic acids are formed, the theoretical yield

of levulinic acid from hexoses and cellulose has been reported to reach 64.5

wt % and 71.5 wt % only, respectively [32]. It should, however, be stated that

the real yields of levulinic acid are typically lower than the theoretical yields

due to the harsh reaction conditions required for hydrolysis of

macromolecules. Thus, the formation of humins is hardly to be avoided and

throughout all kinetic experiments, black insoluble-substances known as

humins/tar were also found on the reactor walls and the amount was visibly

increased with reaction time. These are well known products of side-

reactions upon acid-catalyzed decomposition of pulp. Girisuta et al. [94]

confirmed by elemental analysis that the solids are a mixture consisting

mainly of humins (typical composition: C, 63.1; H, 4.2) and some unreacted

cellulose (C, 42.2; H, 6.1). In addition to the anticipated products, Girisuta et

al. [42] also detected small amounts of glucose-reversion products (e.g.,

levoglucosan, isomaltose and gentiobiose). The formation of the reversion

products was also observed in this study, but the maximum amount of these

products was very low and not accurately quantified in this work. A typical

example of an HPLC chromatogram of a sample taken in the beginning

during a reaction is given in Fig. 6.

Fig. 6. Typical HPLC chromatogram for the acid-catalyzed hydrolysis of soft wood pulp.

Other possible by-products of the acid-catalyzed hydrolysis of cellulose are

gas-phase components from thermal degradation (eventually catalyzed by

the autoclave walls) reactions of reactants and/or products. To gain insights

into the extent of these reactions, a GC study was carried out to analyze the

gas phase composition during an experiment performed at 180 °C without

Page 41: Formic and Levulinic Acid from Cellulose via Heterogeneous

32

any gas addition. CO, CO2 and H2 are typically detected. The amounts

observed were less than 0.1 wt% of the cellulose intake. This implies that the

formation of gas-phase compounds is only a minor reaction pathway under

the investigated reaction conditions, at lower temperatures. However, we

could not quantify the presence of any formed carbon dioxide since it was

used to pressurize the reactor system during most of our experiments. On

the other hand on the basis of the graphic illustrations of the kinetic

experiments, we can clearly see that formic acid decomposes at higher

temperatures. Most likely increases the levels of CO and CO2 with increasing

reaction temperatures, but no conclusive investigations were performed.

Formic acid decomposes primarily to CO and H2O in the gas phase

(dehydration), but to CO2 and H2 (decarboxylation) in the aqueous phase

[95] and indicating an increase in the total pressure of the reaction vessel.

The influence of reaction conditions and raw material on product

yields

Comparison between different wood pulps

In Fig. 7, one can see the results from the two experiments performed with

different wood pulps. One can clearly see that the yield of formic acid

compared to levulinic acid in the wood pulp from Metsä-Serla is much

higher than in the wood pulp from Domsjö. The underlying reason lies in the

high hemicelluloses content found in the wood pulp from Metsä-Serla. The

main difference when comparing the catalytic transformation of pulp from

Domsjö and pulp from Metsä-Serla over Amberlyst 70 is the absence of

xylose and, consequently, lower formation of furfural from the former raw

material.

Fig. 7. The comparison of formic and levulinic acid yields for soft wood pulp from Domsjö

(sulfite pulp) and Metsä-Serla (sulfate pulp) (the reaction atmosphere is CO2, the catalyst and

substrate amount is 3 g respectively and the temperature is 180 ◦C).

0

10

20

30

40

50

60

0 10 20 30 40 50 60

Yie

ld /

mo

l-%

Reaction time / h

LA Domsjö

FA Domsjö

LA Metsä-Serla

FA Metsä-Serla

Page 42: Formic and Levulinic Acid from Cellulose via Heterogeneous

33

Reaction with D-glucose

The conversion of wood pulp to LA and FA is a complex multistep process. It

is generally believed that the polymer chains of cellulose are broken down

into low molecular weight fragments and finally to glucose. Then, glucose is

decomposed to HMF, which is further converted into LA and FA in a molar

ratio of 1:1 [9]. However, HMF has also been reported to react into furfural

(carbonylation) which is further converted to formic acid [78]. Fructose is

also a known intermediate in acid catalyzed decomposition of glucose, but

could not be detected in this study. The epimerization step is very rapid and,

consequently, fructose was omitted as an individual chemical species. To

find out if we also have the same plausible reaction mechanism for the

Amberlyst 70 catalyzed conversion of wood pulp we did an experiment with

glucose. We can clearly see (Fig. 8) that LA and FA are not formed in a molar

ratio of 1:1. It is even clearer if we take into account the decomposition of FA

during time. All above-mentioned products (except fructose) were detected

in this study, which means we also detected furfural in our product samples

taken from the reactor during reaction with glucose. Furfural formed in

experiments with D-glucose originated most probably from the

decarbonylation of HMF, but also most certainly from the transformation of

some of the impurities, C5 sugars, in the D-glucose.

Fig. 8. Experiment with d-glucose (the reaction atmosphere is CO2, the catalyst and substrate

amount is 3 g respectively and the temperature is 180 ◦C).

The influence of reaction temperature and catalyst to substrate

dosage

Hydrolysis of wood pulp by Amberlyst 70 is highly dependent upon the

reaction temperature. The effect of reaction temperature on the cellulose

conversion into LA and FA was investigated, and the experiments were

conducted at 150, 170, 180, 185, 190 and 200 °C, respectively. At lower

temperatures (150 °C), there is little conversion of cellulose to levulinic acid

as well as to formic acid. Note however, that the experiments at 150 and 170

0

10

20

30

40

50

60

70

80

90

100

0 5 10 15 20 25 30

Yie

ld /

mo

l-%

Reaction time / h

D-glucose

Levulinic acid

Formic acid

Page 43: Formic and Levulinic Acid from Cellulose via Heterogeneous

34

°C were interrupted before all the wood pulp was consumed because of low

reaction rate. As the temperature was increased, the conversion rates

improved very rapidly. The increased conversion is likely related to increased

solubility of cellulose at higher temperatures. The conversion at 190 and 200

°C is very fast as can be seen in Fig. 9.

Fig. 9. The influence of reaction temperature on levulinic and formic acid yields (the reaction

atmosphere is CO2 and the catalyst and substrate amount is 3 g respectively).

It has been reported that when the reaction temperature was increased

above 200 °C, the yield of LA began to fall. Higher temperatures could thus

accelerate the rate of chemical reactions, but unwanted side reactions also

appear at the same time. For example, LA is unstable above 230 °C, which

would be dehydrated to unsaturated lactones. Additionally, more humins

may occur during the reaction [10]. Therefore, an increase in the

temperature is unfavorable for the extent of reaction, and, consequently, the

reaction temperature was limited to 200 °C in this work. Another reason that

the maximum reaction temperature was limited to 200 °C is that the

manufacturer of Amberlyst 70 recommends a maximum reaction

temperature of 190 °C. However, it has been reported that the catalyst could

be considered as stable in some conditions even above this temperature [84].

Fig. 9 also shows the effect of the temperature on the decomposition of FA. It

can be observed that temperature has a significant effect on the

decomposition rate of FA. When the reaction temperature was increased

from 180 to 200 °C, there was a huge increase in the decomposition rate of

FA. The amount of FA decreased quickly very likely forming carbon

monoxide, carbon dioxide and hydrogen gas [95] indicated by a small but

notable rise of the pressure in the autoclave. The effect of catalyst dosage on

the yields of products is shown in Fig. 10.

0

10

20

30

40

50

60

0 10 20 30 40 50 60 70 80 90 100

ƳL

A /

mo

l-%

Reaction time / h

200 C 190 C

185 C

180 C

170 C

150 C

0

10

20

30

40

50

60

0 10 20 30 40 50 60 70 80 90 100

ƳF

A /

mo

l-%

Reaction time / h

200 C 190 C 185 C180 C

170 C

150 C

Page 44: Formic and Levulinic Acid from Cellulose via Heterogeneous

35

Fig. 10. The effect of temperature and catalyst amount on levulinic and formic acid yields (the

reaction atmosphere is CO2 and substrate amount is 3 g).

The amount of catalyst in the reaction mixture was varied from 2 to 4 g at

180 °C and 1 to 3 g at 185 °C, respectively. As can be seen, the amount of

catalyst affected the extent of cellulose conversion, but not as much as the

reaction temperature does. Fig. 11 illustrates the maximum levulinic acid

yields obtained upon kinetic experiments conducted at different acid and

substrate concentrations. Theoretically, a higher liquid-to-solid ratio should

result in a higher yield of levulinic acid. However, it was further observed

that higher substrate loadings had an adverse effect in terms of the final

levulinic acid yields detected in the reaction mixture. These observations are

in line with other studies on the acid-catalyzed decompositions of HMF and

glucose, whereupon low substrate concentrations favored high LA yields.

Generally, lower concentration of cellulose resulted in higher yields of

levulinic acid. For each levulinic acid molecule generated, two molecules of

water are needed; otherwise 5-hydroxymethylfurfural (HMF) may be

converted into humins [42]. As also depicted by Fig. 11, evidently there exists

an optimal catalyst loading. Consequently, the addition of excess catalyst has

little or even negative effect on the LA yield. In fact, Yan et al also observed

that low concentration of hydrochloride acid resulted in too low

concentration of H+ ions thus being unable to facilitate cellulose hydrolysis

and, consequently, a low reaction rate was observed. Using low acid

concentrations is problematic, because the lignin and other impurities

present in the biomass consumes a fraction of the acid during the reaction.

In our case a load of 9 g substrate and 1 g catalyst resulted in only 5.9 mol %

yield of levulinic acid, at a reaction temperature of 190 °C.

0

10

20

30

40

50

60

0 10 20 30 40 50 60 70

ƳL

A /

mo

l-%

Reaction time / h

185 C, 3 g cat180 C, 3 g cat

180 C, 2 g cat

0

10

20

30

40

50

60

0 10 20 30 40 50 60 70

ƳF

A /

mo

l-%

Reaction time / h

185 C, 3 g cat

180 C, 3 g cat

180 C, 2 g cat

Page 45: Formic and Levulinic Acid from Cellulose via Heterogeneous

36

Figure 11. Maximum levulinic acid yields formed at 180 °C and 190 °C at different acid and

substrate concentrations (the reaction atmosphere is CO2).

The influence of reaction gas and pressure

In this study, carbon dioxide was used as the reaction atmosphere. However,

to investigate whether various gases affect the conversion of wood pulp,

some tests with other gases and pressures were also performed. Fig. 12

illustrate the results of these experiments.

Fig. 12. The influence of reaction atmosphere and pressure on levulinic and formic acid yields

(the catalyst and substrate amount is 3 g respectively and the temperature is 180 ◦C).

As can be observed, argon as the reaction atmosphere seems to promote the

reaction rates, whereas the presence of carbon dioxide acted as a mild

inhibitor. This was surprising since we expected the significantly better

soluble CO2 to yield carbonic acid that would enhance the performance of

our system. However, the much less soluble and supposedly inert argon (by a

factor of roughly 103 [96]) seemed to yield better performance. The reason

for this is at present not clear to us. Our hypothesis is that the argon gas and

formic acid are capable of forming relatively strong argon–HCOOH

complexes [97] resembling hydrogen bonding networks. Thus, the

equilibrium in between the gas and the liquid phase is presumably shifted

toward the gas phase (low solubility of noble gases in water) and,

consequently, any possible product inhibition contribution to the additional

13

6

9

40

50

60

70

13

69

51.5

67.8

60.2

43.1

57.4

65.4

59.9

55.5

46.9

Substrate [g]

Levu

lin

ic a

cid

yie

ld [m

ol %

]

Catalyst [g]180 °C

13

6

9

40

50

60

70

13

69

62.1

66.1

62.7

43.9

54.3

61.2

57.0

51.7

59.2

5.9

48.250.7

Substrate [g]

Levu

lin

ic a

cid

yie

ld [m

ol %

]

Catalyst [g]190 °C

0

10

20

30

40

50

60

0 10 20 30 40 50 60 70 80

ƳL

A /

mo

l-%

Reaction time / h

CO2 50 bar

Ar 50 bar

No gas additionAr 3 bar

0

10

20

30

40

50

60

0 10 20 30 40 50 60 70 80

ƳF

A /

mo

l-%

Reaction time / h

Ar 3 bar

Ar 50 bar

CO2 50 bar

No gas addition

Page 46: Formic and Levulinic Acid from Cellulose via Heterogeneous

37

formation of HCOOH from cellulose (in liquid phase) is diminished,

resulting in increased amounts of the final desired products. Similar

phenomena might also contribute in the case of levulinic acid. In addition it

is well known that CO2 can be hydrogenated to formic acid (reversible

reaction). Further, particularly at higher temperatures, formic acid tends

easily to decompose. Thus, it might be that since these reactions are typically

metal catalyzed that our reactor setups (stainless steel and zirconium) could

to certain extent interact with the species emerging. A system containing at

least water, CO2 and HCOOH, in the presence of metals could provide

conditions that inhibit the initial HCOOH formation from cellulose.

Selectivity

Overall, we can conclude that the catalysis reaction is highly selective, and

there are very little by-products formed (except humins). Nevertheless, if we

compare our main products formic and levulinic acid, we can perhaps

discern small differences among the selectivity under different reaction

conditions. The molar yield of formed levulinic acid as a function of the

molar yield of produced formic acid in the catalytic transformation of wood

pulp over Amberlyst 70 resin is presented in Fig. 13, respectively.

Fig. 13. The influence of the reaction atmosphere, temperature and catalyst amount on the

molar yield of levulinic acid as a function of the molar yield of formic acid (the reaction

atmosphere is CO2, the substrate and catalyst amount is 3 g respectively and the temperature

is 180 ◦C).

0

10

20

30

40

50

60

0 10 20 30 40 50 60

LA

/ m

ol-

%

FA / mol-%

No gas addition

CO2 50 bar

Ar 50 bar

Ar 3 bar

0

10

20

30

40

50

60

0 10 20 30 40 50 60

LA

/ m

ol-

%

FA / mol-%

200 C

190 C

185 C

180 C

170 C

150 C

0

10

20

30

40

50

60

0 10 20 30 40 50 60

LA

/ m

ol-

%

FA / mol-%

185 C, 3 g cat

185 C, 1 g cat

180 C, 4 g cat

180 C, 3 g cat

180 C, 2 g cat

Page 47: Formic and Levulinic Acid from Cellulose via Heterogeneous

38

The most striking difference is perhaps that an atmosphere with low

pressure of argon seems to favor formic acid transformation. However, up to

40% yield level, the differences are extremely small. It seems likely that in

the end of a reaction the decomposition of formic acid affects to a greater

degree the selectivity. Ultimately, more experiments are needed to draw

more accurate conclusions.

Kinetic modelling

Upon modelling of the reaction kinetics, Modest® was used [91]. The

software uses the backward difference method to solve the reactor mass

balances (a system of ordinary differential equations, ODEs) and a hybrid

method involving Simflex and Levenberg-Marquardt methods for the

estimation of the kinetic parameters (rate constants and activation energies)

defined in the kinetic model. The Levenberg-Marquardt method [98], is a

more effective modification of the well-known Newton-Raphson method and

the Simplex algorithm is based on the Simplex method introduced by

Spendley et al. [99]. If the two methods are compared, one can conclude that

the Simplex algorithm is slightly better at finding the way to the optimum

from a poor initial guess of parameters, while the sharp optimum is easier

and faster achieved by the Levenberg-Marquardt method, provided that the

initial guess is sufficiently close to the optimum. Consequently, the rate

equations for the proposed elementary reaction kinetic model and batch

reactor model were implemented in the software which also provides the

graphical interface.

Page 48: Formic and Levulinic Acid from Cellulose via Heterogeneous

39

O

OH O

O

O

OOH

OHOH

OH OH

O

OHCH3

O

+ OHO

Levulinic acid Formic acid

Wood pulp

A1 / A

2

O OH

OHOH

OH OH

Glucose Xylose

HMFFurfural

H2O H2O

2 H2O

3 H2O 3 H2O

Humins

(or Tar)

decarbonylation

OHO

Formic acid

r1

r2

r3

r4

r5 r

6

Decomp. prod.

r14

Figure 14. The reaction network upon production of the platform chemicals levulinic and

formic acid from cellulose as well as the reactions used in the mathematical model.

For the parameter estimation the following objective function was

minimized

𝑄 =∑

𝑡

∑(𝑐𝑖,𝑡,𝑒𝑥𝑝𝑖

− 𝑐𝑖,𝑡,𝑚𝑜𝑑𝑒𝑙)2𝑤𝑖,𝑡 (5)

where ci,t,exp and ci,t,model are the experimentally recorded concentrations and

the concentrations predicted by the model, respectively. The weight factors,

wi, were at this time set equal to unity (1) for all experimental data.

The reaction system was described with the following reactions

𝐴1𝑟1→ 𝐺𝑙𝑢𝑐𝑜𝑠𝑒

𝑟3→𝐻𝑀𝐹

𝑟5→𝐿𝑒𝑣𝑢𝑙𝑖𝑛𝑖𝑐 𝑎𝑐𝑖𝑑 + 𝐹𝑜𝑟𝑚𝑖𝑐 𝑎𝑐𝑖𝑑 (6)

𝐴2𝑟2→𝑋𝑦𝑙𝑜𝑠𝑒

𝑟4→𝐹𝑢𝑟𝑓𝑢𝑟𝑎𝑙

𝑟6→ 𝐹𝑜𝑟𝑚𝑖𝑐 𝑎𝑐𝑖𝑑 (7)

Page 49: Formic and Levulinic Acid from Cellulose via Heterogeneous

40

𝐹𝑜𝑟𝑚𝑖𝑐 𝑎𝑐𝑖𝑑𝑟14→ 𝐷𝑒𝑐𝑜𝑚𝑝𝑜𝑠𝑒𝑑 𝑝𝑟𝑜𝑑𝑢𝑐𝑡𝑠 (8)

In the initial approach, only the experimental runs at higher temperature

range, i.e. 180–200 °C, were concerned since the reaction proceeds

extremely sluggishly at lower temperatures and batch experiments lasting

for several days had to be carried out. Nevertheless, some preliminary

modeling revealed that a first order reaction kinetics model was able to

describe the reactions in a reasonable and simple way. Thus, the reaction

rates were determined as follows:

𝑟𝑗 = 𝑘𝑗𝑐𝑖 (9)

where rj denote the individual reaction rates for the reaction j, kj denote the

reaction rate constants and ci the concentration of component i. On the other

hand, the temperature influence on the reaction rate was taken into account

with the Arrhenius equation

𝑘𝑗 = 𝑘0𝑗𝑒−𝐸𝐴𝑗𝑅(1𝑇−

1𝑇𝑚𝑒𝑎𝑛

) (10)

and thus, kj denote the reaction rate constants, koj the frequency factor, R is the ideal gas constant, T is the temperature and EAj the activation energy. The modeling approach chosen verified that the experimental results

obtained at 180–200 °C, in the heterogeneously catalyzed reaction displayed

prominent Arrhenius-type temperature dependence demonstrated for the

reaction products identified.

For the reactor, the ideal batch reactor model was used.

𝑑𝑐𝑖𝑑𝑡= 𝑟𝑖 𝑚𝑐𝑎𝑡 (11)

where ci denote the concentration of component i, ri the generation rate for

component i and mcat the catalyst amount.

The parameters (frequency factors k0j, activation energies EAj) were

estimated using the parameter estimation software. Estimation of the

Page 50: Formic and Levulinic Acid from Cellulose via Heterogeneous

41

reaction kinetics parameters were split into two parts, whereupon first only

parameters for the reactions r2, r4 and r6 were estimated, after which the

parameters for r2, r4 and r6 were fixed and the rest of the parameters were

estimated. Initially an estimation of all parameters simultaneously was tried

but it did not give good results for the reactions r2 and r4 (formation of

xylose and furfural), mainly due to large differences in the concentration

levels of the compounds present thus leading to unstable behavior of the

solver. The values of the estimated parameters are enlisted in Table 1.

Table 1. Numerical values for the estimated parameters (degree of explanation 97.8 %).

Parameter / (for units see

below)

Value Parameter error / %

k01 2.7 1.5

k02 15.3 11.8

k03 18.9 4.3

k04 30.8 8.4

k05 291 130

k06 40.7 14.2

k014 0.429 2.7

EA1 163 2.8

EA2 33.3 82.7

EA3 163 7.2

EA4 64.5 34.1

EA5 8.83 2705

EA6 16.7 203

EA14 134 7

Tmean = 185 °C, Frequency factor (koj) in h-1 and activation energy (EAj) in kJ/mol.

Fit of the model to experimental data

The most common measure for the goodness of fit is the degree of

explanation, the R2 value. The idea is to compare the residuals given by the

model to the residuals of the simpliest model one may think of, the average

value of all the data points. The R2 value is given by the expression

𝑅2 = 100(𝑦𝑚𝑜𝑑𝑒𝑙 − 𝑦𝑒𝑥𝑝𝑒𝑟𝑖𝑚𝑒𝑛𝑡)

2

(𝑦𝑚𝑜𝑑𝑒𝑙 − �̅�𝑒𝑥𝑝𝑒𝑟𝑖𝑚𝑒𝑛𝑡)2 (12)

where the y values denotes the estimated vs. measured values of the

respective parameters. The closer the R2 value is 1, the better is the fitting of

the model. For the parameter estimation the degree of explanation of 97.8 %

was obtained.

Page 51: Formic and Levulinic Acid from Cellulose via Heterogeneous

42

The distribution and sensitivity of the parameters were controlled with the

Markov Chain Monte Carlo method (MCMC) [100], see Figure 16. These

figures also show low correlations between parameters and confirm well

identified optimal values.

Page 52: Formic and Levulinic Acid from Cellulose via Heterogeneous

43

Figure 15. Sensitivity analysis for the parameters.

Page 53: Formic and Levulinic Acid from Cellulose via Heterogeneous

44

Figure 16. Fit of model to experimental data for one of the experiments performed at 190 °C

with 6 g catalyst and 6 g substrate. Levulinic acid (○), formic acid (+), glucose (*), 5-HMF (□),

xylose (∆) and furfural (◊) vs. time.

Figure 17. Fit of model to experimental data for one of the experiments performed at 180 °C

with 6 g catalyst and 9 g substrate. Levulinic acid (○), formic acid (+), glucose (*), 5-HMF (◊),

xylose (∆) and furfural (□) vs. time.

Figure 18. Fit of model to experimental data for one of the experiments performed at 200 °C

with 1 g catalyst and 3 g substrate. Levulinic acid (○), formic acid (+), glucose (*), 5-HMF (◊),

xylose (∆) and furfural (□) vs. time.

The figures depicting the fit of model to experimental data (Figs. 16-18) show

that the model can predict the behavior of the reaction rather well at

0 5 10 15 20 25 30 35 40 450

5

10

15

20

25

Glucose

Levulinic acid

Formic acid

time (h)

mm

ole

0 5 10 15 20 25 30 35 40 450

0.1

0.2

0.3

0.4

0.5

0.6

0.7

HMF

Xylose

Furfural

time (h)

mm

ole

0 5 10 15 20 25 300

1

2

3

4

5

6

7

Glucose

Levulinic acid

Formic acid

time (h)

mm

ole

0 5 10 15 20 25 300

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

HMFXyloseFurfural

time (h)

mm

ole

Page 54: Formic and Levulinic Acid from Cellulose via Heterogeneous

45

different conditions as displayed in the parity plot Figure 19. Nevertheless,

the fit could be better for the components present in low concentrations.

Figure 19. Fit of model to all experimental data (levulinic acid (○), formic acid (+) and

glucose (*)).

Page 55: Formic and Levulinic Acid from Cellulose via Heterogeneous

46

Conclusions

Levulinic acid can be successfully produced from biomass using either

homogeneous or heterogeneous catalysts albeit with variations in

performance. Besides more traditional approaches like taking advantage of

mineral acids, in particular hydrochloric and sulfuric ones, many solid acids

have shown comparable performance. New, unorthodox processing

alternatives such as application of microwave dielectric heating or various

mechanical commutator designs have also been demonstrated for this

process. In general, the beneficial conditions involve relatively low acid

concentrations, low initial concentrations of feedstock and moderate

temperatures.

In this thesis solid acid catalyzed degradation of bleached Nordic industrial

dissolving pulp (a mixture of softwoods Norway Spruce and Scots Pine) was

investigated under mild batch wise reaction conditions and the goal was to

obtain versatile platform chemicals – levulinic and formic acids. The optimal

reaction conditions were determined and a simple, first-order kinetic

approach was developed to describe the experimental data with satisfactorily

accuracy. Sulphite-cellulose, being lean on pentose sugar containing

hemicelluloses was revealed as a good raw material compared to sulphate

pulp (birch kraft pulp), thus resulting in about 59 and 68 mol % theoretical

yields of levulinic vs. formic acid, respectively, obtained at a reaction

temperature of 180 °C and at a low initial cellulose concentration. It was also

observed that excessively high reaction temperatures contribute to thermal

degradation of the formed formic acid, since formic acid tends to degrade to

carbon dioxide, carbon monoxide, hydrogen and water, which in turn led to

lower overall yields of the desired products. Further it was observed that a

low substrate load favors a high levulinic acid yield and that the catalyst

loading has an optimum, i.e. whether the amount of catalyst is too high or

too low reduces the yield. One can also observe that a low acid

concentrations is problematic, because the lignin and other impurities

present in the biomass consumes a fraction of the acid during the reaction.

Consequently, the loss of sulfonic groups as a result of extended reaction

times leads to the lower activity of the catalyst. Nevertheless, an acid

treatment should at least partially restore the catalytic performance, but it

has never been examined in practice.

More surprisingly was that argon as the reaction atmosphere seems to

promote the reaction rates, whereas the presence of carbon dioxide acted as

a mild inhibitor. We expected the significantly better soluble CO2 to yield

carbonic acid that would enhance the performance of our system. The reason

for this is at present not clear to us.

Page 56: Formic and Levulinic Acid from Cellulose via Heterogeneous

47

Also we have observed very little auto-catalysis occurring during our

experiments. What could be observed was that the hemicellulose and a very

small part of the cellulose were converted to intermediates and final

products before the reaction subsided. It appeared as if the solution was not

sufficiently acidic to react the crystalline part of the cellulose, but enough

acidic to react some of the amorphous part. In an experiment with catalyst

(heterogeneous catalysis), the contribution of auto-catalysis in the system is

likely to be very negligible.

Page 57: Formic and Levulinic Acid from Cellulose via Heterogeneous

48

Acknowledgements

First of all, I would like to express my deepest gratitude to my supervisor and

closest colleague, Prof. Jyri-Pekka Mikkola for giving me the opportunity to

work at the laboratory.

I would like to give special thanks to Prof. Johan Wärnå for his contribution

in the computational modelling and Dr. William Larsson for his assistance

with laboratory equipment.

Also my assistant supervisors Prof. Leif Jönsson and Doc. Torbjörn

Lestander are acknowledged for their invaluable advices and discussions.

Markku Reunanen and Dr. Carina Jonsson for carrying out the GC–MS

analysis and Per Hörstedt are acknowledged for carrying out the SEM

analysis.

I also would like to thank all the co-authors and co-workers and all the rest

of my collages.

Finally, I would like to express gratitude to my parents Gunvor and Runar

for their love and encouragement throughout my education. Special thanks

go also to my two sisters, Karin and Ingrid, and their families for their

support.

It would not have been possible without you all!

The Bio4Energy programme, Knut and Alice Wallenberg and Kempe

Foundations are gratefully acknowledged for the financial support. Further,

COST action CM0903 (UBioChem) and the Academy of Finland are

acknowledged and greatly appreciated. Collaboration with Processum

Biorefinery Initiative AB is acknowledged. This work is also a part of

activities at the Åbo Akademi Process Chemistry Centre, a center of

excellence financed by the Åbo Akademi University.

Page 58: Formic and Levulinic Acid from Cellulose via Heterogeneous

49

References

[1] P. Gallezot, Process options for converting renewable feedstocks to bioproducts, Green Chemistry, 9 (2007) 295-302.

[2] D. Mohan, C.U. Pittman, P.H. Steele, Pyrolysis of wood/biomass for bio-oil: A critical review, Energy & Fuels, 20 (2006) 848-889.

[3] S.N. Naik, V.V. Goud, P.K. Rout, A.K. Dalai, Production of first and second generation biofuels: A comprehensive review, Renewable & Sustainable Energy Reviews, 14 (2010) 578-597.

[4] G.W. Huber, S. Iborra, A. Corma, Synthesis of transportation fuels from biomass: Chemistry, catalysts, and engineering, Chemical Reviews, 106 (2006) 4044-4098.

[5] R. Rinaldi, F. Schuth, Acid hydrolysis of cellulose as the entry point into biorefinery schemes, Chemsuschem, 2 (2009) 1096-1107.

[6] W.L. Faith, Development of the scholler process in the United-States, Industrial and Engineering Chemistry, 37 (1945) 9-11.

[7] A.A. Efremov, G.G. Pervyshina, B.N. Kuznetsov, Production of levulinic acid from wood raw material in the presence of sulfuric acid and its salts, Chem Nat Compd, 34 (1998) 182-185.

[8] C. Chang, P. Cen, X. Ma, Levulinic acid production from wheat straw, Bioresource Technology, 98 (2007) 1448-1453.

[9] B. Girisuta, L.P.B.M. Janssen, H.J. Heeres, Kinetic study on the acid-catalyzed hydrolysis of cellulose to levulinic acid, Industrial & Engineering Chemistry Research, 46 (2007) 1696-1708.

[10] L. Yan, N. Yang, H. Pang, B. Liao, Production of levulinic acid from bagasse and paddy straw by liquefaction in the presence of hydrochloride acid, CLEAN – Soil, Air, Water, 36 (2008) 158-163.

[11] J.-P. Lange, W.D. van de Graaf, R.J. Haan, Conversion of Furfuryl Alcohol into Ethyl Levulinate using solid acid catalysts, Chemsuschem, 2 (2009) 437-441.

[12] D.W. Rackemann, W.O.S. Doherty, The conversion of lignocellulosics to levulinic acid, Biofuels, Bioproducts and Biorefining, 5 (2011) 198-214.

Page 59: Formic and Levulinic Acid from Cellulose via Heterogeneous

50

[13] Amberlyst 70 high temperature strongly acidic catalyst http://www.amberlyst.com/literature/a4/70.pdf, (accessed 6.11.2011).

[14] R. Rinaldi, R. Palkovits, F. Schüth, Depolymerization of cellulose using solid catalysts in ionic liquids, Angewandte Chemie International Edition, 47 (2008) 8047-8050.

[15] I. Agirrezabal-Telleria, A. Larreategui, J. Requies, M.B. Güemez, P.L. Arias, Furfural production from xylose using sulfonic ion-exchange resins (Amberlyst) and simultaneous stripping with nitrogen, Bioresource Technology, 102 (2011) 7478-7485.

[16] X. Hu, C.-Z. Li, Levulinic esters from the acid-catalysed reactions of sugars and alcohols as part of a bio-refinery, Green Chemistry, 13 (2011) 1676-1679.

[17] J.F. Saeman, Kinetics of wood saccharification-hydrolysis of cellulose and decomposition of sugars in dilute acid at high temperature, Industrial & Engineering Chemistry, 37 (1945) 43-52.

[18] R.D. Fagan, H.E. Grethlein, A.O. Converse, A. Porteous, Kinetics of the acid hydrolysis of cellulose found in paper refuse, Environmental Science & Technology, 5 (1971) 545-547.

[19] D.R. Thompson, H.E. Grethlein, Design and evaluation of a plug flow reactor for acid hydrolysis of cellulose, Industrial & Engineering Chemistry Product Research and Development, 18 (1979) 166-169.

[20] J.-P. Franzidis, A. Porteous, J. Anderson, The acid hydrolysis of cellulose in refuse in a continuous reactor, Conservation & Recycling, 5 (1982) 215-225.

[21] M. Green, S. Kimchie, A. Malester, B. Rugg, G. Shelef, Utilization of municipal solid wastes (MSW) for alcohol production, Biological wastes, 26 (1988) 285-295.

[22] I.A. Malester, M. Green, G. Shelef, Kinetics of dilute acid hydrolysis of cellulose originating from municipal solid wastes, Industrial & Engineering Chemistry Research, 31 (1992) 1998-2003.

[23] Q. Fang, M.A. Hanna, Experimental studies for levulinic acid production from whole kernel grain sorghum, Bioresour Technol, 81 (2002) 187-192.

[24] T. Frost, E. Kurth, Tappi, 1951.

Page 60: Formic and Levulinic Acid from Cellulose via Heterogeneous

51

[25] Q. Jing, X. LÜ, Kinetics of non-catalyzed decomposition of glucose in high-temperature liquid water, Chinese Journal of Chemical Engineering, 16 (2008) 890-894.

[26] J. Shen, C.E. Wyman, Hydrochloric acid-catalyzed levulinic acid formation from cellulose: data and kinetic model to maximize yields, AIChE Journal, 58 (2012) 236-246.

[27] A. Onda, T. Ochi, K. Yanagisawa, Selective hydrolysis of cellulose into glucose over solid acid catalysts, Green Chemistry, 10 (2008) 1033-1037.

[28] G.J. Mulder, J. Prakt. Chem., 21 (1840).

[29] T. Werpy, G. Petersen, A. Aden, J. Bozell, J. Holladay, J. White, A. Manheim, D. Eliot, L. Lasure, S. Jones, in, Top value added chemicals from biomass. Volume 1-Results of screening for potential candidates from sugars and synthesis gas, DTIC Document, 2004.

[30] A.M.R. Galletti, C. Antonetti, V. De Luise, D. Licursi, N. Nassi, Levulinic acid production from waste biomass, BioResources, 7 (2012) 1824-1835.

[31] B. V. Timokhin, V. A. Baransky, G. D. Eliseeva, Levulinic acid in organic synthesis, Russian Chemical Reviews, 68 (1999) 73-84.

[32] R.H. Leonard, Levulinic acid as a basic chemical raw material, Industrial & Engineering Chemistry, 48 (1956) 1330-1341.

[33] S.W. Fitzpatrick, Production of levulinic acid from carbohydrate-containing materials (US Patent 5,608,105, 1997).

[34] D.B. Bevilaqua, M.K.D. Rambo, T.M. Rizzetti, A.L. Cardoso, A.F. Martins, Cleaner production: levulinic acid from rice husks, Journal of Cleaner Production, 47 (2013) 96-101.

[35] M. Kang, S.W. Kim, J.-W. Kim, T.H. Kim, J.S. Kim, Optimization of levulinic acid production from Gelidium amansii, Renewable Energy, (2012).

[36] D.-m. Lai, L. Deng, Q.-x. Guo, Y. Fu, Hydrolysis of biomass by magnetic solid acid, Energy & Environmental Science, 4 (2011) 3552-3557.

[37] P. Wang, S. Zhan, H. Yu, Production of levulinic acid from cellulose catalyzed by environmental-friendly catalyst, Advanced Materials Research, 96 (2010) 183-187.

Page 61: Formic and Levulinic Acid from Cellulose via Heterogeneous

52

[38] L. Peng, L. Lin, J. Zhang, J. Zhuang, B. Zhang, Y. Gong, Catalytic conversion of cellulose to levulinic acid by metal chlorides, Molecules, 15 (2010) 5258-5272.

[39] B. Girisuta, B. Danon, R. Manurung, L.P.B.M. Janssen, H.J. Heeres, Experimental and kinetic modelling studies on the acid-catalysed hydrolysis of the water hyacinth plant to levulinic acid, Bioresource Technology, 99 (2008) 8367-8375.

[40] V.S. Minina, A.E. Sarukhanova, K.U. Usmanov, Preparation of fural and levulinic acid by hydrolysis of pressed cotton stems, Fiz. i Khim. Prirodn. i Sintetich. Polimerov, Akad. Nauk Uz. SSR, Inst. Khim. Polimerov, 1962, pp. 87-93.

[41] A. Szabolcs, M. Molnar, G. Dibo, L.T. Mika, Microwave-assisted conversion of carbohydrates to levulinic acid: an essential step in biomass conversion, Green Chemistry, 15 (2013) 439-445.

[42] B. Girisuta, L. Janssen, H. Heeres, Green Chemicals: A kinetic study on the conversion of glucose to levulinic acid, Chemical Engineering Research and Design, 84 (2006) 339-349.

[43] A. Thompson, Method of making levulinic acid (US patent 2,206,311, 1940).

[44] W.W. Moyer, Preparation of levulinic acid (US patent 2,207,328, 1942).

[45] Z. Yang, H. Kang, Y. Guo, G. Zhuang, Z. Bai, H. Zhang, C. Feng, Y. Dong, Dilute-acid conversion of cotton straw to sugars and levulinic acid via 2-stage hydrolysis, Industrial Crops and Products, 46 (2013) 205-209.

[46] H. Lin, J. Strull, Y. Liu, Z. Karmiol, K. Plank, G. Miller, Z. Guo, L. Yang, High yield production of levulinic acid by catalytic partial oxidation of cellulose in aqueous media, Energy & Environmental Science, 5 (2012) 9773-9777.

[47] L.J. Carlson, Process for the manufacture of levulinic acid (US patent 3,065,263, 1962).

[48] H. Mehdi, V. Fábos, R. Tuba, A. Bodor, L. Mika, I. Horváth, Integration of homogeneous and heterogeneous catalytic processes for a multi-step conversion of biomass: From sucrose to levulinic acid, γ-valerolactone, 1,4-pentanediol, 2-methyl-tetrahydrofuran, and alkanes, Top Catal, 48 (2008) 49-54

Page 62: Formic and Levulinic Acid from Cellulose via Heterogeneous

53

[49] Y. Takeuchi, F. Jin, K. Tohji, H. Enomoto, Acid catalytic hydrothermal conversion of carbohydrate biomass into useful substances, J Mater Sci, 43 (2008) 2472-2475.

[50] A.A. Efremov, G.G. Pervyshina, B.N. Kuznetsov, Thermocatalytic transformations of wood and cellulose in the presence of HCl, HBr, and H2SO4, Chem Nat Compd, 33 (1997) 84-88.

[51] W.O.S. Doherty, P. Mousavioun, C.M. Fellows, Value-adding to cellulosic ethanol: Lignin polymers, Industrial Crops and Products, 33 (2011) 259-276.

[52] R.A. Schraufnagel, H.F. Rase, Levulinic acid from sucrose using acidic ion-exchange resins, Product R&D, 14 (1975) 40-44.

[53] J. Jow, G.L. Rorrer, M.C. Hawley, D.T.A. Lamport, Dehydration of d-fructose to levulinic acid over LZY zeolite catalyst, Biomass, 14 (1987) 185-194.

[54] K. Lourvanij, G.L. Rorrer, Dehydration of glucose to organic acids in microporous pillared clay catalysts, Applied Catalysis A: General, 109 (1994) 147-165.

[55] J. Hegner, K.C. Pereira, B. DeBoef, B.L. Lucht, Conversion of cellulose to glucose and levulinic acid via solid-supported acid catalysis, Tetrahedron Letters, 51 (2010) 2356-2358.

[56] R. Weingarten, W.C. Conner, G.W. Huber, Production of levulinic acid from cellulose by hydrothermal decomposition combined with aqueous phase dehydration with a solid acid catalyst, Energy & Environmental Science, 5 (2012) 7559-7574.

[57] R.W. Thomas, H.A. Schuette, Studies on levulinic acid. I. It’s preparation from carbohydrates by digestion with hydrocloric acid under pressure, Journal of the American Chemical Society, 53 (1931) 2324-2328.

[58] D.M. Alonso, S.G. Wettstein, M.A. Mellmer, E.I. Gurbuz, J.A. Dumesic, Integrated conversion of hemicellulose and cellulose from lignocellulosic biomass, Energy & Environmental Science, 6 (2013) 76-80.

[59] J.Y. Cha, M.A. Hanna, Levulinic acid production based on extrusion and pressurized batch reaction, Industrial Crops and Products, 16 (2002) 109-118.

Page 63: Formic and Levulinic Acid from Cellulose via Heterogeneous

54

[60] H. Chen, B. Yu, S. Jin, Production of levulinic acid from steam exploded rice straw via solid superacid, Bioresource Technology, 102 (2011) 3568-3570.

[61] Y. Bin, C. Hongzhang, Effect of the ash on enzymatic hydrolysis of steam-exploded rice straw, Bioresource Technology, 101 (2010) 9114-9119.

[62] V. Chang, M. Holtzapple, Fundamental factors affecting biomass enzymatic reactivity, M. Finkelstein, B. Davison (Eds.) Twenty-First Symposium on Biotechnology for Fuels and Chemicals, Humana Press, 2000, pp. 5-37.

[63] V.M. Ghorpade, M.A. Hanna, Method and apparatus for production of levulinic acid via reactive extrusion (US Patent 5,859,263, 1999).

[64] S. Van de Vyver, J. Thomas, J. Geboers, S. Keyzer, M. Smet, W. Dehaen, P.A. Jacobs, B.F. Sels, Catalytic production of levulinic acid from cellulose and other biomass-derived carbohydrates with sulfonated hyperbranched poly(arylene oxindole)s, Energy & Environmental Science, 4 (2011) 3601-3610.

[65] G. Banerjee, S. Car, T. Liu, D.L. Williams, S.L. Meza, J.D. Walton, D.B. Hodge, Scale-up and integration of alkaline hydrogen peroxide pretreatment, enzymatic hydrolysis, and ethanolic fermentation, Biotechnology and Bioengineering, 109 (2012) 922-931.

[66] G. Banerjee, S. Car, J.S. Scott-Craig, D.B. Hodge, J.D. Walton, Alkaline peroxide pretreatment of corn stover: effects of biomass, peroxide, and enzyme loading and composition on yields of glucose and xylose, Biotechnol Biofuels, 4 (2011) 16.

[67] M.F. Zaranyika, M. Madimu, Heterogeneous dilute acid hydrolysis of cellulose: A kinetic model for the hydrolysis of the difficultly accessible portion of cellulose based on donnan's theory of membrane equilibria, Journal of Polymer Science Part A: Polymer Chemistry, 27 (1989) 1863-1872.

[68] L. Liu, J. Sun, C. Cai, S. Wang, H. Pei, J. Zhang, Corn stover pretreatment by inorganic salts and its effects on hemicellulose and cellulose degradation, Bioresource Technology, 100 (2009) 5865-5871.

[69] J. Potvin, E. Sorlien, J. Hegner, B. DeBoef, B.L. Lucht, Effect of NaCl on the conversion of cellulose to glucose and levulinic acid via solid supported acid catalysis, Tetrahedron Letters, 52 (2011) 5891-5893.

Page 64: Formic and Levulinic Acid from Cellulose via Heterogeneous

55

[70] S.G. Wettstein, D.M. Alonso, Y. Chong, J.A. Dumesic, Production of levulinic acid and gamma-valerolactone (GVL) from cellulose using GVL as a solvent in biphasic systems, Energy & Environmental Science, 5 (2012) 8199-8203.

[71] I.T. Horvath, H. Mehdi, V. Fabos, L. Boda, L.T. Mika, [gamma]-Valerolactone-a sustainable liquid for energy and carbon-based chemicals, Green Chemistry, 10 (2008) 238-242.

[72] J.-P. Lange, J.Z. Vestering, R.J. Haan, Towards 'bio-based' Nylon: conversion of [gamma]-valerolactone to methyl pentenoate under catalytic distillation conditions, Chemical Communications, (2007) 3488-3490.

[73] M. Chalid, H.J. Heeres, A.A. Broekhuis, Ring-opening of γ-valerolactone with amino compounds, Journal of Applied Polymer Science, 123 (2012) 3556-3564.

[74] J.Q. Bond, D.M. Alonso, D. Wang, R.M. West, J.A. Dumesic, Integrated catalytic conversion of γ-valerolactone to liquid alkenes for transportation fuels, Science, 327 (2010) 1110-1114.

[75] J.C. Serrano-Ruiz, D.J. Braden, R.M. West, J.A. Dumesic, Conversion of cellulose to hydrocarbon fuels by progressive removal of oxygen, Applied Catalysis B: Environmental, 100 (2010) 184-189.

[76] S.K.R. Patil, C.R.F. Lund, Formation and growth of humins via aldol addition and condensation during acid-catalyzed conversion of 5-hydroxymethylfurfural, Energy & Fuels, 25 (2011) 4745-4755.

[77] V.G. Devulapelli, H.-S. Weng, Synthesis of cinnamyl acetate by solid–liquid phase transfer catalysis: Kinetic study with a batch reactor, Catalysis Communications, 10 (2009) 1638-1642.

[78] M. Käldström, N. Kumar, D.Y. Murzin, Valorization of cellulose over metal supported mesoporous materials, Catalysis Today, 167 (2011) 91-95.

[79] A. Sundberg, K. Sundberg, C. Lillandt, B. Holmbom, Determination of hemicelluloses and pectins in wood and pulp fibres by acid methanolysis and gas chromatography, Nordic Pulp and Paper Research Journal, 11 (1996) 216-219.

[80] P. Strunk, T. Öman, A. Gorzsás, M. Hedenström, B. Eliasson, Characterization of dissolving pulp by multivariate data analysis of FT-IR and NMR spectra, Nordic Pulp and Paper Research Journal, 26 (2011) 398.

Page 65: Formic and Levulinic Acid from Cellulose via Heterogeneous

56

[81] S. Willför, A. Sundberg, A. Pranovich, B. Holmbom, Polysaccharides in some industrially important hardwood species, Wood Science and Technology, 39 (2005) 601-617.

[82] A. Sundberg, S. Emet, P. Rhen, l. Vähäsalo, B. Holmbohm, Estimation of proportions of different pulp types in recycled paper and deinked pulp by acid methanolysis, Proc. of 13th International Symposium on Wood, Fibre and Pulping Chemistry, Auckland, New Zealand, 2005, pp. 361-365.

[83] P.F. Siril, H.E. Cross, D.R. Brown, New polystyrene sulfonic acid resin catalysts with enhanced acidic and catalytic properties, Journal of Molecular Catalysis A: Chemical, 279 (2008) 63-68.

[84] R. Bringué, M. Iborra, J. Tejero, J.F. Izquierdo, F. Cunill, C. Fité, V.J. Cruz, Thermally stable ion-exchange resins as catalysts for the liquid-phase dehydration of 1-pentanol to di-n-pentyl ether (DNPE), Journal of Catalysis, 244 (2006) 33-42.

[85] Z.P. Xu, K.T. Chuang, Effect of internal diffusion on heterogeneous catalytic esterification of acetic acid, Chemical Engineering Science, 52 (1997) 3011-3017.

[86] D. Harvey, Modern analytical chemistry, McGraw-Hill Boston, (2000).

[87] C. Corradini, Liquid Chromatography, (2000).

[88] Willard, H.H., Dean, J.A., Settle, F.A., and Merritt, L.L., Instrumental methods of analysis, 7th ed.; Wadsworth: Belmont, CA (1988).

[89] A. Kraj, D.M. Desiderio, N.M. Nibbering, Mass spectrometry: instrumentation, interpretation, and applications, John Wiley & Sons, (2008).

[90] Silylating agents, http://www.sigmaaldrich.com/content/dam/sigma-aldrich/docs/Aldrich/General_Information/silylation_overview.pdf, (accessed 10.2.2014).

[91] H. Haario, Modest user´s guide, Profmath Oy, Helsinki, 1994.

[92] M. Saeed, M. Depala, D. Craston, T. Lynch, A standard addition method for the quantitative determination of succinic acid and levulinic acid in processed acid stream extracts employing indirect UV capillary electrophoresis, Chromatographia, 47 (1998) 709-715.

Page 66: Formic and Levulinic Acid from Cellulose via Heterogeneous

57

[93] A. Gildberg, J. Raa, Properties of a propionic acid/formic acid preserved silage of cod viscera, Journal of the Science of Food and Agriculture, 28 (1977) 647-653.

[94] B. Girisuta, L.P.B.M. Janssen, H.J. Heeres, A kinetic study on the decomposition of 5-hydroxymethylfurfural into levulinic acid, Green Chemistry, 8 (2006) 701-709.

[95] N. Akiya, P.E. Savage, Role of water in formic acid decomposition, AIChE Journal, 44 (1998) 405-415.

[96] W. Hayduk, W. Landie, A.I.Ch.E.J, 19 (1973) 1233.

[97] J.J. Panek, P.K. Wawrzyniak, Z. Latajka, J. Lundell, Interaction energy decomposition analysis for formic acid complexes with argon and krypton atoms, Chemical Physics Letters, 417 (2006) 100-104.

[98] D.W. Marquardt, An algorithm for least-squares estimation of nonlinear parameters, Journal of the Society for Industrial & Applied Mathematics, 11 (1963) 431-441.

[99] W. Spendley, G.R. Hext, F. Himsworth, Sequential application of simplex designs in optimisation and evolutionary operation, Technometrics, 4 (1962) 441-461.

[100] N. Metropolis, A.W. Rosenbluth, M.N. Rosenbluth, A.H. Teller, E. Teller, Equation of state calculations by fast computing machines, The journal of chemical physics, 21 (1953) 1087.