Transcript

Characterization of the White-rot Fungus, Phanerochaete carnosa,

through Proteomic Methods and Compositional Analysis of Decayed Wood Fibre

By

Sonam Mahajan

A thesis submitted in conformity with the requirements for the degree of Doctorate of Philosophy

Department of Chemical Engineering and Applied Chemistry University of Toronto

© Copyright by Sonam Mahajan 2011

ii

Characterization of the white-rot fungus, Phanerochaete carnosa,

through proteomic methods and compositional analysis of decayed wood fibre

Sonam Mahajan

Doctorate of Philosophy

Department of Chemical Engineering and Applied Chemistry

University of Toronto

2011

Abstract

Biocatalysts are important tools for harnessing the potential of wood fibres since they can

perform specific reactions with low environmental impact. Challenges to bioconversion

technologies as applied to wood fibres include low accessibility of plant cell wall polymers and

the heterogeneity of plant cell walls, which makes it difficult to predict conversion efficiencies.

White-rot fungi are among the most efficient degraders of plant fibre (lignocellulose), capable of

degrading cellulose, hemicellulose and lignin. Phanerochaete carnosa is a white-rot fungus that,

in contrast to many white-rot fungi that have been studied to date, was isolated almost

exclusively from fallen coniferous trees (softwood). While several studies describe the

lignocellulolytic activity of the hardwood-degrading, model white-rot fungus Phanerochaete

chrysosporium, the lignocellulolytic activity of P. carnosa has not been investigated.

iii

An underlying hypothesis of this thesis is that P. carnosa encodes enzymes that are particularly

well suited for processing softwood fibre, which is an especially recalcitrant feedstock, though a

major resource for Canada. Moreover, given the phylogenetic similarity of P. carnosa and P.

chrysosporium, it is anticipated that the identification of pertinent enzymes for softwood

degradation can be more easily conducted. In particular, this project describes the

characterization of P. carnosa in terms of the growth conditions that support lignocellulolytic

activity, the effect of enzymes secreted by P. carnosa on the chemistry of softwood feedstocks,

and the characterization of the corresponding secretome using proteomic techniques. Through

this study, cultivation methods for P. carnosa were established and biochemical assays for

protein activity and quantification were developed. Analytical methods, including FTIR and

ToF-SIMS were used to characterize wood samples at advancing stages of decay, and revealed

preferential degradation of lignin in the early stages of growth on all softwoods analyzed.

Finally, an in depth proteomic analysis of the proteins secreted by P. carnosa on spruce and

cellulose established that similar sets of enzyme activities are elicited by P. carnosa grown on

different lignocellulosic substrates, albeit to different expression levels.

iv

Acknowledgements

I would like to express my deepest gratitude to Dr. Emma Master for accepting me as her PhD

student 5 some years ago, in spite of the sufficient lack of background that I brought with me in

this demanding field. Over the years, I have acquired technical skills and developed the ability to

learn newer information better - it would certainly not have been possible without her kind

patience throughout my learning stages. I am very grateful for her constant guidance,

intellectually stimulating discussions, understanding of my personal and technical challenges,

and encouragement during the trying times.

I am also very grateful to Dr. Dragica Jeremic for all her technical assistance, and more

importantly, for her friendly guidance throughout the later years of my PhD. It would have been

very difficult for me to see the end of my PhD if it were not for her un-ending support.

I am thankful to the members of my committee, Dr. Elizabeth Edwards and Dr. Krishna

Mahadevan, for the feedback and direction they provided for this project.

I am also grateful to Dr. Robyn Goacher for her technical assistance and contribution to the fibre

characterization studies, and to Peter Brodersen for his help in the early phases of fibre

characterization experiments with ToF-SIMS. I would also like to express my thanks to Dr. Eric

Yang from Sunnybrooke for his collaboration through the Proteomic Studies and to Dr. Tony

Ung for his assistance in sugar analysis and kind offering of all technical resources whenever

required.

I will always remember Jacqueline, for our friendly PhD pep-talks and for the technical and

moral support in designing the mammoth fibre characterization experiments. Finally, my huge

appreciation for all the members of the Master Lab and BioZone for their cooperation, and for

accepting me as a relatively inert member of the group, especially during the final stages of my

PhD.

v

I am very grateful to my parents, Dr. Ravi Mahajan and Dr. Kalpana Mahajan, who inspired me

to begin this marathon journey in life. Over the last 5 years, I’ve learnt a significant amount and

in spite of the challenges, I would have never taken it up, if it were not for them. My loving

thanks to my sister, Samridhi for adding a fresh breathe of non-academic humor to my

sometimes humdrum life.

I am indebted to my husband, Ateet, for his never-ending patience, gentle encouragement, kind

technical assistance and acceptance of all sloppy standards at home. And to all my dearest

friends, whose warmth created for me a home away from home, whose company made dull

moments bright, and whose smiles, encouraging emails, texts and generous home visits with

food, kept me alive through the most trying times…we did it!

Finally, my heartfelt gratitude to all my spiritual teachers and mentors, who have carved on my

heart determination and faith in the absolute will, the ability to discern the temporary from the

eternal, and the desire to serve with perfection…

Things that are very difficult to do become easy to execute if one somehow or other simply

remembers Lord Caitanya Mahaprabhu. But if one does not remember Him, even easy things

become very difficult. To this Lord Caitanya Mahaprabhu I offer my respectful obeisances.

Caitanya Caritamrita Adi Lila 14.1

vi

Table of Contents

Table of Contents ......................................................................................................................... vi

List of Figures .................................................................................................................................x

List of Abbreviations ................................................................................................................... xi

Chapter 1 : Overview.....................................................................................................................1

Chapter 2 : Literature Review ......................................................................................................7

2.1 Lignocellulose: A Valuable Resource ...............................................................................7

2.2 Composition and Structure of Plant Cell Walls ..............................................................8

2.3 Bioconversion of Lignocellulose........................................................................................9

2.4 Lignocellulose-degrading Bacteria ...................................................................................9

2.5 Lignocellulose-degrading Fungi .....................................................................................10

2.5.1 White-Rot Fungi...................................................................................................11

2.5.2 Brown-Rot Fungi .................................................................................................11

2.6 Lignocellulose Active Enzymes .......................................................................................12

2.6.1 Carbohydrate Active Enzymes ...........................................................................12

2.6.1.1 Cellulases ..............................................................................................................13

2.6.1.2 Hemicellulases ......................................................................................................13

2.6.1.3 Fungal Oxidative Lignin Enzymes .....................................................................14

2.7 Investigative Approaches to Improve Lignocellulose Bioconversion ..........................14

2.7.1 Analytical Characterization of Wood Fibre ......................................................15

2.7.2 Biochemical Characterization of Wood-degrading Fungi ...............................18

Chapter 3 : Effect of Cultivation Conditions on the Expression of Cellulolytic Activity by Phanerochaete species P. chrysosporium and P. carnosa ................................................30

3.1 Abstract .............................................................................................................................30

3.2 Introduction ......................................................................................................................30

vii

3.3 Review of P. chrysosporium Cultivation and Lignocellulolytic Activity .....................32

3.4 Materials and Methods ....................................................................................................41

3.4.1 Microorganism and Materials ............................................................................41

3.4.2 Cultivation Conditions.........................................................................................41

3.4.3 Growth Measurements ........................................................................................42

3.4.4 Biochemical Assays ..............................................................................................43

3.5 Results ...............................................................................................................................45

3.5.1 Effect of Cultivation Condition on the Extent of P. carnosa Growth on Microcrystalline Cellulose. ..............................................................................................45

3.5.2 Effect of Nitrogen Concentration and Agitation on Lignocellulolytic Expression by P. carnosa. ................................................................................................48

3.6 Discussion..........................................................................................................................51

Chapter 4 : Mode of Coniferous Wood Decay by the White Rot Fungus Phanerochaete carnosa as Confirmed by FT-IR and ToF-SIMS ..................................................................56

4.1 Abstract .............................................................................................................................56

4.2 Introduction ......................................................................................................................56

4.3 Materials and Methods ....................................................................................................59

4.3.1 Fungal Strain and Cultivation Conditions ........................................................59

4.3.2 Fourier Transform Infrared Spectroscopy .......................................................59

4.3.3 Transmission Electron Microscropy - Energy-Dispersive X-ray Analysis (TEM-EDXA) ...................................................................................................................60

4.3.4 Time of Flight Secondary Ion Mass Spectroscopy ............................................60

4.3.5 Statistical Analysis ...............................................................................................61

4.4 Results ...............................................................................................................................61

4.4.1 FTIR Spectroscopy ..............................................................................................62

4.4.2 TEM-EDXA ..........................................................................................................68

4.4.3 ToF-SIMS .............................................................................................................68

viii

4.5 Discussion..........................................................................................................................71

Chapter 5 : Proteomic Characterization of Lignocellulose-degrading Enzymes Secreted by Phanerochaete carnosa Grown on Spruce and Microcrystalline Cellulose ..................79

5.1 Abstract .............................................................................................................................79

5.2 Introduction ......................................................................................................................79

5.3 Materials and Methods ....................................................................................................81

5.3.1 Cultivation Conditions.........................................................................................81

5.3.2 Protein Extraction ................................................................................................81

5.3.3 Protein Preparation and Analysis by Mass Spectrometry ...............................82

5.3.4 Peptide Sequence Annotation .............................................................................82

5.4 Results ...............................................................................................................................83

5.4.1 Preparation and Analysis of Peptide Samples...................................................83

5.4.2 Cellulases and Hemicellulases in Cellulose and Spruce Cultivations .............84

5.4.3 Oxidoreductases in Cellulose and Spruce Cultivations ....................................91

5.4.4 Carbohydrate Esterases, Proteases and Other Glycoside Hydrolases ............95

5.4.5 Peptides Annotated to Hypothetical Proteins Identified in Proteomic Analyses of P. chrysosporium ........................................................................................100

5.5 Discussion........................................................................................................................101

Chapter 6: Synthesis and Conclusions .....................................................................................107

Chapter 7: Engineering Relevance ...........................................................................................110

Appendix 1 : B3 Medium ..........................................................................................................113

Appendix 2: Supplemental Information for Chapter 4 ..........................................................115

Appendix 3: Supplemental Information for Chapter 5 ..........................................................119

Appendix 4: Substrate recognition and hydrolysis by a fungal xyloglucan-specific family 12 hydrolase ...............................................................................................................126

Appendix 6: Additional Information Collected During ResearchError! Bookmark not defined.

ix

List of Tables

Table 3.1 Summary of carbohydrate active enzyme production by Phanerochaete chrysosporium

Table 3.2 Expression of lignocellulolytic activity by Phanerochaete chrysosporium

Table 3.3 Cultivation of P. carnosa in B3 medium at 27˚C with 1 % microcrystalline cellulose, variable nitrogen, pH 4.5 Table 3.4 Protein concentration in P. carnosa culture supernatant during growth on microcrystalline cellulose

Table 3.5 Chitin content of P. carnosa cultivations after 24 days

Table 3.6 Endoglucanase activity in P. carnosa culture supernatant during growth on microcrystalline cellulose

Table 3.7 β-glucosidase activity in P. carnosa culture supernatant during growth on microcrystalline cellulose Table 3.8 Total cellulolytic activity in P. carnosa culture supernatant during growth on microcrystalline cellulose Table 3.9 MnP activity in P. carnosa culture supernatant during growth on microcrystalline cellulose Table 4.1 Assignment of FTIR peaks that distinguish control and decayed wood samples Table 5.1 Analysis of peptides from extracellular filtrates of P. carnosa grown on cellulose and spruce that correspond to predicted cellulases and hemicellulases Table 5.2 Analysis of peptides from extracellular filtrates of P. carnosa grown on cellulose and spruce that correspond to extracellular oxidases and proteins involved in lignin degradation Table 5.3 Analysis of peptides from extracellular filtrates of P. carnosa grown on cellulose and spruce that correspond to extracellular carbohydrate-active enzymes Table 5.4 Peptides from extracellular filtrates of P. carnosa grown on cellulose that were annotated to proteins from P. chrysosporium with unknown function

x

List of Figures Figure 3.1 SDS-PAGE of P.carnosa culture supernatants Figure 4.1 Spruce decay described by FTIR analysis Figure 4.2 Fir decay described by FTIR analysis Figure 4.3 Pine decay described by FTIR analysis Figure 4.4 PCA of normalized and mean-centered ToF-SIMS spectra: A) Fir, B) Pine, and C) Spruce Figure 4.5 PCA of mean-centered ToF-SIMS image of decayed pine (time point 5) Figure 5.1 Distribution of peptide annotations from proteins produced by P. carnosa grown on (A) crystalline cellulose, and (B) spruce wood chips Figure 6.1 Synthesis of information flow between the three key objectives of this thesis

xi

List of Abbreviations FTIR - Fourier Transform InfraRed ToF-SIMS - Time-of-Flight Secondary Ion Mass Spectrometry TEM-EDXA - Transmission Electron Microscope with Energy Dispersive X-ray Analysis SEM - scanning electron microscopy CAZymes - Carbohydrate Active Enzymes FOLymes - Fungal Oxidative Lignin enzymes GH - Glycoside Hydrolase CE - Carbohydrate esterase LiP - Lignin peroxidase MnP - Manganese peroxidase ABTS - 2,2'-azino-bis(3-ethylbenzthiazoline-6-sulphonic acid) EGase - Endoglucanase CBH - Cellobiohydrolase CBD - Cellobiose Dehydrogenase BGL - beta-glucosidase FPU - Filter Paper Units XYN - Xylanase MAN - Mannanase PCR - Polymerase chain reaction Ribosomal ITS sequences – ribosomal internal transcribed spacer sequences BSA - Bovine Serum Albumin LN - Low nitrogen, stationary cultivations LNS - Low nitrogen cultivations with shaking HN - Low nitrogen, stationary cultivations PCA - Principal Component Analysis PLS Toolbox - Partial Least Squares Toolbox PDMS - Polydimethylsiloxane MWCO – Molecular weight cut off

1

Chapter 1 : Overview

Terrestrial plants synthesize nearly 200 billion tons of lignocellulosic biomass per year (Zhang

2008). Lignocellulose comprises the structural component of plant cell walls and is composed of

cellulose, hemicellulose, and lignin, as well as relatively low amounts of aliphatic and alicyclic

compounds, phenolic compounds, and proteins. The “lignocellulose biorefinery” aims to produce

renewable fuels, chemicals and new polymers from all the constituents of lignocellulose. As the

non-consumable portion of plant biomass, bioproducts from lignocellulose do not directly

compete with food resources. Moreover, lignocellulose can be supplied as a residual of

agricultural and forest industries.

Despite the advantages of lignocellulose, the complex and variable composition of this feedstock

creates significant technical barriers that can be broadly grouped into 1) the production of

fermentable sugars and technical polymers, and 2) the conversion of mixed sugars to fuels and

chemicals. Current research towards efficient production of fermentable sugars and polymeric

materials is aimed at improving the integration of pretreatment technologies and downstream

enzyme applications, and minimizing the production of compounds that inhibit fermentation

processes (Kabel et al. 2006). Still, the yield of fermentable sugars from enzyme treatments is

variable and incomplete, and the effect of commercially available enzymes often depends on the

nature of the substrate, rather than advertised enzyme activity (Lynd et al. 2008).

White-rot fungi are the most efficient degraders of lignocellulose as they can degrade cellulose

and hemicellulose, as well as lignin (Schmidt 2006). Many white-rot fungi have been isolated

from hardwoods. By contrast, Phanerochaete carnosa is a white-rot fungus that has been

isolated predominantly from softwood (Burdsall 1985). While several studies describe the

lignocellulolytic behaviour of the hardwood-degrading, model white-rot fungus P.

chrysosporium (Sato et al. 2007; Vanden Wymelenberg et al. 2005a; Vanden Wymelenberg et al.

2005b), the lignocellulolytic activity of P. carnosa has not been investigated. As a softwood

degrader, P. carnosa might encode enzymes that are particularly well suited for processing

softwood fibre, which consists of different lignin and hemicellulose composition, compared to

hardwood. Given the abundance and comparatively high recalcitrance of softwood fibre (Mosier

2

et al. 2005), enzymes that effectively process this material would benefit forest sectors in

northern countries, including Canada, where boreal forests occupy 77% of the forest land

(Natural Resources Canada 2011).

Accordingly, the main aim of this PhD thesis is to characterize the secretome of P. carnosa

during growth on lignocellulosic substrates, and to evaluate the pattern of wood decay by this

fungus over time. In addition to increasing our fundamental understanding of bioprocesses that

have evolved to transform recalcitrant biomass, the analyses described herein assesses the

potential of P. carnosa to be used for biopulping and as a supply of industrially relevant

enzymes.

In particular, the following specific research hypotheses were tested:

1. Cultivation conditions that promote the expression of lignocellulolytic enzymes in P.

chrysosporium will also induce the expression of lignocellulolytic enzymes in P. carnosa.

2. P. carnosa will exhibit selective degradation of lignin at early stages of softwood decay, and

having been isolated from softwood species, P. carnosa will efficiently degrade guaiacyl

lignin, which dominates in softwood.

3. The profile of proteins secreted by P. carnosa during growth on lignocellulose will depend

on the source of the substrate, and could be used to predict enzyme formulations that

optimally transform a particular biomass feedstock.

As summarized below, the research hypotheses listed above were tested as described in chapters

three, four and five, respectively.

Chapter 3: Effect of cultivation conditions on the expression of cellulolytic activity by

Phanerochaete species, P. chrysosporium and P. carnosa

This study describes the first reported effort to cultivate and biochemically characterize

Phanerochaete carnosa. Accordingly, the primary aim of this study was to establish growth

conditions that induce the expression of lignocellulolytic enzymes by this organism, and to

investigate the effect of nitrogen concentration, temperature and agitation on the production of

cellulolytic and lignin-degrading activities. Protein concentration in culture supernatants, fungal

3

growth, and biochemical assays for cellulases, lignin peroxidases and manganese peroxidases,

were measured. By comparing the effect of cultivation condition on the production of

lignocellulolytic activity by P. carnosa and P. chrysosporium, differences in enzyme activities

and their regulation were also evaluated.

Chapter 4: Mode of coniferous wood decay by the white-rot fungus Phanerochaete carnosa

as elucidated by FTIR and ToF-SIMS (a version of this chapter has been accepted for

publication)

The objective of this study was to investigate the effect of conifer species and stage of

degradation on the mode of softwood decay by P. carnosa (i.e., selective vs. simultaneous

decay). The three wood species investigated were white spruce, lodgepole pine and balsam fir,

all of which are abundant in the North American boreal forests, and comprise common mixed

feedstocks for pulp and paper industries. FTIR and ToF-SIMS were used to monitor changes in

lignocellulose composition resulting from fungal activity. Additionally, TEM and SIMS imaging

were used to evaluate the selectivity of the degradation across cell walls by determining the

localization of lignin prior and post-decay.

Chapter 5: Proteomic analysis of the secretome of P. carnosa grown on microcrystalline

cellulose and spruce wood chips (a version of this chapter was published in Applied and

Microbial Biotechnology)

The objective of this research was to describe the secretome of P. carnosa while growing on

cellulose and spruce. A comparative study of the extracellular proteins expressed in both

cultivation conditions was performed to determine whether a model cellulose substrate elicited a

similar profile of lignocellulose-active enzymes as a complex lignocellulosic substrate. Since the

genome sequence of P. carnosa was not available, the proteomic protocols were modified to

accommodate de novo sequencing and manual validation of peptides along with sequence-

similarity based annotations. In this way, the regulation of lignocellulose-active enzyme

expression in P. carnosa was investigated, as was the possibility to define enzyme formulations

that are optimal for particular feedstocks. Finally, the proteomic data resulting from this study

were analyzed to predict enzyme activities that may have evolved in P. carnosa to promote

softwood degradation.

4

As summarized in Chapters 6 and 7, the results of the research described herein elucidate the

growth conditions that support lignocellulolytic activity of P. carnosa, the differential effects of

P. carnosa on the chemistry of softwood feedstocks, and the expression profile of enzymes

secreted by P. carnosa during growth on cellulose and a lignocellulosic feedstocks. As a result,

the softwood-degrading activity of P. carnosa was confirmed and the applied potential of P.

carnosa and encoded enzymes was explored.

Summary of Scholarly Contributions: A. Peer-reviewed Publications: MASc 1. S. Mahajan, S.K. Konar and D.G.B. Boocock. Standard Biodiesel from Soybean Oil by a Single Chemical Reaction. Journal of American Oil Chemists’ Society. 2006. 84: 641-644. 2. S. Mahajan, S.K. Konar and D.G.B. Boocock. Determining acid number for biodiesel. Journal of American Oil Chemists’ Society. 2006. 83: 567-570. 3. S. Mahajan, S.K. Konar and D.G.B. Boocock. Variables Affecting the Production of Standard Biodiesel. Journal of American Oil Chemists’ Society. 2007. 84: 189-195. PhD 4. J. Powlowski, S. Mahajan, M. Schapira and E.R. Master. Substrate Recognition and Hydrolysis by a Fungal Xyloglucan-Specific Family 12 Hydrolase. Carbohydrate research. 2009. 344(10):1175-9. (see Appendix 4). (Contribution: Performed enzymatic digestion for

xyloglucans and detected products through HPLC) 5. S. Mahajan and E.R. Master. Proteomic Characterization of Lignocellulose-degrading Enzymes Secreted by Phanerochaete carnosa Grown on Spruce and Microcrystalline Cellulose. Applied Microbiology and Biotechnology. 2010. 86(6):1903-14. (see Chapter 5) 6. S. Mahajan, D. Jeremic, R. E. Goacher and E. R. Master. Mode of coniferous wood decay by the white-rot fungus Phanerochaete carnosa as elucidated by FTIR and ToF-SIMS. 2011 (accepted with revisions; see Chapter 4). (Contribution: established growth conditions for P.

carnosa on softwood species; harvested samples and performed FTIR analysis; prepared

5

samples for ToF-SIMS and performed data analyses in collaboration with D. Jeremic and R.E.

Goacher.) B. Conference Presentations (Presenter is underlined): MASc 1. S. Mahajan, S.K. Konar and D.G.B. Boocock. ‘Production of Standard Biodiesel from Soybean Oil: A one-step process using NaOH’ (speaker + poster). 55th Canadian Chemical Engineering Conference, Toronto, ON, October 2005 PhD 2. S. Mahajan and E.R. Master. ‘Proteomic Analysis of Softwood-degrading Fungi’ (speaker). 56th Canadian Chemical Engineering Conference, Sherbrooke, QC, October 2006 3. S. Mahajan and E.R. Master. ‘Proteomic Analysis of Softwood-degrading Fungi’ (poster). 56th Canadian Chemical Engineering Conference, Sherbrooke, QC, October 2006 4. S. Mahajan. ‘Unit Operations Laboratory: Problem based learning’ (speaker). 56th Canadian Chemical Engineering Conference, Sherbrooke, QC, October 2006 5. S. Mahajan and E.R. Master. ‘Proteomic Analysis of Softwood-degrading Fungi towards Biomimetic Enzyme Applications’ (poster). The World Congress on Industrial Biotechnology and Bioprocessing, Toronto, ON, July 2006 6. S. Mahajan and E.R. Master. ‘Proteomic Analysis of Softwood-degrading Fungi’ (poster). 107th General Meeting of the American Society for Microbiology.Toronto, ON May 2007 7. S. Mahajan and E.R. Master. ‘Refining models of Lignocellulose Biotransformation towards Biomimetic Enzyme Applications’ (poster). 10th International Congress on Biotechnology in the Pulp and Paper Industry, Madison WI, June 2007 8. J. MacDonald, S. Mahajan and E. R. Master ‘Transcription Profiles and Proteomic Analysis of P. carnosa Grown on Softwood Feedstocks for Tailored Applications of Hydrolytic Enzymes.’ MIE BioForum Mie, Japan. Sep 2008 9. S. Mahajan and E.R. Master. ‘Proteomic Analysis of the Secretome of Softwood-Degrading Fungi’ (poster). 4th annual conference US HUPO North Bethesda, MD, March 2008 10. S Mahajan, E Yang and E.R. Master. ‘The secretome of the softwood-degrading Basidiomycete, Phanerochaete carnosa, grown on cellulose and spruce.’

6

Gordon Research Conference (GRC): Cellulosomes, Cellulases & Other Carbohydrate Modifying Enzymes. Proctor Academy, NH, USA. July 26-31, 2009. 11. J. MacDonald, S. Mahajan and E. R. Master. ‘Proteomic and Transcriptomic Analysis of Lignocellulose-degrading enzymes in softwood-degrading fungus, Phanerochaete carnosa’ Lignobiotech One Symposium, Reims, France, March 2010 Chapter 1 References Burdsall HHJ (1985) A contribution to the taxonomy of the genus Phanerochaete (Corticiaceae,

Aphyllophorales). Mycological Memoir No. 10. Braunschweig. Kabel MA, van der Maarel MJEC, Klip G, Voragen AGJ, Schols HA (2006) Standard assays do

not predict the efficiency of commercial cellulase preparations towards plant materials. Biotechnology and Bioengineering 93 (1):56-63.

Lynd LR, Laser MS, Bransby D, Dale BE, Davison B, Hamilton R, Himmel M, Keller M,

McMillan JD, Sheehan J, Wyman CE (2008) How biotech can transform biofuels. Nature Biotechnology 26 (2):169-172

Mosier N, Wyman C, Dale B, Elander R, Lee YY, Holtzapple M, Ladisch M (2005) Features of

promising technologies for pretreatment of lignocellulosic biomass. Bioresource Technology 96 (6):673-686

Sato S, Liu F, Hasan K, Tien M (2007) Expression analysis of extracellular proteins from

Phanerochaete chrysosporium grown on different liquid and solid substrates. Microbiology 153:3023-3033

Schmidt O (2006) Wood and Tree Fungi. Biology, Damage, Protection and Use. Springer-

Verlag, New York Vanden Wymelenberg A, Sabat G, Martinez D, Rajangam AS, Teeri TT, Gaskell J, K KPJ,

Cullen D (2005a) The Phanerochaete chrysosporium secretome: Database predictions and initial mass spectrometry peptide identifications in cellulose-grown medium Journal of Biotechnology 118 (1):17-34

Vanden Wymelenberg A, Sabat G, Martinez D, Rajangam AS, Teeri TT, Gaskell J, Kersten PJ,

Cullen D (2005b) The Phanerochaete chrysosporium secretome: database predictions and initial mass spectrometry peptide identifications in cellulose-grown medium. Journal of Biotechnology 118:17-34

Zhang YH (2008) Reviving the carbohydrate economy via multi-product lignocellulose

biorefineries. Journal of Industrial Microbiology Biotechnology 35 (5):367-375.

7

Chapter 2 : Literature Review

2.1 Lignocellulose: A Valuable Resource

Lignocellulose is the major structural component of woody plants. It is a network of lignin,

cellulose and hemicellulose that is chemically bonded through non-covalent forces and covalent

cross-linkages (Perez et al. 2002).

Terrestrial plants synthesize nearly 200 billion tons of lignocellulosic biomass per year (Zhang

2008). Forests, which are the primary source of lignocellulosic biomass, cover approximately

30% of the total land area, and comprise a significant source of energy, timber, and pulp (Global

Forest Resources Assessment 2005). Protective functions of forests include soil and water

conservation, biological diversity and mitigation of climate change. While conventional research

in wood fibre structure and composition focused on properties that improve timber, pulp and

paper qualities, present research has expanded to include bioconversion of wood fibre to

platform sugars that can be fermented to fuels and chemicals, and bioprocessing techniques that

increase the sustainability of processes used to convert wood fibre to high-quality pulp and other

value-added materials.

Large amounts of lignocellulosic residues (6 to 3849 million t and from 64 to 561 million t, for

US and Canada) are generated through forestry and agricultural practices, with the most common

disposal being combustion (Magdalena 2008). This is not only an environmental concern but

also represents an under-utilization of a renewable resource that can be converted to energy and

biochemicals. For instance, over 75% of organic chemicals are produced from five primary base

chemicals: ethylene, propylene, benzene, toluene and xylene. The aromatic compounds could be

produced from lignin, while the low molecular weight aliphatic compounds could be obtained

from alcohols that result from fermentation of sugars generated from bioconversion of

hemicellulose and cellulose (Howard et al. 2003). Moreover, the alcohols produced could be

utilized as a biofuel. And chemicals like vanillin, xylitol, and furfural from lignocellulosic wastes

can be used in industrial products including herbicides, pharmaceuticals, and household products

(Howard et al. 2003; Ragauskas et al. 2006).

8

2.2 Composition and Structure of Plant Cell Walls

The main components of plant cell walls are cellulose, lignin, and hemicellulose. Cellulose and

hemicellulose are polysaccharides, whereas lignin is an aromatic polymer synthesized from

phenylpropanoid precursors (Sjostrom 1983). Wood from different tree species is typically

composed of 40–50% α-cellulose, 20–35% hemicellulose and 15–35% lignin (Perez et al. 2002).

With the exception of cellulose, these polymers are synthesized inside the cell and then

organized outside the cell membrane. The primary cell wall is deposited first and is characterized

by relatively amorphous cellulose structure. During cell differentiation, the primary cell wall

expands and elongates, then secondary cell walls are synthesized, which are characterized by

cellulose microfibrils with higher crystallinity and altered hemicellulose content (Ding and

Himmel, 2006). Differentiated xylem cells with a secondary cell wall are called secondary xylem

or wood.

Cellulose is the most abundant organic polymer on earth, and is composed of repeating

cellobiose subunits that consist of β-1,4-linked glucose. The degree of polymerization (DP) of

wood-derived cellulose is typically greater than 10,000 and adjacent cellulose polymers interact

through hydrogen bonds, forming highly stable structures that contain both amorphous and

crystalline regions (Lerouxel et al. 2006; McCann and Carpita 2008). Cellulose is synthesized

by a cellulose synthase complex that is located within the cytoplasmic membrane of plant cells.

In plant cells, the cellulose synthase is comprised of many enzymes that include 36 cellulose

synthase enzymes assembled as rosette structure (Taylor et al. 2000). Accordingly, 36 cellulose

molecules are thought to emerge from each rosette structure.

Like cellulose, hemicellulose backbones are composed of β-1,4-linked sugars. However, the

particular sugar composition of hemicellulose depends on the source of the polysaccharide. For

instance, xyloglucan is the main hemicellulose of primary cell walls and consists of a β-1,4-

linked glucose backbone that is decorated with xylose, galactose and sometimes fucose

branching sugars (Hayashi and Kaida 2011). By contrast, the main hemicellulose in secondary

cells walls of hardwoods and softwoods is xylan and galactoglucomannan, respectively. Xylan is

composed of β-1,4-linked xylose that can be substituted by arabinose and glucuronic acid; xylan

9

can also be acetylated. By contrast, wood-derived mannans are composed of β-1,4-linked

mannose and glucose that can be substituted by galactose (Stenius and Vuorinen 1999). These

polymers have a lower DP than cellulose, averaging between 100 and 200 and have a lower

crystallinity (Rowell 2005a). While cellulose is synthesized by cellulose synthases located within

the cytoplasmic membrane, hemicelluloses are generated in the Golgi complex, and then secreted

to the plant cell wall (Keegstra 2010).

Lignin is a complex polymer that is composed of phenylpropane units linked together by carbon-

carbon (C-C) and ether (C-O-C) linkages. Lignin provides structural support to plant fibres, as

well as resistance against microbial attack and improved water transport in xylem and phloem

tissues. Precursors of lignin biosynthesis are p-coumaryl alcohol, coniferyl alcohol, and sinapyl

alcohol, and while p-coumaryl alcohol predominates in grasses to form H-lignin, coniferyl

alcohol is the main monolignol in softwood G-lignin, and hardwood G/S-lignin contains both

sinapyl and coniferyl monolignols (Humphreys and Chapple 2002).

Besides carbohydrates and lignin, the cell wall contains hydroxyproline-rich glycoproteins,

arabinogalactan proteins, glycine-rich proteins, and proline-rich proteins (Showalter 2001). Other

extraneous compounds that do not contribute to the structure of cell walls are grouped as

extractives and ash (Pettersen 1984).

2.3 Bioconversion of Lignocellulose

Physical and chemical processes have been developed to pretreat wood fibres to separate

cellulose, hemicellulose, and lignin; however, these processes can decrease the quality of the

polymers and create by-products that inhibit the fermentation of resulting sugars. Alternatively,

biocatalysts (enzymes) can perform specific reactions under comparatively mild reaction

conditions, and could be used to improve the pretreatment process (Lynd et al. 2008).

2.4 Lignocellulose-degrading Bacteria

Evidence of bacterial degradation of lignin is sparse. Streptomycete species have been shown to

degrade low levels of lignin (Crawford, 1978; Watanabe et al. 2003), and multicopper oxidases

10

with laccase activity have been isolated from bacteria, although these enzymes are expected to

mainly participate in sporulation (Claus 2004; Malherbe and Cloete 2002).

Accordingly, bacteria are generally considered secondary lignocellulose degraders, and can

degrade cellulose and hemicellulose both aerobically and anaerobically (Walker and Wilson

1991). Examples of aerobic (hemi)cellulose degraders include Thermobifida fusca and

Cellulomonas composti, as well as several other bacteria (Béguin and Aubert 1994). In these

cases, degradation is initiated by the concerted activity of cell-associated and free extracellular

cellulases and hemicellulases (Walker and Wilson 1991). By contrast, anaerobic cellulose-

degrading bacteria typically anchor relevant hydrolase activities to the cell through a cellulosome

complex. Cellulosomes are multienzyme complexes that bind to the bacterial cell wall and

promote the uptake of solubilized sugars by the hydrolytic organism (Bayer et al. 2008).

Clostridium thermocellum and C. cellulolyticum are among the best-studied anaerobic, cellulose-

degrading bacteria (Bayer et al. 2008). Notably, these organisms have been extensively

evaluated for their potential to promote simultaneous saccharification and fermentation (SSF) or

consolidated bioprocessing (CBP) (Lynd et al. 2002).

2.5 Lignocellulose-degrading Fungi

Fungi are the primary degraders of lignocellulose (Rabinovich et al. 2002; Sanchez 2009). In

addition to secreting enzymes that are critical to lignocellulose decomposition, fungal growth on

lignocellulose is promoted by the formation of mycelia that allow filamentous fungi to transport

nutrients, including nitrogen and iron, to the carbon-rich lignocellulosic substrate (Hammel

1997). Many fungi are also more resistant to wood-derived biocides that limit bacterial growth.

These compounds include tannins and various phenolic compounds (terpenes, stilbenes,

flavonoids and tropolones) that are particularly abundant in the heartwood of fallen trees.

The majority of wood-degrading fungi that have been characterized to date are members of the

phylum Basidiomycota and are characterized by either brown-rot or white-rot decay.

11

2.5.1 White-Rot Fungi

Fungi causing white-rot decay secrete enzymes that degrade lignin, hemicellulose and cellulose,

leaving the residual wood fibre with a bleached appearance. Two main patterns of white-rot

decay have been distinguished by microscopic and ultra structural investigations (Liese 1970).

Simultaneous white-rot (“corrosion rot”) is exemplified by combined degradation of

carbohydrates and lignin at early and late stages of decay. Examples of fungi that elicit

simultaneous white-rot include Fomes fomentarius, Phellinus robustus, and Trametes versicolor

(Blanchette 1984, Blanchette 1994). By contrast, selective (sequential) white-rot is exemplified

by early degradation of lignin and hemicelluloses followed by cellulose degradation.

Ceriporiopsis subvermispora and Phlebia radiata are perhaps the best studied fungi to elicit

selective white-rot decay (Ander and Eriksson, 1977; Blanchette, 1991; Fackler et al., 2007).

Often, the sequential white-rot fungi “selectively” degrade lignin and hemicellulose in small

elongated cavities within a wood tissue such that decayed regions are surrounded by tissue that

appears sound (Blanchette 1984). With advancing decay and delignification of the middle

lamella and primary cell wall, the wood sample typically acquires a fibrous texture (Schmidt

2006). Importantly, elicitation of simultaneous or sequential white-rot decay can depend on the

wood that is being degraded, the stage of wood degradation, and the particular fungal strain used

in the study (Messner and Srebonik, 1994). For instance, certain strains of Phanerochaete

chrysosporium (e.g. BKM-F-1767), cause selective decay of deciduous wood samples, while

many other strains cause simultaneous wood decay (Blanchette 1992).

2.5.2 Brown-Rot Fungi

Brown-rot decay is characterized by rapid degradation of cellulose and hemicellulose, and

retention of a modified lignin residue (Yelle et al. 2008). To date, brown-rot fungi were mainly

isolated from coniferous (softwood) trees and represent approximately 7 % of isolated wood-

rotting basidiomycetes (Martínez et al. 2005). Brown-rot decay can cause rapid loss in the

structural strength of construction wood (Green III and Highley 1997), and decayed woods are

typically characterized by reddish brown color and dry, crumbly and brittle consistency

(Martínez et al. 2005). Examples of brown-rot fungi include Fomitopsis lilacino-gilva,

Laetiporus portentosus, Postia placenta, Gloeophyllum trabeum and Serpula lacrymans. In

contrast to the numerous enzymes secreted by white-rot fungi, brown-rot fungi appear to initiate

12

the degradation of wood polysaccharides using Fenton chemistry, whereby diffusible hydroxyl

radicals penetrate the wood surface and depolymerize cellulose and hemicelluloses while leaving

lignin intact (Jensen Jr. et al. 2001).

Studies on the phylogeny and substrate preference of wood decaying fungi suggest that the

brown-rot fungi have evolved multiple times from white-rot fungi. Hibbet et al. (2001) predict

that white-rot, tetrapolar mating systems, and ability to degrade conifers and hardwoods is a

primitive characteristic of wood-degrading fungi, whereas brown-rot, bipolar mating systems,

and exclusive decay of conifers, evolved subsequently from these ancestral characteristics

(Hibbett and Donoghue 2001).

2.6 Lignocellulose Active Enzymes

Given the polymeric and complex composition of lignocellulose, it is not surprising that the

initial attack of this substrate is achieved through the concerted activity of several extracellular

microbial enzymes. Lignocellulose-active enzymes that are produced by white-rot fungi are

particularly valuable for biomass conversion, since they can be used to selectively transform both

lignin and polysaccharides (Kirk and Cullen 1998). The enzymes that contribute to this activity

can be broadly classified as Carbohydrate-Active enzymes (CAZymes) and Fungal Oxidative

Lignin enzymes (FOLymes) (Cantarel et al. 2009; Levasseur et al. 2008). The following is a

summary of the main enzymes involved in these degradation processes.

2.6.1 Carbohydrate Active Enzymes

A sequence-based classification scheme for carbohydrate-active enzymes was developed in

1991, called the CAZy database (CArbohydrate enZYme database) (Cantarel et al. 2009;

Henrissat 1991). At present, this database is comprised of 125 glycoside hydrolase families, 92

glycoside transfer families, 22 polysaccharide lyase families and 16 carbohydrate esterase

families. Glycoside hydrolases hydrolyze the glycosidic bonds between α-linked or β-linked

sugars, using a retaining or inverting mechanism (Davies and Henrissat 1995). Polysaccharide

lyases cleave polysaccharide chains via a β-elimination mechanism resulting in the formation of

a double bond at the newly formed non-reducing end, whereas carbohydrate esterases catalyze

the deacetylation and demethylation of substituted polysaccharides. Efficient degradation of

13

polysaccharides requires cooperation or synergistic interactions between enzymes responsible for

cleaving the different linkages. Significant research has been done to demonstrate and

understand synergy between various isolated enzymes for degradation of microcrystalline

cellulose (Avicel) and commercial xylans (de Vries and Visser 2001). For instance, hydrolysis

of xylan by an Aspergillus xylanase was increased in the presence of accessory enzymes that

catalyze the hydrolysis of xylan side chains (Paszczynski et al. 1988).

2.6.1.1 Cellulases

Hydrolysis of cellulose to glucose requires the activity of endoglucanases (endo-cellulases),

cellobiohydrolases and β-glucosidases. While endo-cellulases hydrolyze glycosidic linkages at

internal positions within cellulose molecules, cellobiohydrolases release cellobiose from the

reducing or non-reducing end of cellulose and β-glucosidases hydrolyze cellobiose to glucose.

These enzymes work synergistically to degrade both amorphous and crystalline cellulose, and

endo-cellulases and cellobiohydrolases are often associated with cellulose-binding modules to

promote their activity on polymeric substrates (Kirk and Cullen 1998). Of the 125 GH families,

fungal cellulases belong to GH families 5, 6, 7, 9, 12, 44, 45, 48, 61 and 74 (Dashtban et al.

2009). In addition to the hydrolytic enzymes, oxidative enzymes also participate in cellulose

degradation (Kirk and Cullen 1998). For example, quinone oxidoreductase (cellobiose

dehydrogenase) reduces quinones and phenoxy radicals in the presence of cellobiose, which is

oxidized to cellobiono-δ-lactone. And cellobiose oxidase uses molecular oxygen to oxidize

cellobiose and longer cello-oligomers to corresponding acids.

2.6.1.2 Hemicellulases

Wood hemicelluloses include xylan, (galacto)glucomannan, and xyloglucan. These

polysaccharides contain a β-1,4-linked sugar backbone that can be acetylated or substituted by

sugar branches (Scheller and Ulvskov 2010). Given the diversity of hemicelluloses, many

glycoside hydrolases, as well as carbohydrate esterases participate in their degradation. For

instance, xylan degradation requires the activity of xylanases, xylosidases, arabinofuranosidases,

galactosidases, deacetylases, glucuronidases, glucuronyl esterases, and feruloyl esterases. Like

cellulases, many of these activities function synergistically and are associated with carbohydrate-

binding modules that promote enzyme activity on polymeric substrates (de Vries et al. 2000;

14

Hervé et al. 2010). So far, fungal hemicellulases were identified in nineteen GH families: 1, 2, 3,

5, 10, 11, 26, 27, 36, 39, 43, 51, 53, 54, 62, 67, 74,115, and 116, and nine CE families: 1, 2, 3, 4,

5, 6, 12, 15 and 16.

2.6.1.3 Fungal Oxidative Lignin Enzymes

Similar to carbohydrate-active enzymes, enzymes involved in lignin catabolism have also been

grouped into sequence-based families and integrated in a database, named the Fungal Oxidative

Lignin enzymes (FOLy) (Levasseur et al. 2008).

Laccases and peroxidases are extracellular, lignolytic enzymes that are key to enzymatic lignin

degradation (Perez et al. 2002; ten Have and Teunissen, 2001). The peroxidases include lignin

peroxidase (LiPs) and manganese-dependent peroxidase (MnP). Both LiP and MnP oxidize

corresponding substrates through two consecutive one-electron oxidation steps with

intermediate cation radical formation (Sanchez 2009). LiP degrades non-phenolic lignin units (up

to 90% of the polymer) whereas MnP generates Mn3+, which acts as a diffusible oxidant of

phenolic or non-phenolic lignin units likely mediated through lipid peroxidation reactions

(Cullen and Kersten 2004; Moen and Hammel, 1994).

Laccases are blue copper oxidases that catalyze the one-electron oxidation of phenolics, aromatic

amines, and other electron-rich substrates with the concomitant reduction of O2 to H2O (d'Souza

et al. 1999). Like Mn(III) chelates, laccases oxidize the phenolic units in lignin to phenoxy

radicals, which can lead to aryl-C cleavage (Kawai et al. 1988). Laccase can also oxidize non-

phenolic substrates in the presence of certain auxiliary substrates such as 2,2´-azino-bis-3-

ethylthiazoline-6- sulfonate (Call and Muncke 1997).

2.7 Investigative Approaches to Improve Lignocellulose Bioconversion

Challenges associated with bioconversion of lignocellulose can be attributed to the distinct

composition and structure of different plant cell walls. For instance, the accessibility of enzymes

to fibre polymers can vary, and the enzymes required to depolymerize cell wall components

depends on the source of fibre. Moreover, since plant cell walls are a composite of different

15

polymers, multiple enzymes are necessary to process fibres, and the enzyme requirement is

likely to change as the fibre is treated.

Methods for increasing the efficiency of biocatalysts include screening for organisms with novel

enzymes, improvement of existing industrial strains, enzyme engineering and development of

enzyme cocktails that are tailored to particular lignocellulosic feedstocks. Considerable progress

has been made concerning the biochemistry of purified, fungal enzymes that participate in wood

degradation. However, the effect of enzyme activities on natural lignocellulosic feedstocks is

limiting. It is also unclear how the profile of enzymes secreted by wood-degrading fungi

depends on the composition of the lignocellulosic feedstock.

It is anticipated that by characterizing the effect of lignocellulose composition on the expression

of lignocellulolytic enzymes by fungi, and enhanced analytical tools to characterize the impact of

these enzymes on the lignocellulose substrate, the development of more efficient bioprocesses

can be developed. The following is a summary of some of the analytical and molecular biology

tools that have recently been used to increase our understanding of lignocellulose bioconversion

by fungi.

2.7.1 Analytical Characterization of Wood Fibre

Degradation patterns associated with advanced stages of white-rot and brown-rot decay can be

identified macroscopically and microscopically. However, a precise analysis of the effect of

degradation requires chemical assessment of the residual fibre. Lignin in wood is traditionally

measured by the Klason method, which is based on total acid hydrolysis of polysaccharides and

gravimetric estimation of the precipitated lignin. This, however, is a time consuming method and

not particularly suited to characterizing partially decayed or modified wood (Martínez et al.

2005). Alternatively, the direct characterization of fibre chemistry facilitates the analysis of

fungal decay, since these methods require minimal sample preparation and quantity (Stenius and

Vuorinen 1999). These methods typically use dry fibre samples and enable either surface

analysis or bulk analysis of fibre chemistry. Examples of analytical techniques for fibre surface

analysis include X-ray photoelectron spectroscopy (XPS) and secondary ion mass spectroscopy

(SIMS), as well as atomic force microscopy (AFM) and attenuated total reflectance (ATR).

16

Common analytical techniques for bulk analysis of fibre chemistry include nuclear magnetic

resonance (NMR), Fourier transform infrared (FTIR), Raman and ultraviolet/visible (UV/Vis)

spectroscopy, as well as pyrolysis gas chromatography (Py-GC) (Stenius and Vuorinen 1999).

Additionally, microscopic methods, including transmission electron microscopy (TEM), have

been used to detect morphological changes in wood cell walls that occur during wood decay. In

the present thesis, FTIR, ToF-SIMS and TEM were used to analyze residual fibre composition

and anatomy, and so these methods will be further described below.

ToF-SIMS is a powerful technique that provides chemical information about a solid surface by

ionizing the sample surface with a primary ion beam and determining the mass-to-charge ratio of

the secondary ion fragments (Benninghoven 1994). ToF-SIMS has been used to characterize

pulp fibre (Rowell 2005), as well as spatially resolve different components on the surface of

wood, providing compositional information across cell walls (Jung et al. 2010). Further, Kangas

et al., (2004) compared different types of fines separated from thermomechanical pulp based on

the chemical composition as determined by ToF-SIMS (Kangas and Kleen 2004). In another

study, Saito et al., (2008) used ToF-SIMS to investigate the distribution of trace elements and

lignin in contiguous rings to highlight differences between sapwood and heartwood of

Chamaecyparis obtusa. ToF-SIMS imaging was also used to conclude that the ray parenchyma

cells play a role in defining the state of elements during the transition from sapwood to

heartwood (Saito et al. 2008). Another application of this technology has been to determine the

surface distribution of lignin, carbohydrates and metal for aspen and spruce wood tissue-sections

(Tokareva et al. 2007). Additionally, effects of processing techniques like refining and pulping

on fibre surfaces have also been determined using ToF-SIMS (Fardim and Duran 2003).

Importantly, in a recent publication, researchers from our own group conducted a detailed

characterization of wood samples to improve ToF-SIMs analysis of lignocellulosic materials by

substantially increasing the annotation of corresponding secondary ions (Goacher et al. 2011).

FTIR has been used to distinguish angiosperms and gymnosperms (Schultz et al. 1985),

determine the lignin, glucose and xylose content of different woods (McDonald 1976), and

characterize the variation in fibre chemistry of natural and transgenic plants (Mouille et al.

2006). FTIR has also been used to detect compositional changes in cell wall polymers as a result

17

of chemical extraction (McCann et al. 1992). FTIR is especially suited to rapid chemical

characterization of small quantities of wood with minimum sample preparation and has been

used to analyze changes in wood fibre chemistry after degradation by different brown and white-

rot fungi. The recently published investigation of spruce degradation by the brown-rot fungi G.

trabeum and P. placenta, describes for the first time, the application of FTIR imaging

microscopy coupled with multivariate analysis (Fackler et al. 2010). In their study, Fackler et al

(2010) concluded that the brown-rot decay was most significant in the outer cell wall regions in

the early stages of decay. This was accompanied by loss in glycosidic bonds and pectic

substances.

Standard wavelengths that describe chemical bonds characteristic of lignin and cellulose have

been reported (Faix 1992; Faix et al. 1991; Kacurakova et al. 2000; Kotilainena et al. 2000;

Labbe et al. 2005; Pandey and Nagveni 2007; Pandey and Pitman 2003; Schwanninger et al.

2004). Commonly, visual comparison of spectra and lignin to carbohydrate peak area ratios have

been used to characterize the activity of brown-rot and white-rot fungi on wood fibre (Pandey

and Nagveni 2007). However, the utilization of chemometric tools would allow more efficient,

complete and un-biased assessment of large sets of FTIR data.

Typically, delignification is not uniformly distributed throughout wood samples, and so bulk

chemical methods to characterize the mode of fungal decay (i.e., selective or simultaneous) can

be misleading. Conventional SEM and TEM have been used as effective tools to confirm the

ability of fungi to degrade wood cell walls, and specific morphological features observed during

selective and simultaneous wood decay have been established. Simultaneous white-rot results in

thinning of wood cell walls along the circumference of the lumen, while selective white-rot leads

to gradual delignification of middle lamella and cell walls without any observable changes in cell

morphology (Blanchette and Reid 1986; Blanchette et al. 1985; Ruel et al. 1981; Ruel et al.

1986). Visualization of lignin by staining wood samples with either Bromine or KMnO4 has

greatly enhanced the capability of TEM to decipher the type and extent of wood decay caused by

microbes (Fromm et al. 2003; Saka et al. 1978). Notably, an extensive review on the utility of

microscopic techniques to characterize wood biodegradation has been previously published

(Daniel 1994; Daniel 2003).

18

2.7.2 Biochemical Characterization of Wood-degrading Fungi

2.7.2.1. Techniques to Measure Fungal Growth.

Monitoring fungal growth is often used to assess degradation potential. Because fungal

mycelium is difficult to separate from solid substrates like wood, rapid methods for measuring

fungal growth are by necessity indirect, and measure either constituents of fungal cells or factors

relating to their metabolic processes (Gottlieb and van Etten 1964). These indirect methods

however, have their own limitations. While visual inspection is the simplest method to assess

fungal growth, poor visibility of hyphae limits its application for detecting small differences.

Additionally, hyphae can be obscured by substrate and living mycelium can not be differentiated

from the dead (Boyle and Kropp 1992). In their study to develop methods to measure growth of

filamentous fungi on wood, Boyle et al., (1992) compared visual inspection, substrate dry weight

loss, rate of fluroescein diacetate hydrolysis, extractable protein content and chitin content of the

colonized substrate. The authors concluded that each assay measured a different aspect of the

growth and chitin gave the best measure of biomass.

There are several methods available for determination of chitin content (Ekblad and Nisholm

1996; Zamani et al. 2008). An assay of glucosamine based on acidic hydrolysis is considered

sensitive and to have low susceptibility to interfering compounds. Hydrolysis of chitin in 6 M

HCl for 16 hours at 80˚C is approximately 90 % complete and yields acetyl and glucosamine

residues. Alternative methods include freeze-drying and evaporation under reduced pressure,

which are comparatively arduous. Further, alkaline hydrolysis typically releases only 50 % of

glucosamine residues and requires numerous washings to remove excess base. However as a

cautionary note, the slight release of other sugars by HCl used in this assay can create minor

interferences during the color reaction, leading to slight over estimations of chitin concentration.

This disadvantage could be avoided by using alkaline hydrolysis, but it appears to be largely

compensated by simplicity of this method (Plassard et al. 1982).

19

2.7.2.2. Techniques to Characterize the Secreted Enzymes that Participate in Lignocellulose

Degradation.

Traditionally, biochemical assays have been used to characterize lignocellulose degrading

enzymes. These include assays that use model and often chromogenic substrates to determine

specific cellulolytic, hemicellulolytic and ligninolytic activities (Himmel et al. 1997).

Biochemical characterization of enzymes through activity assays requires previous knowledge of

expected activities, and is often limited by the availability of commercial substrates. As a result,

biochemical characterization of culture supernatant through conventional enzyme activity studies

limits the possibility to discover new enzymes and proteins, as well as novel protein

combinations that promote lignocellulose degradation.

Phanerochaete chrysosporium was the first lignocellulose degrading fungus to have its genome

sequenced (Martinez et al. 2004). Since then, the genomes of over 25 basidiomycetes have been

sequenced (http://genome.jgi-psf.org/). With the sequencing of genomes from lignocellulose

degrading microorganisms, advances in instrumentation for protein separation, and protein

identification by mass-spectrometry, new proteins that contribute to lignocellulose

transformation can now be identified (Vanden Wymelenberg et al. 2005b). For example,

proteomic analyses of Gloeophyllum trabeum grown on spruce sawdust identified secreted

enzymes and proteins that were not distinguished by previous biochemical assays of culture

supernatant (Abbas et al. 2005; Varela et al. 2003). Notably, less than 15 % of the proteins

secreted by G. trabeum had been previously identified by biochemical or sequence data. This

potential of genomic and proteomic approaches to identify new enzymes is further highlighted

by the systematic analysis of the P. placenta genome, transcriptome and proteome (Martinez et

al. 2009). While oxidases predicted to generate hemicellulases and extracellular Fe2+ and H2O2

were identified, perhaps most striking was the relatively low number of cellulolytic glycoside

hydrolases (GH) encoded by the P. placenta genome, and apparent lack of GHs containing

family 1 carbohydrate-binding modules.

Genomic and proteomic analyses of the white-rot basidiomycete Phanerochaete chrysosporium

have revealed many glycoside hydrolases and other carbohydrate-active proteins that were not

differentiated by previous biochemical assays of culture supernatant (Vanden Wymelenberg et

20

al. 2006; Vanden Wymelenberg et al. 2005a). Notably, more than 65% of the proteins secreted

by P. chrysoporium grown on crystalline cellulose corresponded to previously uncharacterized

glycoside hydrolases, or proteins with unknown function. Genomic analyses indicate that P.

chrysosporium is distinguished from other eukaryotic genomes by encoding more than twice as

many carbohydrate hydrolyzing enzymes than synthesis enzymes. Since less than 20 % of

predicted glycoside hydrolases encoded by P. chrysosporium were characterized before the

genome sequence was published, this example clearly illustrates the potential for genome

sequencing to increase the arsenal of industrially relevant enzymes (Martinez et al. 2004; Vanden

Wymelenberg et al. 2006).

As might be predicted from biomass composition, the characterization of extracellular proteins

from P. chrysosporium grown on cellulose and wood substrates consistently detect the

expression of GH3 β-glycosidases, GH6 and GH7 cellobiohydrolases, GH10 xylanases, and

GH12 endoglucanases; polygalacturonases, α-galactosidase, aspartic acid proteases, and GH88

d-4,5-unsaturated glucuronyl hydrolases are also frequently detected (Abbas et al. 2005;

Ravalason et al. 2008; Sato et al. 2007; Vanden Wymelenberg et al. 2009; Vanden Wymelenberg

et al. 2005a). Notably, lignin peroxidases and manganese peroxidases were often missed in

proteomic analyses of P. chrysosporium grown on cellulose and wood substrates, although

detectable by RT-PCR (Sato et al. 2009; Vanden Wymelenberg et al. 2009). Proteomic analyses

of P. chrysosporium have also revealed substrate dependent expression profiles of hypothetical

proteins, identified secreted proteins not predicted from initial bioinformatics analyses, and have

improved gene models for this organism (Vanden Wymelenberg et al. 2006). More recently,

pyrosequencing technology was used to sequence an expression library from P. chrysosporium

grown on oak (Sato et al. 2009). These data were consistent with previously reported proteomic

analyses, and underscore the significance of gene expression levels in microbial responses to

substrate composition.

For several years, a main focus in lignocellulose degradation research has been to isolate and

identify organisms that are hyper-producers of stable lignocellulolytic enzymes. Now, the

application of genomic and proteomic techniques can be used to reveal new enzymes from un-

cultured organisms, or enzymes that are difficult to detect using conventional biochemical

21

assays. The challenge, however, is to couple these advanced molecular techniques with

analytical methods that characterize the effects of new enzymes and enzyme mixtures on

complex, industrially relevant substrates and predict which proteins directly participate in

lignocellulose utilization.

Chapter 2 References

Abbas A, Koc H, Liu F, Tien M (2005) Fungal degradation of wood: initial proteomic analysis of extracellular proteins of Phanerochaete chrysosporium grown on oak substrate. Current Genetics 47 (1):49-56.

Ander, P. and Eriksson, K.E. (1977) Selective degradation of wood components by white-rot fungi. Physiol. Plant. 41:239-248.

Bayer EA, Lamed R, White BA, Flint HJ.(2008) From cellulosomes to cellulosomics. The Chemical Record. (6):364-77.

Béguin P, Aubert J-P (1994) The biological degradation of cellulose. FEMS Microbiology Reviews 13 (1):25-58

Benninghoven A (1994) Chemical analysis of inorganic and organic surfaces and thin films by static Time-of-Flight Secondary Ion Mass Spectrometry (TOF-SIMS). Angewandte Chemie International Edition in English 33 (10):1023-1043.

Blanchette RA (1984) Screening wood decayed by white rot fungi for preferential lignin degradation. Applied and Environmental Microbiology. 48 (3):647-653

Blanchette RA, Otjen L, Effland MJ, Eslyn WE (1985) Changes in structural and chemical components of wood delignified by fungi. Wood Science and Technology 19:35-46

Blanchette R, Reid ID (1986) Ultrastructural aspects of wood delignification by Phlebia

(Merulius) tremellosus. Appl Environ Microbiol 52:239-245 Blanchette RA (1991) Delignification by wood-decay fungi. Annual Reviews of Phytopathology

29:381-398 Blanchette, R.A., Burnes, T.A., Eerdmans, M.M. and Akhtar, M. (1992) Evaluating isolates of

Phanerochaete chrysosporium and Cerzporiopsis subz:errnispora for use in biological pulping processes. Holzforschung 46, 109-115. 645-662.

Blanchette, R.A, Obst JR, Timell, TE (1994) Biodegradation of comparession wood and tension

wood by white and brown-rot fungi. Holzforschung 48: 34-42.

22

Boyle DC, Kropp BR (1992) Development and comparison of methods for measuring growth of filamentous fungi on wood. Canadian Journal of Microbiology 38:1053-1060

Call HP, Muncke I (1997) History, overview and applications of mediated lignolytic systems, especially laccase-mediator systems (lignozyme(R)-process). Journal of Biotechnology 53:163-202

Cantarel BL, Coutinho PM, Rancurel C, Bernard T, Lombard V, Henrissat B (2009) The Carbohydrate-Active EnZymes database (CAZy): an expert resource for Glycogenomics. Nucleic Acids Research 37 (suppl 1):D233-D238.

Claus H (2004) Laccases: structure, reactions, distribution. Micron 35 (1-2):93-96

Crawford DL.(1978) Lignocellulose decomposition by selected Streptomyces strains.Appl Environ Microbiol. 35(6):1041-5.

Cullen D, Kersten PJ (2004) Enzymology and molecular biology of lignin degradation. In: Brambl R, Marzluf GA (eds) The Mycota III Biochemistry and Molecular Biology. Springer-Verlag, Berlin-Heidelberg,

d'Souza TM, Merritt CS, Reddy AC (1999) Lignin modifying enzymes of the white rot basidiomycete Ganoderma lucidum. Applied and Environmental Microbiology 65 (12):5307-5313

Daniel G (1994) Use of electron microscopy for aiding our understanding of wood biodegradation. FEMS Microbiology Reviews 13:199-233

Daniel, G. (2003) Microview of wood under degradation by bacteria and fungi.In: Wood

Deterioration and Degradation. Advances in Our Changing World. B. Goodell, D. D. Nicholas and T. P. Schultz, ACS Symposium Series. 845:34 - 72

Dashtban M, Schraft H, Qin W (2009) Fungal bioconversion of lignocellulosic residues - opportunities and perspectives. International Journal of Biological Sciences 5:578-595

Davies G, Henrissat B (1995) Structures and mechanisms of glycosyl hydrolases. Structure 3:853-859

de Vries RP, Kester HCM, Poulsen CH, Benen JAE, Visser J (2000) Synergy between enzymes from Aspergillus involved in the degradation of plant cell wall polysaccharides. Carbohydrate Research 327 (4):401-410

de Vries RP, Visser J (2001) Aspergillus enzymes involved in degradation of plant cell wall polysaccharides. Microbiology and Molecular Biology Reviews 65 (4):497-522

Ding SY, Himmel ME. (2006) The maize primary cell wall microfibril: a new model derived from direct visualization. Journal of Agriculture and Food Chemistry.54(3):597-606.

23

Ekblad A, Nisholm T (1996) Determination of chitin in fungi and mycorrhizal roots by an improved HPLC analysis of glucosamine. Plant and Soil 178:29-35

Fackler K, Stevanic J, Ters T, Hinterstoisser B, Schwanninger M, Salmén L (2010) Localisation and characterisation of incipient brown-rot decay within spruce wood cell walls using FTIR imaging microscopy. Enzyme and Microbial Technology 47 (6):257-267

Faix O (ed) (1992) Fourier Transform Infrared Spectroscopy. Methods in lignin chemistry. Springer, Berlin-Heidelberg

Faix O, Bremer J, Schmidit O, Stevanovic T (1991) Monitoring of chemical changes in white-rot degraded beech wood by pyrolysis-gas chromatography and Fourier transform infrared spectroscopy. Journal of Analytical and Applied Pyrolysis 21:147-162

Fardim P, Duran N (2003) Modification of fibre surfaces during pulping and refining as analysed by SEM, XPS, and TOF-SIMs. Colloids and Surfaces A; Physiochemical Engineering Aspects 223:263-276

34r cc J, Rockel B, Lautner S, Windeisen E, Wanner G (2003) Lignin distribution in wood cell walls determined by TEM and backscattered SEM techniques. Journal of Structural Biology 143 (1):77-84

Global Forest Resources Assessment (2005). Food and Agriculture Organization of the United Nations, Rome

Goacher R, Jeremic D, Master ER (2011) Expanding the library of secondary ions that distinguish lignin and polysaccharides in ToF-SIMS analysis of wood. Analytical Chemistry. 83(3):804-12

Gottlieb D, Van Etten JL (1964) Biochemical changes during the growth of fungi. Journal of Bacteriology 88 (1):114-121

Green III F, Highley TL (1997) Mechanism of brown-rot decay: Paradigm or Paradox. International Biodeterioration and Biodegradation 39 (2-3):113-124

Hammel KE (1997) Fungal Degradation of Lignin. In: Cadisch G, Gillier KE (eds) Driven by Nature: Plant litter quality and decomposition. CAB International, pp 33-45

Hayashi T, Kaida R (2011) Hemicelluloses as recalcitrant components for saccharification in wood. Routes to Cellulosic Ethanol 2:45-52

Henrissat B (1991) A classification of glycosyl hydrolases based on amino acid sequence similarities. Biochem J 280 (2):309-316

Hervé C, Rogowski A, Blake A, Marcus S, Gilbert H, Knox J (2010) Carbohydrate-binding modules promote the enzymatic deconstruction of intact plant cell walls by targeting and proximity effects. Proc Natl Acad Sci U S A 107 (34):15293-15298

24

Hibbett DS, Donoghue MJ (2001) Analysis of character correlations among wood decay mechanisms, mating systems, and substrate ranges in homobasidiomycetes. Systems Biology 50:215-242

Himmel ME, Adney William S, Baker John O, Elander R, McMillan James D, Nieves Rafael A, Sheehan John J, Thomas Steven R, Vinzant Todd B, Zhang M (1997) Advanced Bioethanol Production Technologies: A Perspective. In: Fuels and Chemicals from Biomass, vol 666. ACS Symposium Series. American Chemical Society, pp 2-45.

Howard RL, Abotsi E, Jansen van Rensburg EL, Howard S (2003) Lignocellulose biotechnology: issue of bioconversion and enzyme production. African Journal of Biotechnology 2 (12):602-619

Humphreys JM, Chapple C (2002) Rewriting the lignin roadmap. Current Opinion in Plant Biology 5 (3):224-229

Jensen Jr. KA, Houtman CJ, Ryan ZC, Hammel KE (2001) Pathways for extracellular Fenton chemistry in the brown rot basidiomycete Gloeophyllum trabeum. Applied and Environmental Microbiology 67 (6):2705-2711

Jung S, Foston M, Sullards CM, Ragauskas AJ (2010) Surface characterization of dilute acid pretreated Populus deltoides by TOF-SIMS. Energy Fuels 24:1347-1357

Kacurakova M, Capek P, Sasinkova V, Wellner N, Ebringerova A (2000) FTIR study of plan cell wall model compounds: pectic polysacchardes and hemicelluloses. Carbohydrate polymers 43:195-203

Kangas H, Kleen M (2004) Surface chemical and morphological properties of mechanical pulp fines. Nordic Pulp and Paper Research Journal 19 (2):191-199

Kawai S, Umezawa T, Higuchi T (1988) Degradation mechanisms of phenolic P-1 lignin substructure model compounds by laccase of Coriolus versicolor. Archives of Biochemistry and Biophysics 262 (1):99-110

Keegstra K (2010) Plant Cell Walls. Plant Physiol 154 (2):483-486. doi:10.1104/pp.110.161240

Kirk TK, Cullen D (1998) Enzymology and molecular genetics of wood degradation by white-rot fungi. In: Young RA, Akhtar M (eds) Environmentally Friendly Technologies for the Pulp and Paper lndustry. John Wiley & Sons, Inc,

Kotilainena RA, Toivanena T-J, Aléna RJ (2000) FTIR monitoring of chemical changes in softwood during heating. Journal of Wood Chemistry and Technology 20 (3):307-320

Labbe N, Rials TG, Kelley SS, Cheng Z-M, Kim J-Y, Li Y (2005) FTIR imaging and pyrolysis-molecular beam mass spectrometry:new tools to investigate wood tissues. Wood Science and Technology 39:61-77

25

Lerouxel O, Cavalier DM, Liepman AH, Keegstra K (2006) Biosynthesis of plant cell wall polysaccharides - a complex process. Current Opinion in Plant Biology 9 (6):621-630

Levasseur A, Piumi F, Coutinho P, Rancurel C, Asther M, Delattre M, Henrissat B, Pontarotti P, Asther M, Record E (2008) FOLy: an integrated database for the classification and functional annotation of fungal oxidoreductases potentially involved in the degradation of lignin and related aromatic compounds. Fungal Genetic Biology 45 (5):638-645

Liese W (1970) Ultrastructural aspects of woody tissue disintegration as reported in wood and tree fungi by Schmidt Olaf; 2006. Ann Rev Phytopath 8:231-258

Lynd LR, Laser MS, Bransby D, Dale BE, Davison B, Hamilton R, Himmel M, Keller M, McMillan JD, Sheehan J, Wyman CE (2008) How biotech can transform biofuels. Nature Biotechnology 26 (2):169-172

Lynd LR, Weimer PJ, van Zyl WH, Pretorius IS (2002) Microbial cellulose utilization: fundamentals and biotechnology. Microbiol Mol Biol Rev 66 (3):506-577.

Lynd LR, Wyman CE, Gerngross T (1999) Biocommodity Engineering. Biotechnology Progress 15 (5):777-793

Magdalena G (2008) An evaluation of U.S. and Canadian lignocellulosic biomass supply and ethanol production costs. University of Toronto (dissertation)

Malherbe S, Cloete TE (2002) Lignocellulose biodegradation: Fundamentals and applications. Reviews in Environmental Science and Biotechnology 1 (2):105-114.

Martínez ÁT, Speranza M, Ruiz-Dueñas FJ, Ferreira P, Camarero S, Guillén F, Martínez MJ, Gutiérrez A, del Río JC (2005) Biodegradation of lignocellulosics: microbial, chemical, and enzymatic aspects of the fungal attack of lignin. International Microbiology 8:195-204

Martinez D, Challacombe J, Morgenstern I, Hibbett D, Schmoll M, Kubicek CP, Ferreira P, Ruiz-Duenas FJ, Martinez AT, Kersten P, Hammel KE, Vanden Wymelenberg A, Gaskell J, Lindquist E, Sabat G, Splinter BonDurant S, Larrondo LF, Canessa P, Vicuna R, Yadav J, Doddapaneni H, Subramanian V, Pisabarro AG, LavÃn JL, Oguiza JA, Master E, Henrissat B, Coutinho PM, Harris P, Magnuson JK, Baker SE, Bruno K, Kenealy W, Hoegger PJ, KÃes U, Ramaiya P, Lucas S, Salamov A, Shapiro H, Tu H, Chee CL, Misra M, Xie G, Teter S, Yaver D, James T, Mokrejs M, Pospisek M, Grigoriev IV, Brettin T, Rokhsar D, Berka R, Cullen D (2009) Genome, transcriptome, and secretome analysis of wood decay fungus Postia placenta supports unique mechanisms of lignocellulose conversion. Proceedings of the National Academy of Sciences 106 (6):1954-1959

Martinez D, Larrondo LF, Putnam N, Sollewijn Gelpke MD, Huang K, Chapman J, Helfenbein KG, Ramaiya P, Detter JC, Larimer F, Coutinho PM, Henrissat B, Berka R, Cullen D, Rokhsar D (2004) Genome sequence of the lignocellulose degrading fungus Phanerochaete chrysosporium strain RP78. Nature Biotechnology 22:695-700

26

McCann MC, Carpita NC (2008) Designing the deconstruction of plant cell walls. Current Opinion in Plant Biology 11 (3):314-320

McCann MC, Hammouri M, Wilson R, Belton P, Roberts K (1992) Fourier Transform Infrared Microspectroscopy is a new way to look at plant cell walls. Plant Physiol 100 (4):1940-1947.

McDonald RS (1976) Infrared spectrometry. Analytical Chemistry 48 (5):196-216

Moen MA, Hammel KE (1994) Lipid peroxidation by the manganese peroxidase of Phanerochaete chrysosporium is the basis for phenanthrene oxidation by the inact fungus. Applied and Environmental Microbiology. 60: 1956-61.

Mouille G, Witucka-Wall H, Bruyant M-P, Loudet O, Pelletier S, Rihouey C, Lerouxel O, Lerouge P, Hofte H, Pauly M (2006) Quantitative trait loci analysis of primary cell wall composition in Arabidopsis. Plant Physiol 141 (3):1035-1044.

Pandey KK, Nagveni HC (2007) Rapid characterisation of brown and white rot degraded chir pine and rubberwood by FTIR spectroscopy European Journal of Wood and Wood Products 65 (6):477-481

Pandey KK, Pitman AJ (2003) FTIR studies of the changes in wood chemistry following decay by brown-rot and white-rot fungi. International Biodeterioration and Biodegradation 52:151-160

Paszczynski A, Crawford RL, Blanchette RA (1988) Delignification of wood chips and pulps by using natural and synthetic porphyrins: models of fungal Decay. Applied and Environmental Microbiology 54 (1):62-68

Perez J, Munoz-Durado J, de la Rubia T, Martinez J (2002) Biodegradation and biological treatment of cellulose, hemicellulose and lignin : an overview. International Microbiology 5:53-63

Pettersen RC (1984) The Chemical Composition of Wood. Advances in Chemistry Series 207:57-126

Plassard CS, Mousain DG, Salsaca LE (1982) Estimation of mycelial growth of basidiomycetes by means of chitin determination. Phytochemistry 21 (2):345-348

Rabinovich ML, Melnik MS, Bolobova AV (2002) Microbial Cellulases- A Review. Applied Biochemical Microbiology 38:305-321

Ragauskas AJ, Williams CK, Davison BH, Britovsek G, Cairney J, Eckert CA, Frederick WJ Jr, Hallett JP, Leak DJ, Liotta CL, Mielenz JR, Murphy R, Templer R, Tschaplinski T. (2006) The path forward for biofuels and biomaterials. Science. 311(5760):484-9.

Ravalason H, Jan G, Mollé D, Pasco M, Coutinho P, Lapierre C, Pollet B, Bertaud F, Petit-Conil M, Grisel S, Sigoillot J-C, Asther M, Herpoël-Gimbert I (2008) Secretome analysis of

27

Phanerochaete chrysosporium strain CIRM-BRFM41 grown on softwood. Applied Microbiology and Biotechnology 80 (4):719-733.

Rowell RM (2005) Handbook of wood chemistry and wood composites. In. CRC Press,

Ruel K, Barnoud F, Eriksson K-E (1981) Micromorphological and ultrastructural aspects of spruce wood degradation by wild-type Sporotrichum pulverulentum and its cellulase-less mutant Cel 44. Holzforschung 25:157-171

Ruel K, Joseleau J-P, Johnsrud SC, Eriksson K-E (1986) Ultrastructural aspects of birch degradation by Phanerochaete chrysosporium and two of its cellulase deficient mutants. Holzforschung 40:5-9

Saito K, Mitsutani T, Imai T, Matsushita Y, Yamamoto A, Fukushima K (2008) Chemical differences between sapwood and heartwood of Chamaecyparis obtusa detected by ToF-SIMS. Applied Surface Science 255 (4):1088-1091

Saka S, Thomas RJ, Gratzl JS (1978) Lignin distribution: Determination by energy-dispersive analysis of X-rays. Tappi 61 (1):73-76

Sanchez C (2009) Lignocellulosic Residues - Biodegradation and bioconversion by fungi. Biotechnology Advances 27:185-194

Sato S, Liu F, Hasan K, Tien M (2007) Expression analysis of extracellular proteins from Phanerochaete chrysosporium grown on different liquid and solid substrates. Microbiology 153:3023-3033

Sato S, Feltus F, Iyer P, Tien M (2009) The first genome-level transcriptome of the wood-degrading fungus Phanerochaete chrysosporium grown on red oak. Current Genetics 55 (3):273-286.

Scheller HV, Ulvskov P (2010) Hemicelluloses. Annual Review of Plant Biology 61 (1):263-289.

Schmidit O (2006) Wood and Tree Fungi. Biology, Damage, Protection and Use. Springer-Verlag, New York

Schultz TP, Templeton MC, McGinnis GD (1985) Rapid determination of lignocellulose by diffuse reflectance Fourier transform infrared spectrometry. Analytical Chemistry 57 (14):2867-2869.

Schwanninger, M., Rodrigues, J., Pereira, H., Hinterstoisser, B. (2004) Effects of short-time vibratory ball milling on the shape of FTIR spectra of wood and cellulose. Vibrational Spectroscopy 36:23-40

Showalter AM (2001) Introduction: plant cell wall proteins. Cellular and molecular life sciences 58 (10):1361-1362

28

Sjostrom E (1983) Wood Chemistry, Second Edition: Fundamentals and Applications. Academic Press,

Stenius P, Vuorinen T (1999) Direct characterization of chemical properties of fibres. In: Sjostrom E, Alen R (eds) Analytical Methods in Wood Chemistry, Pulping, and Papermaking, vol 2. Springer -Verlag,

Taylor NG, Laurie S, Turner SR (2000) Multiple cellulose synthase catalytic subunits are required for cellulose synthesis in Arabidopsis. Plant Cell 12 (12):2529-2540.

ten Have R, Teunissen P (2001) Oxidative mechanisms involved in lignin degradation by white-rot fungi. Chemical Reviews 11:3397-3414

Tokareva EN, Fardim P, Pranovich aV, Fargerholm H-P, Daniel G, Holmbom B (2007) Imaging of wood tissue by TOF-SIMS: Critical evaluation and development of sample preparation techniques. Applied Surface Science 253:7569-7577

Vanden Wymelenberg VA, Sabat G, Martinez D, Rajangam AS, Teeri TT, Gaskell J, K KPJ, Cullen D (2005a) The Phanerochaete chrysosporium secretome: Database predictions and initial mass spectrometry peptide identifications in cellulose-grown medium Journal of biotechnology 118 (1):17-34

Vanden Wymelenberg VA, Sabat G, Martinez D, Rajangam AS, Teeri TT, Gaskell J, Kersten PJ, Cullen D (2005b) The Phanerochaete chrysosporium secretome: database predictions and initial mass spectrometry peptide identifications in cellulose-grown medium. Journal of Biotechnology 118:17-34

Vanden Wymelenberg AV, Minges P, Sabat G, Martinez D, Aerts A, Salamov A, Grigoriev I, Shapiro H, Putnam N, Belinky P, Dosoretz C, Gaskell J, Kersten P, Cullen D (2006) Computational analysis of the Phanerochaete chrysosporium v2.0 genome database and mass spectrometry identification of peptides in ligninolytic cultures reveal complex mixtures of secreted proteins. Fungal Genetics and Biology 43 (5):343-356

Vanden Wymelenberg VA, Gaskell J, Mozuch M, Kersten P, Sabat G, Martinez D, Cullen D (2009) Transcriptome and secretome analysis of Phanerochaete chrysosporium reveal complex patterns of gene expression. Appl Environ Microbiol. 75(12): 4058-68.

Varela E, Mester T, Tien M (2003) Culture conditions affecting biodegradation components of the brown-rot fungus Gloeophyllum trabeum. Archives of Microbiology 180 (4):251-256.

Walker LP, Wilson DB (1991) Enzymatic hydrolysis of cellulose: An overview. Bioresource Technology 36 (1):3-14

Watanabe Y, Shinzato N, Fukatsu T. (2003) Isolation of actinomycetes from termites' guts. Biosci Biotechnol Biochem. 67(8):1797-801.

Yelle, D. J., J. Ralph, F. C. Lu, and K. E. Hammel (2008) Evidence for cleavage of lignin by a

29

brown rot basidiomycete. Environmental Microbiology. 10:1844-1849.

Zamani A, Jeihanipour AJ, Edebo L, Niklasson C, Taherzadeh MJ (2008) Determination of glucosamine and N-acetyl glucosamine in fungal cell walls. Journal of Agricultural and Food Chemistry 56 (18):8314-8318

Zhang YH (2008) Reviving the carbohydrate economy via multi-product lignocellulose biorefineries. Journal of Industrial Microbiology Biotechnology 35 (5):367-375.

30

Chapter 3 : Effect of Cultivation Conditions on the Expression of Cellulolytic Activity by Phanerochaete species

P. chrysosporium and P. carnosa

3.1 Abstract

This chapter describes the first biochemical and growth study of the white-rot fungus

Phanerochaete carnosa. A range of simple compounds (glucose, glycerol, cellulose) and

lignocellulosic substrates (coniferous softwoods: pine, fir, spruce; and deciduous hardwoods:

maple and poplar) were validated as potential substrates for P. carnosa. In-depth characterization

of P. carnosa cultivation was pursued using microcrystalline cellulose. Culture conditions and

activity studies indicate that P. carnosa grows optimally at temperatures between 27ºC and 30ºC,

and confirmed that lignocellulolytic activity is present in extracellular filtrates of cultures grown

on cellulose. Low nitrogen (2.5 mM), stationary cultures led to highest growth measurements

(as determined by chitin content) and final protein concentration in culture supernatant.

However, while high nitrogen condition (25 mM) was consistently correlated to low specific

cellulase and MnP activity, shaking at 150 rpm did not appear to significantly effect the secretion

of lignocellulose-degrading enzyme in P. carnosa.

3.2 Introduction

Efficient biotransformation of wood components is critical for the production of renewable

energy and chemicals from this abundantly available resource. Fungi are the major decomposers

of higher plants, and the largest group of wood-degrading fungi is the Basidiomycetes. The terms

brown-rot and white-rot have been used to classify two types of decay by these fungi. White-rot

fungi enzymatically degrade all major components of wood (i.e., hemicellulose, cellulose and

lignin). By contrast, brown-rot fungi typically initiate rapid degradation of cellulose and

hemicellulose through Fenton chemistry, and modify the residual lignin without substantial

depolymerization. A different type of decay that is similar in macroscopic appearance to brown-

rotted wood is caused by soft-rot fungi. This form of wood decay is usually important in

extremely wet environments, or environments that undergo frequent wet-dry cycles, where more

aggressive wood degrading fungi cannot survive. Of all these, the white-rot fungi are known to

cause most wide-spread and significant decay (Eriksson 1990; Otjen et al. 1987).

31

Within the white-rot fungi, there are selective and simultaneous degraders, that is, fungi that

preferentially degrade lignin before attacking cellulose and hemicellulose (e.g. Ceriporiopsis

subvermispora) and those that degrade all components at similar rates (e.g. Trametes versicolor),

respectively. The fungi that can selectively remove lignin without extensive cellulose

degradation are of prime interest specifically for the pulping process as well as bioremediation

due to their established ability to degrade the recalcitrant lignin.

Phanerochaete chrysosporium is commonly used to study the production and regulation of

lignocellulolytic enzymes by white-rot fungi. This model white-rot fungus belongs to the wood

inhabiting Agaricomycete class of the Basidiomycota phylum. There are over 90 species in this

genus that are currently described all over the world (Wu et al. 2010). P. chrysosporium was

first isolated from wood chip piles and was given the name Chrysosporium lignorum (Bergman

and Nilsson 1966). This name was later changed to Sporotichum pulverulentum and finally the

present name was assigned to it based on its perfect state (the stage at which spores are formed

after nuclear fusion). The fungus is characterized as thermotolerant and has an optimum growth

rate at around 38-39˚C (Eriksson 1978).

Phanerochaete carnosa is a comparatively poorly characterized member of the family

Phanerochaete sensu strictum, and was previously isolated from balsam fir (Abies balsamea) as

well as other softwoods (de Koker et al. 2003). The phylogenetic analysis reported by de Koker

et al. (2003) groups P. carnosa with P. arizonica, P. burtii and P. sanguinea, and reveals that the

ribosomal ITS sequences of P. carnosa and P. chrysosporium share 89% identity. However, the

isolation of P. carnosa being primarily from softwoods is different from other well-studied

white-rot fungi like P. chrysosporium, which were primarily isolated from hardwoods.

Accordingly, it is possible that P. carnosa has evolved to express unique lignocellulolytic

activities that effectively transform softwoods, which would constitute a valuable source of

enzymes, given the typical recalcitrance and abundance of this biomass feedstock.

Since the cultivation and biochemical characterization of P. carnosa has not been described, the

primary aim of this first study was to establish growth conditions that induce the expression of

lignocellulolytic enzyme by this organism. This information will provide a basis for further

investigation of lignocellulose conversion by P. carnosa to utilize and enhance its potentially

32

unique lignocellulolytic activity. Moreover, by comparing the effect of cultivation conditions on

the production of lignocellulolytic activity by P. carnosa and P. chrysosporium, differences in

enzyme activities and their regulation can be evaluated.

Given the phylogenetic similarity of P. carnosa and P. chrysosporium, growth conditions that

promote lignocellulolytic activity in P. chrysosporium were predicted to also promote

lignocellulolytic activity by P. carnosa. Therefore, in order to establish the growth conditions

and investigate the lignocellulolytic enzyme profiles of P. carnosa in the present study,

published literature on P. chrysosporium was used as a basis for defining initial experimental

conditions.

3.3 Review of P. chrysosporium Cultivation and Lignocellulolytic Activity

General maintenance and sporulation of P. chrysosporium has been achieved using potato

dextrose agar (PDA). Tien and Kirk (1988) reported a medium containing 1% glucose, 1% malt

extract, 0.2% peptone, 0.2% yeast extract, 0.1%, aspargine, 0.2% KH2PO4, 0.1% MgSO4.7H2O,

0.0001% thiamine, and 2% agar for maintenance and sporulation. In P. chrysosporium, cellulose

degradation occurs during primary metabolism, while lignin mineralization has been

characterized as a secondary metabolic event that is triggered by carbon, nitrogen or sulfur

starvation (Jeffries et al 1981; Kirk and Farrell 1987).

The P. chrysosporium genome sequence reveals that this organism encodes at least 10 LiPs, 5

MnPs and 6 copper radical oxidases, in addition to several other gene products that might

contribute to lignin degradation, including glyoxal oxidases, (Martinez et al. 2004). Growth

conditions that trigger the production of lignin degrading enzymes by P. chrysosporium were

recently reviewed (Singh and Chen 2008). To summarize, the expression of lignin-degrading

enzymes by submerged liquid cultures of P. chrysosporium is promoted by limiting carbon,

nitrogen and/or sulfur availability, as well as increasing organic acid (α-hydroxy acids) and trace

element concentration. Increased agitation, temperature and pH appeared to reduce lignin

degrading activity. The effect of agitation was more pronounced in large flasks with high

shaking, which was predicted to generate shear forces that can denature secreted enzymes. It has

also been proposed that enzyme inactivation may occur due to higher level of dissolved oxygen

33

in agitated cultures than in stationary cultures. This has been attributed to the higher level of

active oxygen content in the culture which can oxidize, fragment, and cross-link proteins (Wolff

et al. 1986). Further, high oxygen concentrations can inhibit fungal growth at low glucose

concentrations (Leisola et al. 1984).

Decades worth of studies now mean that P. chrysosporium is typically cultivated between pH

4.2-6.2 (optimal below 6.0) at 25-39˚C (optimal 37˚C) in either stationary or shaking cultivations

(up to 300 rpm; optimal below 200 rpm). The growth medium is also amended with 0.02%

ammonium tartarate and 1% glucose or glycerol, as well as trace metals including calcium,

manganese, iron, zinc, copper and magnesium. Notably, adding detergents such as Tween 80 or

non-reactive foam-stabilizing agents such as polyethylene glycol (PEG) to the culture medium

can also increase lignolytic activity in culture supernatants (Asther et al. 1987; Venkatadri and

Irvine 1993). Moreover, the highest LiP production of 3,800 U/ml was obtained by

immobilizing fungal mycelia in polyurethane foam, whereas the highest MnP production

observed was 1,375 U/ml with solid-state fermentation using steam-exploded straw as substrate.

Some effective repressors of lignin-degrading enzyme production are glutamic acid, glutamine

and ammonium ion.

While the effect of cultivation conditions on the expression of lignolytic activity by P.

chrysosporium was recently reviewed, corresponding effects on the expression of cellulolytic

activities are more dispersed in the published literature. Similar to lignolytic enzymes,

carbohydrate active enzyme expression is highly dependent on the nitrogen and carbon source

supplied to the culture medium, pH, temperature, and presence of inhibitors (Table 3.1 and Table

3.2). For instance, early studies of white-rot fungi indicated that cellulases are induced in the

presence of cellulosic substrates, and that cellulase expression is regulated through catabolite

repression (Eriksson 1978; Jensen 1971; Johansson 1966; Smith and Gold 1979). These studies

employed viscometric measurements to determine endoglucanase activity on

carboxymethylcellulose, and colorimetric assays to measure glucosidase activity on p-

nitrophenyl-D-glycopyranosides. In general, early biochemical analyses of P. chrysosporium

confirmed that cellulose is a good substrate to investigate cellulolytic and hemicellulolytic

activity by this fungus (Bao et al. 1994; Eriksson and Rzedowski 1969). Early studies also

34

determined that asparagine is an effective nutrient source and that aeration with oxygen enhances

the activity of the cellulase systems (Ander and Eriksson 1977; Levonen-Munoz and Bone

1984). No clear impact of agitation on cellulolytic activity has been reported, and the pH

optimum for cellulolytic activity in basidiomycetes is typically between 3.5 to 5.5 and the

temperature optima are between 35-50˚C (Baldrian and Valaskova 2008; Eriksson 1978; Smith

and Gold 1979; Eriksson and Hamp 1978).

More recently, the effect of substrate, inducing compounds, and cultivation conditions on

lignocellulolytic expression by P. chrysosporium has been evaluated using molecular techniques,

including quantitative PCR and proteomic methods. PCR is particularly useful for monitoring

microbial activity when corresponding enzymes are not easily accessible or distinguished by

activity alone. One of the early studies to use PCR to investigate P. chrysosporium activity grew

the fungus on four different carbon sources: ball-milled straw, microcrystalline cellulose

(Avicel), high glucose concentration (2 %), and low glucose concentration (0.2 %); all under

nitrogen limiting conditions (0.23 g/L) (Broda et al. 1995). This research revealed differential

expression of cellulase gene families, depending on the substrate and stage of cultivation.

Specifically, the expression of genes that encode lignin-degrading and cellulose-degrading

enzymes was observed on straw and microcrystalline cellulose, suggesting lignocellulose-

dependent expression of these enzymes (Broda et al. 1995).

RT-PCR was also used to investigate the regulation of LiPs and MnPs by P. chrysosporium,

(Janse et al. 1998), and several more recent studies have used RT-PCR to identify inducers and

repressors of lignocellulolytic genes. For example, Suzuki et al., (2008) used real-time

quantitative RT-PCR to investigate carbon catabolite repression in P. chrysosporium. They

concluded that cellobiohydrolase genes (cel6A, cel7d and cdh) were drastically influenced by

increased levels of glucose concentration, whereas no significant changes were observed in β-

glucosidase gene (bgl3A) (Suzuki et al. 2008). Recent advances in RT-PCR techniques that

improve the quantification of gene transcripts produced at low levels, were also used to quantify

transcript levels of glycoside hydrolases genes (cel6A and cel7A to cell 7F/G) in cultures

containing glucose, cellulose and cellooligosaccharides (Suzuki et al. 2009). Cellulose,

cellotriose and cellotetraose demonstrated clear upregulation of transcript levels as compared to

35

Table 3.1. Summary of carbohydrate active enzyme production by Phanerochaete chrysosporium

Strain Cultivation Conditions

Agitation Substrate Enzyme Activity (Units per mL)a,b Ref

EGase CBH CBD FPU BGL XYN/ MAN

ATCC 32629)

Submerged 0.25% aspargine 1 L or 250 mL 11 days and 14 days

Yes (no details)

1 % Cellulose powder

6.8 8.9 77 105 ND ND ND ND (Ander and Eriksson 1977)

23.0 46.3 227 263

Submerged 0.25% NH4H2PO4 1 L or 250 mL 11 days and 14 days

Yes (no details)

1 % Cellulose powder

0.9 1.1 8.5 11.7 ND ND ND ND (Ander and Eriksson 1977) 2.2 1.6 23 50

Submerged 0.25% aspargine 18 days and 25 days

Yes (no details)

1 % pine wood meal

0.40

12.5 ND ND ND ND (Ander and Eriksson 1977)

0.50 19.4

Submerged 0.25% NH4H2PO4 18 days and 25 days

Yes (no details)

1 % pine wood meal

0.06

5.4 ND ND ND ND (Ander and Eriksson 1977)

0.04 4.0

MTCC 787

Submerged 7 days and 15 days

No Paddy straw 0.11 ND ND 0.40 0.33 0.50 c (Mishra et al. 2007) 0.43 0.88 1.2 0.67c

NCIM 1073

Submerged 7 days and 15 days

No Paddy straw 0.22

ND ND 0.44 0.34 0.65 c (Mishra et al. 2007)

0.37 0.56 1.1 0.64 c

36

Strain Cultivation Conditions

Agitation Substrate Enzyme Activity (Units per mL)a,b Ref

EGase CBH CBD FPU BGL XYN/ MAN

NCIM 1106

Submerged 7 days and 15 days

No Paddy straw 0.10 ND ND 0.30 0.38 0.43 c (Mishra et al. 2007)

0.15 0.36 1.1 0.47 c

NCIM 1197

Submerged 7 days and 15 days

No Paddy straw 0.16 ND ND 0.31 0.37 0.46 c (Mishra et al. 2007)

0.18 0.41 0.98 0.54 c

OGC 101

Submerged 500 mL 14 days succinate buffer protein concentration: 939 ± 24.5 mg

150 rpm Cotton linters ND 138±5 74± 10.0

ND 237± 20.5

ND (Bao et al. 1994)

OGC 101

Submerged 500 mL 14 days succinate buffer protein concentration: 694 ± 53.7 mg

150 rpm Microcrystalline cellulose

ND 75±9 90±7.1 ND 82± 15.5

ND (Bao et al. 1994)

OGC 101

Submerged 500 mL 14 days succinate buffer protein concentration: 583 ± 25.5 mg

150 rpm Filter-paper ND 29±5 50.7± 12.0

ND 155± 10.8

ND (Bao et al. 1994)

37

Strain Cultivation Conditions

Agitation Substrate Enzyme Activity (Units per mL)a,b Ref

EGase CBH CBD FPU BGL XYN/ MAN

OGC 101

Submerged 500 mL 14 days succinate buffer protein concentration: 421 ± 39.3 mg

150 rpm Acid-treated cellulose

ND 35±4 8.9± 12.6

ND 8.9± 2.5

ND (Bao et al. 1994)

OGC 101

Submerged 14 days succinate (pH 4.5) protein concentration: 1.08± 0.14g/l

150 rpm Microcrystalline cellulose

ND 141±16 53±17 ND 237±21 Max on day 14

ND (Bao et al. 1994)

OGC 101

Submerged 14 days acetate (pH 4.5) protein concentration 0.59 ± 0.02 g/l

150 rpm Microcrystalline cellulose

ND 20 ± 1.4 184 ±20

ND 23 ±1 Max on day 6

ND (Bao et al. 1994)

OGC 101

Submerged 14 days phosphate (pH 5.8) protein concentration 0.31 ± 0.01 g/l

150 rpm Microcrystalline cellulose

ND 6 ± 1 75 ± 15

ND 7 ± 1 Max on day 3

ND (Bao et al. 1994)

OGC 101

Submerged 14 days phosphate

150 rpm Microcrystalline cellulose

ND 58 ± 6 58 ± 2 ND 45 ± 9 Max on day

ND (Bao et al. 1994)

38

Strain Cultivation Conditions

Agitation Substrate Enzyme Activity (Units per mL)a,b Ref

EGase CBH CBD FPU BGL XYN/ MAN

(pH 7.0) protein concentration 0.64 ± 0.17 g/l

3

OGC 101

Submerged 14 days phosphate (pH 8.0) protein concentration 1.39 ± 0.29 g/l

150 rpm Microcrystalline cellulose

ND 77±3 129± 20

ND 81±6 Max on day 9

ND (Bao et al. 1994)

a. EG: Endoglucanase, CBH: cellobiohydrolase, CDH: cellobiose dehydrogenase, BGL: beta-glucosidase, XYN: xylanase,

MAN: mannanase. b. Split cells indicate activities in 1L and 250 mL cultures (left to right) and first and second time point (top to

bottom), respectively. c. Activity was expressed as IU/ml (one unit = 0.1 unit change in OD per min).

39

Table 3.2. Expression of lignocellulolytic activity by Phanerochaete chrysosporium

Strain Cultivation Conditions

Agitation Substrate Summary of Resultsa Ref

n/a 300 mL 13 days conical shake flasks modified Norkrans’ medium

Yes Powdered cellulose

No induction of BGL activity was observed when glucose was used as the carbon source. When cellobiose and cellulose were used, BGL activity was detected in fungal cell wall fractions after 15-20 min. Extracellular enzyme activity was detected with cellulose as the sole carbon source. pH optima 4-4.5. Localization of BGL activity depended on the carbon source.

(Deshpande et al. 1978)

ME 446

Submerged 1 L modified Vogel medium

150 rpm

0.25 % cotton or CMC

Extracellular BGL was induced by cellulose but repressed by glucose. Intracellular BGL was induced by cellobiose but repressed by cycloheximide. Optimum temperature was 45˚C. pH optima for intracellular and extracellular activities were pH 7.0 and 5.5, respectively. Strongest inducers of extracellular BGL were cellulosic polymers (highest for CMC – 14.9 U/mg), wherease cellobiose was a very poor inducer. Intracellular BGL activity was 27.5 U/mg protein in cellobiose cultivations. The soluble intracellular enzyme represented approximately 90 and 75% of the total BGL activity in cells induced with cellobiose for 10 and 25 hours, respectively

(Smith and Gold 1979b)

Strain K3

Investigation with purified enzymes

No

Bacterial micro-crystalline cellulose

Cellobiose acted as a competitive inhibitor of cellobiohydrolase. Decomposition of bacterial microcrystalline cellulose by CBH I was enhanced in the presence of CDH/ferricyanide redox-system. It was concluded that the reason for the enhancement of CBH I activity in the presence of CDH redox system was that it relieves competitive inhibition of cellobiose by its oxidation to cellobionolactone.

(Igarashi et al. 1998)

40

Strain Cultivation Conditions

Agitation Substrate Summary of Resultsa Ref

ATCC 32629

Modified Norkrans medium (pH 5.0)

200 rpm

Glucose (5g/L)

The effect of several monosaccharides and disaccharides in different concentrations on expression of EGs was investigated. Induction of EGs occurred at cellobiose concentrations as low as 1 mg/mL, catabolite repression was not obtained until the glucose or mannose concentration reached 50 mg/L. Xylose, galactose and arabinose did not repress EG expression even at concentrations as high as 1g/L.

(Eriksson and Hamp 1978)

n/a

Modified Norkrans medium 27˚C

No

Powdered cellulose/ xylan/guar powder

Cellulase, xylanase, mannanase, BGL and aryl-β-glucosidase activities were detected in cultures containing cellulose. The production of cellulase and mannanase was not observed in xylan cultivations, and only traces of all enzymes were found in mannan cultivations.

(Eriksson and Rzedowski 1969)

BKM-F-1767

500 ml conical flasks 100 ml mineral medium (pH 5.0) 37˚C

200 rpm

Sugarcane bagasse

The optimum pH values were 4.6, 4.2, 5.0 and 5.0 for EG, BGL, xylanase, and β-xylosidase activities, respectively; optimum temperatures were 60, 70, 65, and 60 ˚C, respectively. Significant loss in activity was reported after 5 hours heating at 60˚C. Addition of 1% glucose inhibited the production of cellulolytic and xylanolytic enzymes at the beginning of the cultivation; activity increased after the total consumption of glucose.

(Khalil 2002)

a. EG: Endoglucanase, CBH: cellobiohydrolase, CDH: cellobiose dehydrogenase, BGL: beta-glucosidase

41

glucose for most genes, at all time points, while cellobiose was a weak inducer of cellulase gene

transcription by comparison (Suzuki et al. 2010).

The specific objective of the current study was to establish conditions and methods to cultivate

P. carnosa, as well as monitor the growth and the profile of proteins secreted by this fungus. In

addition to enabling subsequent investigations and growth studies on wood, comparison of

optimum cultivation conditions to those defined for other wood-degrading Basidiomycetes, will

help elucidate evolutionary adaptations among the white-rot fungi.

3.4 Materials and Methods

3.4.1 Microorganism and Materials

P. carnosa strains (FP-135508-Sp, HHB-10118-Sp, RLG-7412-Sp) were obtained from the US

Department of Agriculture (USDA) Forest Products Laboratory, Madison, WI. All strains were

maintained on YMPG agar media; yeast (2 % w/v), maltose (10 % w/v), peptone (2 % w/v),

glucose (1 % w/v) at 27°C. Strain HHB-10118 was selected for detailed cultivation studies,

given its consistent growth rate and phenotype. The internal transcribed spacer (ITS) region was

sequenced to confirm the identity of the fungus (de Koker et al. 2003). Avicel (Fluka 11365),

which is a commercial preparation of wood-derived microcrystalline cellulose, and contains

approximately 30 % of amorphous structure, was used as the carbon source in liquid cultivations,

and ammonium tartarate (Sigma A4767) was used as the nitrogen source.

3.4.2 Cultivation Conditions

Liquid cultures were prepared in 200 mL B3 medium (Appendix 1) (Hon, D.N-S 1984; Jai, M.,

et. al., 1996). Liquid media were then inoculated with 0.1 mg/mL blended mycelia from 10-day

liquid YMPG cultures. Cultivation temperatures were 22°C, 27°C, 30°C, and 37°C, and cultures

were incubated for up to 18 days under stationary conditions or with shaking (150 rpm). The

impact of nitrogen concentration on the secretion of lignocellulolytic activity was tested in

stationary cultivations grown for 27 days at 27 °C with 1 % Avicel and 2.5 or 25 mM ammonium

tartrate (Table 3.2).

42

3.4.3 Growth Measurements

Chitin measurement was used to compare the extent of growth in all cultivation conditions

(Plassard et al. 1982). Briefly, three replicate cultivations were harvested after five time points

(Table 3.3), and the mycelia were separated from the culture supernatant using Mira Cloth. The

mycelia were then oven dried overnight and the final dry weight of the mycelia was recorded.

Precisely weighed amounts (~0.2 g) of each mycelial sample were transferred to a glass tube

with a screw-cap lid, incubated in 5 mL of 6 M HCI at room temperature for 3 h with the lid

open, capped and then hydrolyzed for 16 h at 80°C. Aliquots of 1 mL hydrolysate were

transferred to 5 mL of 1.25 M NaOAc to bring the pH of the solution to approximately pH 3, and

then further processed at room temperature. Specifically, 1 mL of the solution at pH 3 was mixed

with 1 mL of 5 % KHSO4 and 1 mL of 5 % NaNO2. After shaking for 5 min, the mixture was

allowed to stand for 15 min to allow for deamination, and then 1 mL of 12.5 % ammonium

sulfonate was added to release excess nitric acid (note: This step is accompanied by vigorous

mixing to prevent over-flow of the solution from reaction vial due to effervescence.) After

shaking for 5 min, 1 mL of 0.5 % MBTH (Methylbenzthiazolinone-2-hydrazone) was added. The

reaction was allowed to proceed for 1 h without any shaking. For colorimetric detection of the

anhydromannose-MBTH-FeCl3 complex at 653 nm; 1 mL of 0.5 % FeCl3 was added and the

solution was left to develop for 30 min. The absorbance was stable up to 20 h.

For protein concentration measurements, culture supernatants from the three biological replicates

prepared for each cultivation condition were pooled to obtain adequate amounts of protein for all

biochemical assays. Prior to pooling, the biological reproducibility of protein profiles was

determined using SDS-page gels. The pooled samples were then centrifuged for 20 min at 4500

rpm and concentrated at 4 °C using Pall’s Jumbosep (FD010K65) centrifugal units at a cutoff of

10 kDa. Protein concentration was measured using the Bio-rad Protein Assay according to the

manufacturer’s instructions. The absorption measurements were performed at 595 nm. A BSA

stock solution of 2 mg/ml was used to generate a standard curve from 0.05-0.5 mg/mL of

protein.

43

3.4.4 Biochemical Assays

The standard endoglucanase assay contained 50 mM sodium acetate (pH 4.5) and 0.3 wt % of

carboxymethyl cellulose (CMC). The final reaction volume was 100 µL and reactions were

incubated at 27°C for up to 1 h. The reaction products were developed using the BCA assay for

reducing ends (Doner et al. 1992), and measured at 565 nm. Glucose was used to generate a

standard curve.

β-glucosidase activity was determined by measuring the amount of p-nitrophenol released from

p-nitrophenol-β-D-glucoside (pNPG) using 0.6 % pNPG by weight in 50 mM sodium acetate

(pH 4.5) (Bowers et al. 1980). The reaction products were developed using 100 µL of 1M

Na2CO3 and measured at 410 nm; pNP (0-0.05µmol) was used to generate a standard curve.

Filter paper units (FPUs) were measured using Whatman No.1 filter paper. An office hole-

puncher was used to produce equal sized filter paper samples (~ 1 mm2), which were submerged

in 50 mM sodium acetate (pH 4.5). The total reaction volume was 200 µL and reactions were

incubated at 27°C for up to 1 hour. The reaction products were again developed using the BCA

assay, and glucose was used to generate a standard curve (Ghose 1987).

Lignin peroxidase activity was measured by monitoring oxidation of 3,5-dimethoxybenzyl

alcohol to veratraldehyde at 310 nm (Tien and Kirk 1984). An extinction coefficient of 9.3

mmol-1 cm-1 was used to determine product formation (Kapich et al. 2004). Manganese

peroxidase (MnP) activity was measured by monitoring the oxidation of ABTS at 415 nm

(Katagiri et al. 1995). The reaction mixture contained 7.8 mM ABTS, 0.1 mM MnSO4, 50 mM

sodium lactate, 0.1 mM H2O2, and 20 mM sodium succinate buffer (pH 4.5). An extinction

coefficient of 3.6 mmol-1 cm-1 was used to determine product formation. The assay was

performed in the presence and absence of MnSO4 to determine Mn-dependent and Mn-

independent activity.

All assays were performed in triplicate at the temperature and pH of the cultivation. Enzyme

activities are expressed as units (U) per ml of culture supernatant, where one unit of enzyme

activity is defined as the amount of enzyme generating 1 µmol of product per minute.

44

Table 3.3 Cultivation of P. carnosa in B3 medium at 27˚C with 1 % microcrystalline cellulose, variable nitrogen, pH 4.5

Nomenclature Nitrogen (Ammonium Tartarate)

Agitation Incubation Time

Analyses

LN (low nitrogen)

2.5 mM No 3 days, 6 days, 12 days, 18 days and 24 days

Mycelial weight, protein concentration in culture supernatant, protein profile (SDS-PAGE). Cellulase (CMC, pNP and FPU), LiP and MnP activity for each time point (at pH 4.5 and 27 °C).

HN (high nitrogen)

25 mM No 3 days, 6 days, 12 days, 18 days and 24 days

Mycelial weight, protein concentration in culture supernatant, protein profile (SDS-PAGE). Cellulase (CMC, pNP and FPU), LiP and MnP activity for each time point (at pH 4.5 and 27 °C).

LNS (low nitrogen with shaking)

2.5 mM 150 rpm 3 days, 6 days, 12 days, 18 days and 24 days

Mycelial weight, protein concentration in culture supernatant, protein profile (SDS-PAGE). Cellulase (CMC, pNP and FPU), LiP and MnP activity for each time point (at pH 4.5 and 27 °C).

45

3.5 Results

Initial growth studies performed to evaluate the range of compounds that could be used to

cultivate P. carnosa confirmed that P. carnosa grew on commercial substrates including

glycerol, glucose and microcrystalline cellulose (Avicel), as well as chips of lodgepole pine,

white spruce, balsam fir, maple and poplar, submerged in B3 medium. To characterize the

growth of P. carnosa in more detail, cultivations were prepared using Avicel as the carbon

source.

3.5.1 Effect of Cultivation Condition on the Extent of P. carnosa Growth on

Microcrystalline Cellulose.

Preliminary cultivations of P. carnosa were grown on micro-crystalline cellulose (Avicel) at

22˚C, 27°C, 30°C and 37°C. From visual inspection of cultivations and weight of collected

mycelia, it was concluded that cultures incubated at 27°C led to the highest fungal growth. The

air-dried weight of mycelia collected after 5 weeks at 22˚C, 27˚C, 30˚C and 37˚C were 0.99g,

0.95g, 0.70 and 0.74g, respectively. In contrast to P. chrysosporium, which grows optimally at

37°C, growth of P. carnosa was not visible at 37°C,

As reviewed above, nitrogen concentration and agitation affects the expression of

lignocellulolytic enzymes by P. chrysosporium. Therefore, the impact of these two parameters

on the secretion of lignocellulolytic activity by P. carnosa grown at 27°C on Avicel was

investigated (Table 3.3). P. carnosa grew well in low and high nitrogen conditions as well as in

cultivations agitated at 150 rpm. At each of the time points indicated in Table 3.3, the extent of

fungal growth was determined by protein concentrations in culture supernatants, and by

measuring chitin content (Plassard et al. 1982).

The final protein concentrations for high nitrogen (HN) and low nitrogen with shaking (LNS)

cultivations were similar, while the low nitrogen, stationary (LN) cultivations had the highest

protein concentration at the last time point (Table 3.4). After 12 days of cultivation, protein

concentrations in both HN and LNS cultivations appeared to decrease, suggesting proteolysis. By

constrast, the protein concentration in LN cultivations increased steadily over 24 days.

46

Table 3.4 Protein concentration in P. carnosa culture supernatant during growth on

microcrystalline cellulose

Days

(protein concentrations are indicated as µg/ml)

Cultivation 3 6 12 18 24

LN 1.0 ± 0.3 1.8 ± 0.2 4.0 ± 0.1 4.1 ± 0.1 8.0 ± 0.0 LNS n/a 2.3 ± 0.3 3.4 ± 0.8 1.7 ± 0.1 3.2 ± 0.4 HN 3.4± 0.7 2.6 ± 0.1 8.3 ± 0.6 3.1 ± 0.1 4.5 ± 0.0

LN: low nitrogen, stationary; LNS: low nitrogen, shaking; HN: high nitrogen, stationary. n = 3;

errors correspond to standard deviations.

The effect of growth condition and cultivation time on the profile of proteins secreted by P.

carnosa was evalutated by SDS gel electrophoresis (Fig. 3.1). The protein profiles from the

different cultivations appeared similar at early time points, and the complexity of the protein

profiles from stationary cultivations appeared to increase with time (Fig 3.1). Consistent with

protein concentration measurements (Table 3.4), the intensity of proteins in the LNS cultivations

decreased after time point 3.

Figure 3.1 Silver stained 10% SDS-PAGE gel of P.carnosa culture supernatants. LN: low

nitrogen, stationary; LNS: low nitrogen, shaking; HN: high nitrogen, stationary. Numbers refer

to time points specified in Table 3.3. Sample volumes were 10 µL, and ranged from 0.01 to 0.08

µg.

kDa 250 150 100 75 50 37

25

1 2 3 4 5

47

Indirect methods including visual inspection, substrate dry weight loss, rate of diacetate

hydrolysis, extractable protein content and chitin content have previously been used to quantify

fungal growth on substrates like wood chips, where separation between mycelia and substrate is

challenging. In a comparative study of these methods, Boyle and Kropp (1992) concluded that

chitin (a polymer of N-acetyl-p-glucosamine, the major cell wall component in fungi) is the most

suitable means of measuring fungal growth. Ergosterol, which is a fungus specific membrane

component, is another means to quantify fungal growth. While ergosterol measurements can give

an estimate of the living fungal biomass at time of harvest, chitin content is a measure of both

living and dead fungal biomass, and so can be used to estimate total growth (Ekblad and

Nisholm 1996). Literature cites several methods to determine chitin content (Ekblad and

Nisholm 1996; Francois 2006; Zamani et al. 2008). The method developed by Plassard et al.

(1982) has been used to determine chitin content in similar cultivations, and so was used in the

current study (Papinutti and Lechner 2008; Levin et al. 2008). As anticipated, low nitrogen

stationary cultivations resulted in the highest fungal growth and protein concentration. FTIR of

the mycelial samples was also investigated as a means to obtain relative mycelial growth.

However, due to the low mycelial density, the FTIR spectra of the test samples were very similar

to the controls containing Avicel alone.

While chitin was initially measured from 0.3-30 mg (dry weight) of mycelia recovered from

cellulose cultivations, the mycelial sample had to be increased to approximately 200 mg before

reliable absorption measurements could be obtained. As a result, chitin was measured in oven

dried mycelial samples collected from 24-day cultures, which corresponded to comparatively

high visible growth. Microcrystalline cellulose that remained in the 24-day P. carnosa

cultivations was aggretated with fungal mycelia, and so could not be removed prior to chitin

measurements. Notably, up to 0.8 mg of cellulose was found to contribute to the calorimetric

reading. However, since cellulose in exceess of 0.8 mg did not increase background readings,

the absorption of 0.8 mg of microcrystalline cellulose was used as a reference in the assay. Chitin

measurements were then used to estimate the growth of P. carnosa in the different cultures; this

analysis was consistent with protein concentrations measured in the culture media, where the

highest chitin content was observed for LN cultures (Table 3.5).

48

Table 3.5. Chitin content of P. carnosa cultivations pooled after 24 days

Cultivation Total dry weight (mg) Total chitin (mg)

Negative control (cellulose only) 0.8 0.4 LN 2.9 2.8 LNS 2.1 2.1 HN 1.6 1.6

LN: low nitrogen, stationary; LNS: low nitrogen, shaking; HN: high nitrogen, stationary.

Taken together, the protein concentration measurements and the chitin measurements suggest

that stationary cultivations containing low nitrogen resulted in the highest fungal growth. The

impact of cultivation conditions on the expression lignocellulolytic activity was then assessed

using standard enzyme activity assays.

3.5.2 Effect of Nitrogen Concentration and Agitation on Lignocellulolytic Expression by

P. carnosa.

All cellulytic activities (CMC, pNP and FPU activities) were observed in all cultivation

conditions at all time points, while LiP activity was not observed at any time point (Tables 3.6 to

3.8). MnP activity (both Mn dependent and independent) was observed at most time points, but

there was a considerable variation over the time course (Table 3.9). The apparent lack of LiP

activity is distinct from cultivation studies using P. chrysosporium (Kirk et al. 1978), but

consistent with recent transcriptional analysis of P. carnosa grown on various wood and simple

substrates (MacDonald et al. 2011). As will be discussed in Chapter 4, this difference in enzyme

expression compared to P. chrysosporium, could reflect distinct lignin-degrading activity by

these fungi.

Specific endoglucanase activity was initially highest in the LN cultivations, and then decreased

with increasing cultivation time (Table 3.6). This could be explained by increasing protein

concentrations in the culture supernatant that resulted from lysis of aging cells. The highest

specific endoglucanase activity was detected in LNS cultivations, and the lowest was initially

detected in HN cultivations.

49

Table 3.6. Endoglucanase activity in P. carnosa culture supernatant during growth on

microcrystalline cellulose

Cultivation Days

(specific activities are indicated as nmol reducing ends/ µµµµg/min) 3 6 12 18 24

LN 20.9 ± 0.6 10.2±0.2 1.3 ± 0.03 3.4 ± 0.1 1.5 ± 0.05 LNS n/a 7.7 ± 0.2 2.4 ± 0.05 16.2 ± 0.4 9.9 ± 0.1 HN 3.5±0.2 0.9 ± 0.3 1.3 ± 0.01 2.4 ± 0.02 3.9 ± 0.2

LN: low nitrogen, stationary; LNS: low nitrogen, shaking; HN: high nitrogen, stationary. n = 3;

errors correspond to standard deviations.

Other researchers have shown that product inhibition of cellulases by cellobiose is minimized by

the presence of β-glucosidase. Accordingly, β-glucosidase activity was also measured. Similar

to cellulase profiles, β-glucosidase activity was highest in LNS cultivations (Table 3.7). Further,

β-glucosidase activity in P. carnosa was generally lower than endo-glucanase activity, which is

consistent with recently published proteomic and transcriptomic analyses of P. carnosa

(MacDonald et al, 2011; Mahajan and Master, 2010). Total cellulolytic activity was then

measured using the Filter Paper Assay (Table 3.8).

Table 3.7. ββββ-glucosidase activity in P. carnosa culture supernatant during growth on

microcrystalline cellulose

Cultivation Days

(specific activities are indicated as nmol reducing ends/ µµµµg/min)

3 6 12 18 24

LN 0.9 ± 0.01 0.3 ± 0.01 0.2 ± 0.01 0.3 ± 0.001 0.1 ± 0.01 LNS n/a 0.6 ± 0.01 0.5 ± 0.01 1.4 ± 0.01 2.1 ± 0.01 HN 0.04 ± 0.01 0.1 ± 0.001 0.14 ± 0.01 0.09 ± 0.01 0.03 ± 0.01

LN: low nitrogen, stationary; LNS: low nitrogen, shaking; HN: high nitrogen, stationary. n = 3;

errors correspond to standard deviations.

50

Table 3.8. Total cellulolytic activity in P. carnosa culture supernatant during growth on

microcrystalline cellulose

Cultivation Days

(specific activities are indicated as nmol reducing ends/µµµµg/min) 3 6 12 18 24

LN 0.67 ± 0.01 0.81±0.01 0.26±0.01 0.15±0.02 0.01 ± 0.001 LNS n/a 1.10±0.01 0.15±0.01 0.16±0.17 0.73 ± 0.01 HN 0.21 ± 0.01 1.08±0.01 0.33±0.01 0.98±0.06 0.75 ± 0.001

LN: low nitrogen, stationary; LNS: low nitrogen, shaking; HN: high nitrogen, stationary. n = 3;

errors correspond to standard deviations.

MnP activities were detected in both the presence and absence of MnSO4. The values obtained

were not significantly different, implying that manganese is not required as an inducer for MnP

expression in P. carnosa cultivations grown on cellulose. The highest specific activity was

observed in LNS cultivations (Table 3.9), although MnP was initially higher in LN and HN

cultivations. The initial rise and the decline in MnP activity is consistent with recent

transcriptional analysis of P. carnosa where transcripts predicted to encode MnP activity were

quantified using RT-PCR. Specifically, transcripts predicted to encode MnP activity were

highest in mycelia collected from early cultivations of P. carnosa grown on various wood

samples, and then decreased at later growth stages (personal communication with J. MacDonald,

a Ph.D.candidate in E. Master’s lab).

Table 3.9. MnP activity in P. carnosa culture supernatant during growth on

microcrystalline cellulose

Cultivation Days

(specific activities are indicated as nmol reducing ends/ µµµµg/min)

3 6 12 18 24

LN 0.3 ± 0.01 n/a n/a n/a 0.01 ± 0.001 LNS n/a n/a 2.4 ± 0.01 n/a 0.2 ± 0.001 HN 0.05 ± 0.02 n/a 0.01 ± 0.01 n/a 0.03 ± 0.001

LN: low nitrogen, stationary; LNS: low nitrogen, shaking; HN: high nitrogen, stationary. n = 3;

errors correspond to standard deviations.

51

3.6 Discussion

Similar to P. chrysosporium, the cellulolytic enzyme activity expressed by P. carnosa was

lowest in HN (high nitrogen) stationary cultivations (Barclay et al. 1993). While β-glucosidase

activity in P. chrysosporium was typically higher than endoglucanase activity (Tables 3.1 and

3.2), the opposite trend was observed in P. carnosa cultivations (Tables 3.6 and 3.7). As

described in Chapter 5, comparatively low β-glucosidase activity in P. carnosa cultures is

consistent with the comparatively few peptides corresponding to GH3 enzymes detected in P.

carnosa culture supernatants.

The highest lignin degrading activity was measured in LNS cultivations using the MnP assay.

Low-nitrogen concentration also induces the expression of lignolytic activity by P.

chrysosporium, where nutrient limiting conditions are thought to induce secondary metabolism

(Kirk and Farrell 1987). Decreased MnP activity in later time points sampled from P. carnosa

cultivations was consistent with recent transcription analyses performed by J. MacDonald (a PhD

candidate in E. Master’s lab), where it was observed that MnP expression was highest during

early growth on wood samples, and then declined (personal communication with J. MacDonald).

Shaking at 150 rpm did not have a detrimental effect on lignocellulolytic activity, and in fact

increased the activities measured in this study. Notably, early studies with P. chrysosporium

revealed that agitation (rpm = 125, 150) of submerged cultures can suppress lignin degrading

activity (Kirk et al. 1978), perhaps as a result of pellet formation (Shimada 1981). However,

other studies report increased lignin degrading activity in P. chrysosporium cultivations shaking

at 150 rpm, due to increased diffusion of oxygen in the culture medium (Leisola et al. 1984).

Providing oxygen while reducing shear stress might be key (Venkatadri and Irvine 1993).

It is evident that final lignocellulolytic activity measurements were low, despite the complex

extracellular enzyme profiles observed by SDS-PAGE (Fig 3.1). This can be attributed to several

factors, including the presence of proteases or enzyme inhibitors. Based on the review of P.

chrysosporium, it is possible that changes to cultivation conditions including pH, nitrogen

source, carbon source, or addition of inducers including oligosaccharides and cellobiose oxidase,

could increase the production of lignocellulolytic enzymes by P. carnosa. Further, the addition

52

of surfactants like Tween, or organic acids is likely to impact production of lignin degrading

peroxidases. Still, it is important to remember that biochemical assays of enzyme activity are

limited to the availability of appropriate substrates, and so can underestimate the range of

lignincellulolytic activities present in culture supernatants. Accordingly, as described in

Chapters 4 and 5, further characterization of lignocellulose conversion by P. carnosa applied

analytical tools that could retrieve more detailed descriptions of enzymes secreted by P. carnosa

during growth on lignocellulosic substrates, and the corresponding impact on wood fibre

composition and structure.

Chapter 3 References

Ander P, Eriksson KE (1977) Selective degradation of wood components by white-rot fungi. Physiology Plant 41:239-248

Asther M, Corrieu G, Drapron R, Odier E (1987) Effect of Tween 80 and oleic acid on ligninase production by Phanerochaete chrysosporium INA-12. Enzyme and Microbial Technology 9 (4):245-249

Baldrian P, Valaskova V (2008) Degradation of cellulose by basidiomycetous fungi. FEMS Microbiology Reviews 32:501-521

Bao W, Lymar E, Renganathan V (1994) Optimization of cellobiose dehydrogenase and β−glucosidase production by cellulose-degrading cultures of Phanerochaete chrysosporium. Applied and Microbial Biotechnology 42:642-646

Barclay CD, Legge RL, Farquhar GF (1993) Modelling the growth kinetics of Phanerochaete

chrysosporium in submerged static culture. Appl Environ Microbiol 59 (6):1887-1892

Bergman O, Nilsson T (1966) Research Notes. As published by Erikkson et. al., 1978.

Bowers G, N. Jr., McComb RB, Christensen RC, Schaffer R (1980) High-purity 4-nitrophenol: purification, characterization and specifications for use as a spectrophotometric reference material. Clinical chemistry 26 (6):724-729

Boyle DC, Kropp BR (1992) Development and comparison of methods for measuring growth of filamentous fungi on wood. Canadian Journal of Microbiology 38 (10):1053-1060

Broda P, Birch PR, Brooks PR, Sims PF (1995) PCR-mediated analysis of lignocellulolytic gene transcription by Phanerochaete chrysosporium: substrate-dependent differential expression within gene families. Appl Environ Microbiol 61 (6):2358-2364

53

De Koker TH, Nakasone KK, Haarhof J, Burdsall jr HH, Janse BJH (2003) Phylogenetic relationships of the genus Phanerochaete inferred from the internal transcribed spacer region. Mycological Research 107 (9):1032-1040

Deshpande V, Eriksson K-EL, Pettersson B (1978) Production, purification and partial characterization of 1,4-β-glucosidase enzymes from Sporotrichum pulverulentum. European Journal of Biochemistry 90:191-198

Ekblad A, Nisholm T (1996) Determination of chitin in fungi and mycorrhizal roots by an improved HPLC analysis of glucosamine. Plant and Soil 178:29-35

Eriksson KE, Rzedowski W (1969) Extracellular enzyme systems utilized by the fungus Chrysosporium lignorum for the breakdown of cellulose. Archives of Biochemistry and Biophysics 129:683-688

Eriksson KE (1978) Enzyme mechanisms involved in cellulose hydrolysis by the rot fungus Sporotrichum pulverulentum. Biotechnology and Bioengineering 20:317-322

Eriksson KEL, Hamp SG (1978) Regulation of endo-1,4-β-glucanase production in Sporotrichum pulverulentum. European Journal of Biochemistry 90:183-190

Eriksson KEL, Blanchette RA, Ander P. (1990) Microbial and Enzymatic Degradation of Wood and Wood Components. Springer-Verlag. Berlin, Germany.

Francois JM (2006) A simple method for quantitative determination of polysaccharides in fungal cell walls. Nature Protocols 1 (6): 2995-3000

Ghose TK (1987) Measurement of cellulase activities. Pure and Applied Chemistry 59:257-268

Hibbett D (2007) After the gold rush, or before the flood? Evolutionary morphology of mushroom-forming fungi (Agaricomycetes) in the early 21st century. Mycology Research 111:1001-1018

Igarashi K, Sampy JD, Eriksson K-EL (1998) Cellobiose dehydrogenase enhances Phanerochaete chrysosporium cellobiohydrolase I activity by reliveing product inhibition. European Journal of Biochemistry 253:101-106

Janse BJH, Gaskell J, Akhtar M, Cullen D (1998) Expression of Phanerochaete chrysosporium genes encoding lignin peroxidases, manganese peroxidases, and glyoxal oxidase in wood. Appl Environ Microbiol 64 (9):3536-3538

Jeffries TW, Choi S, Kirk TK (1981) Nutritional regulation of lignin degradation by Phanerochaete chrysosporium. Appl Environ Microbiol.42(2):290-6.

Jensen KF (1971) Cellulolytic enzymes of Stereum gausapatum. Phytopathology 61:134-138

Johansson M (1966) A comparison between the cellulolytic activity of white and brown rot fungi. I. The activity on insoluble cellulose. Physiology Plant 19:709-722

54

Khalil AI (2002) Production and characterization of cellulolytic and xylanolytic enzymes from the ligninolytic white-rot fungus Phanerochaete chrysosporium grown on sugarcane bagasse. World Journal of Microbiology and Biotechnology 18:753-759

Kirk TK, Schlutz E, Connors WJ, Lorenz LF, Zeikus JG (1978) Influence of culture parameters on lignin metabolism by Phanerochaete chrysosporium. Archives of Microbiology 117:277-285

Kirk TK, Farrell RL (1987) Enzymatic 'combustion': The microbial degradation of lignin. Annual Reviews Microbiology 41:465-505

Leisola MSA, Ulmer DC, Fiechter A (1984) Factors affecting lignin degradation in lignocellulose by Phanerochaete chrysosporium. Archives of Microbiology 137 (2):171-175.

Levin L, Herrmann C, Papinutti VL (2008) Optimization of lignocellulolytic enzyme production by the white-rot fungus Trametes trogii in solid-state fermentation using response surface methodology. Biochemical Engineering Journal 39 (1):207-214

Levonen-Munoz E, Bone DH (1984) Effect of different gas environments on bench-scale solid state fermentation of oat straw by white-rot fungi. Biotechnology and Bioengineering 27:382-387

MacDonald J, Doering M, Canam T, Gong Y, Guttman D, Campbell M, Master ER (2011) Transcriptomic analysis of the softwood-degrading white-rot fungus Phanerochaete

carnosa reveals concerted gene expression responses to growth on coniferous and deciduous wood. Applied and Environmental Microbiology. [Epub ahead of print]

Martinez D, Larrondo LF, Putnam N, Sollewijn Gelpke MD, Huang K, Chapman J, Helfenbein KG, Ramaiya P, Detter JC, Larimer F, Coutinho PM, Henrissat B, Berka R, Cullen D, Rokhsar D (2004) Genome sequence of the lignocellulose degrading fungus Phanerochaete chrysosporium strain RP78. Nature Biotechnology 22:695-700

Mishra BK, Pandey AK, Lata (2007) Lignocellulolytic enzyme production from submerged fermentation of paddy straw. Indian Journal of Microbiology 47:176-179

Otjen L, Blanchette R, Effland M, Leathem G (1987) Assessment of 30 white rot basidiomycetes for selective lignin degradation. Holzforschung. 41 (6):343-349

Papinutti L, Lechner B (2008) Influence of the carbon source on the growth and lignocellulolytic enzyme production by Morchella esculenta strains. Journal of Industrial Microbiology and Biotechnology 35 (12):1715-1721.

Plassard CS, Mousain DG, Salsaca LE (1982) Estimation of mycelial growth of basidiomycetes by means of chitin determination Phytochemistry 21 (2):345-348

55

Shimada M, Nahatsubo F, Kirk TK, Higuchi T (1981) Biosynthesis of the secondary metabolite veratryl alcohol in relation to lignin degradation in Phanerochaete chrysosproium. Archives of Microbiology. 129: 321-324

Singh D, Chen S (2008) The white-rot fungus Phanerochaete chrysosporium: conditions for the production of lignin-degrading enzymes. Applied Microbiology and Biotechnology 81:399-417

Smith MH, Gold MH (1979) Phanerochaete chrysosporium ß-glucosidases: Induction, cellular localization, and physical characterization. Applied and Environmental Microbiology 37 (5):938-942

Suzuki H, Igarashi K, Samejima M (2008) Real-time quantitative analysis of carbon catabolite derepression of cellulolytic genes expressed in the basidiomycete Phanerochaete

chrysosporium. Applied Microbiology and Biotechnology 80 (1):99-106

Suzuki H, Igarashi K, Samejima M (2009) Quantitative transcriptional analysis of the genes encoding glycoside hydrolase family 7 cellulase isozymes in the basidiomycete Phanerochaete chrysosporium. FEMS Microbiology Letters 299 (2):159-165.

Suzuki H, Igarashi K, Samejima M (2010) Cellotriose and cellotetraose as inducers of the genes encoding cellobiohydrolases in the Basidiomycete Phanerochaete chrysosporium. Appl Environ Microbiol 76 (18):6164-6170.

Vanden Wymelenberg VA, Sabat G, Martinez D, Rajangam AS, Teeri TT, Gaskell J, K KPJ, Cullen D (2005) The Phanerochaete chrysosporium secretome: Database predictions and initial mass spectrometry peptide identifications in cellulose-grown medium Journal of Biotechnology 118 (1):17-34

Venkatadri R, Irvine RL (1993) Cultivation of Phanerochaete chrysosporium and production of lignin peroxidase in novel biofilm reactor systems: Hollow fiber reactor and silicone membrane reactor. Water Research 27 (4):591-596

Wolff SP, Garner A, Dean RT (1986) Free radicals, lipids and protein degradation. Trends in Biochemical Sciences 11 (1):27-31

Zamani A, Jeihanipour AJ, Edebo L, Niklasson C, Taherzadeh MJ (2008) Determination of glucosamine and N-acetyl glucosamine in fungal cell walls. Journal of Agricultural and Food Chemistry 56 (18):8314-8318

56

Chapter 4 : Mode of Coniferous Wood Decay by the White Rot Fungus Phanerochaete carnosa as Confirmed by FT-IR

and ToF-SIMS

4.1 Abstract

The softwood degrading white-rot fungus, Phanerochaete carnosa was investigated for its ability

to degrade three coniferous woods: balsam fir, white spruce and lodgepole pine. P. carnosa

grew similarly on these three wood species, and like the hardwood-degrading white-rot fungi

Ceriporiopsis subvermispora and Phanerochaete chrysosporium BKM-F-1767, P. carnosa

demonstrated selective degradation of lignin, as observed by FTIR and ToF-SIMS analyses.

Lignin degradation across-the-wall was determined by the ratio of lignin concentration between

distinct cell wall regions as measured by TEM and by ToF-SIMS, and was shown to be uniform.

This study evaluates the ability of a white-rot fungus to utilize softwood lignin and reveals the

industrial potential of the lignocellulolytic activity elicited by this fungus.

4.2 Introduction

Basidiomycetes are among the most important degraders of lignocellulosic biomass (Lundell et

al. 2010). The two most common modes of wood degradation by basidiomycetes are brown-rot

and white-rot decay. Brown-rot decay is characterized by extensive degradation of cellulose and

hemicellulose with only limited modification of lignin. Examples of well-studied brown-rot

basidiomycetes include Gloeophyllum trabeum, Serpula lacrymans, and Postia placenta

(Hogberg et al. 2006; Jensen et al. 2001; Martinez et al. 2009). These fungi appear to initiate the

depolymerization of wood polysaccharides using Fenton chemistry, whereby diffusible hydroxyl

radicals penetrate the wood surface, leaving the lignin as a modified residue (Jensen et al. 2001;

Yelle et al. 2008). By contrast, white-rot fungi secrete numerous glycoside hydrolases,

carbohydrate esterases and lignin-degrading peroxidases that target all of the main components

of wood, including cellulose, hemicellulose and lignin, leaving residual wood with a bleached

appearance (Hatakka 1994; Kersten and Cullen 2007; Tuor et al. 1995).

57

Certain white-rot fungi, such as Trametes versicolor, appear to simultaneously degrade lignin

and polysaccharides, while other white-rot fungi, including Ceriporiopsis subvermispora, appear

to selectively degrade lignin and then only slowly hydrolyze cellulose (Blanchette 1991;

Blanchette et al. 1997; Fackler et al. 2006; Pandey and Pitman 2003). The ability of fungi to

selectively degrade lignin has been used to pretreat fibre and defibrillate wood in the production

of mechanical pulp (Fackler et al. 2006; Scott et al. 2002).

While many white-rot fungi characterized to date have been isolated from hardwoods (Hibbett

and Donoghue 2001), Phanerochaete carnosa, studied here, was isolated almost exclusively

from softwood species (Burdsall 1985). As boreal forests comprise more than 45% of the North

America’s forested area and softwoods are comparatively recalcitrant sources of lignocellulosic

biomass (Birdsey et al. 2006), the glycoside hydrolases and lignin-degrading peroxidases

produced by P. carnosa constitute an important source of industrially relevant enzymes. The

secretome and transcriptome of the P. carnosa grown on lignocellulosic substrates were recently

described, and show that similar sets of enzymes are elicited by P. carnosa grown on different

lignocellulosic substrates although to different expression levels, suggesting that adaptations to

degrade lignocellulose might result from fine-tuning the expression of similar enzyme families

(Mahajan and Master 2010; MacDonald et al., 2011). Although the production of

lignocellulolytic activities by P. carnosa has been studied, the pattern of wood decay by this

fungus over time is not known. By characterizing the progression of chemical and morphological

changes of softwood decayed by P. carnosa, the aim of the current study was to predict whether

the expression of secreted hemicellulases and cellulases likely follows initial lignin degradation

by this organism. That is, to determine if P. carnosa can be classified as a selective or

simultaneous degrader.

Previous reports describe several analytical methods that can be used to investigate anatomical

and chemical changes introduced in wood fibres upon microbial decay. Electron microscopy

techniques, including scanning electron microscopy (SEM) and transmission electron

microscopy (TEM) have been used to determine micro-morphological degradation patterns in

plant cell walls (Blanchette 1991). Lignin-specific dyes such as phloroglucinol and safranin are

used for the visualization of lignin by light microscopy (Hakala et al. 2004). Changes in fibre

58

chemistry have been monitored directly using Fourier transform infrared spectroscopy (FTIR),

nuclear magnetic resonance spectroscopy (NMR), pyrolysis mass spectrometry (Pyr-MS), ion

chromatography, and thermochemolysis (Blanchette 1995; Davis et al. 1994; Faix et al. 1991;

Irbe et al. 2006; Kelley et al. 2002). FTIR is especially suited to rapid chemical characterization

of small quantities of wood with minimum sample preparation and has been used to analyze

changes in wood fibre chemistry after degradation by different brown and white-rot fungi (Faix

et al. 1991; Pandey and Nagveni 2007; Pandey and Pitman 2003). For example, Pandey and

Pitman (2003) use FTIR to analyze the chemical composition of pine (softwood) and beech

(hardwood) following treatment with the brown-rot Coniophora puteana, the non-selective

white-rot T. versicolor and the selective white-rot P. chrysosporium. In their study, brown-rot

decay was indicated by a relative increase in the lignin-to-carbohydrate peak area ratio, selective

white-rot decay was demonstrated by a relative decrease in lignin and xylan content, and the

simultaneous decay led to similar reduction in carbohydrate and lignin peaks.

Time-of-flight secondary ion mass spectrometry (ToF-SIMS) is another versatile technique that

has previously been applied to characterize surfaces of pulp, fibres and wood samples (Rowell

2005). In addition to parallel analysis of organic and inorganic constituents of wood fibre, ToF-

SIMS provides chemical images with a mass spectrum at each pixel. As a result, ToF-SIMS is

capable of spatially resolving different components on the surface of wood, providing

information about the composition across cell walls (Jung et al. 2010). ToF-SIMS analysis of

lignin and carbohydrate polymers in wood had been limited to a few wood peaks assigned in

literature until the recent identification of approximately 40 additional confirmed lignocellulose-

descriptive peaks through ToF-SIMS analysis of pine species (Goacher et al. 2011).

The current study investigates the effect of conifer species and stage of degradation on the mode

of softwood decay by P. carnosa. Specifically, P. carnosa was grown on wood chips prepared

from white spruce, lodgepole pine and balsam fir. These three wood species are not only of

prime commercial value in the North American boreal forests, but they also comprise common

mixed feedstocks for pulp and paper industries. Specifically, FTIR was employed to monitor

temporal changes in lignocellulose composition resulting from fungal activity. To evaluate the

mode of degradation across cell walls and regions of lignin degradation (i.e. cell wall or middle

59

lamella), TEM was then used to determine the localization of lignin prior and post-decay.

Finally, ToF-SIMS analysis was performed to simultaneously describe the chemical composition

and spatial resolution of lignocellulose components and to validate results obtained using FTIR

and TEM.

4.3 Materials and Methods

4.3.1 Fungal Strain and Cultivation Conditions

Phanerochaete carnosa strain HHB-10118-sp was obtained from the US Department of

Agriculture (USDA) Forest Products Laboratory, Madison, WI, and was maintained on YMPG

medium containing 2 g yeast extract, 10 g malt extract, 2 g peptone, 10 g glucose, 2 g KH2PO4, 1

g MgSO4·7H2O, 1 g asparagine per 1 L in H2O. Wood cultivations were prepared by using a

Waring blender to grind balsam fir (Abies balsamea), lodgepole pine (Pinus contorta), white

spruce (Picea glauca), and then sifting air-dried samples through US standard mesh size 6 (3.35

mm2) and 14 (1.5 mm2) sieves. Wood particles that were retained on mesh size 14 were

recovered, and 4 g samples were transferred to a total of sixty (3 species x 5 time points x 3 test

replicates and one control) 500 mL beakers containing 10 mL of B3 buffer (Appendix 1; Kirk et

al. 1978). Thiamine-HCl, ammonium tartrate, and 2,2- dimethylsuccinic acid were added as

filter-sterilized solutions, while all other media were steam sterilized for 20 to 30 min.

Each culture medium was inoculated with an 11 mm circular agar plug taken from the growing

edge of P. carnosa cultivated on solid YMPG. Cultivations were incubated under stationary

conditions at 27 °C until the diameter of the surface mycelial mat was 1 cm, 4 cm, 6 cm and 8

cm; the final sample was taken after 12 weeks of cultivation. At each growth point, a metal cutter

with 2.8 cm diameter was used to isolate residual wood particles from the centre of each

cultivation. Wood samples were rinsed with milli-Q water to remove mycelia and were air dried

before analysis.

4.3.2 Fourier Transform Infrared Spectroscopy

A minimum of three wood slivers from each time point for each wood species were randomly

chosen and powdered using a bench-top ball mill (WIG-L-BUG model, Dentsply Rinn). A small

amount of the milled sample (~ 1 mg) was mixed with KBr (200 mg) and the mixture pelletized

60

using a die (1.3 cm diameter) and a hydraulic press. A Bruker Tensor 27 FTIR was used to

record the absorbance between 4000 and 400 cm-1 with a resolution of 4 cm-1. The spectra

representing the average of 32 scans were corrected for atmospheric vapor compensation;

baseline was corrected using the rubber band method (Opus software, v. 5.0). Spectra were

normalized for unit-vector and mean-centred prior to the principal component analysis (PCA) in

Matlab v. 7.8.0 software.

The following compounds were run as standards to assist lignocellulose FTIR assignments for

unreported peaks: microcrystalline cellulose (Avicel PH-101), tamarind xyloglucan, birchwood

xylan, arabinogalactan, guar (galactomannan), pectin, and mannose were obtained from Sigma-

Aldrich; konjac glucomannan was obtained from Megazyme, and Klason lignin was prepared by

Dr. D. Jeremic (PDF in E. Master’s lab) from white spruce according to TAPPI standards (T222

om-83). The FTIR spectra were obtained and analyzed as described above.

4.3.3 Transmission Electron Microscropy - Energy-Dispersive X-ray Analysis (TEM-

EDXA)

Lignin is usually visualized with TEM-EDXA by labeling lignin with KMnO4 or bromine

(Sjöström 1983; Stein et al. 1992). In this study, six late decay wood samples (time point 5) and

six early decay wood samples (time point 1) from spruce cultivations were brominated following

the protocol of Saka et al (1978). The brominated samples were then embedded in low viscosity

embedding medium (Modification E) (Spurr 1968). Thin sections (100 nm) of the embedded

samples were obtained by using a Leica EM UC 6 microtome, and Diatome Ultra 35˚ knife (MC

12734), and analyzed using an HD-2000 transmission electron microscope (Hitachi High

Technologies America, Inc) equipped with an Oxford Instruments Inca X-sight detector.

4.3.4 Time of Flight Secondary Ion Mass Spectroscopy

Since samples observed under TEM-EDXA showed very similar patterns of lignin distribution

across the cell walls, cross-section surfaces of a fibre sample from each wood species

representing growth points one and five were prepared using a microtome (Leica EM UC 6).

ToF-SIMS spectra of these cross-sections were collected by Dr. R. Goacher (PDF in C. Mims

and E. Edwards’ lab) using a ToF-SIMS IV (IonTof, Muenster, Germany). 25 keV Bi3+ primary

61

ions were utilized to generate high mass resolution spectra in the bunched mode (0.2 pA pulsed

current) and high spatial resolution images in burst alignment mode. Charge compensation was

applied with a low energy electron beam. Base vacuum was 5 × 10-9 mbar. Spectra were

calibrated to the CH3+, H3O

+, C2H3+ and C3H5

+ ions using IonSpec v.4.1 software. At m/z 91, the

mass resolution (M/∆M) was ~5,000 in bunched mode and ~250-300 in burst alignment mode;

depending on sample roughness. To determine local variations in degradation, one spot on each

sample (100×100 to 200×200 µm2) was analyzed and split into three to four regions of interest to

generate replicates. Prior to PCA, peak intensities were Poisson-corrected to compensate for

minor detector saturation (Stephan et al. 1994). A peak list consisting of 36 definitive

lignocellulosic m/z ratios was employed for the analyses in both spectral and spatial models.

These peaks comprised groups 1 and 2 of Goacher et al. (2011), minus peaks at m/z 19 and 31

due to suspicion of moisture effects and at m/z 45, 59, 73, 131, and 147 because of PDMS

(polydimethylsiloxane) interferences. ToF-SIMS high mass resolution spectra were normalized

to unit area and mean-centred. ToF-SIMS images were mean-centred without normalization.

4.3.5 Statistical Analysis

PLS Toolbox v. 5.8.2 (Eigenvector Research, Inc.) was used with MATLAB v. 7.8.0 (The

MathWorks, Inc.) to perform PCA of spectra obtained from FTIR and ToF-SIMS. MIA Toolbox

(Eigenvector Research, Inc.) was used for PCA analysis of ToF-SIMS nominal mass resolution

ion images. PCA loadings were analyzed to determine the specific lignocellulose components

that changed significantly over time.

4.4 Results

Radial extension of fungal mycelia has been used to measure growth and compare growth

profiles on solid substrates. Previous reports also describe that visual estimates of fungal growth

on wood chips are consistent with indirect methods including substrate dry weight loss, rate of

diacetate hydrolysis, or extractable protein content and chitin content, where separation of

mycelia and substrate is challenging (Boyle and Kropp 1992; Marin et al. 2005). To monitor

growth in the present study, radial extension of P. carnosa from the agar plug was measured, and

wood samples from the centre of each cultivation were harvested at pre-defined radial

measurements. The time required to reach each radial extension was similar between cultivation

62

replicates and across wood species, indicating a similar growth progression on all the three

species selected (Appendix 2, Fig. S4.1). Gradual darkening of wood chips was observed in all

cultivations. This darkening has been correlated to chemical changes due to lignin decay (Fackler

et al. 2007; Fackler et al. 2003). These decay products and compositional changes induced in the

material were investigated using FTIR and ToF-SIMS spectral analysis and the abundance of

lignin in different cell wall layers was monitored by TEM-EDXA and ToF-SIMS imaging.

4.4.1 FTIR Spectroscopy

The PCA of wood FTIR spectra in the fingerprint region (800-1800 cm-1) clearly separated the

advanced decayed samples from the controls (Fig. 4.1 to 4.3). Considering the natural variability

of wood samples, Hotelling’s T-square statistics were used to initially identify samples from the

same time point and species that generated scans with high leverage as well as residual variance

(i.e., dangerous outliers). Despite the removal of the obvious outliers, the initial principal

component (PC1) of the analysis of spruce and fir identified variability among the data due to

sample origin.

PC1 of pine and subsequent PCs of fir and spruce (PC2 and PC3, respectively) revealed

progressive lignin decay by P. carnosa (Fig. 4.1 to 4.3). All control samples exhibited relatively

high peak intensity at ~1512 and ~1262 cm-1 as determined by PCA (Table 4.1). The 1512 cm-1

peak corresponds to aromatic skeletal vibrations of lignin in softwoods (Faix and Beinhoff 1988;

Pandey and Pitman 2003). The PCA loading at 1262 cm-1 does not correspond to an actual peak

in the spectra, but rather the difference in the slope of the peak with 1268 cm-1 maxima (guaiacyl

ring breathing). It has been previously shown that the 1263 cm-1 peak appears in a first derivative

IR spectra of lignin, confirming its suggested relation to 1268 cm-1 peak (Faix 1992). Therefore,

for all of the examined species, the significant aromatic assignments for lignin were

distinguished by PCA as higher in control samples than the decayed samples.

63

Peaks assigned in literature (* sh = shoulder) SPRUCE PINE FIR

Wavenumber (cm-1)

Peak description PCA

loading actual peak

PCA loading

actual peak

PCA loading

actual peak

1768 1761 1767

1728

1722 1800-1700

Unconjugated carbonyl stretches; characteristic of both lignin and carbohydrates (Faix 1992)

1711 1709 1711

1696 1692

1683 1700-1550 Conjugated carbonyl stretches, conjugated and nonconjugated C=C stretches; characteristic of both lignin and carbohydrates (Faix 1992) 1646

1400 1401

1511 Aromatical skeletal vibrations (Faix and Beinhoff, 1988)

1512 1511 1512 1511 1512 1511

1493, 1498 Aromatical skeletal vibrations (Faix and Beinhoff, 1988)

1496 1495

1460-1470 C-H deformations, asym in -CH3 and -CH2 (Faix and Beinhoff, 1988, Faix 1992)

1466 1466 sh*

1453 Seen as deconvoluted wavelengths for 1465 cm-1 peak (Faix and Beinhoff, 1988)

1453 1453

1427-1428 Aromatic skeletal vibrations combined with C-H in-plane deformations (Faix and Beinhoff, 1988, Faix 1992)

1427 1426 sh

1400 1401

1268-1273 Guaiacyl ring breathing with carbonyl stretching (Faix and Beinhoff, 1988)

1261 1269 1261 1268 1263 1268

1223-1226 Common to lignin and cellulose (Hobro et al. 2010)

1228 1233

1057-1058

C-O stretching from the pyranosidic ring in carbohydrates and lignin (Denes et al. 1999, Pandey and Pitman 2003, Faix and Beinhoff, 1988)

1061 1061

1031-1033

C-H in-plane deform. guaiacyl ring and C-O deform. primary alcohols and possibly influence of non-conjugated C=O groups (Faix and Beinhoff, 1988)

1033 1031-1034

Table 4.1 Assignment of FTIR peaks that distinguish control and decayed wood samples

64

Fig. 4.1 Spruce decay described by

FTIR analysis. A) Grouping on PCs 2

and 3 for normalized and mean-

centered FTIR data. Closed symbols

represent test data; open symbols

represent control data. Symbol shapes

represent time progression: time point

1 (O), 2 (), 3 (�), 4 ( ), 5 (◊). B)

Loadings for PC 3, separating more

and less decayed samples. Horizontal

dotted lines at magnitude |0.05|

represent thresholds for loading

significance. Peaks with highest

loadings for the control samples

denotes their loss in the decayed

samples. C) Mean baseline corrected

FTIR data normalized to total intensity

between 800 and 1800 cm-1: control

time point 1 (thick black line), test time

point 2 (thin black line) and test time

point 5 (thin grey line). Grey vertical

dashed lines are visual guides to

compare significant loadings in (B)

with FTIR spectra in (C).

65

Fig. 4.2 Fir decay described by

FTIR analysis. A) Grouping on PCs

1 and 2 for normalized and mean-

centered FTIR data. B) Loadings for

PC 2, separating more and less

decayed samples. C) Mean baseline

corrected FTIR data for control and

test data. Symbols and lines as

described in Fig. 4.1.

66

Fig. 4.3 Pine decay described

by FTIR analysis. A) Grouping

on PCs 1 and 2 for normalized

and mean-centered FTIR data. B)

Loadings for PC 1, separating

more and less decayed samples.

C) Mean baseline corrected

FTIR data for control and test

data. Symbols and lines as

described in Fig 4.1.

67

Nearly all of the peaks that corresponded to other significant loadings (>|0.05|) in control

samples were previously reported in milled-wood lignin (MWL) and/or G and GS lignin

dehydrogenation polymers by Faix and Beinhoff (1988). Exceptions include certain peaks in the

fir control samples that are not exclusive to lignin (Table 4. 1). For example, the peak at 1228

can be attributed to both lignin and hemicelluloses (Hobro et al. 2010). Still, the only peak not

associated with lignin that showed high peak loading in a control sample, was at 1400 cm-1 for

spruce. This peak was also highest in the 1480-1300 cm-1 region of non-cellulosic carbohydrates,

including arabinogalactan, guar galactomannan, mannose, pectin, tamarind xyloglucan and

birchwood xylan (Appendix 2, Table S4.1).

The PC loadings that describe the decayed samples indicate either bonds that were not attacked

by the fungus or else were generated through biological modification of the substrate. With the

exception of the loading at 1400 cm-1 that described decayed pine, PCA of all samples showed

high loadings in the 1800-1700 cm-1 region and the 1700-1550 cm-1 region. The PCA loading at

~1767 cm-1 and ~1711 cm-1 described decayed spruce, pine and fir, but were not detected in raw

spectra of the samples or standards and instead were correlated to shoulders (Fig. 4.1 to 4.3;

Appendix 2, Fig. S4.2). Similarly, the loading at 1722 cm-1 in decayed pine samples and loading

at 1728 cm-1 in decayed fir corresponded to peaks that were not detected in the raw spectra. All

of these loadings describe slopes of peaks in the 1800-1700 cm-1 region of the corresponding

spectra, where the 1734-1732 cm-1 region defines the peak maximum. This region describes

unconjugated carbonyl stretches that can be formed either in lignin or hemicelluloses (Faix

1992). For instance, three peaks in the 1737-1710 cm-1 region (1737, 1720 and 1710 cm-1) are

detected in lignin spectra after deconvolution (Faix and Beinhoff 1988) and like the peak at 1228

cm-1, a peak maximum at 1734 cm-1 has also been attributed to both lignin and hemicelluloses

(Hobro et al. 2010; Pandey and Pitman 2003)

The decayed samples were also characterized by significant loadings in the 1700-1550 cm-1

region, which were identified as slopes of peaks that are characteristic of conjugated carbonyl

stretches (Table 4.1). For instance, aromatic rings of milled wood lignin absorb at ~1595 cm-1

(Colom et al. 2003; Faix and Beinhoff 1988; Pandey and Pitman 2003), while Klason lignin

displays carbon stretching at 1682 cm-1 and aromatic skeletal vibration breathing with C=O

68

stretching at 1603 cm-1 (Kubo and Kadla 2005). Although outside of the range of important

loadings that describe the data reported here, conjugated carbonyl stretch in degraded cellulose

has been reported at 1620 cm-1(Lojewska et al. 2005).

In summary, the PCA of FTIR spectra described the decayed samples by peaks characteristic of

lignin and carbohydrates, while the same analysis distinguished the control samples primarily by

lignin-specific peaks that were lost upon degradation. Accordingly, the PCA of FTIR spectra

indicated that progressive decay by P. carnosa was mainly described by change in lignin

composition. TEM and ToF-SIMS were performed next to evaluate the localization of lignin

decay and to characterize the biochemical transformations in more detail.

4.4.2 TEM-EDXA

Lignin concentration in different regions of the cell wall differs considerably, with the highest

reported concentration being in cell corners and middle lamellae (Fromm et al. 2003; Saka et al.

1982). Here, the spatial degradation of lignin was investigated to describe microscopic

progression of P. carnosa attack on the softwood species.

Spruce samples from growth point one and five were selected for TEM-EDXA image analyses

given the clear compositional transformations detected in these samples as characterized by

FTIR. In this study, line scans of brominated samples across the cell wall and cell corners

demonstrated that the ratio of lignin in the cell walls to lignin in the middle lamella did not

change significantly with decay (Appendix 2, Table S4.1). No apparent change in lignin

concentration between cell walls suggests that P. carnosa simultaneously degraded lignin present

in the cell walls and middle lamella. Further, wood decay by P. carnosa did not appear to cause

morphological changes in the decayed samples.

4.4.3 ToF-SIMS

ToF-SIMS analysis was used to validate the FTIR results of lignin degradation and also the

lignin concentration pattern across fibre cell walls as observed by TEM-EDXA. Chemical

changes due to decay were assessed using an extended list of peak annotations for lignocellulosic

materials (Goacher et al. 2011).

69

A clear progression of lignin degradation in each of the three wood species tested was revealed

when comparing ToF-SIMS spectra of samples from early and late P. carnosa cultivations (Fig

4.4). The lignin peaks described by Goacher et al. (2011) characterized the samples at early

growth points (except m/z 67 for spruce and fir), while peaks characteristic of polysaccharides

were more prevalent in the later growth points.

However, the contributions of each lignin and polysaccharide peak depended on the wood

species. For example, degraded spruce samples exhibited lower enhancement of high-mass

polysaccharide peaks, suggesting that bioconversion activity had progressed further in these

samples. Also, the m/z 137 peak characteristic of guaiacyl lignin (Saito et al. 2005) was

particularly influential in separating degraded spruce and fir samples from corresponding

controls, but less influential for the pine samples. Finally, the prevalence of the lignin peak at

m/z 67 at the later time points of decayed spruce and fir samples may indicate a degradation

product that arose at this mass.

For all wood species, the PCA models of ToF-SIMS cross-sections of control and decayed

samples demonstrated higher concentrations of lignin in middle lamella and cell corners relative

to the intensities in the cell walls (Fig. 4.5, Appendix 2 Fig. S4.3 and S4.4), further supporting

the TEM-EDXA observations that P. carnosa simultaneously degrades lignin in the cell walls

and middle lamella.

70

Fig 4.4 PCA of normalized and mean-centered ToF-SIMS spectra: A) Fir, B) Pine, and C)

Spruce. Left column depicts scores on PC1 separating controls (squares) and test samples

(diamonds). Right column depicts corresponding loadings with peaks identified by Goacher et al.

(2011) labeled for polysaccharides (stars) and lignin (triangles). Grey horizontal lines at loading

magnitudes |0.05| represent thresholds for loading significance

71

Fig. 4.5 PCA of mean-centered ToF-SIMS image of decayed pine (time point 5): image of

scores (left) and loadings (right) for PC 2. Image dimensions are in pixels where 50 pixels equal

19.1 microns.

4.5 Discussion

P. carnosa grew on wood particles prepared from lodgepole pine, balsam fir and white spruce; in

all cases, radial growth of 8 cm was reached on average in 45-50 days. The darkening of wood

chips observed during the decay by P. carnosa was similarly reported for growth of white-rot

fungi C. subservmispora, P. chrysosporium and T. versicolor on spruce, and has been correlated

to the formation of free lignin radicals and corresponding decay products (Fackler et al. 2007).

Significant changes in fibre chemical composition were detected by FTIR analysis after growth

point three, which was reached after approximately 25 days of cultivation. The time required to

reach this stage of decay might be reduced in cultivations that do not contain additional nutrient

medium (Singh and Chen 2008). In general, ToF-SIMS and FTIR analyses revealed the

degradation of lignin accompanied by partial hemicellulose degradation, reminiscent of selective

white-rot fungi (Blanchette et al. 1985).

PCA of FTIR spectra indicated the disappearance of lignin in decayed samples by revealing the

dominance of vibrations at 1512 and 1267 cm-1 in the control samples, which correspond to

softwood lignin. While the PCA loadings descriptive of control samples pointed to actual peaks

observed in the spectra and could be attributed either to lignin or hemicelluloses, the loadings of

72

decayed samples were not as clear and described changes in slopes in regions between 1800 and

1590 cm-1 overall suggesting changes in residual lignin and hemicelluloses. Notably, the PCA of

FTIR and ToF-SIMS also revealed that the progression of decay was most aggressive in spruce

samples, since similar levels of decay were achieved after the third growth point. Further, the

PCA of ToF-SIMS spectra composed of previously assigned peaks more clearly separated early

and late decay by lignin degradation. Given the clarity of the ToF-SIMS data and the capability

of this technique to give both spectral and spatial information of the samples, these analyses

reinforce the utility of ToF-SIMS for chemical analysis of complex biomass samples.

During delignification or selective decay, cell walls retain their morphology, but gradually

become un-reactive to lignin stains and increasingly reactive to cellulose stains as the lignin is

removed (Kirk and Cullen 1998). For instance, morphological alterations of cell walls were not

detected in pinewood chips degraded by the selective white-rot C. subvermispora, even though

staining with uranyl acetate indicated that the permeability of the secondary wall had changed

(Blanchette et al. 1997). Similarly, changes in the cell morphology and the ratio of lignin in cell

walls and middle lamellae were not detected when comparing early and late decayed samples by

TEM-EDXA and ToF-SIMS in the present study.

Most white-rot fungi that have been characterized to date typically degrade syringyl (S) lignin in

preference to guaiacyl (G) lignin (Faix et al. 1985). For instance, the secondary cell wall layer in

hardwoods is primarily composed of syringyl units and is typically more receptive to white-rot

attack than the cell corners, which are rich in guaiacyl units (Blanchette et al. 1987). By contrast,

the current results indicated that P. carnosa is proficient at attacking and modifying the guaiacyl

lignin component in softwood (as denoted by ToF-SIMS as m/z 137, and by FTIR as 1512 and

1268 cm-1). Though not detected in the samples characterized here, syringyl lignin is described

by a peak at 1128-1125 cm-1 in FTIR spectra and m/z 167 and 181 in ToF-SIMS spectra (Saito et

al. 2005)..

The degradation of guaiacyl lignin by P. carnosa is consistent with recent transcriptomic

analyses of P. carnosa grown on woody substrates (MacDonald et al. 2011). MacDonald et al.

(2011) revealed that the expression of transcripts encoding manganese peroxidases (MnP) in P.

73

carnosa were higher than those predicted to encode lignin peroxidases (LiP), and that the P.

carnosa genome encoded more genes that were predicted to encode MnPs than LiPs

(MacDonald et al. 2011). This is in contrast to P. chrysosporium, which encodes 10 LiPs and 5

MnPs (Martinez et al. 2004). Since the oxidants produced by MnP, including Mn3+, are smaller

than LiPs and associated oxidants, it is possible that products of MnP activity can diffuse more

easily through denser guaiacyl lignin structures present in softwood fibre, and promote softwood

decay by P. carnosa (Fackler et al. 2007; Hammel and Cullen 2008). As described in the next

Chapter, the similar pattern of decay observed for lodgepole pine, balsam fir and white spruce is

also consistent with the proteomic analysis of P. carnosa, which highlighted the secretion of

similar lignocellulolytic enzymes during growth on spruce and cellulose, implying the elicitation

of a similar decay strategy for different lignocellulosic substrates (Mahajan and Master 2010).

To summarize, P. carnosa can degrade guaiacyl lignin moieties and elicits selective white-rot

decay of lodgepole pine, balsam fir and white spruce. These results suggest that P. carnosa could

benefit bioprocess treatments used to reduce the recalcitrance of softwood species that

predominate in many boreal forests. Still, it is important to bear in mind that the pattern of white-

rot decay (i.e., selective or simultaneous decay) by wood-degrading fungi can vary depending on

cultivation conditions. For example, Ganoderma applanatum elicits simultaneous decay when

grown on Acer saccharum, Populus alba, Populus tremuloides, as does Ischnaderma resinosum

when grown on Populus deltoides (Blanchette 1984). However, both of these organisms were

correlated with selective decay of other tree species (Blanchette 1984). Accordingly, although P.

carnosa elicited selective degradation of lignin in lodgepole pine, balsam fir and white spruce, it

is plausible that the mode of decay may differ when grown on significantly different feedstocks,

including hardwood species.

Chapter 4 References

Birdsey RA, Jenkins JC, Johnston M, Huber-Sannwald E (2006) Chapter 11: North American Forests. In: Climate Change Science Program (CCSP) Product 2.2. USA.

Blanchette RA (1984) Screening wood decayed by white rot fungi for preferential lignin degradation Applied and Environmental Microbiology 48 (3):647-653

74

Blanchette RA, Otjen L, Effland MJ, Eslyn WE (1985) Changes in structural and chemical components of wood delignified by fungi. Wood Science and Technology 19:35-46

Blanchette RA, Otjen L, Carlson MC (1987) Lignin distribution in cell walls of birch wood decayed by white rot Basidiomycetes. Phytopathology 77 (5):684-690

Blanchette RA (1991) Delignification by wood-decay fungi. Annual Reviews of Phytopathology 29:381-398

Blanchette RA (1995) Degradation of the lignocellulosic complex in wood. Canadian Journal of Botany 73 (Supplemental 1):S999 - S1010

Blanchette RA, Krueger EW, Haight JE, Akhtar M, Akin DE (1997) Cell wall alterations in loblolly pine wood decayed by the white-rot fungus, Ceriporiopsis subvermispora

Journal of Biotechnology 53 (2-3):203-213

Boyle DC, Kropp BR (1992) Development and comparison of methods for measuring growth of filamentous fungi on wood. Canadian Journal of Microbiology 38:1053-1060

Burdsall HHJ (1985) A contribution to the taxonomy of the genus Phanerochaete (Corticiaceae,

Aphyllophorales). Mycological Memoir No. 10. J. Cramer, Braunschweig.

Colom X, Carrillo F, Nogués F, Garriga P (2003) Structural analysis of photodegraded wood by means of FTIR spectroscopy. Polymer Degradation and Stability 80 (3):543-549

Davis MF, Schroeder HA, Maciel GE (1994) Solid state 13C nuclear magnetic resonance studies of wood decay. II White rot decay of paper birch. Holzforschung 48:186-192

Fackler K, Schwanninger M, Hinterstoisser B, Messner K (2003) Bio-modification of spruce wood by Ceriporiopsis subvermispora: Comparison of the effects of three different strains. In: 12th International Symposium on Wood and Pulping Chemistry, Madison USA. pp 291-294

Fackler K, Gradinger C, Hinterstoisser B, Messner K, Schwanninger M (2006) Lignin degradation by white rot fungi on spruce wood shavings during short-time solid-state fermentations monitored by near infrared spectroscopy. Enzyme and Microbial Technology 39 (7):1476-1483

Fackler K, Gradinger C, Schmutzer C, Tavzes C, Burgert I, Schwanninger M, Hinterstoisser B, Watanabe T, Messner K (2007) Biotechnological wood modification with selective white-rot fungi and its molecular mechanisms. Food Technology and Biotechnology 45 (3):269-276

Faix O, Mozuch MD, Kirk TK (1985) Degradation of gymnosperm (guaiacyl) vs. angiosperm (syringyl/guaiacyl) lignins by Phanerochaete chrysosporium. Holzforschung - International Journal of the Biology, Chemistry, Physics and Technology of Wood 39 (4):203-208

75

Faix O, Beinhoff O (1988) FTIR spectra of milled wood lignins and lignin polymer models (DHP's) with enhanced resolution obtained by deconvolution. Journal of Wood Chemistry and Technology 8 (4):505-522

Faix O, Bremer J, Schmidt O, Stevanovic T (1991) Monitoring of chemical changes in white-rot degraded beech wood by pyrolysis-gas chromatography and Fourier transform infrared spectroscopy. Journal of Analytical and Applied Pyrolysis 21:147-162

Fromm J, Rockel B, Lautner S, Windeisen E, Wanner G (2003) Lignin distribution in wood cell walls determined by TEM and backscattered SEM techniques. Journal of Structural Biology 143 (1):77-84

Goacher R, Jeremic D, Master ER (2011) Expanding the library of secondary ions that distinguish lignin and polysaccharides in ToF-SIMS analysis of wood. Analytical Chemistry. 83:804-812.

Hakala TK, Maijal P, Konn J, Hatakka A (2004) Evaluation of novel wood-rotting polypores and corticioid fungi for the decay and biopulping of Norway spruce (Picea abies) wood Enzyme and Microbial Technology 34 (3-4):255-263

Hammel KE, Cullen D (2008) Role of fungal peroxidases in biological ligninolysis. Current Opinion in Plant Biology 11 (3):349-355

Hatakka A (1994) Lignin-modifying enzymes from selected white-rot fungi: production and role from in lignin degradation. FEMS Microbiology Reviews 13 (2-3):125-135

Hibbett DS, Donoghue MJ (2001) Analysis of character correlations among wood decay mechanisms, mating systems, and substrate ranges in homobasidiomycetes. Systems Biology 50:215-242

Hobro A, Kuligowski J, Döll M, Lendl B (2010) Differentiation of walnut wood species and steam treatment using ATR-FTIR and partial least squares discriminant analysis (PLS-DA). Analytical and Bioanalytical Chemistry 398 (6):2713-2722

Högberg N, Svegården IB, Kauserud H (2006) Isolation and characterization of 15 polymorphic microsatellite markers for the devastating dry rot fungus, Serpula lacrymans. Molecular Ecology Notes 6 (4):1022-1024

Irbe I, Andersons B, Chirkova J, Kallavus U, Andersone I, Faix O (2006) On the changes of pinewood (Pinus sylvestris L.). Chemical composition and ultrastructure during the attack by brown-rot fungi Postia placenta and Coniophora puteana International Biodeterioration and Biodegradation 57 (2):99-106

Jensen KA, Jr, Houtman CJ, Ryan ZC, Hammel KE (2001) Pathways for extracellular Fenton chemistry in the brown rot basidiomycete Gloeophyllum trabeum. Applied Environmental Microbiology 67:2705-2711

76

Jung S, Foston M, Sullards CM, Ragauskas AJ (2010) Surface characterization of dilute acid pretreated Populus deltoides by TOF-SIMS. Energy Fuels 24:1347-1357

Kelley SS, Jellison J, Goodell B (2002) Use of NIR and pyrolysis-MSMS coupled with multivariate analysis for detecting the chemical changes associated with brown-rot biodegradation of spruce wood. FEMS Microbiology Letters 209 (1):107-111

Kersten P, Cullen D (2007) Extracellular oxidative systems of the lignin-degrading Basidiomycete Phanerochaete chrysosporium. Fungal Genetics and Biology 44 (2):77-87

Kirk TK, Schultz E, Connors WJ, Lorenz LF, Zeikus JG (1978) Influence of culture parameters on lignin metabolism by Phanerochaete chrysosporium. Archives of Microbiology 117:277-285

Kirk TK, Cullen D (1998) Enzymology and molecular genetics of wood degradation by white-rot fungi. In: Young RA, Akhtar M (eds) Environmentally Friendly Technologies for the Pulp and Paper lndustry. John Wiley & Sons, Inc,

Kubo S, Kadla JF (2005) Hydrogen bonding in lignin: A Fourier Transform infrared model compound study. Biomacromolecules 6 (5):2815-2821

Lojewska J, Miskowiec P, Lojewski T, Proniewicz LM (2005) Cellulose oxidative and hydrolytic degradation: In situ FTIR approach. Polymer Degradation and Stability 88 (3):512-520

Lundell TK, Mäkelä MR, Hildén K (2010) Lignin-modifying enzymes in filamentous basidiomycetes – ecological, functional and phylogenetic review. Journal of Basic Microbiology 50 (1):5-20.

MacDonald J, Doering M, Canam T, Gong Y, Guttman D, Campbell M, Master ER (2011) Transcriptomic analysis of the softwood-degrading white-rot fungus Phanerochaete

carnosa reveals concerted gene expression responses to growth on coniferous and deciduous wood. Applied and Environmental Microbiology. [Epub ahead of print]

Mahajan S, Master ER (2010) Proteomic characterization of lignocellulose-degrading enzymes secreted by Phanerochaete carnosa grown on spruce and microcrystalline cellulose. Applied Microbiology and Biotechnology 86 (6):1903-1914

Marin S, Ramos AJ, Sanchis V. (2005) Comparison of methods for the assessment of growth of food spoilage moulds in solid substrates. International Journal of Food Microbiology 99:329-341

Martinez D, Larrondo LF, Putnam N, Sollewijn Gelpke MD, Huang K, Chapman J, Helfenbein KG, Ramaiya P, Detter JC, Larimer F, Coutinho PM, Henrissat B, Berka R, Cullen D, Rokhsar D (2004) Genome sequence of the lignocellulose degrading fungus Phanerochaete chrysosporium strain RP78. Nature Biotechnology 22:695-700

77

Martinez D, Challacombe J, Morgenstern I, Hibbett D, Schmoll M, Kubicek CP, Ferreira P, Ruiz-Duenas FJ, Martinez AT, Kersten P, Hammel KE, Vanden Wymelenberg A, Gaskell J, Lindquist E, Sabat G, Splinter BonDurant S, Larrondo LF, Canessa P, Vicuna R, Yadav J, Doddapaneni H, Subramanian V, Pisabarro AG, Lavin JL, Oguiza JA, Master E, Henrissat B, Coutinho PM, Harris P, Magnuson JK, Baker SE, Bruno K, Kenealy W, Hoegger PJ, Kües U, Ramaiya P, Lucas S, Salamov A, Shapiro H, Tu H, Chee CL, Misra M, Xie G, Teter S, Yaver D, James T, Mokrejs M, Pospisek M, Grigoriev IV, Brettin T, Rokhsar D, Berka R, Cullen D (2009) Genome, transcriptome, and secretome analysis of wood decay fungus Postia placenta supports unique mechanisms of lignocellulose conversion. Proceedings of the National Academy of Sciences 106 (6):1954-1959

Pandey KK, Pitman AJ (2003) FTIR studies of the changes in wood chemistry following decay by brown-rot and white-rot fungi. International Biodeterioration and Biodegradation 52:151-160

Pandey KK, Nagveni HC (2007) Rapid characterisation of brown and white rot degraded chir pine and rubberwood by FTIR spectroscopy. European Journal of Wood and Wood Products 65 (6):477-481

Rowell RM (2005) Handbook of Wood Chemistry and Wood Composites. CRC Press,

Saito K, Kato T, Tsuji Y, Fukushima K (2005) Identifying the characteristic secondary ions of lignin polymer using TOF-SIMS. Biomacromolecules 6:678-683

Saka S, Whiting P, Fukazawa K, Goring DAI (1982) Comparative studies on lignin distribution by UV microscopy and bromination combined with EDXA. Wood Science and Technology 16:269-277

Scott GM, Akhtar M, R.E S, Houtman CJ (2002) Recent developments in biopulping technology In: R L, L V (eds) Biotechnology in the pulp and paper industry, Madison WI, Elsevier Science

Singh D, Chen S (2008) The white-rot fungus Phanerochaete chrysosporium: conditions for the production of lignin-degrading enzymes. Applied Microbiology and Biotechnology 81:399-417

Sjöström E (1983) Wood Chemistry, Second Edition: Fundamentals and Applications. Academic Press,

Spurr AR (1968) A low-viscosity epoxy resin embedding medium for electron microscopy Journal of Ultrastructure Research 26 (1-2):31-43

Stein BD, Klomparens KL, Hammerschmidt R (1992) Comparison of bromine and permanganate as ultrastructural stains for lignin in plants infected by the fungus Colletotrichurn lagenariurn. Microscopy research and technique 23:201-206

78

Stephan T, Zehnpfennig J, Benninghoven A (1994) Correction of dead time effects in time-of-flight mass spectrometry. Journal of Vacuum Science and Technology A 12:405-410

Tuor U, Winterhalter K, Fiechter A (1995) Enzymes of white-rot fungi involved in lignin degradation and ecological determinants for wood decay. Journal of Biotechnology 41 (1):1-17

79

Chapter 5 : Proteomic Characterization of Lignocellulose-degrading Enzymes Secreted by Phanerochaete carnosa

Grown on Spruce and Microcrystalline Cellulose

5.1 Abstract

Proteins secreted by the white-rot, softwood degrading fungus Phanerochaete carnosa during

growth on cellulose and spruce were analyzed using tandem mass spectrometry and de novo

sequencing. Homology driven proteomics was applied to compare P. carnosa peptide sequences

to proteins in P. chrysosporium using MS BLAST and non-gapped alignment. In this way, 665

and 365 peptides from cellulose and spruce cultivations, respectively, were annotated. Predicted

activities included endoglucanases from glycoside hydrolase (GH) families 5, 16, and 61,

cellobiohydrolases from GH6 and GH7, GH3 β-glucosidases, xylanases from GH10 and GH11,

GH2 β-mannosidases, and de-branching hemicellulases from GH43 and CE15. Peptides

corresponding to glyoxal oxidases, peroxidases, and glycopeptides that could participate in lignin

degradation were also detected. Overall, predicted activities detected in extracellular filtrates of

spruce cultures represented a subset of those identified in cellulose cultures, suggesting that the

adaptation of P. carnosa to growth on lignocellulose might result from fine-tuning the expression

of similar enzyme families.

5.2 Introduction

Genomic and proteomic analyses of basidiomycetes are providing more complete descriptions of

enzyme combinations that play a critical role in carbon-cycling and can be applied in the

production of energy and chemicals from lignocellulosic biomass. Progress in this area is

exemplified by recent genomic and proteomic analyses of the white-rot fungus Phanerochaete

chrysosporium and the brown-rot fungus Postia placenta (Martinez et al. 2004; Vanden

Wymelenberg et al. 2009; Martinez et al. 2009). For instance, less than 20 % of predicted

glycoside hydrolases encoded by P. chrysosporium were characterized before the genome

sequence was published. And while proteomic analysis of P. chrysosporium grown on cellulose

and wood substrates consistently detect the expression of GH3 β-glycosidases, GH6 and GH7

80

cellobiohydrolases, GH10 xylanases, and GH12 endoglucanases, more than 65 % of the proteins

secreted by P. chrysoporium grown on crystalline cellulose corresponded to previously

uncharacterized glycoside hydrolases, esterases and proteases, as well as proteins with unknown

function (Abbas et al. 2005; Ravalason et al. 2008; Sato et al. 2007; Wymelenberg et al. 2005;

Vanden Wymelenberg et al. 2006; Vanden Wymelenberg et al. 2009). By contrast, genomic and

proteomic analysis of P. placenta identified oxidases predicted to generate extracellular Fe2+ and

H2O2 and relatively few cellulolytic glycoside hydrolases (GH) (Martinez et al. 2009). The

apparent lack of GHs encoded by P. placenta that contain carbohydrate-binding modules was

particularly striking.

As described in previous chapters, Phanerochaete carnosa is a white-rot fungus isolated from

balsam fir (Abies balsamea) as well as other softwoods, and shown to belong to the family

Phanerochaete sensu strictum (de Koker et al. 2003). Given the abundance of softwood and the

unique challenges associated with its hydrolysis, enzymes that have evolved to efficiently

transform softwood feedstocks are particularly relevant both industrially and for fundamental

investigation. To ascertain the diversity and regulation of carbohydrate-active and lignolytic

enzymes secreted by P. carnosa, a proteomic approach was applied to compare the secretome of

this fungus while growing on microcrystalline cellulose and spruce wood chips. This study

represents the first detailed characterization of the carbohydrate-active and lignin-degrading

enzymes produced by P. carnosa. Further, comparative proteomics was performed to reveal

enzymes and proteins secreted by P. carnosa that can accelerate the bioconversion of softwood

feedstocks to fermentable sugars. Since the P. carnosa genome sequence was not available

when this study was performed, the genome sequence of P. chrysosporium was used to facilitate

the annotation of peptide sequences retrieved. In this way, this study describes a proteomic

approach to investigate the enzymatic basis for feedstock preference, while also contributing to

the still limited number of investigations that have applied homology driven proteomics to

annotate peptides from organisms that lack genome sequence information (Junqueira et al. 2008).

81

5.3 Materials and Methods

5.3.1 Cultivation Conditions

As previously described, Phanerochaete carnosa strain HHB-10118 was obtained from the US

Department of Agriculture (USDA) Forest Products Laboratory, Madison, WI. The strain was

grown on liquid or solid medium under stationary conditions at 27 °C, and was maintained on

YMPG agar slants (yeast extract (2 % w/v), maltose (10 % w/v), peptone (2 % w/v), glucose (1

% w/v) and agar (2 %w/v)) at 4 °C.

Cellulose medium was prepared in 1 L flasks and contained 1 % wood-derived microcrystalline

cellulose (Avicel PH-101, Fluka) submerged in 200 mL B3 medium (low nitrogen, 0.1 mg L-1

thiamine HCl, 0.2 mg L-1 ammonium tartrate, 10 mM dimethyl succinate, pH 4.5) (Kirk et al.

1978). Wood cultivations were prepared in autoclavable 22 X 28 cm polypropylene bags

containing a gas exchange filter (Unicorn Imp & Mfg Corp TX USA) (Sato et al. 2007).

Briefly, chips of white spruce (Picea glauca) measuring approximately 0.1 cm X 0.2 cm X 1.2

cm were washed with milliQ water, steam sterilized for 30 min, and 225 g of the chips were

submerged in 350 mL of B3 medium. The inoculum was obtained from P. carnosa cultivations

grown for 7 days at 27 °C in liquid YMPG medium. The mycelium was filtered through sterile

Miracloth, rinsed with B3 buffer (pH 4.5), and then blended three times for 10 s using a sterile

Waring blender. The mycelium suspension was transferred to each culture to obtain an initial

mycelium concentration of 40 mg (fresh weight) 100 mL-1. Cellulose and spruce cultivations

were grown at 27 °C under stationary conditions for 4 and 12 weeks, respectively. Triplicate

cultivations were prepared for each growth condition; cellulose and spruce growth medium

without fungal inoculation served as abiotic controls. Extents of fungal growth on cellulose was

determined by measuring final dry weight and total chitin in each cultivation (Plassard et al.

1982).

5.3.2 Protein Extraction

Following vacuum filtration of cultivation supernatants through Miracloth, mycelium was rinsed

with B3 buffer and the filtrate was recovered then pooled with the corresponding culture

supernatant. Liquid fractions were centrifuged at 5300 x g for 20 min at 4 °C, and then filtered

using 0.45 µm syringe filters. The filtrate was concentrated approximately 20 times using

82

Amicon Centrifugal devices (5000 MWCO polyethersulfone membrane; Centricon-20 cat:

UFC2BCC24). Protein concentration was measured by the Bio-Rad Bradford assay, and

concentrated samples were stored at -80 ˚C for proteomic analysis.

5.3.3 Protein Preparation and Analysis by Mass Spectrometry

Protein preparations (20 µg each) were fractionated by 1-D SDS-PAGE and each lane was

excised into 10 fractions using a clean razor blade. In-gel trypsin digestion was then performed

by treating each gel segment with 40 µL of 50 mM NH4HCO3 and 12.5 ng µL-1 of modified

trypsin (Promega, sequencing grade) (Shevchenko et al. 1996). Peptides were extracted from the

gel by shaking for 20 min at room temperature with 50 µL of 5 % formic acid twice, and then

with 50 µL of 5 % formic acid in 30 % acetonitrile. The peptide extracts were concentrated to

roughly 20 µL by vacuum centrifuge. Concentrated peptide samples were loaded onto a

HPLC_Chip (160 nL high capacity sample enrichment column and 75-µm x 150 mm SB-C18

separation column, Agilent Technologies, Santa Clara, CA, USA) and separated by flow rate at

300 nL min-1, with solvent A (0.2 (v/v) % formic acid in water) and solvent B (100 %

acetonitrile) and the following gradients: 3 %, 35 %, 80 %, 100 % of solvent B at 0, 50, 54 and

56 min after injection, respectively. The LC-MS/MS analysis was carried out by an Agilent 1100

HPLC-chip and 6340 ion trap system with MS scan range from 300 to 1,300 m/z. Three

precursor ions were selected from each MS scan for back-to-back collision induced dissociation

and electron transfer dissociation (CID/ETD). A 30 s dynamic exclusion period was applied to

the precursor previously selected twice for MS/MS, i.e., two rounds, each with CID/ETD.

(SunnyBrook Health Science Centre, Proteomics Core Facility).

5.3.4 Peptide Sequence Annotation

Since the P. carnosa genome sequence was not available at the time of analysis, it was necessary

to identify peptide spectra through de novo sequencing. The SHERENGA de novo sequencing

tool (version A.03.03.084, Agilent Technologies) was used to explicitly read peptide sequences

from high-quality fragment ion spectra (Dancik et al. 1999). Peptide sequences that received a

SHERENGA score greater than 100 were matched to all predicted proteins from Phanerochaete

chrysosporium (v.2.1) using MS BLAST ((Shevchenko et al. 2001),

http://genetics.bwh.harvard.edu/msblast/index.html). To increase the confidence of protein

83

matches based on a homology-driven approach, only alignments receiving an MS BLAST score

of 55 or above were considered, which is consistent with other scoring schemes based on MS

BLAST analyses (Shevchenko et al. 2001, Junqueira et al. 2008). Peptide sequences were also

compared to predicted proteins from Phanerochaete chrysosporium using BLASTp and the

following parameter settings: no gaps, word length 5, and gap extension cost 100. The PAM30

substitution matrix was selected given the short peptide sequences that were queried in the

BLAST search. Initial analyses were automated using freely available BioPython source code

(Cock et al. 2009), and peptide matches were manually inspected. Percent identity was computed

as the number of contiguous amino acid matches divided by the total peptide length. To avoid

over prediction of protein hits, protein sequences receiving only one peptide match were ignored

if the corresponding peptide aligned to another protein with additional peptide matches. The

MS/MS spectra of peptides that were annotated were then manually interpreted to verify the

validity of de novo sequences predicted using SHERENGA. Non-gapped direct alignments using

BLASTp was also performed to compare peptide sequences from extracellular filtrates of

cellulose and spruce cultivations.

5.4 Results

5.4.1 Preparation and Analysis of Peptide Samples

Cellulose and spruce cultivations prepared for proteomic analyses of P. carnosa extracellular

filtrates were grown at 27 °C. While cellulose cultivations were harvested after four weeks of

incubation, spruce cultivations were grown for an additional eight weeks given the comparative

recalcitrance of the substrate.

Protein secreted from Phanerochaete carnosa grown on cellulose and spruce was fractionated by

SDS-PAGE to confirm that concentrated proteins from triplicate cellulose and spruce

cultivations generated reproducible protein profiles. Approximately 5.4 mg and 9.8 mg of protein

were recovered from 200 mL of cellulose-grown and 350 mL of spruce-grown culture

supernatant, respectively. Proteins were fractionated by one-dimensional gel electrophoresis,

each lane was cut into at least ten segments, and each section was processed by in-gel trypsin

digestion. Peptides eluted from each gel segment were then analyzed by LC-MS/MS. This

procedure enabled the characterization of proteins that might not have been detected by two-

84

dimensional gel electrophoresis. This approach also avoided reported difficulties associated with

proteins extracted from wood-grown cultivations, which produce streaks in electrophoresis gels

that mask protein spots (Sato et al. 2007).

De novo sequencing was performed to explicitly read peptide sequences. In total, 3222 and 1763

peptide sequences were recovered from cellulose and spruce cultivations of P. carnosa,

respectively. Of these, 665 and 365 could be annotated based on sharing at least 70 % sequence

identity to predicted proteins from the white-rot fungus Phanerochaete chrysosporium. Notably,

direct non-gapped alignments also indicated that only 16 % (284) of the peptides obtained from

extracellular filtrates of spruce cultures shared greater than 70 % identity to peptides extracted

from cellulose cultivations.

Peptide sequences with best matches to predicted glycoside hydrolases, esterases,

oxidoreductases, monooxygenases, and proteins with unknown function were manually

inspected; all hits reported here correspond to proteins with predicted secretory signal sequences

and that lack predicted transmembrane sequences (Fig. 5.1). In most cases, non-gapped direct

alignments were able to validate the MS BLAST results, since many of the same peptide matches

were predicted in both cases (Tables 5.1 to 5.3). However, the percent identity of peptide and

protein sequences from P. carnosa and P. chrysosporium, respectively, did not correlate well to

MS BLAST score. Therefore, matches to enzyme families identified through both MS BLAST

and non-gapped alignment are reported in Tables 5.1 and 5.2, while enzyme families that were

predicted using non-gapped alignment only are summarized in the supplemental material

(Appendix 3, Table S5.1). Since P. carnosa peptide sequences were annotated using predicted

protein sequences from a different, albeit related, organism, P. carnosa peptide annotations are

discussed below in terms of predicted enzyme activity rather than expression of a specific

protein.

5.4.2 Cellulases and Hemicellulases in Cellulose and Spruce Cultivations

Cellulose degradation by white-rot fungi is mediated by the concerted activity of three

cellulolytic enzymes: endoglucanases, cellobiohydrolases and β-glucosidases (Baldrian and

Valaskova 2008). Evidence of these cellulolytic activities was found in the secretome

85

B.

Fig. 5.1. Distribution of peptide annotations from proteins produced by P. carnosa grown

on (A) crystalline cellulose, and (B) spruce wood chips. Unique peptide sequences annotated

to proteins with predicted signal sequences using both MS BLAST and non-gapped alignment

are summarized.

of P. carnosa grown on cellulose (Table 5.1). While over ten unique peptide sequences

corresponding to GH7 cellobiohydrolases were detected, only one peptide corresponding to a

GH6 cellobiohydrolase was detected. Most of the peptide matches to family 7 cellobiohydrolases

Glycoside Hydrolase (16, 16.5%)

Carbohydrate Esterase, Lipase(7,7%)

Peptidase (19, 20%)

Oxidoreductase (19, 20 %) Glycosyl Transferase (3, 3 %)

Monooxygenase (3, 3%)

Extracellular - Structural (5, 5 %)

Transport (11, 11%)

Unknowns (14, 14.5 %)

Glycoside Hydrolase (49, 21%)

Carbohydrate Esterase, Lipase (10, 4 %)

Peptidase (26, 11%)

Oxidoreductase (20, 9 %)

Glycosyl Transferase (7, 3 %)

Monooxygenase (8, 3%) Extracellular - Structural (17, 7%)

Transport (37, 16%)

Unknowns (56, 24%)

A.

86

were distributed over conserved regions of this GH family. This result confounds the ability to

predict the number of unique GH7 enzymes secreted by P.carnosa. However, since only those

protein hits containing at least one unique peptide match are listed in Table 5.1, the higher

number of peptides corresponding to a GH7 cellobiohydrolase likely results from the secretion of

multiple GH7 isozymes. MS BLAST analyses identified three unique peptide sequences

corresponding to GH3 glycoside hydrolases in the culture of P. carnosa grown on cellulose. GH

family 3 enzymes exhibit a broad range of substrate specificity, including β-glucosidases that

hydrolyze cellobiose to glucose.

Comparatively few peptide sequences corresponding to GH families with uniquely endo-

cellulolytic activities were identified in extracellular filtrates of P. carnosa grown on cellulose or

spruce. Non-gapped alignments identified peptide sequences that were matched to GH5 and

GH74 enzymes in P. chrysosporium; so it is conceivable that GH5 and GH74 enzymes secreted

by P. carnosa catalyze additional cellulolytic activity (Appendix 3 Table S5.1). Five peptide

sequences corresponding to GH61 endoglucanases were also identified in cellulose cultivations

(two by both MS BLAST and BLASTp). The Cel61A enzyme from Aspergillus kawachii has

endoglucanase activity, and so it is tempting to conclude that GH61 enzymes could play a role in

cellulose degradation by P. carnosa. However, the biological function of GH61 enzymes

remains unclear. Recent structural analyses of a GH61 enzyme could not unambiguously assign

amino acid residues involved in catalysis and substrate binding (Karkehabadi et al. 2008).

87

Table 5.1. Analysis of peptides from extracellular filtrates of P. carnosa grown on cellulose and spruce that correspond to

predicted cellulases and hemicellulases.

Predicted Activity

GH

Substrate

MS BLAST

Non-gapped Alignment

Pchr IDa

P. carnosa Peptide Sequence

Scoreb Pchr IDa

P. carnosa Peptide Sequence

Pchr Sequence Match

% ID

CBH II GH6 Cellulose 133052 WGDWCNI 66 133052 WGDWCNIQ WGDWCNI 100

CBH I GH7 Cellulose

126964

DTDFFSQHGGI 321

126964

CTAGTGFCV

CTSNTGFCD

75

GLCDADGCDFNSFR

NTGICDGDGCDFDS

YGTGYCDSQC

Cellulose/ Spruce

127029

FGDTDFFSQHGGIc 540

127029

FGDTDFFSQHGGc FGDTNYFAQHGG 75

FVTGSNVGc TDNFCTQQQc TDNFCSQQ 87

LMSDDSHYQMIQc LMSDDSHYQMc LMADDTHYQM 80

GLCDADGCDFNSFRc GLFGDGSVDFNSFd

GLCDADGCDFNSFR

57

NTGICDGDGCDFDSc TDNFCTQQc

YGTGYCDSQCc

137216

FGDTDFFSQHGGIQc 461

137216

YGQGAGTQSc YGQQAGTQT 87

GTQTPETHPc GTQTPETHPSLSc

GTQTAETHPQLT

75

GLCDADGCDFNSFRc

FVTGSNVGc

NTGLCDGDGCDFDSc

YGTGYCDSQCc

VTQDTSVVIDGNd

GTQTPETHPSLSXQd

137372 GLCDGDGCDFNSFR 520 137372 GLCDGDGCDFNSFRc GLCDGDGCDFNSF 100

88

Predicted Activity

GH

Substrate

MS BLAST

Non-gapped Alignment

Pchr IDa

P. carnosa Peptide Sequence

Scoreb Pchr IDa

P. carnosa Peptide Sequence

Pchr Sequence Match

% ID

R

FVTGSNVG NTGLCDGDGCDFDSRc

78

NTGLCDGDGCDFDS VDGDGCDFDSc 72

LMSDDSHYQMI FVTGSNVd FVTGSNV 100

FGDTDFFSQHGGI

TDNFCTQQ

YGTGYCDSQC

Possible Endo-glucanase

GH61

Cellulose

41650 QTDVTGLSWF 66 41650 QTDVTGLSWFQ QTDVTGLKWFK 81

122129 DDTYTTSWAVDR 89 122129 DDTYTTSWAVDR DTYTTSWAVDR 91

41563

QTDVTGLSWF

59

129325 VGPSGPTGE PSGPTG 100

Glyco-sidase / BGL

GH3

Cellulose

139063

ITFSVGAS 189

139063

ITFSVGAS ITFSVGAS 100

GLCLEDSPLG GLCLEDSPLG GLCLEDSPL 100

HYINNEQE

HYINNEQEFDR HYINNEQE 100

PIAQGSTT PIAEGSTT 87

GNIPAI GNIPAI 100

9257 IVCNPSADVV VCNASADPV 77

Spruce 133018 WEAEGFN WESEGFD 71

β-manno-sidase

GH2

Cellulose

135385

ITEPLLGINEFTQR 336

135385

ITEPLLGINEFTQR ITEPLLGINEFTQR 100

SPVFIEESSLEI SPVFIEESSLEI SPIFIEESSLEIS 91

LESAVLSGQNMLR LESAVLSGQNMLR LESAVLSGQNMLR 100

89

Predicted Activity

GH

Substrate

MS BLAST

Non-gapped Alignment

Pchr IDa

P. carnosa Peptide Sequence

Scoreb Pchr IDa

P. carnosa Peptide Sequence

Pchr Sequence Match

% ID

TPGFIYGNTDR

TPGFIYGNTDR TAGFIYGNSER 72

YSPPVIY YSPVVIY 85

Mannan-ase

GH5

Cellulose

5115 EPTILGWELANE 82 5115 EPTILGWELANE EPTILAWELANE 91

140501

DIAGAGSTTVR 126

140501

DIAGAGSTTVR DIANAGSTVVR 82

EPTILGWELANE EPTILGWELANE EPTIMAWELSNE 75

Xylanase

GH10

Cellulose 125669 LPSTPALLQ 58

139732

DVCNEMQNEDGT 145

INDYNLDSANA

7852 VSVWGVSR VSVWGVS 100

Cellulose/Spruce

138345

GVDGIAIGFa PAYDGIAIGF 77

PGAAGIAIGFa PAYDGIAIGF 70

ENIYNIEYAb INEYNIEYA 87

GH11

Cellulose

133788

TSDGSSYDVYE

70

133788

TSDGSSYDVYE

TSDGATYDVYE

81

Endo-1,3(4)-β-glucanase

GH16 Cellulose 3846 YDNAYFEVR 59 3846 YDNAYFEVR YDEAYFEIR 77

Arabino-furanosi-dase

GH43

Cellulose

4822

NDGAIEASVIW

78

4822

NDGAIEASVIWQ

NNGAIEASVIW

90

Cellulose/Spruce

333

ITGGGGIGASNSa ITGTGGIGTS 80

DIVISGFAPb IISGFAP 85

90

Predicted Activity

GH

Substrate

MS BLAST

Non-gapped Alignment

Pchr IDa

P. carnosa Peptide Sequence

Scoreb Pchr IDa

P. carnosa Peptide Sequence

Pchr Sequence Match

% ID

NA

Cellulose

3651

IAETLVEMG 133

3651

IAETLVEMG IAEALVEMG 88

PTMESSSVSAAFMT NLEGQTTATR NLGQTTATR 100

Spruce 38548 YENEVAIR 57 38548 YENEVAIR YENEVSIR 87

91

a Additional information for protein from P. chrysosporium can be accessed by ending the

following URL with the model number, e.g. http://genome.jgi-psf.org/cgi-

bin/dispGeneModel?db=Phchr1&id=6482; bAnnotation score from MS BLAST; c Detected in

extracellular filtrates of cellulose cultivations; d Detected in extracellular filtrates of spruce

cultivations. Figure S5.1 summarizes Table 5.1 as pie graphs.

Both MS BLAST and non-gapped alignments matched peptide sequences from P. carnosa to

Pchr5115 and Pchr140501, which are predicted to encode mannanase activity and were detected

in P. chrysosporium cultivations grown on Wood’s medium and spruce (Ravalason et al. 2008;

Vanden Wymelenberg et al. 2009). Both methods also identified a GH2 mannosidase in P.

carnosa cultures grown on cellulose. While a GH2 mannosidase was detected in extracellular

filtrates of P. chrysosporium maintained on replete medium (Vanden Wymelenberg et al. 2009),

GH2 enzymes have not been recovered by proteomic analyses of P. chrysosporium grown on

cellulose or woody substrates.

Seven peptides sequences matching GH10 xylanases were detected in both cellulose and spruce

cultivations. By contrast, only one peptide sequence from cellulose cultures was matched to a

GH11 xylanase. Both MS BLAST and non-gapped alignments matched a similar number of

peptide sequences to α-arabinofuranosidases, supporting other studies that emphasize the

importance of debranching enzymes in the hydrolysis of lignocellulosic substrates (de Vries et al.

2000).

5.4.3 Oxidoreductases in Cellulose and Spruce Cultivations

Peptides recovered from both cellulose and spruce cultivations were matched to Pchr4636 (Table

5.2), a manganese peroxidase (MnP) that was previously detected in P. chrysosporium

cultivations grown on spruce (Ravalason et al. 2008). Non-gapped alignment also identified

peptides from spruce cultivations that matched Pchr131707, a lignin peroxidase (LiP) that is

encoded by a gene located on scaffold 19 of the P. chrysosporium genome (Appendix 3 Table 1).

While the expression cellobiose dehydrogenease activity (CDH) was not detected, peptides from

extracellular filtrates of P. carnosa grown on spruce were matched to a putative multicopper

92

oxidase and Cir1, a CBM-containing reductase that is related to CDH1 (Table 5.2, Appendix 3

Table 1). Multicopper oxidases are not implicated in lignin degradation by P. chrysosporium;

instead, they may function as extracellular ferroxidases with potential to modulate Fenton

reactions (Kersten and Cullen 2007; Larrondo et al. 2003). Notably, the majority of peptide

matches to oxidoreductases were to enzymes thought to be involved in extracellular peroxide

generation, including glyoxal oxidases and GMC oxidoreductases (Table 5.2). Further, a peptide

sequence matching Pchr1923 was detected in culture filtrates of P. carnosa grown on cellulose.

Pchr1923 shares over 70 % sequence identity to deduced amino acid sequence of glp2, which

encodes a glycopeptide implicated in the production of hydroxyl radicals involved in lignin and

carbohydrate degradation (Baldrian and Valaskova 2008; Tanaka et al. 2007).

Peptides corresponding to cytochrome P450s and monooxygenases were also detected in both

cellulose and spruce cultivations (Table 5.2, Appendix 3 Table 1). Notably, Pchr37971, Pchr663,

Pchr38849 and Pchr139445 are CYP2E-type enzymes, a class of P450 monooxygenases that

participate in the bioconversion of exogenous aromatic compounds. Although only a few

peptides corresponding to monooxygenase activity were detected in spruce cultivations, the

dependence of monooxygenase expression on biomass feedstock composition will be interesting

to pursue given the potential of this enzyme family to degrade oleoresins in softwood.

93

Table 5.2. Analysis of peptides from extracellular filtrates of P. carnosa grown on cellulose and spruce that correspond to

extracellular oxidases and proteins involved in lignin degradation.

Predicted Activity

Substrate

MS BLAST Non-gapped Alignment

Pchr IDa

P. carnosa Peptide Sequence

Scoreb Pchr IDa

P. carnosa Peptide Sequence

Pchr Sequence Match

% ID

Monooxygenase P450

Cellulose

37971

EEQSAVTEEIN

58

663 GFIFEMIGWR IFERIGW 85

38849 QTDVSAISWFQ DVSPISWF 87

139445 IFITVA IFITVA 100

138612 IVIGDGLPD IVIGDGL 100

Spruce 5081 SVGSHVATE GSHVAT 100

GMC

Cellulose

6270

GGSSSIDGAAW

71

6270

GWGWSGER GWGWSG 100

AFIATCQ AFIATC 100

WVVIVGHTA VIVGGGTA 66

11098 ATYLQTAL ATYLQTAL 100

Spruce

9508 RETAHTQATF AHTQAT 100

6017 GIVGGGTSGS IVGGGTSG 100

Cellulose/Spruce

37188

VGTAATDADIVA 183

37188

VGTAATDADIVc VGTAESDAEIVAA 72

GTAATDADIVAA GTAATDADIVAAc VGGGTAGNVVA

75

VGGGGTAGNIVA PVGTAATDADIVAc 75

VGTAATDADIVAAd

67

SSVGGGGTAGNIVAc

83

PVGTAATDADIVAAd 76

Glyoxal oxidase

Cellulose

8882

IVAATSSTH 311

8882

TVAATSSTH VAATSSTH 100

VAATSSTHGN VAATSSTHGNPA VAATSSTHGN 100

ACYVDNA ACYVDNAD

ACYVDNA

100

IGGWSLQSTQGVR

VGYYNEAR

94

Predicted Activity

Substrate

MS BLAST Non-gapped Alignment

Pchr IDa

P. carnosa Peptide Sequence

Scoreb Pchr IDa

P. carnosa Peptide Sequence

Pchr Sequence Match

% ID

Cellulose/ Spruce

11068

PANSFEFFc

63

11068

PANSFEFF PANSFEFF 100

TFTAPPDc,d TFTAPP 85

Peroxidase

Cellulose/ Spruce

4636 LQSDFALARc,d 61 4636 QPATFIAADGPQd PATFPATKGPQ 72

Spruce 36058 SGRSHAVAI GRSHAV 100

Multicopper oxidase

Spruce 132237 LDIPQAIID 61 132237 LDIPQAIIDY LDIPQAIVD 88

Glycopeptides

Cellulose

1932 TVTFVNNCGFGT 77 1932 VTFVDNCGFGT

TVTFVDICGFGT a Additional information for protein from P. chrysosporium can be accessed by ending the following URL with the model number, e.g.

http://genome.jgi-psf.org/cgi-bin/dispGeneModel?db=Phchr1&id=37971; bAnnotation score from MS BLAST; c Detected in

extracellular filtrates of cellulose cultivations; d Detected in extracellular filtrates of spruce cultivations.

95

5.4.4 Carbohydrate Esterases, Proteases and Other Glycoside Hydrolases

Approximately 10 % of the peptide sequences identified from P. carnosa culture filtrates were

matched to predicted esterases in P. chrysosporium. Lipase and esterase enzymes typically

demonstrate broad substrate specificity, and have been implicated in the hydrolysis of lignin-

carbohydrate complexes. For instance, glucuronoyl esterases were recently identified and shown

to hydrolyze linkages between lignin and glucuonate branching sugars of xylan (Spanikova et al.

2007). Both MS BLAST and non-gapped alignment identified peptides from cellulose and

spruce cultivations that matched Pchr6482, a predicted glucuronoyl esterase from P.

chrysosporium. Non-gapped alignment identified three additional sequences from cellulose

cultivations with greater than 75 % identity to Pchr130517, a second putative glucuronoyl

esterase encoded by P. chrysosporium. Notably, in a separate study, a full-length cDNA

transcript isolated from P. carnosa was predicted to encode a protein that shares over 50 % and

80 % sequence identity to Pchr130517 and Pchr6482, respectively (data not shown).

Peptide sequences that were matched to proteolytic activities accounted for 25 % and 40 % of the

peptides analyzed from cellulose and spruce cultures, respectively (Fig. 5.1). Secreted proteases

could participate in nitrogen cycling and acquisition through hydrolysis of fungal proteins, as

well as plant proteins that are bound to lignocellulose. The relative susceptibility of

lignocellulose degrading enzymes to secreted fungal proteases remains to be elucidated.

In both cellulose and spruce cultivations, the majority of peptide sequences annotated to

extracellular glycoside hydrolases were matched to enzymes involved in cellulose and

hemicellulose degradation. Still, extracellular filtrates of P.carnosa grown on cellulose and

spruce also contained several peptides that were matched to amylases involved in starch

degradation (Table 5.3). MS BLAST and non-gapped alignment also identified peptides

sequences corresponding to enzymes involved in pectin degradation, namely a CE8 pectin

esterase and a GH28 pectinase, respectively (Table 5.3, Appendix 3, Table S5.1).

Similar to P. chrysosporium, P. carnosa appears to secrete a GH88 candidate d-4,5-unsaturated

β-glycosidase during growth on cellulose and spruce (Ravalason et al. 2008; Vanden

Wymelenberg et al. 2009) (Appendix 3, Table S5.1). GH88 d-4,5-unsaturated β-glycosidases

96

have been implicated in the degradation of unsaturated oligosaccharides, likely produced by

polysaccharide lyases (Nankai et al. 1999). Notably, peptides corresponding to a PL14 lyase

were also detected in cellulose cultivations (Table 5.3). Eight unique peptides from P.carnosa

culture filtrates were matched to GH92 α-1,2-mannosidases; only one peptide sequence was

matched to a GH47 α-mannosidase. While GH47 α-mannosidases are thought to participate in

post-translational modification of secreted proteins (Yoshida et al. 1993), most GH92 α-1,2-

mannosidases that have been identified and characterized to date are of bacterial origin.

Nevertheless, P. chrysosporium is predicted to encode four GH92 enzymes, and the Pchr1930

and Pchr133585 protein matches identified in my study correspond to genes that are upregulated

in P. chrysosporium during growth on cellulose (Vanden Wymelenberg et al. 2009).

97

Table 5.3. Analysis of peptides from extracellular filtrates of P. carnosa grown on cellulose and spruce that correspond to

extracellular carbohydrate-active enzymes.

Predicted Activity

GH Family

Substrate MS BLAST Non-gapped Alignment

Pchr IDa P. carnosa Peptide Sequence

Scoreb

Pchr IDa

P. carnosa Peptide Sequence

Pchr Sequence Match

% ID

exo-1,3-glucanase

GH55

Cellulose

8072

NPDGFADTITAWTR

103

8072 NPDGFADTITAW

NPNGFADTITAW

92

α-amylase

GH13

Cellulose

38357

GQTSDQALI

56

38357

GQTSDQALIQ GQTSDQSLI 88

ASDWQGR ASDWQSR 85

Spruce

130190

HDIDVLIDA

56

130190 HDIDVLIDAEQFIN

HGIDVLIDA 89

127674 VEADINGIR VEAEIDGER 66

GH15 Cellulose 138813 IDQYTSGQDTS 62 138813 IDQYTSGQDTS VDQYAAGQDTS 72

GH31

Cellulose

35408

EQYTDVI 169

35408

EQYTDVI

EQYTDVI

100

GHWLGDNY

LETMWTDID

125462

DGGVGTIQTMWAR

233

125462

DGGVGTIQTMWAR

DGGVGTVQTMWAR

92

GHWLGDNY IYGLGEVVASSG IYGLGEVVASSG 100

IYGLGEVVASSG

GDDGGFGTIQTFW

IGTDGGVGTVQTMW

69

PEQGAISQQSN EQGAISQ 87

968

TVDPDYF 186

968

TVDPDYF TVDPDYF 100

VTYETNTR VTYETNTR VTYETNTR 100

LETMWTDI SLETMWTDIDY LETMWTDI 100

135833 GHWLGDNY 185

98

Predicted Activity

GH Family

Substrate MS BLAST Non-gapped Alignment

GTIQTMWAR

IYGLGEVVASS

G Chitinase

GH18

Cellulose

39872 LSIGGWTGAR 63 39872 LSIGGWTGAR LSIGGWTGGR 88

128098 FIAEQGLR 58 128098 FIAEQGLR FIAEQGLR 100

β-galactosidase

GH35

Cellulose

9466

NPDTGAQFVIVR

272

9466

DTGAQFVIVR NPDTGAQFIIVR 90

INEGGLFGER GAQFVIVR NEGGLFGER

83

IFVDGWQY NPDEGAQFVIVR 83

LADYTFGSPA

INEGGLFGER 100

NPDTGAQFVNVR

91

NPDTGAQFVIVR 91

IFVDGWQY 85

DTGAQF DTGAQF 100

Spruce 134404 GITGAGGTVI GITGAG 100

α-mannosidasae

GH47

Cellulose

4550

MSVAWVGS

62

4550

MSVAWVGS MSVAWVGS 100

SAIQSFQQ SAIQSFQK 87

GH92

cellulose

1930 VGIDVDEVPR 65 1930 VGIDVDEVPR VGIDVDETPR 90

133585

EYAFEDFAIR

80

133585

EYAIGDFAVR EYAFEDFAIR 70

EYAFEDFAIR 100

35714

FNAGTGFMEAR 200

SFISVDQAR

YNDYAVSV

3431

AFDQWDELL 234

EYAFEDFAI

ISPADYTDAN

YANQPGLSTQ

99

Predicted Activity

GH Family

Substrate MS BLAST Non-gapped Alignment

Glucuronoyl-esterase

CE15

Cellulose

6482

MAGAFEER 120

6482

VIEVTPAAHV VIEVTPAAHV 100

IEVTPAAHV MAGAFEER MAGAFEER 100

130517

WVWGVSR WAWGVSR 85

PFIFNDGT PFVFNDGS 75

Spruce 6482 MAGAFEER 62

Pectin esterase CE8 cellulose 132137 AYFYGNTIAT 59

Polysaccharide lyase

PL14 cellulose 964 YTAYFPASFD 57

bAnnotation score from MS BLAST; c Detected in extracellular filtrates of cellulose cultivations; d Detected in extracellular filtrates of

spruce cultivations.

100

5.4.5 Peptides Annotated to Proteins with Unknown Function, Identified in Proteomic

Analyses of P. chrysosporium

An important outcome of proteomic analysis is the detection of proteins with unknown function

whose expression can be correlated to specific cultivation conditions. Seven unique peptide

sequences from cellulose cultivations were assigned to six P. chrysosporium proteins that have

predicted signal sequences but unknown function (Table 5.4). Matches to Pchr3328 and

Pchr138739 were to proteins previously identified in culture supernatant of P. chrysosporium

grown on spruce and cellulose (Vanden Wymelenberg et al. 2006; Ravalason et al. 2008; Vanden

Wymelenberg et al. 2009). As noted by Vanden Wymelenberg et al. (2006), Pchr138739 shares

homology to proteins also found in filamentous Ascomycetes. Further, the carboxy-terminus of

Pchr138739 contains a series of proline residues and is rich in serine and threonine, reminiscent

of linker sequences that connect catalytic domains and carbohydrate binding modules.

Table 5.4. Peptides from extracellular filtrates of P. carnosa grown on cellulose that were

annotated to proteins from P. chrysosporium with unknown function.

Pchr IDa P. carnosa Peptide Sequence Scoreb

5038 SNIEDLTIR 57

4562 IPGAVQVYGIES 61

3328

ISDTDFSQR 126

QVIEWTDID

3053 TIIDGDDFT 56

1934 EDDVPSMSA 57

138739 NPADTSETDGIR 69

a Additional information for protein from P. chrysosporium can be accessed by ending the

following URL with the model number, e.g. http://genome.jgi-psf.org/cgi-

bin/dispGeneModel?db=Phchr1&id=5038; bAnnotation score from MS BLAST.

101

5.5 Discussion

Proteomic analyses of wood-degrading microorganisms reveal a rich repertoire of biomass-

degrading proteins, many of which are yet to be characterized. In addition to identifying new

families of carbohydrate-active enzymes, correlating substrate composition to profile of secreted

proteins can inform the design of enzyme formulations that are most suitable for processing

specific biomass feedstocks. Boreal forests represent one of the largest biomes on Earth, and

bioprocessing softwood remains a significant challenge. Accordingly, the aim of this study was

to characterize the secretome of P.carnosa grown on spruce and microcrystalline cellulose to i)

uncover potential adaptations that allow this white-rot fungus to efficiently grow on softwood

feedstocks, and ii) evaluate substrate-dependent protein secretion by this organism. Given the

phylogenetic similarity of P. carnosa and P. chrysosporium, the genome sequence of P.

chrysosporium could be used to annotate sequenced peptides from P.carnosa; comparative

proteomics was also enabled.

A comparison of peptide annotations from both cellulose and spruce cultivations of P. carnosa to

proteomic data for P. chrysosporium reveals their potential to secrete similar families of

cellulolytic and lignin-degrading enzymes. For example, cellobiohydrolases, endoglucanases,

and β-glycosidases were identified in P. carnosa secretomes, as were peptide sequences

corresponding oxidative enzymes involved in lignin degradation. The relative abundance of

GH7 peptide sequences compared to GH6 peptide sequences that were identified in the P.

carnosa secretome suggests that like P. chrysosporium, GH7 isozymes are more highly

represented in P. carnosa than GH6. Also similar to P. chrysosporium, xylanolytic enzymes,

including GH10 and GH11 xylanases, debranching hydrolases, and glucuronoyl esterases were

identified in extracellular filtrates of P.carnosa cultivations. Notably, while GH2 β-

mannosidases have not been detected in extracellular filtrates of P. chrysosporium grown on

cellulosic substrates, a GH2 β-mannosidase was predicted in extracellular filtrates of P. carnosa

grown on cellulose and spruce. Since mannan is the main hemicellulose in softwood fibre, this

result raises the question whether up-regulation of mannan-degrading enzymes could contribute

to preferential growth of P. carnosa on softwood.

102

P. carnosa peptide sequences were also matched to a multicopper oxidase, Cir1, and two distinct

glycopeptides. These proteins have been implicated in lignocellulose degradation, but have not

been identified in previous proteomic analyses of P. chrysosporium. Cir1 and certain membrane-

bound glycopeptides are likely involved in generating hydroxyl radicals that participate in

carbohydrate and lignin degradation. Given the lack of peptide sequences from P. carnosa

extracellular filtrates that were matched to exclusively endo-acting cellulases, an intriguing

possibility is that oxidative enzymes play a significant role in cellulose degradation by P.

carnosa.

The range of predicted glycoside hydrolases detected in the culture medium of P. carnosa grown

on spruce was represented in the culture medium of P. carnosa grown on cellulose. These

results are consistent with those reported by Sato et al. (2007) and Vanden Wymelenberg et al.

(2005) who detect similar profiles of cellulases and hemicellulases in culture supernatants of P.

chrysosporium grown on cellulose and wood substrates. Moreover, these results support the

possibility that like P. chrysosporium, P. carnosa optimizes its growth on different

lignocellulosic substrates by modifying the relative abundance of a similar pool of glycoside

hydrolases and oxidative enzymes. Transcriptional analyses are currently underway to test this

hypothesis. Quantitative transcription profiles will also complement proteomic analyses by

narrowing the set of proteins with unknown function that are most important to lignocellulose

degradation. Still, 1229 peptides were unique to spruce cultivations and could not be annotated

based on identity to predicted proteins from P. chrysopsorium. Many of these will correspond to

intracellular proteins that are detected in culture supernatants due to cell lysis. However, some

will likely correspond to secreted or intracellular proteins involved in the hydrolysis or

metabolism of softwood components that enable the growth of P. carnosa on this feedstock. For

comparison, MS BLAST was used to align peptide sequences from P. carnosa to P. placenta

protein models. This analysis did not lead to the identification of additional GH families or

lignin-degrading enzymes in extracellular filtrates of P. carnosa grown on cellulose or spruce.

Notably, peptides from cellulose cultivations were most frequently matched to predicted GH31

enzymes in P. placenta, while peptides from spruce cultivations were most frequently matched

to predicted proteases. While the number of peptide matches to a particular protein family

provides some indication of the relative abundance of that family in culture supernatants,

103

differences in digestibility by trypsin, and subsequent ionization and detection by LC/MS limits

quantificative comparisons of proteins families detected in the current analysis.

In summary, proteomic analysis of extracellular filtrates of P. carnosa grown on cellulose and

spruce revealed a similar range of glycoside hydrolases and oxidative activities. The distribution

of cellulases and xylanases predicted in extracellular filtrates of P. carnosa were also similar to

those identified in cultures of P. chrysosporium. However, the profile of secreted proteins

detected in cultures of P. carnosa could be distinguished by the prediction of mannan-degrading

enzymes and oxidative enzymes that were not previously identified in proteomic analysis of P.

chrysosporium.

Clearly, the similarity of P. carnosa and P. chrysoporium was sufficient to apply a homology

driven proteomic approach to identify a core set of carbohydrate-active and lignin-degrading

enzymes that are secreted by P. carnosa grown on cellulosic substrates. However, in December,

2010, the P. carnosa genome sequence was completed and annotated using automated

algorithms, and so the proteomic data collected for P. carnosa was re-analyzed to determine

whether additional carbohydrate-active or lignin-degrading enzymes could be identified.

Through this analysis, additional 116 and 11 peptides that were predicted to originate from

lignocellulose-active enzymes were annotated from cellulose and spruce samples, respectively

(Appendix 3, Tables S5.2 and S5.3). Interestingly, new assignments that corresponded to lignin-

degrading activities were comparatively high in peptide samples originating from spruce

cultures. Is is anticipated that proteins identified using the P.carnosa genome but not by

comparison to P. chrysosoporium, comprise a starting point for the discovery of novel enzymes

that could accelerate the performance of a core set of lignocellulose-degrading enzymes on

softwood feedstocks.

Chapter 5 References

Abbas A, Koc H, Liu F, Tien M (2005) Fungal degradation of wood: initial proteomic analysis of extracellular proteins of Phanerochaete chrysosporium grown on oak substrate. Curr Gen 47: 49-56

104

Baldrian P, Valaskova V (2008) Degradation of cellulose by basidiomycetous fungi. FEMS Microbiol Rev 32: 501-521

Cock PJ, Antao T, Chang JT, Chapman BA, Cox CJ, Dalke A, Friedberg I, Hamelryck T, Kauff

F, Wilczynski B, de Hoon MJ (2009) Biopython: freely available Python tools for computational molecular biology and bioinformatics. Bioinformatics 25: 1422-1423

Dancik V, Addona TA, Clauser KR, Vath JE, Pevzner PA (1999) De novo peptide sequencing

via tandem mass spectrometry. J Comput Biol 6: 327-342 de Koker TH, Nakasone KK, Haarhof J, Burdsall Jr HH, Janse BJ (2003) Phylogenetic

relationships of the genus Phanerochaete inferred from the internal transcribed spacer region. Mycol Res 107: 1032-1040

de Vries RP, Kester HCM, Poulsen CH, Benen JAE, Viser J (2000) Synergy between enzymes

from Aspergillus involved in the degradation of plant cell wall polysaccharides. Carbohydrate Res 327: 401-410

Hibbett DS, Donoghue MJ (2001) Analysis of character correlations among wood decay

mechanisms, mating systems, and substrate ranges in homobasidiomycetes. Sys Biol 50: 215-242

Ishida T, Yaoi K, Hiyoshi A, Igarashi K, Samejima M (2007) Substrate recognition by glycoside

hydrolase family 74 xyloglucanase from the basidiomycete Phanerochaete

chrysosporium. FEBS J 274: 5727-5736 Junqueira M, Spirin V, Balbuena TS, Thomas H, Adzhubei I, Sunyaev S, Shevchenko A (2008)

Protein identification pipeline for the homology-driven proteomics. J Proteomics 71: 346-356

Karkehabadi S, Hansson H, Kim S, Piens K, Mitchinson C, Sandgren, M (2008) The first

structure of a glycoside hydrolase family 61 member, Cel61B from Hypocrea jecorina, at 1.6 A resolution. J Mol Biol 383: 144-154

Kersten P, Cullen D (2007) Extracellular oxidative systems of the lignin-degrading

Basidiomycete Phanerochaete chrysosporium. Fungal Genet Biol 44: 77-87 Kirk TK, Schultz E, Connors WJ, Lorenz LF, Zeikus JG (1978) Influence of culture parameters

on lignin metabolism by Phanerochaete chrysosporium. Arch Microbiol 117: 277-285 Larrondo LF, Salas L, Melo F, Vicuna R, Cullen D (2003) A novel extracellular multicopper

oxidase from Phanerochaete chrysosporium with ferroxidase activity. Appl Environ Microbiol 69: 6257-6263

105

Martinez AT, Speranza M, Ruiz-Duenas FJ, Ferreira P, Camarero S, Guillen F, Martínez MJ, Gutiérrez A, del Río JC (2005) Biodegradation of lignocellulosics: microbial, chemical, and enzymatic aspects of the fungal attack of lignin. Int Microbiol 8: 195-204

Martinez D, Larrondo LF, Putnam N, Gelpke MD, Huang K, Chapman J, Helfenbein KG,

Ramaiya P, Detter JC, Larimer F, Coutinho PM, Henrissat B, Berka R, Cullen D, Rokhsar D (2004) Genome sequence of the lignocellulose degrading fungus Phanerochaete chrysosporium strain RP78. Nat Biotechnol 22: 695-700

Martinez D, Challacombe J, Morgenstern I, Hibbett D, Schmoll M, Kubicek CP, Ferreira P, Ruiz-Duenas FJ, Martinez AT, Kersten P, Hammel KE, Vanden Wymelenberg A, Gaskell J, Lindquist E, Sabat G, Bondurant SS, Larrondo LF, Canessa P, Vicuna R, Yadav J, Doddapaneni H, Subramanian V, Pisabarro AG, Lavín JL, Oguiza JA, Master E, Henrissat B, Coutinho PM, Harris P, Magnuson JK, Baker SE, Bruno K, Kenealy W, Hoegger PJ, Kües U, Ramaiya P, Lucas S, Salamov A, Shapiro H, Tu H, Chee CL, Misra M, Xie G, Teter S, Yaver D, James T, Mokrejs M, Pospisek M, Grigoriev IV, Brettin T, Rokhsar D, Berka R, Cullen D (2009) Genome, transcriptome, and secretome analysis of wood decay fungus Postia placenta supports unique mechanisms of lignocellulose conversion. Proc Natl Acad Sci USA 106: 1954-1959

Nankai H, Hashimoto W, Miki H, Kawai S, Murata K (1999) Microbial system for

polysaccharide depolymerization: enzymatic route for xanthan depolymerization by Bacillus sp. strain GL1. Appl Environ Microbiol 65: 2520-2526

Plassard CS, Mousain DG, Salsac LE (1982) Estimation of mycelial growth of basidiomycetes

by means of chitin determination. Phytochem 21: 345-348 Ravalason H, Jan G, Molle D, Pasco M, Coutinho PM, Lapierre C, Pollet B, Bertaud F, Petit-

Conil M, Grisel S, Sigoillot JC, Asther M, Herpoël-Gimbert I (2008) Secretome analysis of Phanerochaete chrysosporium strain CIRM-BRFM41 grown on softwood. Appl Microbiol Biotechnol 80: 719-733

Sato S, Liu F, Koc H, Tien M (2007) Expression analysis of extracellular proteins from

Phanerochaete chrysosporium grown on different liquid and solid substrates. Microbiol 153:3023-3033

Sato S, Feltus FA, Iyer P, Tien M (2009) The first genome-level transcriptome of the wood-

degrading fungus Phanerochaete chrysosporium grown on red oak. Curr Genet 55: 273-286

Shevchenko A, Wilm M, Vorm O, Mann M (1996) Mass spectrometric sequencing of proteins

silver-stained polyacrylamide gels. Anal Chem 68: 850-858 Shevchenko A, Sunyaev S, Loboda A, Shevchenko A, Bork P, Ens W, Standing KG (2001)

Charting the proteomes of organisms with unsequenced genomes by MALDI-quadrupole time-of-flight mass spectrometry and BLAST homology searching. Anal Chem 73: 1917-1926

106

Spanikova S, Polakova M, Joniak D, Hirsch J, Biely P (2007) Synthetic esters recognized by

glucuronoyl esterase from Schizophyllum commune. Arch Microbiol 188: 185-189 Tanaka H, Yoshida G, Baba Y, Matsumura K, Wasada H, Murata J, Agawa M, Itakura S, Enoki,

A (2007) Characterization of a hydroxyl-radical-producing glycoprotein and its presumptive genes from the white-rot basidiomycete Phanerochaete chrysosporium. J Biotechnol 128: 500-511

Vanden Wymelenberg AV, Sabat G, Martinez D, Rajangam AS, Teeri TT, Gaskell J, Kersten PJ,

Cullen D (2005) The Phanerochaete chrysosporium secretome: Database predictions and initial mass spectrometry peptide identifications in cellulose-grown medium. J. Biotechnol 118: 17-34.

Vanden Wymelenberg A, Minges P, Sabat G, Martinez D, Aerts A, Salamov A, Grigoriev I,

Shapiro H, Putnam N, Belinky P, Dosoretz C, Gaskell J, Kersten P, Cullen D (2006) Computational analysis of the Phanerochaete chrysosporium v2.0 genome database and mass spectrometry identification of peptides in ligninolytic cultures reveal complex mixtures of secreted protein. Fungal Genet Biol 43(5):343-356

Vanden Wymelenberg A, Gaskell J, Mozuch, M., Kersten, P., Sabat, G., Martinez, D., Cullen, D.

(2009) Transcriptome and secretome analyses of Phanerochaete chrysosporium reveal complex patterns of gene expression. Appl Environ Microbiol 75: 4058-4068

Yoshida M, Igarashi K, Wada M, Kaneko S, Suzuki N, Matsumura H, Nakamura N, Ohno H,

Samejima M (2005) Characterization of carbohydrate-binding cytochrome b562 from the white-rot fungus Phanerochaete chrysosporium, Appl Environ Microbiol 71: 4548-4555

Yoshida T, Inoue T, Ichishima E (1993) 1,2-alpha-D-mannosidase from Penicillium citrinum:

molecular and enzymic properties of two isoenzymes. Biochem J 290 ( Pt 2): 349-354

107

Chapter 6: Synthesis and Conclusions

Four hypotheses were primary addressed through this study:

1. Cultivation conditions that promote the expression of lignocellulytic enzymes in P.

chrysosporium will also induce the expression of lignocellulytic enzymes in P. carnosa.

2. P. carnosa will exhibit selective decay of lignin and the extent of this decay will depend on

wood fibre composition.

3. Because of its isolation from softwood species, P. carnosa will efficiently degrade guaiacyl

lignin moiety, which dominates softwood lignin.

4. The extracellular protein activities secreted by P. carnosa to degrade various lignocellulosic

feedbstocks will differ depending on the particular chemistry of the lignocellulosic substrate.

Figure 6.1 describes schematically the synthesis of information flow between the three research

objectives.

To characterize the impact of P.carnosa growth on various lignocellulose substrates and to

understand the mode of lignocellulose decay adopted by P. carnosa, there was a need to initially

establish the growth conditions for P. carnosa. Through a series of experiments, it was

established that P. carnosa grows well at 27˚C and is capable of lignocellulolytic activity

expression under nitrogen limiting conditions. These experimental conditions were then used to

investigate the impact of P. carnosa growth on various wood fibres, specifically to confirm that

P.carnosa decays softwood and if so, then how.

Through FTIR and ToF-SIMS investigations, it was validated that P. carnosa initiates softwood

decay through targeting G-lignin. In a recent transcriptomic analyses, the ability of P. carnosa to

degrade softwood was explained by the relatively high number of MnP genes than LiP genes in

P. carnosa as compared to P. chrysosporium (which is essentially a hardwood degrading white-

rot fungus) (MacDonald et al. 2011). The biochemical assays in Chapter 3 also imply similar

results, where significant MnP activity but no LiP activity was detected on various P. carnosa

cultivations on cellulose.

108

Limitations associated with activity-based enzyme discovery particularly in culture supernatants

that contain multiple enzymes with similar activity, was discussed in Chapter 3. Accordingly, a

proteomic method was used to compare the profiles of enzymes secreted by P. carnosa during

growth on cellulose and spruce wood fibre. Notably, differences in the profile of enzymes

secreted by P. carnosa grown on cellulose and spruce were subtler than initially hypothesized.

In particular, activities detected in spruce cultivations were also identified in cellulose

cultivations, and the profile of enzymes secreted by P. carnosa grown on cellulose was more

complex than that secreted by P. carnosa grown on spurce. This result implies that P. carnosa

might regulate the relative abundances of a similar set of lignocellulose-active enzymes rather

than the presense of particular activities.

Figure 6.1 Conclusions and information flow between the three key objectives of this thesis.

Optimal temperature: 27°C Cellulolytic activity is expressed. MnP activity is expressed. LiP activity is not detected.

G-lignin is degraded. Growth of P. carnosa on balsam fir, lodgepole pine and white spruce reveals similar degradation patterns. P.carnosa selectively degrades lignin prior to wood polysaccharides.

Proteins secreted during growth on spruce were also detected in cellulose cultivations. Cellulase enzymes appear to be more abundant in culture supernatant than β-glucosidases. Glyoxal oxidases appear to be more abundant than lignin-degrading peroxidases. An additional 127 peptides were annonated to predicted CAZymes or FOLymes when the P.carnosa genome was used in the analysis.

Chapter 3: Growth on microcrystalline cellulose

Chapter 4: Growth on coniferous wood

Chapter 5: Comparative proteomic analysis of proteins secreted by P.carnosa grown on cellulose and spruce

Optimal

growth

conditions

used

Optimal

growth

conditions

used

109

P. chrysosporium, which is a model white-rot fungus, was used as a benchmark for results obtain

through studies using P. carnosa. Cultivation conditions published for P. chrysosporium were

used to define the initial experimental conditions for P. carnosa (Chapter 3) and the genome

sequence for P. chrysosporium was used to assist with the protein annotation based on sequence

similarity (Chapter 5). While P. carnosa selectively degrades lignin like certain strains of P.

chrysosporium, these organisms differed in terms of optimal growth conditions and expression of

lignin-degrading and carbohydrate-active enzymes. For example, the optimal growth temperature

of P. carnosa is 27˚C, while that of P. chrysosporium is 37˚C. Unlike P. chrysosporium, P.

carnosa’s lignocellulolytic enzyme expression is not severely hampered by high nitrogen and/or

shaking cultivation conditions. P. carnosa is also distinguished by its ability to effectively

degrade softwood lignin. In addition to lignin degradation, the proteomic analyses also suggest

that P. carnosa was actively degrading softwood hemicelluloses. Softwood lignin and

hemicellulose degradation by P. carnosa was supported by the detection of mannosidase and

lignin-degading activities in the secretome of P.carnosa (Chapter 5), which has previously not

been reported for P. chrysosporium.

To conclude, P. carnosa has been established as a softwood degrading white-rot fungus, which is

capable of selectively degrading softwood-lignin. The following list summarizes

recommendations for further research on the lignocellulolytic activity of P. carnosa and wood-

degrading basiodiomycetes in general:

1. Systematic investigation of the effect of growth conditions (rpm, pH, oligosaccharide

addition) on P. carnosa growth and lignocellulolytic activity.

2. Annotation of proteins with unknown function, identified through proteomic analysis of the

secretome of P. carnosa on spruce and micro-crystalline cellulose.

3. Differential proteomics for extracellular enzymes expressed by P.carnosa on other simple

substrates and hardwood species to determine if the conclusion of non-substrate specific

protein expression can be extended.

4. Investigation of changes in fibre chemistry for hardwood species when degraded by P.

carnosa to determine explicitly the differences between conventional white-rot degrading

fungi isolated from hardwoods and P.carnosa.

110

Chapter 7: Engineering Relevance

Growing environmental concerns are requiring that more industrial sectors adopt eco-friendly

processes. Further, with increasing competition from industries located in the Southern

hemisphere, North American and European forest industries are realizing the need to diversify

the range of products that can be derived from wood fibre. Accordingly, there is an opportunity

for technology that harnesses the full potential of biomass for the production of fuels and other

biodegradable materials.

Enzymes are renewable biocatalysts, and several enzymes are already being used by forest

industries to reduce chemical consumption. For instance, approximately 10% of North American

pulp and paper industries use microbial xylanases to boost fibre bleaching processes (Kenealy

and Jeffries 2003; Subramaniyan and Prema 2002). Laccases and manganese peroxidases have

also shown promise for improving pulp bleaching; cellulases have been used during fibre

pulping and to recover fibre from recycled paper products, whereas lipases and pectinases have

been used to remove pitch deposited on paper manufacturing equipment (Bajpai 2004; Kenealy

and Jeffries 2003). The synergistic action of glycosyl hydrolases, including endo and exo-

glucanases and glucosidases have been used to transform cellulosic material to glucose which

can be fermented for production of bioethanol. Enzymes have also been used to alter the

functionality of wood polymers. For example, Gustavsson et al. (2005) used a lipase from

Candida antartica to link chemically modified oligosaccharides to xyloglucan (Gustavsson et al.

2005). The modified xyloglucans hydrogen bond with cellulose, thereby functionalizing

cellulose microfibrils. In this way, chemically reactive and bioactive side groups were grafted to

filter paper.

Despite their potential as industrial catalysts, associated costs have limited the utility of enzymes

for production of commodities. These costs can be attributed to enzyme production, typically

low availability of enzymes to insoluble substrates like plant fibre, and the variable composition

of plant biomass. In particular, the distinct composition and structure of lignocellulosic substrates

contained in plant fibre presents unique challenges for bioconversion technologies. It is possible

that the efficiency of bioconversion technologies could be improved by tailoring enzyme

applications to particular lignocellulosic feedstocks. One approach is to mimic the profiles of

111

enzymes secreted by fungi grown on different lignocellulosic substrates. Accordingly, it is

anticipated that the proteomic analysis performed as part of the current thesis, will help to

formulate ideal enzyme mixtures for softwood conversion. For instance, the various ligninolytic

enzymes expressed by P. carnosa on the softwoods can be eventually purified for pre-treatment

of softwood fibre used to generate mechanical pulp or otherwise sugars for subsequent

fermentation. Alternatively, since P. carnosa has demonstrated ability to selectively remove

lignin from multiple softwood species, this organism could be used for in-situ conversion of

softwood wastes to fermentable sugars for fuels or chemical production.

In addition to potential industrial relevance, contributions to method development that have

broader scientific relevance include:

1. Proteomic Analysis of Uncharacterized Microorganisms: The methodology

established for differential proteomic analysis of P. carnosa can be transferred to other

investigations where the genome sequence for the organism is not available. In addition,

it can also be applied towards metagenomic studies, where microbial consortia are being

studied using incomplete genomic data.

2. Analytical Techniques for Fibre Characterization: The method developed in this

study for investigating fibre properties of wood chips (as opposed to the commonly used

wood powder) through FTIR and ToF-SIMS makes a significant contribution in

monitoring changes in fibre chemistry. This becomes especially relevant since most

industrial processes deal with wood chips. In addition, the method development for ToF-

SIMS using un-embedded samples will also contribute tremendously for ToF-SIMS

based fibre analysis, while preventing unwanted interference with embedding medium.

Also through this thesis, the use of ToF-SIMS for both spectral and spatial analysis of

degraded wood fibre was established. These results have been validated with those from

FTIR and TEM. The use of ToF-SIMS as a single method to obtain information

regarding changes in fibre chemistry and localization of cell wall constituents, will

improve the efficiency and ease of analyzing treated wood fibre.

112

Chapter 7 References Bajpai P (2004) Biological Bleaching of Chemical Pulps. Critical Reviews in Biotechnology 24

(1):1-58 Gustavsson MT, Persson PV, Iversen T, Martinelle M, Hult K, Teeri TT, Brumer H (2005)

Modification of Cellulose Fiber Surfaces by Use of a Lipase and a Xyloglucan Endotransglycosylase. Biomacromolecules 6 (1):196-203.

Kenealy WR, Jeffries TW (2003) Wood deterioration and preservation: advances in our

changing world. In: Society AC (ed), Washington, DC, Oxford University Press, pp 210-239

Subramaniyan S, Prema P (2002) Biotechnology of microbial xylanases: enzymology, molecular

biology, and application. Critical Reviews in Biotechnology 22 (1):33-64

113

Appendix 1: B3 Medium

10X Buffer Recipe (per 1L) 20 g KH2PO4 5g MgSO4-7H2O 1g CaCl2-2H2O pH to 4.5 add milliQ water up to 1L autoclave for 30 min Mineral Elixir (per 1L): 1.5 g nitrilotriacetate 0.5 g MnSO4 1g NaCl 100 mg FeSO4-7H2O 100mg CoSO4 (= 0.645 mM ≡ 180 mg CoSO4-7H2O 100 mg ZnSO4 (= 0.619 mM ≡ 178 mg ZnSO4-7H2O 10 mg CuSO4-5H2O 10mg AlK(SO4)2 10 mg H3BO3 10mg NaMoO4

autoclave 30 min Vitamins 10 mg/ml thiamine HCl filter sterilize Nitrogen Source 1g/10ml ammonium tartrate filter sterilize Carbon Source

0.5 g/ml selected carbon source (if not wood chips), ρglycerol = 1.27 g/ml autoclave 30 min Other 0.73 g/50 ml 2,2-dimethylsuccinic acid (added as a filter sterilized solution adjusted to pH 4.5 using NaOH)

114

Medium Preparation (1L in Sterilized Bottle containing specified volume of milliQ water) Component Volume (no

Carbon/ low Nitrogen)

Volume (no Carbon/ high Nitrogen)

Volume (Carbon/ low Nitrogen)

Volume (Carbon/ high Nitrogen)

Buffer 100 100 100 100 Mineral 10 10 10 10 Vitamins 50 µl 50 µl 50 µl 50 µl 2,2-dimethylsuccinic acid

50 50 50 50

Carbon Source 0 0 20 20 Nitrogen Source 2 20 2 20 MilliQ Water 838 820 818 800

References:

1.Kenealy and Dietrich. 2004. Enzyme and Microbial Technology. 34:490-498. 2. Kirk et al. 1978. Arch. Microbiol. 117:277-285.

115

Appendix 2: Supplemental Information for Chapter 4

0

10

20

30

40

50

60

70

2 cm 6 cm 8 cm

Growth Diameter

Figure S4.1 P. carnosa cultivation on wood samples. Number of days required to cultivate P.

carnosa on to specified growth diameters on fir (black bars), pine (grey bars) and spruce (white bars). n=3, error bars indicate standard deviations.

116

Figure S4.2. Mean baseline corrected FTIR spectra of nine standard materials, normalized to highest intensity peak. From top to bottom: beech xylan (dark cyan), xyloglucan (orange), pectin (grey), mannose (magenta), lignin (purple), glucomannan (blue), galactomannan (green), cellulose (red), arabinogalactan (black). Dashed vertical lines indicate selected wavelengths significant to PCA grouping of control and decayed samples (see Figs. 4.1-4.3 and Table 4.1)

117

Fig. S4.3 Images of scores for ToF-SIMS analyses of Spruce at time point 1 (left) and 5 (right) resulting from the PCA model built on pine (Fig. 4. 5). Image dimensions are in pixels where 50 pixels equal 7.9 microns

Fig. S4.4 Images of scores for ToF-SIMS analyses of Fir at time point 1 (left) and 4 (right) resulting from the PCA model built on pine (Fig. 5) Image dimensions are in pixels where 50 pixels equal 7.8 microns (left) and 9.7 microns (right)

118

Table S4.1. The Ratio of Bromine Concentration in the Middle Lamella to Cell Wall in Spruce

Samples Analyzed by TEM-EDXA

Ratio of Time point 1 Time point 5 1. cell wall: middle lamella 1.63 1.38

1.25 2.42 1.33 1.33

2. cell corner: middle lamella 1.88 2.25 1.39 1.23 1.23 0.852

* t-test was used to determine the statistical significance; p-value = 0.6590 (1) and 0.4876 (2)

119

Appendix 3: Supplemental Information for Chapter 5

(A) (B)

Figure S5.1 Distribution of peptide sequences corresponding to GH families for cultivations of (A) Cellulose (B) Spruce

120

Table S5.1. Non-gapped alignment of peptides from extracellular filtrates of P. carnosa grown on cellulose and spruce to probable lignocellulose-degrading enzymes from P. chrysosporium. Predicted Activity GH Family Substrate Pchr IDa P. carnosa Peptide Sequence Pchr Sequence Match % ID

Glycoside hydrolase

GH5

Cellulose

3985 IIIEQGSTAV EQGSTA 100 5773 NTGTQTVIFQ TGTPTIIF 75

138313 YNGGAGAPSE YSGGSGAP 75 5607 QITGIGASF VITGIGASG 100

38870 ASTGQIEM TGQIEM 100 126220 PNQFASFI QYASFI 71

Spruce

137948 AENFEVTIQ ENFGVTI 85 ASGGSSAGVSS ASGGSSSG 87

GH74

Cellulose/Spruce

138266

EADSISGAAGAb DSTSGAAGA 88 MGGVGDNGGVc VGDNGG 100

α-D-galactosidase

GH27 Cellulose 134001 FMIGSGNSP FMINTGNS 75 GH28 Cellulose 4449 NSFAQNGAR AQNGAR 100

Polygalacturonase Spruce 3805 LSVNSGAER LTVNSGAER 100 Putative chitinase

Cellulose/ Spruce

444

GGMMTSGSGGb GSMMTYGSG 77 MPSIIAPTQGb IVAPTQG 85 EGAGTSGSGATc TSASAT 87

d-4,5 unsaturated -glucuronyl hydrolase

GH88

Cellulose

1106

VNFDDATIAQ VGFNDATIA 77 IADQSFAQAAR SFAQAAR 100

Spruce 840 MFIDNDXADGIFN DKNDGI 67

CBM, low similarity to glycoside hydrolases

NA Cellulose 131440 ADEDPNYHQY AVEDPNFHQY 87

Monooxygenase

Cellulose

1321 EGSGPQEATE SGPQEST 85 2191

QISAFV QISAFV

100

QISAFVDDV 100

121

2613 SHFTID SHFTID 100 8209 MFGAGPGGGI THFGAGPG 100

Spruce

140125 IYAGNNGIGG YDKNDGIGG 70 8208 SPTAINSR SPTAIN 100

Peroxidase LiP Cellulose 131707 SMSAANAVQT SLSAANAV 87 Catalase Cellulose 128306 YGGEMGAAE VGGEMGSAD 75 Cir1 Spruce 147 SDPADVNS SDPADVDS 87 Copper amine oxidase

Cellulose

4526

QGSVSFAGGH SVSFAG 100 GFIQQGITR IQQKITR 85

Spruce 36058 SGRSHAVAI GRSHAV 100

a Additional information for protein from P. chrysosporium can be accessed by ending the following URL with the model number, e.g.

http://genome.jgi-psf.org/cgi-bin/dispGeneModel?db=Phchr1&id=3985; bDetected in extracellular filtrates of cellulose cultivations;

cDetected in extracellular filtrates of spruce cultivations.

122

P. carnosa Protein ID

Family Predicted Activity Peptide

259050 GH2 β-galactosidase, β-mannosidase IESAVISGQNMIR ITEPIIGINEFTQR PADTSEPT SPVFIEESSIEI TPGFIYGNTDR

262713 GH3 β-glucosidase FDSPAEQPFR IVCNPSAD

266084 GH3 β-glucosidase GICIEDSPIG HYINNEQE

249799 GH3 β-glucosidase ITFSVGAS 247802 GH5 endoglucanase QGVVNAVR 248589 GH5 endoglucanase DIAGAGSTTVR

EPTIIGWEIANE 255866 GH6 cellobiohydrolase WGDWCNI 264060 GH7 cellobiohydrolase GICDADGCDFNSFR 264426 GH7 cellobiohydrolase FGDTDFFSQHGGI

FVTGSNVG GTQTPETHPSI IGAGITVDT NTGICDGDGCDFDS TDNFCTQQ TSSGTAITI YGTGYCDSQC

258037 GH7 cellobiohydrolase IMSDDSHYQM 252890 GH10 xylanase IPSTPAIIQ 249059 GH10 xylanase DVCNEMQNEDGT

INDYNIDSANA TSSSSGIDA

258624 GH11 xylanase TSDGSSYDVYE 261445 GH13 amylase ASDWQGR

GQTSDQAII YQFITDR

261990 GH15 amylase IDQYTSGQDTS 247750 GH15 amylase DPFIFNDGT

MAGAFEER VIEVTPAAHV WVWGVSR

143000 GH16 xyloglucanase VYETGGGII 265415 GH16 xyloglucanase YDNAYFEV

61816 GH18 chitinase ISIGGWTGAR

Table S5.2. Additional peptides from extracellular filtrates of P. carnosa grown on cellulose that

were matched to CAZymes and oxidoreductases using the P. carnosa genome sequence.

123

254594 GH18 chitinase FIAEQGIR 246104 GH31 α-glucosidase, α-xylosidase DGSAIIIHA

EQYTDVI GDDGGFGTIQTFW IENSGTYEVE IETMWTDID

246124 GH31 α-glucosidase, α-xylosidase GHWIGDNY GTIQTMWAR FAGAGAHV TVDPDYF VTYETNTR

95998 GH35 β-galactosidase IADYTFGSPAHS NPDTGAQFVIVR NVITVVQDHM YDYGSTIR

257303 GH35 β-galactosidase DTGAQFVIVR IADYTFGAVIHS IFVDGWQY INEGGIFGER

214333 GH37 α-trehalase IIDIFWDS 192563 GH43 α-L-arabinofuranosidase NDGAIEASVIW 256205 GH43 α-L-arabinofuranosidase, galactan 1,3-

β-galactosidase YTSTDIVNWSR

252166 GH47 α-mannosidase FIESYSAY GGAISAYEIS MSVAWVGS

208092 GH51 α-L-arabinofuranosidase IAETIVEMG PTMESSSVSAAFMT

261448 GH61 cellulose-binding DDTYTTSWAVDR 263097 GH61 cellulose-binding QTDVTGISWF 253139 GH74 xyloglucanase GIVPGIVFN 257835 GH79 β-4-O-methyl-glucuronidase AAQAHAASI 249789 GH88 d-4,5 unsaturated β-glucuronyl

hydrolase IADQSFAQAAR VNFDDATIA

248819 GH92 α-mannosidase FNAGTGFMEAR SFISVDQAR YNDYAVSV

255063 GH92 α-mannosidase AFDQWDEII DSITSIFAR EYAFEDFAI GAASFAASFG ISPADYTDAN YANQPGISTQ

249084 GH92 α-mannosidase EYAIGDFAVR

124

259392 GH95 fucosidase MATDFQNIR 260703 GH115 α-(4-O-methyl)-glucuronidase IGQATCVAF 198942 CE8 pectin methylesterase AYFYGNTIAT 262938 GT15 glycolipid 2-alpha-mannosyltransferase IGYEHNAN

29063 PL14 Polysaccharide Lyase GYTAYFPASFD 97090 Chitin synthase TNILFAMKEDNEK

262183 CBM1 Carbohydrate-Binding NPADTSETDGIR 254017 CBM1 Carbohydrate-Binding EDPNYHQY 261029 Lipase PTSQAEVVR

SAVAIVGDSITD TYITFGDQTR

255713 Lipase DQAAVFDAR IIIEQGSTAV

179599 Glucose-methanol-choline oxidoreductase

DSWNEIITP

GGSSSIDGAAW GWGWSGER WVVIVGHTA

199860 galactose-1-epimerase EVDGDFIPT 206015 Cytochrome P450 AAGSVPAVW 256997 Fungal lignin peroxidase IQSDFAIAR 261553 Multicopper oxidase IINEGSED 260543 Glucose-methanol-choline

oxidoreductase

258261 glyoxal oxidase FMESSVSA PANSFEFF

259359 glyoxal oxidase NEGIASSNI GTAATDADIVAA

IHVIIDNTV PVGTAATDADIVA SSINFITY

263528 glyoxal oxidase ACYVDNA 263533 glyoxal oxidase IGGWSIQSTQGVR

IVAATSSTH VAATSSTHGN VGYYNEAR

266369 FAD linked oxidase NNFGIVTR

125

P. carnosa Protein ID

GH Family Predicted Activity Peptide

260755 GH16 xyloglucanase TSGASTAGDADA 252166 GH47 α-mannosidase FIESYSAY 247750 CE15 4-O-methyl-glucuronoyl methylesterase DVIEVPTAAHVN 100639 Multicopper oxidase DIPQAIID

ISGITVSGTNSDIR 262882 Fungal lignin peroxidase VDCSAVVPV 256980 Fungal lignin peroxidase IAIIGHNR 258261 glyoxal oxidase FMESSVSA 260543 Glucose-methanol-choline oxidoreductase PVGTAATDADIVAA 261029 Lipase NDGDGPNAI

TYITFGDQTR

Table S5.3. Additional peptides from extracellular filtrates of P. carnosa grown on spruce that

were matched to CAZymes and oxidoreductases using the P. carnosa genome sequence.

126

Appendix 4: Substrate recognition and hydrolysis by a fungal xyloglucan-specific family 12 hydrolase

Justin Powlowski, Sonam Mahajan, Matthieu Schapira and Emma R. Master

Abstract

Biochemical studies to elucidate the structural basis for xyloglucan specificity among GH12

xyloglucanases are lacking. Accordingly, the substrate specificity of a GH12 xyloglucanase from

Aspergillus niger (AnXEG12A) was investigated using pea xyloglucan and 12 xylogluco-

oligosaccharides, and data were compared to a structural model of the enzyme. The specific

activity of AnXEG12A with pea xyloglucan was 113 µmol min−1 mg−1, and apparent kcat and Km

values were 49 s−1 and 0.54 mg mL−1, respectively. These values are similar to previously

published results using xyloglucan from tamarind seed, and suggest that substrate fucosylation

does not affect the specific activity of this enzyme. AnXEG12A preferred xylogluco-

oligosaccharides containing more than six glucose units, and with xylose substitution at the −3

and +1 subsites. The specific activities of AnXEG12A on 100 µM XXXGXXXG and 100 µM

XLLGXLLG were 60 ± 4 and 72 ± 9 µmol min−1 mg−1, respectively. AnXEG12A did not

hydrolyze XXXXXXXG, consistent with other data that demonstrate the requirement for an

unbranched glucose residue for hydrolysis by this enzyme.

Graphical abstract

Keywords: Xyloglucan; Oligosaccharides; Xyloglucanase; Glycoside hydrolase; Kinetics

Abbreviations: MALDI, matrix-assisted laser desorption ionization; XEG, xyloglucan endoglucanase; Glc, glucose


Top Related