THE USE OF MEMBRANE FILTRATION TO IMPROVE THE PROPERTIES OF EXTRACTED WOOD COMPONENTS Mikaela Helander Doctoral Thesis Kungliga Tekniska Högskolan, Stockholm 2014 AKADEMISK AVHANDLING Som med tillstånd av Kungliga Tekniska Högskolan i Stockholm framlägges till offentlig granskning för avläggande av teknisk doktorsexamen fredagen den 25 april 2014, kl. 10.00 i sal F3, Lindstedtsvägen 26, KTH, Stockholm. Avhandlingen försvaras på engelska. Fakultetsopponent: Professor Arthur J. Ragauskas, Georgia Institute of Technology. Supervisor Professor Mikael E. Lindström Copyright © 2014 Mikaela Helander All rights reserved Paper I © 2013, BioResources Paper II © 2014, Submitted to Journal of Applied Polymer Science Paper III © 2014, Manuscript Paper IV © 2012, Current Organic Chemistry Paper V © 2012, BioResources TRITA-CHE Report 2014:9 ISSN 1654-1081 ISBN 978-91-7595-046-4 To all the kind people along the way. “Kindness is more important than wisdom, and the recognition of this is the beginning of wisdom.” Theodore Isaac Rubin ABSTRACT The forest is a large and important natural resource in Sweden, and approximately 70% of the country’s land area is woodland. Wood is an excellent raw material for the replacement of oil-based products because it is renewable, biodegradable and carbon neutral. Furthermore, the forest industry is searching for new processes and methods to utilise by-product streams in a so-called integrated biorefinery. A key to the success of producing new products from wood could be pure and homogenous raw materials. Because wood contains a large variety of components with different characteristics and sizes, cross-flow filtration (CFF) will be a key separation technique to obtain homogenous and pure raw materials in the biorefinery concept. Different wood material components have been studied in this thesis. The first part of this work focuses on kraft lignin. Kraft lignin is interesting because approximately 3.5-4 million metric tonnes are produced in Sweden annually (~7 million tonnes of kraft pulp/year in 2012), and today it is mainly used as fuel. The second part of this thesis deals with materials in hot water extract (i.e., galactoglucomannan, but also other components). These extracted materials are interesting because similar materials are extracted in thermomechanical pulping process by-product streams. Finally, nanocellulose has been studied since it is an interesting raw material for future applications. Through CFF, kraft lignin from black liquor has been fractionated into raw material samples with more homogeneous characteristics such as molecular weight, aromatic hydroxyl groups and thermal properties. From dynamic mechanical analysis, low molecular weight fractions were found to have the highest degree softening. To precipitate low molecular weight fractions (<1000 Da cut-off) into a convenient solid, lower temperatures than for high molecular weight fractions were needed. To produce low molecular weight lignin (<5000 Da cut-off) from re-dissolved LignoBoost lignin, lower lignin concentrations and higher pH and ionic strength were found to increase the permeate flux at the tested conditions. Nanocellulose has been produced by a novel process called nanopulping and has subsequently been size fractionated by CFF to obtain more homogenous nanocellulose. CFF and adsorption chromatography can be used to isolate dissolved wood components, yielding several upgraded products: lignin, lignin-carbohydrate complexes, and galactoglucomannan. SAMMANFATTNING Skogen är en viktig naturtillgång för Sverige och omkring 70% av Sveriges landarea består av skog. Då ved är ett förnybart, biologiskt nedbrytbart och koldioxidneutralt material skulle det kunna vara ett utmärkt råmaterial för att ersätta oljebaserade produkter. Samtidigt söker skogsindustrin nya processer och metoder för att utnyttja biprodukter från ved i så kallade integrerade bioraffinaderier. En nyckel till framgång, för att ta fram nya förädlade slutprodukter av ved, kan vara användandet av rena och homogena vedkomponenter. Eftersom ved innehåller en stor variation av komponenter med olika egenskaper och storlekar kan tvärströmsfiltrering (CFF) därmed vara en viktig separationsteknik för att erhålla homogena och rena råmaterial i bioraffinaderi-konceptet. I denna avhandling har olika materialkomponenter från ved studerats. Den första delen av arbetet behandlar sulfatlignin. Sulfatlignin är intressant då det produceras cirka 3.5-4 miljoner ton sulfatlignin årligen i Sverige (baserat på ~ 7 miljoner ton sulfatmassa/år 2012), och i dag används sulfatligninet främst som bränsle. Den andra delen av denna avhandling handlar om vedmaterial från varmvatten-extrakt (dvs. främst hemicellulosan galaktoglukomannan men även andra komponenter). Dessa extraherade vedkomponenter är intressanta då liknande material extraheras som biprodukter i den termomekaniska massatillverkningen. Slutligen har nanocellulosa studerats då råmaterialet har intressanta egenskaper för framtida material. I det praktiska arbetet har CFF använts för att fraktionera sulfatlignin från svartlut till homogenare råmaterial med lägre molekylviktsfördelning, vilket resulterade i enhetligare mängder av aromatiska hydroxylgrupper samt homogenare termiska egenskaper. Dessutom uppvisade lågmolekylära ligninfraktioner de lägsta glastransitionstemperaturerna, som lägst 70°C. För att fälla lågmolekylära ligninfraktioner (<1000 Da cut-off) till en hanterbar ligninsuspension, krävdes lägre temperaturer än för de högmolekylära fraktionerna. Vid produktion av lågmolekylärt sulfatlignin (<5000 Da cut-off) från återupplöst LignoBoost lignin, gynnades permeatfluxet av lägre ligninkoncentrationer, högre pH och jonstyrka vid de testade förhållandena. Nanocellulosa har tagits fram av en ny process som kallas nanopulping, och därefter storleksfraktioneras med CFF för att erhålla homogenare nanocellulosa. CFF och adsorptionskromatografi har använts för att isolera upplösta träkomponenter i flera uppgraderade produkter: lignin, lignin-kolhydratkomplex och galaktoglukomannan. LIST OF PAPERS This thesis is a summary of the following papers: I. II. III. IV. V. "Fractionation of Technical Lignin: Molecular Mass and pH Effects" Mikaela Helander, Hans Theliander, Martin Lawoko, Gunnar Henriksson, Liming Zhang and Mikael E. Lindström, BioResources, 2013, 8 (2), 2270-2282. “Tailoring the molecular and thermo-mechanical properties of kraft lignin by ultrafiltration” Olena Sevastyanova, Mikaela Helander, Sudip Chowdhury, Heiko Lange, Helena Wedin, Liming Zhang, Monica Ek, John F. Kadla, Claudia Crestini and Mikael E. Lindström, Journal of Applied Polymer Science, Submitted. “Parameters effecting cross-flow filtration of LignoBoost kraft lignin” Mikaela Helander, Tuve Mattsson, Hans Theliander, Mikael Lindström, 2014, Manuscript. "A Novel Nano Cellulose Preparation Method and Size Fraction by Cross Flow Ultra-Filtration" Hongli Zhu, Mikaela Helander, Carl Moser, Adam Ståhlkranz, Daniel Söderberg, Gunnar Henriksson and Mikael E. Lindström, Current Organic Chemistry, 2012, 16, 1871-1875. "Separation of Galactoglucomannans, Lignin, and Lignin-Carbohydrate Complexes from Hot-Water-Extracted Norway Spruce by Cross-Flow Filtration and Adsorption Chromatography" Niklas Westerberg, Hampus Sunner, Mikaela Helander, Gunnar Henriksson, Martin Lawoko and Anders Rasmuson, BioResources, 2012, 7 (4), 4501-4516. Author´s contribution to the appended papers: I. Principal author. Took part in outlining the experiments and performed all experimental work at KTH and most of the preparation of the manuscript. II. Second author. Took part in outlining some of the experiments, performed some of the experiments and some of the preparation of the manuscript. Thermal analyses were performed at the University of British Colombia, Vancouver, Canada. III. Principal author. Took part in outlining the experiments, performed all experimental work and some of the preparation of the manuscript. IV. Principal author/equal contributing author. Took part in outlining the experiments and performed some of the experiments and some of the preparation of the manuscript. V. Third author. Took part in outlining the filtration part and its analysis of the experiment and performed the filtration part and its analysis of the experiments and some of the preparation of the manuscript. This thesis also contains unpublished results. Other related materials: Wallenberg Wood Science Center, Theme 1 (2014/2015): Demonstrator, Manuscript. Helander, M., Theliander, H., Lawoko, M., Henriksson, G., Zhang L., and Lindström, M. E. (2012): Lignin for new materials - molar mass and pH effects, 243rd National Meeting of the American Chemical Society & Exposition (“Spring 2012 National Meeting and Exposition”), Division “Cellulose and Renewable Materials, Session “Lignin as Renewable Raw Material”, Mars 25-29, San Diego, California. Helander, M., Sevastyanova, O., Chowdhury, S., Lange, H., Zhang, L., Wedin, H., Crestini, C., Ek, M., Kadla, J. F. and Lindström, M. E. (2013): Evaluation of physicochemical properties and prediction of spinning parameters for high-quality lignins produced by ultrafiltration of industrial kraft pulping liquor, 217th International Symposium on Wood, Fiber and Pulping Chemistry, June 12-14, Vancouver, Canada. Helander, M. and Lindström, M. (2013): Cross-flow filtration for fractionation of wood components, Marcus Wallenberg price awards, September 23-24, Stockholm, Sweden. Helander, M. (2014): Cross-flow filtration for fractionation of wood components, Ekmandagarna, January 28-30, Stockholm, Sweden. Helander, M. (2011): Lignin particles for New materials, SPCI (Svenska Cellulosaoch Pappers Ingenjörsföreningen), May 17-19, Stockholm, Sweden. WALLENBERG WOOD SCIENCE CENTER - WWSC This thesis is part of the research initiated by the Knut and Alice Wallenberg Foundation through the formation of a Swedish research centre, named the Wallenberg Wood Science Center, at the Chalmers University of Technology in Gothenburg and the Royal Institute of Technology in Stockholm. The mission and vision of the WWSC is to create new materials from trees and to develop the processes to produce these materials. During the initial five years, the centre was divided into five themes. This thesis was part of Theme 1 – “From wood chips to material components”. This theme’s focus is on raw materials, and is the first step towards the production of new products that are dependent on the development of efficient and economically feasible novel processes for the separation of wood components. TABLE OF CONTENTS 1. INTRODUCTION ..................................................................................................... 1 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 1.10 BACKGROUND ....................................................................................................... 1 CROSS-FLOW FILTRATION ..................................................................................... 2 .......................................................................................... 5 EXTRACTION/PRODUCTION OF KRAFT LIGNIN ........................................................ 6 FRACTIONATION OF KRAFT BLACK LIQUOR AND KRAFT LIGNIN ............................ 8 NANOCELLULOSE ................................................................................................ 10 PRODUCTION OF NANOCELLULOSE ...................................................................... 10 ENERGY-EFFICIENT METODS FOR NANOCELLULOSE ............................................ 11 HOT WATER EXTRACT FROM SOFTWOOD ............................................................. 13 FRACTIONATION OF HOT WATER EXTRACT FROM SOFTWOOD.............................. 14 KRAFT BLACK LIQUOR 2. AIM OF INVESTIGATION ................................................................................... 15 3. EXPERIMENTAL ................................................................................................... 16 3.1 MATERIALS ......................................................................................................... 16 Black liquor................................................................................................ 16 Kraft lignin ................................................................................................. 16 Kraft pulp ................................................................................................... 16 Softwood .................................................................................................... 16 MEMBRANES AND FRACTIONATION ..................................................................... 17 3.2 4. RESULTS AND DISCUSSION .............................................................................. 18 4.1 PRODUCTION OF LIGNIN FRACTIONS (PAPERS I-III) .............................................. 18 Precipitation and filtration of weak black liquor ....................................... 19 Precipitation and filtration of <1000 Da cut-off permeate ........................ 19 Composition of precipitated lignin ............................................................ 20 Elemental analysis for precipitated lignin ................................................. 21 Yield for <1000 Da cut-off precipitated lignin .......................................... 23 Molecular weight data of precipitated lignin............................................. 24 Chemical characteristics and thermal properties of precipitated lignin ... 25 Cross-flow filtration of dissolved LignoBoost lignin ................................. 29 FRACTIONATION OF NANOCELLULOSE (PAPER IV) ............................................... 37 Energy demand for nanopulping................................................................ 37 Effect of different pretreatment methods .................................................... 38 Fractionation by cross-flow filtration ........................................................ 39 4.2 4.3 FRACTIONATION OF HOT WATER EXTRACT FROM SOFTWOOD (PAPER V) ............. 40 Yield for membrane fractionated hot water extract ................................... 40 Composition of membrane fractionated hot water extract ......................... 41 Molecular weight distribution for fractionated hot water extract ............. 43 Separation by chromatography of fractionated hot water extract ............. 44 5. CONCLUDING REMARKS .................................................................................. 46 6. NOTATIONS ........................................................................................................... 49 7. ACKNOWLEDGEMENTS .................................................................................... 50 8. REFERENCES ........................................................................................................ 51 Introduction 1. INTRODUCTION 1.1 BACKGROUND The forest industry is one of the largest industries in Sweden and contributes a large portion of Sweden’s export income (Skogsindustrin 2012). Today, however, the pulp and paper industry face challenges due to competition from fast-growing trees, digital reading and industrial investments in competing cost-effective countries. For the Swedish forest industry to remain one of the largest, it is necessary to be at the forefront in regard to new processes and products from wood. Therefore, much research has been focused on converting the forest industry’s by-product streams into new products of high value in a so-called integrated biorefinery and also on finding new processes for wood components. For example, in the pulping process, some liquid process streams contain extracted wood components as by-products, which are not being used to generate high value application products. Intensive research has been performed on thermomechanical process waters that mainly contain galactoglucomannan (Willför et al. 2008) and black liquor (Antonsson 2008, Norberg et al. 2012). Although promising, they have not yet reached industrial levels for high value application products. A common feature of these process streams is that they contain inhomogeneous wood components (i.e., different components, characteristics and sizes), and this inhomogeneity may limit opportunities to create new high value application materials. Separation of these process streams 1 Introduction into more homogenous and pure fractions could increase the opportunities for raw materials with different properties to be used in new materials or for chemical modifications (Stewart 2008). 1.2 CROSS-FLOW FILTRATION Cross-flow filtration (CFF) is a dynamic filtration technique that can separate particles and molecules in liquids having different sizes and forms. In the separation, a semipermeable membrane is used. Depending on the pore size of the membrane (cut-off) the process is divided into different levels of filtration: microfiltration, ultrafiltration, nanofiltration, and reverse osmosis (see Figure 1). Bacteria Protein Polysaccharide Salt Water Microfiltration > 0.1 µm Ultrafiltration 10-100 nm Nanofiltration > 1 nm Reverse osmosis < 1 nm Figure 1. Adapted from Nordin (2008). Different levels of filtration. Pore sizes given in µm and nm are rough values (Ripperger et al. 2009). In CFF, the liquid sample (i.e., the feed) flows parallel to the filter media. By introducing a transmembrane pressure (TMP) across the membrane, the liquid is forced towards the membrane, and materials smaller than the membrane’s pores will pass through the membrane as a permeate (see Figure 2). The velocity of the cross-flow, parallel to the filter media, is relatively high, and the flow is turbulent, creating dynamic mechanisms that reduce the build-up of particles and molecules 2 Introduction on the membrane surface. However, there will be an increased concentration of material closer to the membrane surface due to the retained particles; this phenomenon is called concentration polarisation. Tank Feed Retentate Membrane Pump Membrane Retentate Feed Permeate (filtrate) Permeate Figure 2. Illustration of a filtration unit (left picture) and a cross-section of a tubular membrane separation (right picture). The TMP is the driving force for separation (see Equation 1), where Pf is the feed pressure, Pr is the retentate pressure and Pp is the permeate pressure. Equation 1: ( ) The flow velocity though the membrane is defined by Equation 2, where v is the flow velocity in m/s, Q is the feed flow in m3/s, and A is the membrane area in m2. Equation 2: The filtration volume reduction and volume reduction factor can be calculated using Equation 3 and Equation 4, respectively. Where VR is the volume reduction, 3 Introduction VRF is the volume reduction factor, Vp is the volume of the permeate, Vif is the initial feed volume, and Vr is the volume of the retentate. Equation 3: Equation 4: The rejection coefficient is a measure of how efficient the separation between the material in the feed is versus the material in the permeate (see Equation 5), where R is the rejection coefficient, cp is the concentration in the permeate, and cf is the concentration in the feed. Equation 5: The permate flux can be calculated according to Equation 6, here J is the flux, Vp is the volume of the permeate, h is the filtration time in hour, and m2 is the area of the membrane. Equation 6: 4 Introduction 1.3 KRAFT BLACK LIQUOR In the kraft process (also known as kraft pulping or sulphate process), wood chips are digested in an alkaline cooking liquor called white liquor, which contains the cooking chemicals sulphide ions and hydroxide ions. The chemicals in white liquor dissolve wood components at pH > 13 and temperatures between 150 and 170°C (in the so called primary separation). From the digester, a mixture of liberated cellulose fibres and a liquor containing the dissolved wood components and spent cooing chemicals (weak black liquor) are obtained. In weak black liquor (i.e., before the evaporators) the dry content is approximately 15-20% (Theliander 2009). The main component degraded and dissolved into black liquor is lignin, which represents approximately 30-45% of the dry material. Other materials in black liquor are degraded carbohydrate acids (25-35% hydroxyl acids, 5% acetic acid and 3% formic acid), 3-5% extractives (fatty acids and resins), hemicellulose, 3-5% sulphur, and 17-20% sodium. The inorganics in black liquor are sodium hydroxide (NaOH), sodium sulphide (Na2S), sodium carbonate (Na2CO3), sodium sulphate (Na2SO4), sodium thiosulphate (Na2S2O3), and sodium chloride (NaCl) (Grace 1989). Other elements are present in small amounts (< 0.1%), for example, magnesium (Mg), manganese (Mn), aluminium (Al), silicon (Si), iron (Fe), and phosphorus (P) (Theliander 2009). Lignin is a macromolecule that is fragmented and extracted from wood chips during the pulping process. Hence, the lignin found in black liquor has a broad molecularweight distribution, as for example reported by Mörck et al. (1986). The chemical composition in the fractions is also inhomogeneous; it has previously been shown that high molecular weight lignin contains increased amounts of aliphatic groups, while low molecular weight lignins contain an increased amount of phenolic hydroxyl groups (Wada et al. 1962). Today, approximately 3.5-4 million metric tonnes of non-homogenous lignin is being produced each year by the kraft 5 Introduction pulping process in Sweden, from ~7 million tonnes kraft pulp/year (Skogsindustrin 2012). This lignin is mainly burned, generating heat, in the process of recovering pulping chemicals in the recovery boiler. A modern pulp mill, however, has an energy surplus. Furthermore, in many cases, the recovery boilers are the so called bottleneck. Consequently, by extracting a part of the lignin, it is possible to de-bottleneck the recovery boiler, resulting in a pulp production increase and, at the same time, obtaining a new product, the lignin, from the kraft process. 1.4 EXTRACTION/PRODUCTION OF KRAFT LIGNIN To extract lignin from black liquor, the solubility of lignin molecules can be utilised. Because the pH of kraft liquors is ~13-14, the phenolic hydroxyl groups in these liquors are almost completely deprotonated. By lowering the pH of the black liquor with acid, the phenolic hydroxyl groups in lignin can be protonated (Alén et al. 1979, Passinen 1968). The level of protonation depends on the pKa of the lignin structures. Softwood lignin mainly consists of coniferyl alcohol. The pKa value of coniferyl alcohol at room temperature is approximately 10.2 (Norgren and Lindström 2000). There are also some carboxylic groups with lower pKa values (Ragnar et al. 2000); however, the carboxylic acids are often in small amounts (Granata and Argyropoulos 1995, Norgren and Lindström 2000). Furthermore, lignin will coagulate depending on the conditions (Sundin 2000). The pKa depends mainly on the temperature and ionic strength (Ragnar et al. 2000) and on the molecular weight (Norgren and Lindström 2000). The dry content of black liquor also influences the efficiency of precipitation (Alén et al. 1979 and 1985, Lin 1992). In 1872, Tessié du Motay patented a method in which carbon dioxide was injected into hot liquors obtained by boiling woody fibre, also known as spent 6 Introduction liquor from the soda process. The aim of this method was to precipitate “impurities” and consequently obtain a pure liquor that could be recaustified and reused (Tessié du Motay 1872). Other researchers have also studied the use of carbon dioxide to precipitate material from black liquors (Rinman 1911, Scott 1940, Reboulet 1941, Tomlinson and Tomlinson Jr. 1946 and 1948, Pollak et al. 1949). In 1910, William J. Hough patented a method in which lignin and resin were precipitated by acidification of spent liquors from the alkaline pulping process using an acid, such as sulphuric acid. To improve the dead-end filtration (i.e., cake filtration, a filtration process were liquid passes through the membrane and the particles retained building up a filter cake), the precipitated solution was filtered at a high temperature (Hough 1910). A problem that was recently solved by the LignoBoost process is the displacement washing of precipitated lignin to obtain a purified lignin without plugging the filter cake or incurring large material losses. After precipitation, the material is filtered and the filter cake is re-dispersed at a low pH (2-4). The suspension that is formed can then be easily re-filtered and displacement washed (Öhman and Theliander 2006, Öhman et al. 2007). The products of this process are lignin-lean black liquor that is returned to the pulping process and highly pure lignin that can be used for alternative applications. For example, lignin with 0.21% ash content and 1.5-2.8% carbohydrates can be produced by this process using softwood lignin (Tomani 2013, Öhman et al. 2007), a chemical structure characterization has also been reported (Nagy et al. 2010). The LignoBoost process has been a great success with regard to potentially decreasing the load on recovery boilers, which, as previously mentioned, are usually bottlenecks in the kraft pulping process. The LignoBoost process is, however, quite new, and no publications to date have reported any industrial established high value applications. 7 Introduction Another method for the extraction of kraft lignin from black liquor, referred to as “liquid lignin”, has been developed (Lake and Blackburn 2011). In this technique, a phase separation occurs, and decantation is used instead of filtration (Velez and Thies 2013). Lignin extracted with this process is reported to have an ash content of approximately 1-2% (Velez and Thies 2013) and a carbohydrate content of approximately 3%. 1.5 FRACTIONATION OF KRAFT BLACK LIQUOR AND KRAFT LIGNIN Over the years, different techniques have been used to fractionate lignin, including acidification (Wada et al. 1962), Soxhlet extraction (Lindberg et al. 1964), gel permeation chromatography (Mörck et al. 1982), organic solvents (Mörck et al. 1986, Yoshida et al. 1987) and ultrafiltration (Lin and Detroit 1981). Cross-flow ultrafiltration of kraft black liquor was studied in the mid-1980s, and various parameters were adjusted in an attempt to optimise the filtration for concentration and purification of high molecular weight lignin (Woerner and McCarthy 1984). Several studies have examined the fractionation of softwood black liquor using ultrafiltration (Forss and Fuhrmann 1976, Woerner and McCarthy 1984, Griggs et al. 1985, Alén et al. 1986, Drouin and Desrochers 1988, Uloth and Wearing 1989, Wallberg et al. 2003, Keyoumu et al. 2004, Rojas et al. 2006, Brodin et al. 2009). It has been demonstrated that when lignin is fractionated from kraft black liquor by precipitation or filtration, the phenolic hydroxyl contents are increased at lower molecular weights (Wada et al. 1962, Griggs et al. 1985, Lin 1992). These phenolic structures have interesting properties that could be used in various 8 Introduction applications, for example they are possible sites for modification and can be used as an antioxidant (Weber and Weber 2010, Pouteau 2003). A few studies have examined the possibility of separating lignin from the 1000 Da fraction of kraft black liquor by filtration (Keyoumu et al. 2004, Rojas et al. 2006, Elegir et al. 2007, Antonsson et al. 2008, Niemi et al. 2011). Some researchers have focused on utilising the higher content of phenolic hydroxyl groups in low molecular weight lignin to produce new materials (Antonsson et al. 2008); however, a convenient method for extracting the low molecular weight lignin is yet to be found. Little has been reported on CFF of extracted and re-dissolved kraft lignin, although a systematic study could lead to better process understanding and facilitate further process improvements. Woerner and McCarthy (1987) reported studies on filtration of re-dissolved Indulin AT. As mentioned previously, the novel LignoBoost method has gained commercial status for extracting lignin from black liquor. The molecular size distribution of this extracted lignin is, however, rather broad, which is possibly a drawback for certain applications (Stewart 2008, Vishtal and Kraslawski 2011). Hence, it could be of interest to enhance the properties of LignoBoost lignin via membrane fractionation to obtain a more homogeneous lignin. There have been investigations that attempt to use unfractionated and fractionated lignin for the production of carbon fibres (Kadla et al. 2002, Braun et al. 2005, Kubo et al. 2005, Norberg et al. 2012) and investigations into the functional groups and thermal properties of lignin (Rojas et al. 2006, Brodin et al. 2009, Toledano et al. 2010), but a convenient method that could predict spinning properties a priori has not been established. For lignin-derived products that will be processed at elevated temperatures, it could be advantageous to alter the 9 Introduction thermal properties of kraft lignin to achieve the desired performance. By combining an industrially feasible technique such as CFF with kraft lignin, a systematic study on the behaviour of fractionated kraft lignin for the production of materials with tailored thermomechanical properties would be of interest. 1.6 NANOCELLULOSE Cellulose is interesting as a sustainable material due to its good mechanical properties, biodegradability and renewable origin. The traditional ways to utilise cellulose have been as “fibres”, that is more or less intact cell walls with lengths from approximately 0.5 mm to several centimetres, in e.g. paper and textiles. During the last few decades, however, there has been interest in producing and using cellulose as nanosized fibres. This nanocellulose is interesting due to its high surface to volume ratio, light-weight and high specific strength and can be useable in, composites, foams and films (Eichhorn et al. 2010). Nanocellulose has been divided into three types: microfibrillated cellulose (MFC), nanocrystalline cellulose (NCC), and bacterial nanocellulose (BNC). MFC and NCC reported with average diameters of 5-60 nm and 5-60 nm, respectively, and lengths of several micrometres and 100-250 nm (from plant), respectively (Klemm et al. 2011). However, today other abbreviations are commonly used. Over the past ~10 years the numbers of articles and patents concerning nanofibrillar cellulose have increased (Siró and Plackett 2010). 1.7 PRODUCTION OF NANOCELLULOSE In the beginning of the 1980s, the first work on MFC from wood was published (Herrick et al. 1983, Turbak et al. 1983), and in 1982 and 1983, patents were 10 Introduction issued (U.S. 4,341,807, U.S. 4,374,702, U.S. 4,378,381). These works involved a method wherein pulp was treated in several passes through a homogeniser with a high pressure drop (Herrick et al. 1983, Turbak et al. 1983). Since then, other mechanical treatments have been investigated. Super-grinding is a method in which natural fibres pass through a specially designed grinder in cycles to produce nanocellulose (Taniguchi and Okamura 1998). Grinding was also used in a study by Abe et al. (2007), where nanocellulose was produced from chemically treated wood powder. Grinding has also been used after homogenisation to produce nanocellulose from kraft pulp fibres (Iwamoto et al. 2005). By refining kraft pulp to different degrees before homogenisation, Nakagaito et al. (2004) showed that nanocellulose with different properties can be created. In 2004, Zimmerman et al. reported the use of a microfluidizer for the microfibrillation of sulphite pulp. Cryocrushing was a method developed in 2005 by Chakraborty et al., in which wood pulp was refined and placed in liquid nitrogen and then crushed with a mortar. Beads have also been used to fibrillate cellulose particles (Ishikawa and Ide 1993). Similarly, another apparatus with spherical elements was addressed for the production of nanocellulose (Curtolo and Eksteen 2006). Furthermore, ultrasonic treatment of natural fibres in water produces nanocellulose, according to Zhao et al. (2007). In 2011, Chen et al. chemically treated sieved wood powder before ultrasonication and obtained cellulose nanofibres of 5-20 nm in width. 1.8 ENERGY-EFFICIENT METODS FOR NANOCELLULOSE The challenge with producing nanocellulose commercially has been the high energy demand to obtain a homogenous product. The energy input needed to produce nanocellulose can be between 20000 and 30000 kWh/tonne (Siró and 11 Introduction Plackett 2010). Different attempts have been made to reduce the energy required for nanocellulose production. Various pretreatments have been tested, such as oxidation and enzymatic treatments (Saito et al. 2006, Henriksson et al. 2007, Pääkkö et al. 2007). In 2006, Saito et al. reported a new way to produce nanocellulose using 2,2,6,6tetramethylpiperidine-1-oxyl (TEMPO) radical oxidation on never-dried cellulose (i.e., produced cellulose pulp that is never dried after production). By introducing a negative charge on the cellulose, swelling of the cellulose was increased and it was possible to produce nanocellulose by mechanical treatment. In 2007, Henriksson et al. showed that enzymatic pre-treatment with endoglucanase enhanced the separation of nanofibrills produced by mechanical treatment. Compared to acid hydrolysis as a pretreatment, enzymatic hydrolysis resulted in a higher nanocellulose aspect ratio. Enzymatic pretreatment to produce nanocellulose was also performed by Pääkkö et al. (2007). By the use of sulphite and dissolving pulp with different pretreatments prior to homogenisation in a microfluidizer, Ankerfors and Lindström produced nanocellulose with an energy demand of 500-2300 kWh/tonne (Ankerfors 2012), which is approximately tenfold lower than other nanocellulose production methods (as suggested by Siró and Plackett 2010). Although repeated homogenisation produces a relatively homogeneous nanocellulose, other more energy efficient nanocellulose preparation methods might produce homogeneous material. Recently, Madani et al. (2011) presented a study where nanocellulose was separated into different fractions using a hydrocyclone, pressurised screening or centrifugation in a fluid. Researchers have attempted different ways of producing low-cost homogenous nanocellulose, but 12 Introduction further technical progress to increase the numbers of methods available to produce nanocellulose would be interesting. 1.9 HOT WATER EXTRACT FROM SOFTWOOD Hot water treatment of wood chips or sawdust dissolves different wood components, especially hemicelluloses, into water. The degree of hemicellulose dissolution is higher at higher temperatures. Acetyl groups are easily cleaved off at these conditions, forming acetic acid in the extract and consequently lowering the liquid pH to 3-4 (Brasch and Free 1965). These acidic conditions cause hemicellulose degradation by autohydrolysis. Hence, degraded hemicelluloses can be found in thermomechanical pulping process liquids, prehydrolysis liquids and steam explosion extracts. These soluble hemicelluloses are considered a byproduct in the thermomechanical process liquid and could be used for high value application products, such as hydrogels (Lindblad et al. 2001) or gas barrier films (Hartman et al. 2006a,b). Applications for such hemicelluloses are also for the production of bioethanol (Ragauskas et al. 2006). The main hemicellulose in softwood is galactoglucomannan (GGM), which constitutes approximately 15-20% of softwood (Teleman 2009). With softwood as the feedstock, the main component dissolved during hot water extraction is GGM, and approximately 10% of GGM has been shown to be soluble in thermomechanical pulping liquid (Thoron et al. 1994); the high solubility of GGM is attributed to the high degree of acetylation and its comparatively low molar mass (Hartman et al. 2006b). Other hemicelluloses and lignin, are also degraded and dissolved to a lower degree. Under low pH conditions pseudo-lignin might be formed and affect the Klason lignin values (Sannigrahi et al. 2011). Persson et al. (2010) reported the following amounts of wood components from a 13 Introduction liquid stream in the thermomechanical process; 11 kg hemicellulose and 8 kg of aromatic compounds (values per tonne of produced pulp). 1.10 FRACTIONATION OF HOT WATER EXTRACT FROM SOFTWOOD There have been several studies on the isolation of hemicellulose from thermomechanical process liquids and hot water extracts of softwood. Person et al. (2010) and Andersson et al. (2007) studied the use of filtration to fractionate hemicellulose in thermomechanical process liquids, resulting in a maximum hemicellulose purity of approximately 80%. Willför et al. (2003) separated different components present in thermomechanical pulping process liquids by ultrafiltration and resin adsorption, resulting in a GGM with 95% purity. Song et al. studied extraction of GGM by pressurised hot water extraction (2008). However, little work has been reported regarding the combination of membrane filtration and two-step chromatography for the isolation of all dissolved wood components in upgraded fractions that consist of lignin, lignin-carbohydrate complexes (LCC) (i.e., molecules containing both lignin and carbohydrate), and GGM. 14 Aim of Investigation 2. AIM OF INVESTIGATION The objective of this thesis followed the main goal of the Wallenberg Wood Science Center and its subdivision “From Wood Chips to Material Components”, which this work was a part of. The aim of this thesis is to find and study new interesting raw materials and processes from softwood through the use of crossflow filtration (CFF). The focus on the utilisation of CFF to produce more homogenous raw materials is essential because this technique could be critical for the production of high value applications from wood components. Lignin from existing processes, black liquor and re-dissolved LignoBoost lignin, have been studied for potential CFF fractionation to produce products with specific properties. Nanocellulose and hemicellulose have also been used in combination with cross-flow filtration for the production of more homogenous raw materials 15 Experimental 3. EXPERIMENTAL Detailed descriptions of the experimental methods are given in the related papers. 3.1 MATERIALS Black liquor In papers I and II, industrial weak black liquor (pH >13, dry content ~17 %) from kraft pulping of softwood was generously supplied by the Billerud Gruvön pulp and paper mill, Sweden. Kraft lignin In paper III, LignoBoost lignin (lignin extracted from softwood kraft black liquor) was kindly supplied by LignoBoost Demo AB demonstration plant in Bäckhammar, Sweden. Kraft pulp In paper IV, oxygen delignified softwood kraft pulp with kappa number 10.0 was kindly supplied by SCA Östrand Pulp Mill in Sundsvall, Sweden. Softwood In paper V, sawdust of Norwegian spruce wood (from a Swedish sawmill) was used as material for pressurised hot water extraction at 163°C for 60 minutes. 16 Experimental 3.2 MEMBRANES AND FRACTIONATION Paper I: fractionation of weak kraft black liquor was performed with a ceramic membrane with 1 kDa cut-off (Novasep, France), to produce 1 kDa permeate to use for the main aim, i.e., extraction of low molecular lignin. Paper II: fractionation of weak kraft black liquor was performed with ceramic membranes of 1 and 5 kDa cut-off (Novasep, France) and 10 kDa regenerated cellulose (Merck Millipore, USA) membranes, for use in the following steps of precipitation to obtain lignins of different molecular weights. Paper III: cross-flow filtration was used to study fractionation of re-dissolved LignoBoost lignin. The system consists of a 10 L tank, a heating element, a gear pump, a flow meter and a Valisette membrane unit with ceramic membranes (TAMI Industries, Nyons, France). At a cross-flow velocity of 1m/s and a TMP of 3 bar, 5 kDa cut-off single channel membranes with a diameter of 0.6 cm were used. Paper IV: fractionation of nanocellulose was performed with ceramic membranes with 0.45 µm and 0.8 µm cut-off (Novasep, France), to obtain homogenous nanocellulose. Paper V: fractionation of hot water extract was performed with ceramic membranes with 1 and 5 kDa cut-off (Novasep, France) to obtain a size fractionated and purified sample for use in the next step of adsorption chromatography. 17 Results and Discussion 4. RESULTS AND DISCUSSION 4.1 PRODUCTION OF LIGNIN FRACTIONS (PAPERS I-III) In this study, cross-flow filtration (CFF) fractionation of lignin in weak black liquor, and fractionation of re-dissolved LignoBoost lignin for the production of lignin with low molecular-weight dispersity were investigated. Through CFF with several different membrane cut-offs, lignin fractions with specific properties could be produced from weak black liquor. In total, seven samples were obtained (including the unfractionated sample, i.e., reference): > 10 Da, > 5 kDa, 5-10 kDa, 1-5 kDa, 0-5 kDa, and 0-1 kDa. See Figure 3. Furthermore, CFF fractionation of re-dissolved LignoBoost lignin was seemingly affected by the different conditions chosen in the trials. Figure 3. Scheme of obtained CFF fractionations from weak black liquor. 18 Results and Discussion Precipitation and filtration of weak black liquor To extract solid material from the weak black liquor samples (see Figure 3), the unfractionated sample and all fractionated samples, except for the 0-1 kDa sample (see below), were precipitated by adding sulphuric acid (H2SO4 6 M) to pH 9 at ~70°C. The precipitated solids were separated from the liquid, using cake filtration (i.e., a filtration process were liquid passes through the membrane and the particles retained building up a filter cake). The filterability of the precipitated solids was improved by increasing the ionic strength, either by evaporating water or adding salt, which enhanced lignin coagulation. The electrical charge that is formed by ionised hydrophilic groups forms a colloidal lignin, the stability of which is affected by ionic strength (Passinen 1968). Furthermore, Rinman (1911) reported that the precipitation is improved by increasing salt concentration, effectively increasing the ionic strength of the solution. Precipitation and filtration of <1000 Da cut-off permeate The precipitation mentioned in the previous paragraph could be performed at ~70°C for all samples except for the low molecular weight lignin 0-1 kDa. During precipitation of the 0-1 kDa sample, fairly large, rigid particles/agglomerates of lignin material were formed at ~70°C, whereas lowering the pH at ambient temperature resulted in the formation of fine particles of precipitated lignin. This phenomenon was believed to be caused by a lower glass transition temperature (Tg). At temperatures above Tg, an amorphous polymer material becomes softer due to segment movements in the polymer chain. Yoshida et al. (1987) reported Tg temperatures of 32°C and 62°C for weight-average molecular weights lignin of 620 Da and 1300 Da, respectively. Researchers have reported that condensation reactions could take place when the chain mobility of lignin increases (Guigo et al. 2009, Cui et al. 2013). Similar agglomerate phenomena have also been 19 Results and Discussion reported by other research groups, but not for low molecular weight lignin. In those studies, higher temperatures were reported to melt the precipitate or form large tacky clumps (Rinman 1911, Uloth and Wearing 1989). It was suggested that precipitation should be performed just below the coagulation temperature (Merewether 1962). Lignin in black liquor fractions of 15-0 kDa and 15-1 kDa has also been reported to clump at higher temperatures, explained by higher density coagulation (Wallmo and Theliander 2009). Our predictions about lower Tg for the 1000 kDa in Paper I were supported by Tg measurements in Paper II, furthermore, this fraction contained a higher amount of aromatic hydroxyl groups (as shown and discussed later in this thesis). Composition of precipitated lignin Chemical analysis of the unfractionated kraft lignin (i.e., reference) and the fractionated lignin samples was performed by Klason lignin analysis (TAPPI 222 om-02), sugar analysis (SCAN-CM 71:09), acid soluble lignin analysis (according to Dence 1992), and ash content analysis (ISO 1762). For more details, see Papers I-II. From the chemical composition (Table 1), it can be seen that Klason lignin content is quite high for all samples (roughly ~95%), while acid soluble lignin content increases with decreasing molecular weight. In general, the lignin fractions are quite low in carbohydrate content, which is in agreement with previous study by Öhman et al. (2007). The largest amount of carbohydrates is found in higher molecular weight lignin. Because the 5-10 kDa fraction only contains 0.3% carbohydrates, it is concluded that most of the carbohydrates are found in the >10 kDa fraction. These carbohydrates might be in the form of lignin-carbohydrate complexes (LCCs). It has previously been reported that most 20 Results and Discussion of the lignin present in wood, especially in the pulp, exists in the form of LCCs with various types of hemicelluloses (Lawoko et al. 2005, Balakshin et al. 2011). Griggs et al. (1985) also suggested that some xylan was covalently bonded to kraft lignin. Finally, the ash content was quite low in all samples, indicating good washing of the samples. Table 1. Composition of reference and fractionated lignin samples (wt%). Sample (cut-off kDa) Ref. > 10 >5 5-10 1-5 0-5 0-1 Ref. (Paper I) 0-1 (Paper I) Klason lignin Acid-soluble lignin Carbohydrates Ash % 96.9 94.4 94.9 93.8 97.8 94.3 97.2 93.7 98.2 % 2.0 1.0 1.2 2.9 3.7 5.6 5.7 2.0 3.7 % 1.7 5.2 3.4 0.3 0.2 0.3 0.1 2.7 0.2 % 0.1 0.2 0.5 n/a 0.2 1.1 2.2 0.1 0.9 Elemental analysis for precipitated lignin Elemental analysis was performed to investigate the amount of sulphur in the unfractionated (i.e., reference) and the fractionated solid samples (external service provided by the Elemental Analysis Unit of the Santiago de Compostela University, Spain). For more details, see Papers I-II. From the elemental analysis (Table 2), the unfractionated kraft lignin sample (i.e., reference) contained approximately 2% sulphur, a mean of the two reference values in Table 2. Interestingly, the results show increasing sulphur content with decreasing molecular weight, despite the fact that the samples were thoroughly 21 Results and Discussion washed with acidic water, ~pH 2. Low molecular weight lignin seems to be the main material that contributes to the sulphur content of the lignin sample. Due to the high amount of sulphur in the lignin, extraction with organic solvents — toluene and pentane — was performed in an attempt to remove sulphur. A large amount of sulphur was extracted from the lignin samples, but some sulphur was retained, especially in the low molecular weight fraction. An explanation for this could be the reaction mechanism responsible for the lignin dissolution in kraft cooking, where sulphur is incorporated into lignin via thiirane (episulphide) formation (Gierer 1980), which might have remained covalently bound. Therefore, small lignin fragments in the samples could contain a higher ratio of organically bound sulphur. Other researchers have also reported higher sulphur content in low molecular weight lignin, but not the same high content. Wallmo and Theliander (2009) obtained fractions from ultrafiltration of 0-15 and 1-15 kDa black liquor, which contained 4% and 3.2% sulphur, respectively, while the original sample contained 2.9% sulphur. However, if all of the sulphur in the reference samples (Table 2) was bonded to only low molecular weight lignin, the sulphur value for fraction 0-1 kDa would still be too high for all of it being bound to lignin (based on calculations from the lignin yield in the reference and 0-1 kDa fraction, Table 3). Another explanation might be that low molecular weight lignin behaves differently during the washing and filtering stage, and hence, a lesser amount of sulphur is removed during the washing stage. However, it is not clear if this sulphur only comes from the cooking chemicals or is also incorporated into the lignin molecules during kraft cooking. The reason for the high sulphur content observed is not known at this time and needs to be investigated further. 22 Results and Discussion Table 2. Elemental analyses of reference and fractionated lignin samples (wt%). Sample (cut-off kDa) Before extraction (%) After extraction (%) S N C H S N C H Ref. 2.5 0.1 64.9 5.5 1.8 0.1 64.7 5.9 >10 1.7 0.1 62.3 5.1 1.6 0.1 62.1 5.9 >5 1.8 0.1 64.9 5.3 1.6 0.1 63.5 5.8 5-10 2.0 0.1 62.6 5.4 1.7 0.1 65.1 6.0 1-5 7.2 0.1 62.1 5.4 3.0 0.1 68.5 6.4 0-5 23.0 0.1 51.6 4.8 4.0 0.1 66.1 6.3 0-1 27.1 0.1 48.9 4.2 4.8 0.1 63.4 5.9 Ref.(Paper-I) 1.7 0.1 64.9 6.3 0-1 (Paper-I) 8.6 0.1 61.3 5.9 0-1* 25.1 0.1 49.6 4.8 *Extra sample precipitated with CO2 and washed with water/hydrochloric acid (pH ~2). Yield for <1000 Da cut-off precipitated lignin The amount of lignin obtained after precipitation and acidic washing was reported for the unfractionated (i.e., reference) and the 0-1 kDa solid samples (Paper I). To calculate lignin yield, the following analyses were performed on the precipitated lignin sample: Klason lignin (TAPPI 222 om-02), acid-soluble lignin (according to Dence 1992), and precipitated material in the acid-soluble filtrate. For more details, see Paper I. Table 3 outlines the amount of lignin precipitated at pH 9, the material in the filtrate precipitated at pH 2, and the sum of both. However, by extracting the sulphur content from the precipitated lignin values in Table 3, it can be determined that low molecular weight lignin precipitated to pH 9 constitutes approximately 30% of the sample (i.e., dividing 6.8 g/L and 21.3 g/L after subtracting the amount of suphur in each sample). It is well known that the amount of low molecular weight lignin depends on the cooking conditions and 23 Results and Discussion that prolonged cooking at the end of the final phase of degradation can increase the amount of low molecular weight lignin (Dong and Fricke 1995). Table 3. Lignin yield in reference and fractionated liquid samples. Sample (cut-off kDa) Ref. (Paper-I) 0-1 (Paper-I) Lignin precipitated to pH 9 (excl. filtrate) g/L 21.3 6.8 Material in filtrate from pH 9 precipitated to pH 2 (filtrate) g/L 5.8 6.0 Total (incl. filtrate) g/L ~27 ~13 Molecular weight data of precipitated lignin Size exclusion chromatography (SEC) was performed on the unfractionated (i.e., reference) and the fractionated solid samples to determine the effect of membrane separation for the fractionated samples. All samples were acetylated before analysis in the SEC-system with Tertrahydrofuran (THF) as eluent, and polystyrene standards were used for calibration. For more details see Papers I-II. As seen in Table 4, the weight-average (Mw) and number-average (Mn) molecular weights are in fairly good agreement with the membrane cut-offs. The samples in Paper I were acetylated according to Gellerstedt (1992) with ethanol, pyridine and acetic anhydride, while the samples in Paper II were acetylated by acetobromination according to Guerra et al. (2006), which could be an explanation for the differences between the results from Paper I and Paper II, but the trend still shows that more homogeneous materials with narrower molecular weight distributions are obtained with cross-flow fractionation. High-average molecular weight, Mz, for 0-1 kDa sample (Paper I) shows a large decrease in Mz compared to its reference (i.e., unfractionated sample). 24 Results and Discussion Table 4. SEC data for acetylated reference and fractionated lignin samples, including high-average (Mz), weight-average (Mw) and number-average (Mn) molecular weights as well as molecular-weight dispersity (ĐM). THF-SEC Sample (cut-off kDa) Mz Mw Mn ĐM Ref. - 20200 5000 4.1 >10 - 33500 9500 3.5 >5 - 28200 8000 3.5 5-10 - 4900 2300 2.2 1-5 - 4700 2000 2.3 0-5 - 4100 1700 2.4 0-1 - 2700 1200 2.1 Ref. (Paper-I) 13109 3525 847 4.2 0-1 (Paper-I) 2590 1096 508 2.2 Chemical characteristics and thermal properties of precipitated lignin Knowledge of the homogeneity in chemical characteristics and thermal properties of the unfractionated (i.e., reference) and the fractionated solid lignin samples could be an advantage for the development of high value application products. Therefore, aliphatic hydroxyl and aromatic hydroxyl groups were analysed by phosphorous-31 NMR, the glass transition temperature (Tg) was analysed by differential scanning calorimetry (DSC), and the decomposition temperature (Td) was analysed by thermogravimetric analysis (TGA) (see Table 5). For more details, see Paper II. Prior to phosphorous-31 NMR analysis each sample was functionalised by phosphitylation as described in Paper II. From the 31 P NMR data, it can be seen that phenolic content increases with decreased molecular weight and vice versa for aliphatic hydroxyl content, in agreement with previous studies (Wada et al. 25 Results and Discussion 1962, Zhang and Gellerstedt 2001). Due to the cleavage of β-O-4 linkages during kraft cooking, the low molecular weight lignin fractions will have a higher phenolic content. An advantage of higher phenolic hydroxyl content is that these groups have a relatively high reactivity (Weber and Weber 2010), which might be beneficial for future applications. For the use of lignin in thermal processes, it could be advantageous to use kraft lignin fractions with different glass transition temperatures and decomposition temperatures. To determine the thermal stability of kraft lignin samples, Td was determined by TGA, where the temperature of decomposition is defined as the temperature at which 5% weight loss has occurred (Figure 4). In Table 5, it can be seen that Td differs between the lignin fractions. In general, Td increases with molecular weight except for the 5-10 kDa sample, which showed the highest Td. An explanation for this might be the low molecular-weight dispersity (ĐM) for the 5-10 kDa sample (Table 4). In the DSC trials, Tg was determined from the second heat scan, while the first heating scan was performed to eliminate thermal history stored within the polymer. However, it has been reported that cross-linking occurs during the first heating scan, which could increase the observed Tg (Cui et al. 2013). The Tg for the unfractionated kraft lignin sample (i.e., reference) was 144°C, but after fractionation, Tgs between 70 and 170°C were obtained. In higher molecular weight lignin, the movement of the chain mobility of lignin is reduced, resulting in a higher Tg. The dependence of Tg on molecular weight was reported earlier for lignin samples obtained via solvent fractionation (Yoshida et al. 1990). Although a trend can be seen in Tg for fractionated lignin, the exact Tg values are uncertain because Tg is affected by low molecular weight contaminants and irreversible condensation reactions (Guido et al. 2009, Cui et al. 2013). 26 Results and Discussion Figure 4. Thermogravimetric analysis (TGA) of reference and fractionated lignin samples. Decomposition temperature was estimated at 5% of weight loss. Table 5. Functional groups, Tg and Td in the reference and fractionated lignin samples. 31 P NMR Sample (cut-off kDa) Tg by DSC Td by TGA Ph-OH mmol/g Aliph-OH mmol/g °C °C Ref. 3.5 2.2 144 215 > 10 2.8 2.6 170 217 >5 2.9 2.3 159 190 5-10 3.9 1.9 140 240 1-5 3.7 1.4 94 193 0-5 3.6 1.4 82 175 0-1 4.1 1.3 70 163 Dynamic mechanical analysis was used to investigate the viscoelastic response of dry lignin fractions using parallel-plate dynamic torsional rheometry. In Figure 5, the storage (G′) and loss (G″) moduli response is presented for four samples: 27 Results and Discussion unfractionated kraft lignin (reference), 0-5 kDa, 1-5 kDa, and >5 kDa. For the unfractionated (i.e., reference) and > 5 kDa samples, the degree of softening (i.e., the change in G′ from the glassy to the rubbery phase) results in a drop of one to two decades in G′, while the low molecular weight samples — 0-5 kDa and 1-5 kDa — show up to five decades of drop in G′. Differences in the molecular weight clearly affected the degree of softening, and 31P NMR data (Table 5) show that the higher molecular weight fractions had higher proportions of aliphatic-OH groups than the lower molecular weight fractions. Aliphatic-OH groups are more capable of intermolecular hydrogen bonding, resulting in greater mobility restriction (Dallmeyer et al. 2013, Uraki et al. 2012). Except for cross-linking that can occur during heating (Guido et al. 2009, Cui et al. 2013), higher molecular weight kraft lignin is more cross-linked (i.e., condensed) by nature, which might contribute to the lower degree of softening compared to lower molecular weight samples. When G″ > G′, the material is considered to be liquid-like, and for low molecular weight lignin (i.e., 0-5 kDa and 1-5 kDa), the rubbery plateau is reached. A clear rubber plateau is not seen for the high molecular weight samples (i.e., reference and >5 kDa), within chosen temperature interval. From Figure 5, it seems that the low molecular weight fractions (i.e., 0-5 kDa and 1-5 kDa, have better high-temperature flow-like behaviour compared to higher molecular weight fractions. Hence, the low molecular weight fractions could be useful for applications where flow is important, e.g., melt spinning fibres. In a recently published article 15 kDa cut-off permeate (i.e., <15 kDa) softwood kraft lignin produced a smooth carbon fibre (Nordström et al. 2013). Dynamic rheology could be an efficient tool to evaluate the behaviour of lignin samples during the various thermal processing reactions. For example, dynamic rheology could be used to evaluate the spinnability of a lignin sample and to predict the spinning 28 Results and Discussion temperature, as described elsewhere (Dr. Sudip Chowdhury and Prof. John Kadla personal communications, Chowdhury et al. 2012). Figure 5. Dynamic rheological response measured at the first heat responses for unfractionated (Ref.) and fractionated lignin samples. Cross-flow filtration of dissolved LignoBoost lignin As described in the introduction, the LignoBoost process is a commercial process with regard to lignin extraction from the kraft process. To facilitate the use of extracted kraft lignin for high value applications, it could be interesting for future LignoBoost lignin users to fractionate the lignin according to the desired molecular weight to enhance the properties of kraft lignin. Therefore, the aim of this work was to separate already extracted kraft lignin by CFF and to study the results of varying the following parameters: lignin concentration, pH of lignin solution, and ionic strength of the lignin solution. 29 Results and Discussion For fractionation, a ceramic membrane of 5 kDa was used with a transmembrane pressure (TMP) of 3 bar and a velocity of 1 m/s generated by a gear pump. Each dissolved LignoBoost lignin sample was run in the filtration system for ~24 hours with permeate recirculation to the feed tank. Dry contents of the permeate samples were determined according to SCAN-N 22:96. For more details, see Paper III. Table 6 summarises the seven different lignin solutions that were prepared for CFF trials. Three different concentrations, three different pH and two different ionic strengths were chosen as parameters. Because the sample with 20% lignin was dissolved with the highest sodium hydroxide concentration, this sample obtained the highest ionic strength (1.56). Therefore, all other samples were adjusted to an ionic strength of 1.56 with Na2SO4, with one exception: one sample with a lignin concentration of 10% and pH 12.5 was prepared at an even higher ionic strength by adding 30% extra Na2SO4 (marked as >I)*. Each lignin sample was dissolved for at least four days before CFF trials were performed. Table 6. Experimental conditions where X denotes the different lignin solutions prepared. pH >13.5 13.5 13.0 12.5 12.5 Lignin concentration (w/w %) 20% X - 15% X X - 10% X X X X >I* The effect of varying lignin concentration on permeate flux can be seen in Figure 6; higher lignin concentrations decrease the permeate flux. This trend could be explained by the phenomena of a polarisation layer, where a higher concentration 30 Results and Discussion lowers the permeate flux (Berg and Smolders 1990, Bowen and Jenner 1995). That the 20% and 15% samples are relatively close to each other and a bit far from the 10% sample might at least partly be related to the water-lignin ratio. These ratios are 9, 5.67 and 4 kg H2O/kg lignin, where an increase in the amount of water present might affect the polarisation layer, reducing the layer’s resistance. However, other effects cannot be excluded based on this study, and further studies could yield more understanding about this behaviour. Flux [L/h,m2] 15 10 10% lignin pH 13.5 (9 kg H2O/kg lignin) 5 15% lignin pH 13.5 (5.67 kg H2O/kg lignin) 0 20% lignin pH 13.5 (4 kg H2O/kg lignin) 0 10 20 30 40 Time [h] Figure 6. Effect of lignin concentration for lignin dissolved at pH 13.5. Figure 7 presents the effects of different pH on permeate flux, indicating that higher permeate fluxes are achieved at higher pH. An explanation might be that lignin becomes more deprotonated at higher pH, resulting in a greater charge on the lignin molecules. Through a higher repulsion between lignin molecules, the polarisation layer might become more permeable. It has been reported that the pKa value of coniferyl alcohol at room temperature is approximately 10.2 but increases with increasing molecular weight of the lignin molecule (Norgren and Lindström 2000). 31 Results and Discussion Flux [L/h,m2] 15 10% lignin pH 13.5 10 10% lignin pH 13.0 5 10% lignin pH 12.5 15% lignin pH > 13.5 0 0 10 20 30 40 15% lignin pH 13.5 Time [h] Figure 7. Effect of pH for lignin dissolved at 10 and 15% concentrations. Figure 8 presents the effect of increased ionic strength on permeate flux, showing that higher permeate fluxes are achieved at higher ionic strengths. Norgren and Lindström (2000) have shown that an increase in ionic strength lowers the apparent pKa of lignin. This could be an explanation for the increased flux at a higher ionic strength, because lignin with a lower pKa would be more charged. Furthermore, as in the case of the pH effect, this increased charge could generate higher repulsions between molecules and hence reduce the resistance of the polarisation layer. This suggests that pKa had a larger effect on the results than shielding of charges did, which is the other effect of increasing ionic strength. This result might be explained by the Debye length, which shows smaller changes at higher ionic strengths. 32 Results and Discussion Flux [L/h,m2] 15 10 10% lignin pH 12.5, > I 5 10% lignin pH 12.5 0 0 10 20 30 40 Time [h] Figure 8. Effect of ionic strength for lignin dissolved at pH 12.5. Table 7 and Figure 9 shows SEC data for the permeate sample, which has been precipitated, washed and acetylated for THF-SEC analysis. Unfortunately, the sample at 20% lignin concentration most likely condensed during the precipitation trials and was therefore not included. There are no large differences in the SEC data or clear trends between the different samples, indicating that separation of lignin molecules is performed by the membrane. Published molecular weight data (either Mw or Mn) on acetylated LignoBoost lignin report molecular weights of 5500 Da (Moosavifar et al. 2006). This molecular weight data for the starting material indicate that fractionation with the 5 kDa membrane was successful, as the permeate fractions have lower molecular weight values. Table 7. THF-SEC data for permeate samples, weight-average (Mw), number-average (Mn), and molecular-weight dispersity (ĐM). Permeate-Sample 15% pH >13.5 15% pH 13.5 10% pH 13.5 10% pH 13.0 10% pH 12.5 10% pH 12.5, > I Mw 1600 1700 1700 1600 1700 1400 Mn 1000 900 1000 1000 1000 900 33 ĐM 1.54 1.83 1.68 1.67 1.80 1.60 Results and Discussion Figure 9. THF-SEC analysis for permeate samples with UV detection at 280 nm. Table 8 shows rejection coefficients for all samples after CFF. The rejection coefficient, R = 1-(cp/cf), is calculated from the concentration in the permeate (cp) and the concentration in the feed (cf). All trials, except for the experiment with 20% lignin content, have similar rejection coefficients. The fact that the experiments with lignin contents of 10% and 15% have similar rejection coefficient and very similar molecular size distributions in the permeates suggests that it is the properties of the membrane itself that serve as the separating agent for lignin fractionation. The reason why the 20% experiment has a lower rejection coefficient is not fully understood. It should, however, be remembered that the rejection coefficients are based on concentrations and that the permeate flux of the 20% experiment was very low. 34 Results and Discussion Table 8. Rejection coefficients for permeate samples. Permeate Sample ~0 h Rejection coefficient ~9 h ~22 h (%) 39 35 20% pH 13.5 17 15% pH > 13.5 53 54 40 15% pH 13.5 54 58 44 10% pH 13.5 48 46 32 10% pH 13.0 55 49 46 10% pH 12.5 55 58 50 10% pH 12.5, > I 56 54 47 As shown in Table 8, there is a difference in the rejection coefficients of the 20% lignin and samples with lower lignin concentrations, indicating that more lignin, relative to the feed concentration, is passed through the membrane for 20% lignin. Additionally, it was shown that the permeate flux is different for different samples. By combining the rejection coefficient and permeate flux, the actual amount of low molecular weight lignin being produced in ~kg lignin/h,m2 can be calculated (Figure 10 and Figure 11). Figure 10 shows the mass flux of lignin in the permeate. However, this value does not account for feed concentration, and therefore, the actual relative production is unknown. For the actual value for low molecular weight production, the rejection coefficient needs to be included in the calculations. Figure 11 shows the production in the permeate flux relative to the rejection coefficient by multiplying the value in Figure 10 with 1-the rejection coefficient (i.e., cp/cf; where cp is the permeate concentration and cf is the feed concentration). By comparing Figure 10 and Figure 11 it is evident that the calculated production of lignin in the permeate flux changes when the rejection coefficient is included in the calculation. The 35 Results and Discussion smallest change is observed for the sample with the high lignin concentration of 20%, which is the sample with the lowest rejection coefficient (Table 8). For the other samples, there is a larger change when the rejection coefficient is included in the calculations, which reduces the calculated amount of fractionated lignin produced. This result demonstrates that, when optimising the cross-flow separation, it is important to not only investigate the permeate flux but also to investigate the rejection coefficient before concluding appropriate system parameters. In this study, it is shown that production of large amounts of low molecular weight kraft lignin can be accomplished from solutions with a lower lignin concentration (i.e., 10%). Furthermore, higher pH and ionic strength is beneficial to achieve a higher lignin production of low molecular weight lignin. To perform cost estimation other parameters also needs to be considered. These kinds of production costs have been reported for membrane filtration of black cp*(cp/cf) * Flux [~kg lignin/day,m2] liquor (Wallberg et al. 2005, Jönsson and Wallberg 2007). cp * Flux [~kg lignin/day,m2] 20 15 10 5 0 0 Time [h] 20 10. Amount of lignin in the Figure 11. permeates. Based on permeate concentration, cp in wt %. 20 20% pH 13.5 15% pH > 13.5 15 15% pH 13.5 10% pH 13.5 10 10% pH 13.0 10% pH 12.5 5 10% pH 12.5, > I 0 0 Time [h] 20 Figure 11. 10. Amount of lignin in the permeates. Based on cp (wt%) and 1-R. 36 Results and Discussion 4.2 FRACTIONATION OF NANOCELLULOSE (PAPER IV) In the search for new energy efficient methods to produce homogenous nanocellulose, a novel process called nanopulping has been developed and combined with fractionation. In nanopulping cellulose nanoparticles/fibres are manufactured by a disintegration method integrated with a pulping-type process (i.e., pressurised reactor with a valve). In this study, kraft pulp was used as a raw material and pulp pretreatments were performed before nanopulping (i.e. enzymatic pretreatment and TEMPO oxidation). The product from the nanopulping step was then fractionated with CFF to obtain more homogenous nanocellulose. Energy demand for nanopulping The energy demand of the nanopulping process has been estimated, and is ~450 kWh/tonne (~1.6 GJ/tonne ). This energy demand is much lower than that reported for nanocellulose production, which can be between 20000 and 30000 kWh/tonne (Siró and Plackett 2010), and can also be compared to that reported by Ankerfors (2012), who reported an energy demand of 500-2300 kWh/tonne for nanocellulose production. The energy calculation is based on the experimental conditions for nanopulp production (i.e., 1 g dry pulp/L water and 160°C) combined with how the process might be implemented in industry. A pulp consistency of 1 weight percent, which is heated from 90-160°C from the pulp mill’s system, corresponds to an energy demand of ~760 kWh/tonne (~2.7 GJ/tonne). However, by using the existing environment, it is possible to recover energy by flashing from 160-130°C, resulting in a total energy demand for the nanopulping process of ~450 kWh/tonne (~1.6 GJ/tonne). 37 Results and Discussion Effect of different pretreatment methods As a starting raw material, oxygen delignified softwood kraft pulp with original kappa number 10.0 was used in the nanopulping process. Different pretreatments of the kraft pulp were performed before the nanopulping, enzyme pretreatment as well as TEMPO oxidation. After nanopulping, the obtained samples were size fractionated with CFF, resulting in permeate fractions with different concentrations. Table 9. Results for non-pretreated kraft pulp (Reference), endoglucanase and TEMPO pretreated samples: after nanopulping (NP) and after filtration (i.e., in the permeate, Perm.). Endoglucanase Pretreatment Reference Sample TEMPO Pretreatment 1 Conc. after NP (g/l) 1.01 Conc. of Perm. (g/l) 0.22 Conc. after NP (g/l) 1.02 Conc. of Perm. (g/l) 0.26 Conc. after NP (g/l) 1.08 Conc. of Perm. (g/l) 0.14 2 1.02 0.29 0.99 0.28 1.03 0.17 3 1.03 0.24 0.98 0.27 1.02 0.17 Average 1.02 0.23 1.00 0.27 1.05 0.16 STDEV 0.008 0.02 0.02 0.01 0.03 0.02 Reject.coeff. (%) 78 73 85 In Table 9, results from endoglucanase and TEMPO pretreated samples are presented. Enzymatic pretreatment with monocomponent endoglucanase decreased the rejection coefficient from 78% to 73%, which indicates an increase of nanopulped material in the permeate. Engström et al. (2006) reported that monocomponent endoglucanase can be used to increase the reactivity and accessibility of dissolving pulp. Furthermore, it was demonstrated that pulp pretreatment with monocomponent endoglucanase enhance the mechanical 38 Results and Discussion separation of nanofibrils (Henriksson et al. 2007). These effects from endoglucanase could be a reason for the improved efficiency of nanopulping, indicated by the higher content of material in the permeate (Table 9). The high rejection coefficient for the TEMPO pretreated sample is, however, not in agreement with previous results reported by Tsuguyuki et al. (2009), where TEMPO-oxidised pulp treated in a homogeniser showed a large increase in fibrillation. Fractionation by cross-flow filtration After nanopulping, the nanopulped material was fractionated via tubular CFF with ceramic membranes of two different cut-offs: 0.8 µm and 0.45 µm. The nanocellulose in the permeate was thereafter characterised with AFM, where the sample image’s scan size is 5 μm x 5 μm. Before filtration: a) Nanopulped sample After filtration: b) Higher cut-off c) Lower cut-off Figure 12. AFM images of material after nanopulping (a) and after filtration (b and c), scan size 5 μm x 5 μm. In Figure 12, image a), it can be observed that the nanopulped material, contains both long and short nanocellulose with dimensions of approximately 10-20 nm wide and ~0.5 μm long. In images b) and c), the nanocellulose after CFF (i.e., in 39 Results and Discussion the permeate) is presented. It is apparent that nanocellulose can be separated into different fractions with more homogeneous size distribution by CFF. 4.3 FRACTIONATION OF HOT WATER EXTRACT FROM SOFTWOOD (PAPER V) The purpose of this investigation was to study the combination of two downstream separation methods, CFF and chromatography, for fractionation and purification of wood components. Hot water extract of softwood sawdust produced in the lab was used as a model liquid to simulate galactoglucomannan rich process liquids, such as thermomechanical pulping liquid. The target of the separation was to obtain three fractions: galactoglucomannan (GGM), lignincarbohydrate complexes (LCCs), and lignin. The isolation method consisted of CFF, where high and low molecular species were removed, followed by two-step adsorption chromatography to separate the wood species by affinity. Yield for membrane fractionated hot water extract Pressurised hot water extraction of softwood sawdust was performed at 163°C for 60 minutes. The total yield of extracted material, after removing sawdust and fibres, in the filtrate was ~14.4%, based on the initial amount of saw dust. This extract was thereafter fractionated by CFF with 5 kDa and 1 kDa membrane cutoffs, to obtain a purified fraction between 1 and 5 kDa The amount of dissolved material in the 1-5 kDa fraction was 34% (see Table 10). The amount of extracted material in the 1-5 kDa sample was almost three times as much as that in the >5 kDa sample. This result was in accordance with an experimental condition chosen from Song et al. (2008) to produce hot water extracted material within this 40 Results and Discussion molecular weight region. Hence, the 1-5 kDa sample was used for the next step of separation by chromatography. Table 10. Material yield in each filtered fraction (based on dissolved material 14.4%). Samples Yield in CFF fractions wt % Cross-flow filtration: >5 kDa 15 1-5 kDa 34 0-1 kDa 43 Composition of membrane fractionated hot water extract The chemical composition of the unfractionated hot water extract (i.e., reference) and CFF fractionated samples was determined by Klason lignin analysis (TAPPI 222 om-02), sugar analysis (SCAN-CM 71:09) and acid soluble lignin (according to Dence 1992) on a freeze dried aliquot of each liquid sample. The data in Figure 13 show that the >5 kDa and the 1-5 kDa fractions contain greater amounts of carbohydrates and smaller amounts of “other” material compared to the unfractionated reference sample and 0-1 kDa permeate. Klason lignin content was highest for the 0-1 kDa sample. This result might be explained by the low molecular weights of the small amounts of lignin that are extracted. This result is consistent with a previous report by Örså et al. (1997), wherein most of the aromatic species in the hot water extract states to be of low molecular weight. 41 Results and Discussion 100.0 80.0 60.0 40.0 20.0 0.0 Reference Carbohydrates >5 kDa 1-5 kDa Klason lignin 0-1 kDa Acid soluble lignin Other Figure 13. Chemical composition before and after CFF (wt%). Acid soluble lignin was present in very low amounts and is therefore not notable in this table. From carbohydrate analysis in Figure 14 it is obvious that GGM and arabinoglucuronoxylan were extracted and constituted a major part of the extracts. The analysis suggests that this GGM is of higher molecular weight because it is mainly found in the >5 kDa and 1-5 kDa fractions, while arabinose, mainly the side chain of xylan, is cleaved off xylan and enriched in the 0-1 kDa fraction. Reference >5 kDa 1-5 kDa 0-1 kDa Ara % Gal % Glc % Xyl % Man % Other Klasonlignin Figure 14. Distribution of carbohydrates before and after CFF (wt%). 42 Results and Discussion Molecular weight distribution for fractionated hot water extract To analyse the molecular weight distributions of the initial hot water extract and CFF fractionated samples an aliquot of freeze dried samples was dissolved in 0.1 M NaOH and analysed with a water (alkali) SEC system and pullulan standards. The molecular weight distribution curves are presented in Figure 15 and Figure 16. Figure 15 details the detection of all components using a refractive index (RI) detector. Most of the RI signal will reflect the concentration of carbohydrates, because the amount of lignin is fairly small in the samples (Figure 13). The graph shows two effects. First, as expected, the effect of CFF is shown by a shift towards longer retention times for samples with lower molecular weight. Second, for the 0-1 kDa sample, there is a higher signal at retention times of 21-24 minutes, which corresponds to a molecular weight of 2 kDa and lower, whereas the >5 kDa and 1-5 kDa samples show a decrease in RI detection signal for the same retention interval. For hot water extracts performed in similar conditions Song et al. (2008) reported weight-average molecular weight of ~5 kDa, and for TMP process liquids Persson et al. (2010) reported a weight-average molecular weight of ~10 kDa. Figure 16 details the absorbance spectra using a UV detector at a wavelength of 280 nm, which corresponds to aromatic species (i.e., lignin). From the Klason lignin analyses in Figure 13, there is a small amount of lignin in all of the samples, but in the 0-1 kDa sample, the amount of lignin is somewhat increased. When analysing the data from the pullulan standard, the UV peaks had an approximately 0.2 minute shorter retention time than the peaks from the RI detector. Comparing the peaks from the UV and RI graphs, some peaks are overlapping, which might suggest the presence of LCCs. 43 Results and Discussion Figure 15. Water (alkali) SEC analysis with refractive index (RI) detector. Figure 16. Water (alkali) SEC analysis with UV detector at 280 nm. Separation by chromatography of fractionated hot water extract From the first separation by CFF, the 1-5 kDa liquid fraction with a relative high yield and purity was selected as a sample for the second separation by chromatography. Adsorption chromatography was performed in two steps. First, the 1-5 kDa hot water extract was pumped into a hydrophobic sorption column, resulting in two samples/streams: permeate and material adsorbed on the column. Second, the permeate from the first step was pumped into a second column with a higher 44 Results and Discussion selectivity for aromatic structures, resulting in another two samples/streams: permeate and material absorbed on the column. The first hydrophobic column aimed to adsorb relatively pure lignin moieties (here named column 1), while the second column aimed to separate LCCs from GGM due to its very high selectivity for aromatic compounds (here named column 2). Table 11 shows that the isolated LCC fraction contained ~10% aromatics, the upgraded GGM fraction ~1.5% aromatics, and the lignin fraction ~56% aromatics. Based on the detected sugars, the calculated amount of GGM in the upgraded GGM fraction was 79%. The calculated degree of detection, that is the detectable sugars and lignin, was 80-85% in all fractions, and the shortfall was most likely due to a combination of measuring errors. Excluding losses in handling, the obtained products from CFF hot water extract are as follows, based on internal relative values: lignin fraction (adsorbed to column 1) 7%, LCC fraction (adsorbed to column 2) 5%, and GGM (permeate from column 2) 88%. Table 11. Yield and amount of aromatics and galactoglucomannan (GGM). Samples Dissolved starting material Cross-flow filtration: 1-5 kDa Chromatography: Adsorbed to column 1 Adsorbed to column 2 Permeate from column 2 45 Yield based on initial amount of saw dust Aromatics mg/g 144 wt% - 49 5.5 1.8 1.4 23.7 55.7 10.2 1.5 Concluding remarks 5. CONCLUDING REMARKS Process streams from the primary separation of wood components are obtained as complex mixtures of components, which need to be separated and purified. Raw materials with higher homogeneity can be obtained by cross-flow filtration (CFF). For fractionation of kraft lignin, three topics have been examined: characteristics of lignin fractions, extraction of low molecular weight lignin, and the behaviour of re-dissolved lignin during CFF. By filtering weak black liquor with membranes of different cut-offs, more homogenous fractions of lignin could be obtained regarding size, characteristics, and thermal properties. Through fractionation, it is possible to tailor lignin’s molecular weight and consequently the amount of carbohydrates, amount of phenolic hydroxyl groups, and thermal properties. Through dynamic mechanical analysis (DMA), the degree of softening and flow can be monitored, and this analysis might be a useful tool for predicting visco-elastic behavior during thermal processing. To improve the extraction of low molecular weight lignin (<1000 Da cut-off) from black liquor, precipitation needs to be performed at lower temperatures than with other fractions due to the formation of rigid, fairly large lignin 46 Concluding remarks particles/agglomerates. However, lowering the pH at ambient temperature, results in the formation of fine particles of precipitated lignin. As shown by differential scanning calorimetry (DSC), the Tg of low molecular weight lignin could be as low as 70°C or lower if irreversible cross-linking occurred during the first heat scan. This could be an explanation for the low molecular weight behaviour during precipitation at higher temperatures. Due to the recent introduction of the LignoBoost process, which produces kraft lignin of relatively high purity, there might be an interest to fractionate lignin extracted by the LignoBoost process to obtain lignin fractions with more specific properties for the production of high value application products. In this investigation, the possibilities for improved production of low molecular weight lignin were investigated. These results show (during the conditions of this set-up) that lower concentrations of lignin in the feed result in a higher yield of low molecular lignin in the permeate, relative to the feed concentration, and that permeate flux could be increased by a higher pH and ionic strength. For kraft pulp, a novel method for preparing nanocellulose was developed called nanopulping, which appears to be energy efficient. The method is based on a combination of the acceleration of chemical pulp fibres and steam explosion, where the rapid pressure drop causes acceleration. The efficiency of the method might be increased if the pulp is pretreated. To produce a more homogeneous size distribution CFF, was used. Hot water extract was used to demonstrate that a biorefinery process that combines the steric separation of filtration with the selective separation of adsorption chromatography can be used to isolate dissolved wood components in 47 Concluding remarks upgraded fractionations, a process in which several products are obtained, including polymeric lignin, LCCs and GGM. The final separation of wood components using CFF has been successfully used in this thesis to separate lignin fractions according to molar mass or particle size. Furthermore, there is a possibility to combine CFF and adsorption chromatography to obtain hemicellulose fractions rich in lignin or almost completely free of lignin. 48 Notations 6. NOTATIONS CFF cf cp Da ÐM DMA DSC GGM LCC MFC Mw Mn NMR NP ΔP Q R SEC Td Tg TEMPO TGA TMP UV v Vif Vp Vr VR VRF wt% Cross flow filtration Concentration in feed Concentration in permeate Dalton Molecular-weight dispersity Dynamic mechanical analysis Differential scanning calorimetry Galactoglucomannan Lignin carbohydrate complex Microfibrillated cellulose Weight-average molecular weight Number-average molecular weight Nuclear magnetic resonance Nanopulping Transmembrane pressure (symbol) Feed flow Rejection coefficient Size exclusion chromatography Decomposition temperature Glas transition temperature 2,2,6,6-tetramethylpiperidine-1-oxyl Thermogravimetric analysis Transmembrane pressure (abbreviation) Ultraviolet Cross-flow velocity Initial feed volume Permeate volume Retentate volume Volume reduction Volume reduction factor Weight percent 49 Acknowledgements 7. ACKNOWLEDGEMENTS First of all, I want to express my deepest gratitude to Prof. Mikael E. Lindström for accepting me as your PhD student. I have learned a lot, and it has been a wonderful time. Thank you for your scientific support and guidance. You are always so positive, kind and encouraging, and that has been a great help along the way and meant a lot to me. To all my co-supervisors. I am very happy to have had so many carrying and kind cosupervisors who have supported me. Prof. Hans Theliander, I am sincerely grateful for your ambitious supervision and great thoughtfulness. Prof. Gunnar Henriksson, I am very grateful for your enthusiastic supervision and generous support. Ass. Prof. Martin Lawoko, not only my supervisor but also a friend; thank you so much for the great support and all the nice times running together. Dr. Liming Zhang, thank you for always having the door open for nice discussions. To Assoc. Prof. Ulrica Edlund, thank you for taking your time and the valuable inputs for the quality check. The Wallenberg Wood Science Center is gratefully acknowledged for financial support. To all colleges and friends at FPT, you have been many during the years, thank you for a great time, helpfulness and kindness. Sara and Anna, you have been my “rocks” and “lights” along the way. Emma and Nicholas, thank you for being so dear and for all lunchlaughs. Linn and Michaela thank you for other memorable activates outside work. Shoaib, Dimitri, Ramiro, Petri, and, Kasinee thank you for always being so helpful and caring. To all my lovely colleagues, collaborates and friends during the years: Olena, Helena, Göran, Susanne, Louise, Andreas, Farhan, Daniel, Dina, Yan, Ran, Emil, Malin, Jacob, Gisela, Oruc, Lars W, Carl, Adam, Felix, Elisabeth, Monica, Lars Ö, Lars B, Caroline, Anders, Stefan, Joby, Ngesa, Torbjörn, Per, Pontus, Rebecca, Linda, Jiebing, Dongfang, Yadong, Yujia, Raquel and Myriam. Jan-Erik, Mia, Mona, Therese, Ewa, Inga, Kjell, Karin, Bosse and Fredrick thank you for all your help; you are valuable. To my dear colleagues in Gothenburg, you are all amazing: Tuve, Niklas, Susanne, Kerstin and Hampus. Till mina väninnor, tack för att ni är så fina, förstående och härliga. Till min familj. Jag är så tacksam och lycklig över att ha en sådan kärleksfull och hjälpsam familj med underbara föräldrar, en fantastisk syster med familj, mina fina morföräldrar och aunts & unlces med familjer, farföräldar och gammelmormor i hjärtat, och pojkvän av 24k guld <3 Ni betyder allt för mig. 50 References 8.REFERENCES Abe, K., Iwamoto, S., and Yano, H., (2007), Obtaining cellulose nanofibers with a uniform width of 15 nm from wood. Biomacromolecules, 8, 3276-3278. Alén, R., Patja, P., and Sjöström, E., (1979), Carbon dioxide precipitation of lignin from pine kraft black liquor, Tappi Journal, 62, 108-110. Alén, R., Sjöström, E., and Vaskikari, P., (1985), Carbon dioxide precipitation of lignin from alkaline pulping liquors, Cellulose Chemistry and Technology 19, 537-541. Alén, R., Sjöström, E., and Vaskikari, P., (1986), Ultrafiltration studies on alkaline pulping liquors, Cellulose Chemistry and Technology, 20, 417-420. Andersson, A., Persson, T., Zacchi, G., Stalbrand, H., and Jönsson, A.-S., (2007), Comparison of diafiltration and size-exclusion chromatography to recover hemicelluloses from process water from thermomechanical pulping of spruce, Applied Biochemistry and Biotechnology, 137, 971-983. Ankerfors, M., (2012), Microfibrillated cellulose: energy-efficient preparation techniques, Licentiate Thesis, Royal Institute of Technology, KTH, Stockholm, Sweden. Antonsson, S., Henriksson, G., Johansson, M., and Lindström, M. E., (2008), Low Mwlignin fractions together with vegetable oils as available oligomers for novel paper-coating applications as hydrophobic barrier, Industrial Crops and Products, 27, 98-103. Balakshin, M., Capenema, E., Gracz, H., Chang, H. M., and Jameel, H., (2011), Quantification of lignin-carbohydrate linkages with high-resolution NMR spectroscopy Planta, 233, 1097-1110. Berg, G. B., and Smolders, C. A., (1990), Flux decline in ultrafiltration processes, Desalination, 77, 101-133. Bowen, W. R., and Jenner, F., (1995), Theoretical descriptions of membrane filtration of colloids and fine particles: an assessment and review, Advances in Colloid and Interface Science, 56, 141-200. 51 References Brasch, D. J., and Free, K. W., (1965), Prehydrolysis-kraft pulping of pinus radiata grown in New Zealand, Tappi, 48, 245-248. Braun, J. L., Holtman, K. M., and Kadla, J. F., (2005), Lignin-based carbon fibers: Oxidative thermostabilization of kraft lignin, Carbon, 43, 385-394. Brodin, I., Sjöholm, E., and Gellerstedt, G., (2009), Kraft lignin as feedstock for chemical products: The effects of membrane filtration, Holzforschung 63, 290-297. Chakraborty, A., Sain, M., and Kortschot, M., (2005), Cellulose microfibrils: A novel method of preparation using high shear refining and cryocrushing, Holzforschung, 59, 102107. Chen, W., Yu, H., Lui, Y., Chen, P., Zhang, M., and Hai, Y., (2011), Individualization of cellulose nanofibres from wood using high-intensity ultrasonication combined with chemical pretreatments, Carbohydrate Polymers, 83, 1804-1811. Chowdhury, S., Gao, G., and Kadla, J.F., (2012), Investigating the viscoelasticity of ligninpolyacrylonitrile blends using dynamic rheology, Proceedings, 243rd ACS meeting, March 2529, San Diego, California, USA, Abstract: CELL 119. Cui, C., Sadeghifar, H., Sen, S., and Argyropoulos, D.S., (2013), Toward thermoplastic lignin polymers; Part II, BioResources, 8, 864-886. Curtolo, F., and Eksteen, N. C., (2006), Method and apparatus for manufacturing microfibrillated cellulose fiber, US Patent 0006189. Dallmeyer I, Chowdhury S, and Kadla J.F., (2013), Preparation and characterization of kraft lignin-based moisture-responsive films with reversible shape-change capability, Biomacromolecules, 14, 2354-2363. Dence, C. W., (1992), The determination of lignin, Methods in Lignin Chemistry, S. Y. Lin and C. Dence (Eds.), Springer-Verlag, Heidelberg, Germany, 33-61. Dong, D., and Fricke, A. L., (1995), Effects of multiple pulping variables on the molecular weight and molecular weight distribution of kraft lignin, Journal of Wood Chemistry and Technology 15, 369-393. Drouin, M. P., and Desrochers, M. J., (1988), Isolation of lignin from spent kraft liquor by hyper and ultrafiltration, Chemical engineering technology in forest products processing, AIChE forest products division, 2, 58-66. Eichhorn, S. J., Dufresne, A., Aranguren, M., Marcovich, N. E., Capadona, J. R, Rowan, S. J., Weder, C., Thielemans, W., Roman, M., Reneckar, S., Gindl, W., Veigel, S., Keckes, J., Yano, H., Abe, K., Nogi, M., Nakagaito, A. N., Mangalam, A., Simonsen, J., Benight, A. S., Bismarck, A., Berglund, L. A., and Peijs, T., (2010), Review: current international research in cellulose nanofibres and nanocomposites, Journal of Material Science, 45, 1-33. 52 References Elegir, G., Bussini, D., Antonsson, S., Lindström, M. E., and Zoia, L., (2007), Laccaseinitiated cross-linking of lignocellulose fibres using a ultra-filtered lignin isolated from kraft black liquor, Applied Microbiology Biotechnology 77, 809-817. Engström, A.-C., Ek, M., and Henriksson G., (2006), Improved accessibility and reactivity of dissolving pulp for the viscose process – pre-treatment with monocomponent endoglucanase, Biomacromolecules, 7, 2027-2031. Forss, K., and Fuhrmann, A., (1976), Karatex: The lignin-based adhesive for plywood, particle board and fibre board, Paperi ja Puu, 58, 817-824. Gellerstedt, G., (1992), Gel permeation chromatography, Methods in Lignin Chemistry, S. Y. Lin and C. Dence (Eds.), Springer-Verlag, Heidelberg, Germany, 487-497. Gierer, J., (1980), Chemical aspects of kraft pulping, Wood Science Technology, 14, 241-266. Grace, T. M., (1989), Black liquor evaporation, Pulp and paper manufacture - Alkaline pulping, Joint textbook committee of the Paper Industry, 477-530. Granata, A., and Argyropolous, D.S.J., (1995), 1, 2-Chloro-4,4,5,5-tetramethyl-1,3,2dioxaphospholane, a reagent for the accurate determination of the uncondensed and condensed phenolic moieties in lignins, Journal of Agricultural and Food Chemistry, 43, 1538-1544. Griggs, B. F., Gratzl, J. S., Chen, C.-L., and Venica, A., (1985), Chemical characterization of kraft lignin and kraft lignin products, International Symposium on Wood and Pulping Chemistry, August 26-30, Vancouver, Canada, 5-6. Guerra, A., Filpponen, I., Lucia, L. A., and Argyropolous, D.S.J., (2006), Comparative evaluation of three lignin isolation protocols for various wood species, Journal of Agricultural and Food Chemistry, 54, 9696-9705. Guido, N., Mija, A., Vincent, L., and Sbirrazzuoli, N., (2009), Molecular mobility and relaxation process of isolated lignin studied by multifrequency calorimetric experiments, Physical Chemistry Chemical Physics, 11, 1227-1236. Hartman, J., (2006), Hemicellulose as a barrier material, Licentiate Thesis, Royal Institute of Technology, KTH, Stockholm, Sweden. Hartman, J., Albertsson, A. C., Söderqvist Lindblad, M., and Sjöberg, J., (2006), Oxygen barrier materials from renewable sources: Material properties of softwood hemicellulose-based films, Journal of Applied Polymer Science, 100, 2985-2991. Henriksson, M., Henriksson, G., Berglund, L. A., and Lindström, T., (2007), An environmentally friendly method for enzyme-assisted preparation of microfibrillated cellulose (MFC) nanofibers, European Polymer Journal, 43, 3434-3441. 53 References Herrick, F. W., Casedier, R. L., Hamilton, J. K., and Sandberg, K. R., (1983), Microfibrillated cellulose: morphology and accessibility. Journal of Applied Polymer Science: Applied Polymer Symposium, 37, 797-813. Hough, W. J., (1910), Process of treating spent alkaline pulping liquors, United States Patent 949,324. Ishikawa, H., Ide, S., (1993), Method of producing finely divided fibrous cellulose particles. US Patent 5,269,470. Iwamoto, S., Nakagaito, A. N., Yano, H., and Nogi, M., (2005), Optically transparent composites reinforced with plant fiber-based nanofibers, Applied Physics A, 81, 1109-112. Jönsson, A.-S., Wallberg, O., (2009), Cost estimates of kraft lignin recovery by ultrafiltration, Desalination, 237, 254-267. Kadla, J.F., Kubo, S., Venditti, R.A., Gilbert, R.D., Compere, A.L., and Griffith, W. (2002), Lignin-based carbon fibers for composite fiber applications, Carbon, 40, 2913-2920. Keyoumu, A., Sjödahl, R., Henriksson, G., Ek, M., Gellerstedt, G., and Lindström, M. E., (2004), Continuous nano- and ultra-filtration of kraft pulping black liquor with ceramic filters, A method for lowering the load on the recovery boiler while generating valuable side-products, Industrial Crops and Products, 20, 143-150. Klemm, D., Kramer, F., Moritz, S., Lindström, T., Ankerfors, M., Gray, D., and Dorris, A., (2011), Nanocelluloses: A new family of nature-based materials, AngewanteChemie, 50, 5438-5466. Kubo, S., and Kadla, J. F., (2005), Kraft Lignin/poly(ethylene oxide) blends: Effect of lignin structure on miscibility and hydrogen bonding, Applied Polymer Science, 98, 1437-1444. Lake, M. A., and Blackburn, J. C., (2011), Process for recovering lignin, US Patent Application US20110294991 A1. Lawoko, M., Henriksson, G., and Gellerstedt, G., (2005), Structural differences between the lignin-carbohydrate complexes present in wood and in chemical pulps, Biomacromolecules, 6, 3467-3473. Lin, S. Y., (1992), Commercial spent pulping liquors, Methods in Lignin Chemistry, S. Y. Lin and C. Dence (Eds.), Springer-Verlag, Heidelberg, Germany, 75-80. Lin, S. Y., and Detroit, W., (1981), In chemical heterogeneity of technical lignins: Its significance in lignin utilization, Proceedings of the 1st International Symposium on Wood and Pulping Chemistry, Stockholm, Sweden, June 9-12, 1981, 44-52, SPCI (Svenska Cellulosaoch Pappers Ingenjörsföreningen), Stockholm. Lindberg, J. J., Tylli, H., and Majani, C., (1964), Notes on the molecular weight and the fraction of lignins with organic solvents, Paperi ja Puu, 46, 521-526. 54 References Lindblad, M. S., Ranucci, E., and Albertsson, A. C., (2001), Biodegradable polymers from renewable sources. New hemicellulose-based hydrogels, Macromolecular Rapid Communications, 22, 962-967. Madani, A., Kiiskinen, H., Olson, J. A., and Martinez, D. M., (2011), Fractionation of microfibrillated cellulose and its effects on tensile index and elongation of paper, Nordic Pulp and Paper Research Journal, 26, 306-311. Merewether, J.W.T., (1962), The precipitation of lignin from eucalyptus kraft black liquors, Tappi Journal 45, 159-163. Moosavifar, A., Sedin, P., and Theliander, H., (2006), Viscosity and boiling point elevation of black liquor: Consequences when lignin is extracted from the black liquor, Nordic Pulp and Paper Research Journal, 21, 180-187. Mörck, R., Kringstad, K., and Månsson, P., (1982), Structural heterogeneity in softwood kraft lignin as studied by GPC, 13C-NMR, 1H-NMR, Canadian Wood Chemistry Symposium, September 13-15, Niagara Falls, Ontario, Canada, 161-168. Mörck, R., Yoshida, H., and Kringstad, K. P., (1986), Fractionation of kraft lignin by successive extraction with organic solvents, functional groups 13C-NMR-spectra and molecular weight distributions, Holzforschung, 40(supplement issue), 51-60. Nagy, M., Kosa, M., Theliander, H., and Ragauskas, A. J., (2010), Characterization of CO2 precipitated kraft lignin to promote its utilization, Green Chemistry, 12, 31-34. Nakagaito, A. N., Yano, H., (2004), The effect of morphological changes from pulp fiber towards nano-scale fibrillated cellulose on the mechanical properties of high-strength plant fiber based composites, Applied Physics, A78, 547-552. Niemi, H., Lahti, J., Hatakka, H., Kärki, S., Rovio, S., Kallioinen, M., Mänttäri, M., and Louhi-Kultanen, M., (2011), Fractionation of organic and inorganic compounds from black liquor by combining membrane separation and crystallization, Chemical Engineering Technology 34, 593-598. Norberg, I., (2012), Carbon fibres from kraft lignin, Ph.D. Thesis, Royal Institute of Technology, KTH, Stockholm, Sweden. Nordin, A-K., (2008), Process design for ultrafiltration of complex process streams, Ph.D. Thesis, Lund University, Sweden. Nordström, Y., Norberg, I., Sjöholm, E., and Drougge, R. (2013), A new softening agent for melt spinning of softwood kraft lignin, Journal of Applied Polymer Science, 129, 12741279. Norgren, M., and Lindström, B., (2000), Dissociation of phenolic groups in kraft lignin at elevated temperatures, Holzforschung, 54, 519-527. 55 References Passinen, K., (1968), Chemical composition of spent liquors, Proceedings of IUPAC EUCEPA Symposium: Recovery of Pulping Chemicals, Helsinki, Finland, 183-209. Persson, T., Krawczyk, H., Nordin, A.-K., and Jönsson, A.-S., (2010), Fractionation of process water in thermomechanical pulp mills, Bioresource Technology, 101, 3884-3892. Pollak, A. Keilen, J. J., Jr., and Drum, L. F., (1949), Method of producing lignin from black liquor, United States Patent 2,464,828. Pouteau, C., Dole, P., Cathala, B., Averous, L., and Boquillon, N., (2003), Antioxidant properties of lignin in polypropylene, Polymer Degradation and Stability, 81, 9-18. Pääkkö, M., Ankerfors, M., Kosonen, H., Nykänen, A., Ahola, S., Österberg, M., Ruokolainen, J., Laine, J., Larsson, P. T., Ikkala, O., and Lindström, T., (2007), Enzymatic hydrolysis combined with mechanical shearing and high-pressure homogenization for nanoscale cellulose fibrils and strong gels, Biomacromolecules, 8, 1934-1941. Ragauskas, A. J., Williams, C. K., Davison, B. H., Britovsek, G., Cairney, J., Eckert, C. A., Frederick Jr., W. J., Hallett, J. P., Leak, D. J., Liotta, C. L., Mielenz, J. R., Murphy, R., Templer, and R.,Tschaplinski, T., (2006), The path forward for biofuels and biomaterials, Science, 311, 484-489. Ragnar, M., Lindgren, C. T., and Nilvebrant, N. O., (2000), pKa-values of guaiacyl and syringyl phenols related to lignin, Journal of Wood Chemistry and Technology, 20, 277-305. Reboulet, H. J., (1941), Molding material and method of producing the same, United States Patent 2,228,976. Rinman, E. L., (1911), Process of treating waste liquors from soda pulp mills, United States Patent 1,005,882. Ripperger, S., Gösele, W., and Alt, C., (2009), Filtration, 1. Fundamentals, Ullmann's Encyclopedia of Industrial Chemistry, Wiley-VCH Verlag, Weinheim. Rojas, O. J., Song, J., Argyropoulos, D. S., and Bullón, J., (2006), Lignin separation from kraft black liquors by tangential ultrafiltration, La Chimica e I´Industria, 88-95. Saito, T., Nishiyama, Y., Putaux, J-L, Vignon, M., and Isogai, A., (2006), Homogeneous suspensions of individualized microfibrils from TEMPO-catalyzed oxidation of native cellulose, Biomacromolecules, 7, 1687-1691. Sannigrahi, P., Ho Kim, D., Jung, S., and Ragauskas, A., (2011), Pseudo-lignin and pretreatment chemistry, Energy and Environmental Science, 4, 1306-1310. Scott, C. W., (1940), Process of producing molding composition from lignin and formaldehyde, United States Patent 2,201,797. 56 References Siró, I., and Plackett, D., (2010), Microfibrillated cellulose and new nanocomposite materials: a review, Cellulose, 17, 459-494. Skogsindustrin, (2012), En faktasamling 2012 års branschstatistik, Skogsindustrierna, Stockholm, Sweden. Song, T., Pranovich, A., Sumerskiy, I., and Holmbom, B., (2008), Extraction of galactoglucomannan from spruce wood with pressurised hot water, Holzforschung, 62, 659666. Stewart, D., (2008), Lignin as a base material for materials applications: Chemistry, application and economics, Industrial Crops and Products, 27, 202-207. Sundin, J., (2000), Precipitation of kraft lignin under alkaline conditions, Ph.D. Thesis, Royal Institute of Technology, KTH, Stockholm, Sweden. Taniguchi, T., and Okamura, K., (1998), New films produced from microfibrillated natural fibres, Polymer International, 47, 291-294. Teleman, A., (2009), Hemicelluloses and pectins, Pulp and paper chemistry, Wood chemistry and wood biotechnology 1, Ek, M., Gellerstedt, G., Henriksson, G., (Eds.), Walter de Gryter GmbH & Co, Berlin. Tessié du Motay, C. M., (1872), Improvement in recovering waste alkalies used in treating paper-pulp, United States Patent 132,452. Theliander, H., (2009), Recover of cooking chemicals: the treatment and burning of black liquor, Pulp and paper chemistry, Pulping chemistry and technology 2, 1 st edition, Ek, M., Gellerstedt, G., Henriksson, G., (Eds.), Walter de Gryter GmbH & Co, Berlin. Thornton, J., Ekman, R., Holmbom, B., and Örså, F., (1994), Polysaccharides dissolved from Norway spruce in thermomechanical pulping and peroxide bleaching, Journal of Wood Chemistry and Technology, 14, 159-175. Toledano, A., García, A., Mondragon, I., and Labidi, L., (2010), Lignin separation and fractionation by ultrafiltration, Separation and Purification Technology, 71, 38-43. Tomani, P., (2013), Update on LignoBoost lignin and applications, SPCI Convention (Svenska Cellulosa- och Pappers Ingenjörsföreningen), September 25-26, Stockholm, Sweden. Tomlinson, G. H., and Tomlinson Jr., G. H., (1946), Method of treating lignocellulosic material, United States Patent 2,406,867. Tomlinson, G. H., and Tomlinson Jr., G. H., (1948), Method of treating lignocellulosic material, Canadian Patent CA 448,476. Tsuguyuki, S., Masayuki, H., Naoyuki, T., Satoshi, K., Hayaka, F., Laurent, H., and Akira, I., (2009), Individualization of nano-sized plant cellulose fibrils by direct surface 57 References carboxylation using TEMPO catalyst under neutral conditions, Biomacromolecules, 7, 19921996. Turbak, A. F., Snyder, F.W., and Sandberg, K. R., (1982), Food products containing microfibrillated cellulose, U.S. Patent 4,341,807. Turbak, A. F.; Snyder, F. W., and Sandberg, K. R., (1983), Microfibrillated cellulose: A new cellulose product: properties, uses, and commercial potential, Journal of Applied Polymer Science: Applied Polymer Symposium, 37, 815-827. Turbak, A. F., Snyder, F.W., and Sandberg, K. R., (1983), Microfibrillated cellulose, U.S. Patent 4,374,702. Turbak, A. F., Snyder, F. W., and Sandberg, K. R., (1983), Suspensions containing microfibrillated cellulose, U.S. Patent 4,378,381. Uloth, V. C., and Wearing, J. T., (1989), Kraft lignin recovery: Acid precipitation versus ultrafiltration, Part I: Laboratory test results, Pulp and Paper Canada, 90, 67-71. Uraki Y, Sugiyama Y, Koda K, Kubo S, Kishimoto T, and Kadla J. F., (2012), Thermal mobility of β-O-4-type artificial lignin, Biomacromolecules, 13, 867-872. Vishtal, A., Kraslawski, A., (2011), Challenges in industrial application of technical lignins, BioResources, 6, 3547-3568. Wada, S., Iwamida, T., Iizima, R., and Yabe, K., (1962), Fractional precipitation of kraft black liquor from spruce wood, Chemistry of High Polymers (Kabunshi Kagaku), 19, 699-703. Wallberg, O. and Jönsson, A.-S., (2003), Influence of the membrane cut-off during ultrafiltration of kraft black liquor with ceramic membranes, Chemical Engineering Research and Design, 81(A), 1379-1384. Wallberg, O., Jönsson, A.-S., and Wimmerstedt, R., (2003), Fractionation and concentration of kraft black liquor lignin with ultrafiltration, Desalination, 154, 187-199. Wallberg, O., Jönsson, A.-S., and Wimmerstedt, R., (2003), Ultrafiltration of kraft black liquor with a ceramic membranes, Desalination, 156, 145-153. Wallberg, O., Holmqvist, A., and Jönsson, A.-S., (2005), Ultrafiltration of kraft cooking liquors from a continuous cooking process, Desalination, 180, 109-118. Wallmo, H., and Theliander, H., (2009), The influence of hemicelluloses during the precipitation of lignin in kraft black liquor, Nordic Pulp and Paper Research Journal, 24, 165171. Weber, M., and Weber, M., (2010), Phenolic resins: A century of progress, Louis, P. (Ed.), Springer-Verlag Berlin Heidelberg, 9-23. 58 References Willför, S., Sjöholm, R., Laine, C., Roslund, M., Hemming, J., and Holmbom, B, (2003), Characterisation of water-soluble galactoglucomannans from Norway spruce wood and thermomechanical pulp, Carbohydrate Polymers, 52, 175-187. Willför, S., Rehn, P., Sundberg, A., Sundberg, K., and Holmbom, B., (2003), Recovery of water-soluble acetylgalactoglucomannans from mechanical pulp of spruce, Tappi Journal, 2, 27-32. Willför, S., Sundberg, K., Tenkanen, M., and Holmbom, B., (2008), Spruce-derived mannans - A potential raw material for hydrocolloids and novel advanced natural materials, Carbohydrate Polymers, 72, 197-210. Woerner, D. L., and McCarthy, J. L., (1984), Ultrafiltration of kraft black liquor, American Institute of Chemical Engineers, AIChE Symposium Series 80, 25-33. Woerner, D. L., and McCarthy, J. L., (1987), Ultrafiltration of pulp mill liquors, Tappi Journal, 70, 126-129. Yoshida, H., Mörck, R., Kringstad, K. P., and Hatakeyama, H., (1987), Fractionation of kraft lignin by successive extraction with organic solvents, II Thermal properties of kraft lignin fractions, Holzforschung, 41, 171-176. Yoshida, H., Mörck, R., Kringstad, K.P., and Hatakeyama, H. J., (1990), Kraft lignin in polyurethanes. II. Effects of the molecular weight of kraft lignin on the properties of polyurethanes from a kraft lignin–polyether triol –polymeric MDI system, Journal of Applied Polymer Science, 40, 1819-1832. Zhang, L., and Gellerstedt, G., (2001), Lignin fractionation and quantitative structural analysis of the lignin fractions with 13C and HSQC NMR techniques, Proceedings, Postsymposium Workshop of the 11th International Symposium on Wood and Pulping Chemistry, 3:595, Grenoble, France, Zhao, H. P., Feng, X. Q., and Gao, H., (2007), Ultrasonic technique for extracting nanofibres from nature materials, Applied Physics Letters, 90, 073112-1. Zimmermann, T., Pöhler, E., Geiger, T., (2004), Cellulose fibrils for polymer reinforcement, Advanced Engineering Materials, 6, 754-761. Öhman, F., and Theiander, H., (2006), Washing lignin precipitated from kraft black liquor, Paperi ja Puu 88, 287-292. Öhman, F., Wallmo, H., and Theliander, H., (2007), A novel method for washing lignin precipitated from kraft black liquor - Laboratory trials, Nordic Pulp and Paper Research Journal 22, 9-16. Örså, F., Holmbom, B., and Thornton, J., (1997), Dissolution and dispersion of spruce wood components into hot water, Wood Science and Technology, 31, 279-290. 59
© Copyright 2025 Paperzz