978-3-030-11599-9

Download as pdf or txt
Download as pdf or txt
You are on page 1of 429

Green Energy and Technology

Cataldo De Blasio

Fundamentals
of Biofuels
Engineering
and Technology
Green Energy and Technology
Climate change, environmental impact and the limited natural resources urge
scientific research and novel technical solutions. The monograph series Green
Energy and Technology serves as a publishing platform for scientific and
technological approaches to “green”—i.e. environmentally friendly and sustain-
able—technologies. While a focus lies on energy and power supply, it also covers
“green” solutions in industrial engineering and engineering design. Green Energy
and Technology addresses researchers, advanced students, technical consultants as
well as decision makers in industries and politics. Hence, the level of presentation
spans from instructional to highly technical. **Indexed in Scopus**.

More information about this series at http://www.springer.com/series/8059


Cataldo De Blasio

Fundamentals of Biofuels
Engineering and Technology

123
Cataldo De Blasio
Laboratory of Energy Technology
Faculty of Science and Engineering
Åbo Akademi University
Vaasa, Finland

ISSN 1865-3529 ISSN 1865-3537 (electronic)


Green Energy and Technology
ISBN 978-3-030-11598-2 ISBN 978-3-030-11599-9 (eBook)
https://doi.org/10.1007/978-3-030-11599-9

Library of Congress Control Number: 2019934516

© Springer Nature Switzerland AG 2019


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part
of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission
or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt from
the relevant protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the
authors or the editors give a warranty, express or implied, with respect to the material contained herein or
for any errors or omissions that may have been made. The publisher remains neutral with regard to
jurisdictional claims in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
A Daniel Lucas Ettore e Stefan Louis
Emanuel.
Siete la mia vita,
Papà.
Preface

Biomass and waste have become of much interest in recent years for their usage in
power conversion processes. Biomass is considered as a renewable source of
energy, and therefore, it is assumed usually as environmentally friendly to utilize
biomass in power plants or for biofuels production, even if this has some constraints
in reality. The topic is enormously large, and it would require an entire encyclo-
pedia to treat all the related subfields. In this case, the manuscript is driven mainly
by the questions:
• Why biomass would be a suitable choice for energy conversion?
• How biomass is generated and assessed? (This is important to understand that
nothing is coming for free, and time is needed to have suitable yields).
• How biomass is utilized for converting power and producing biofuels?
This first edition is divided into two main parts (Parts I and II) and an additional
part (Part III) which includes some additional material mainly meant as support for
the reader.
The manuscript is addressed at a large audience, and it is suitable for students at
bachelor and master levels of their university studies. Naturally, researchers who
are working in these fields might find this book of interest. In addition, the aim is at
a broad range of readers (engineers, technicians, professionals) who are looking for
some specific topics within the area of biomass and energy. This book would be
suitable as textbook for university courses and lectures given within the field treated
here.
The general outline of the manuscript is to give information following a “ped-
agogical style,” in the same way as lessons are given by this author. At the same
time, the author was trying to give more simplified explanations for some of the
treated arguments, and instead, for some other topics, he is explaining them into
very details. Steps and derivations of equations given normally in the literature,
especially in the case of mass balances, are given here in full. To give a better
understanding of the passages involved, equations are repeated and numbered more
than one time, when needed, to give a better understanding and flow to the
comprehension.

vii
viii Preface

The author of this book would like to welcome the reader to the first edition of
this manuscript. The readers of this book are also encouraged to give any comments
and suggestions which could improve the current version of the manuscript.

Vaasa, Finland Cataldo De Blasio


Acknowledgements

Eng. Marco Simonetti, Eng. Gaetano Lucca, Dr. Mauro Prestipino, and Prof.
Antonio Galvagno are greatly acknowledged for their collaboration and support.
I am proud to work with you.
Many thanks go to my colleagues at the Laboratory of Energy Technology in
Vaasa; Professor Margareta Björklund-Sänkiaho and Dr. Jessica Tuuf, for the great
working environment we have.
The colleagues at the Department of Chemical Engineering in Turku are also
greatly acknowledged. I’d like to thank Prof. Henrik Saxén, Prof. Tapio
Westerlund, and Prof. Frank Petterson for their suggestions, support, and comments
to the manuscript.
The companies UPM, Metsä, and Vaasa Water are acknowledged for giving
permissions to use some of their figures.
Many thanks to my wife Leena Matilda and my children, you are my strength
and joy in this life. Thanks to my sister, Maria Teresa, brother, Luigi, and my
parents Donato and Teresa who are living in the beautiful village of Brienza in
Italy. They are always of great support and sustainment. My mother, Teresa
Giallorenzo, is greatly acknowledged also for providing information on some
agrofuels yield and rate.

ix
About This Book

The scope of this book is to deliver suitable information related to the “generation”
of power and its conversion from biomass sources. This would include also the
production of biofuels. The author of this book wanted to keep a more “teaching
style” during the adopted descriptions, and therefore, this would make the manu-
script suitable to students and persons interested on the subject of biomass, how it is
generated, and how it is converted to get power and fuels. The book is divided
mainly into two parts with an additional part related to the explanation of support
information.
If the intention is to utilize biomass or bioproducts for energy conversion, it
would be beneficial to understand the underlying principles behind the biomass
generation itself.
For this reason, during the first part, the main principles of biomass generation
will be explained with a particular focus on how plants harvest the energy from the
sun. The photosynthetic process is described in some more details to give an
example of the main principles common to many kinds of microorganisms
including micro-algae. These arguments are also beneficial if evaluations should be
done on feasibility studies and process integration which concern biomass.
The main analogies between this kind of biomass generation (CO2 fixation
processes) and processes utilized by microorganisms for reproducing themselves
will be pointed out clearly. After explaining the principal concepts of biomass
generation, some sections of the first part of the book will be dedicated to the
principles of biomass analysis and assessment for the purpose of evaluating its
quality before the utilization in power plants. Particular focus will be on proximate
and ultimate analysis methods.
If the book is utilized in a course at university, the goal of the first part of the
book is to put students in the position of:
• having a sufficient knowledge of the principles behind biomass generation;
• possess sufficient knowledge to understand the principles behind the main types
of chemical analysis done on biomass and biofuels;

xi
xii About This Book

• have knowledge on what is the link between the physics of biomass generation
and some other engineering problems. Recognize that there are many analogies
between the working principles of nature and technical solutions.
The driving questions of this first part of the manuscript are: How biomass is
generated and what are the principles behind it? What kind of feedstock we should
use for energy conversion and to which purposes? How we assess the suitability of
biomass and biofuels made for energy conversion purposes? Bio-organisms are
demonstrated to have many analogies with technical devices converting energy, and
therefore, the thermodynamic efficiencies of these processes can be evaluated. The
most important part here is shown to be the formation of an electrochemical
potential together with a difference in concentration related to the H+ ions across the
membranes. The H+ ions are coming from water oxidation (in the case of photo-
synthesis); clear is the resemblance with the galvanic cells and fuel cells principles.
Aspects related to light harvesting performed by living organisms are also covered
along with the evaluation of the degree of response in relation to the converting
device. The treatment of the analyses done on biomass is covered at first by the
Kjeldahl method. Then, an accurate description of the proximate analysis (calori-
metric bomb) with calculation of the higher heating value is given. The analysis
related to the composition of biomass is done by describing elemental analysis,
thermogravimetric analysis, and chromatography. There are statistical implications
which are descriptive of the methods and the outcomes.
The second part of the manuscript focuses on the production methods for certain
biofuels. The part is mostly focused on giving the basic knowledge required to
understand the details of the biorefinery methods and on technological aspects.
Theoretical aspects are covered as well, especially for the case of biochemical
reactions engineering methods.
The sections start with a preliminary introduction of biorefineries, and then the
principles of biochemical reactor engineering are given in detail. The reader will
differentiate between the different types of reactor configurations and the related
modeling of the kinetics involved.
The main biological pathways of processes like fermentation and anoxic
digestion are discussed along with the main parameters utilized for bioreactors
design. The sections continue with the discussion on biogas, bio-ethanol, biodiesel
production, and Fischer–Tropsch methods.
This part of the book will give the students the possibility to:
• have an understanding of what are the principles behind biochemical reactors
design;
• recognize some of the most important methods used for biofuels production;
• differentiate the main types of reactors used for biofuels production. In addition,
the reader will be capable of analyzing the main related parameters involved;
• distinguish and summarize the different routes performed for biogas,
bio-ethanol, biodiesel production, and Fischer–Tropsch synthesis;
• possess critical thinking about the different technologies.
About This Book xiii

The main focus of the second part is on the questions:


After knowing the main principles of biomass generation, what are the best solu-
tions applied in bioreaction engineering? What are the different routes involved for
biofuels production? How we could utilize the biofuel generated for the purpose of
energy conversion?
Contents

Part I Notions of Biomass Formation and Development


with Some Analytical Methods
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.1 Biomass as a “Almost” Renewable Energy Source . . . . . . . . . 3
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2 Light Harvesting and Biomass Generation . . . . . . . . . . . . . . . . . .. 13
2.1 Preliminary Data Related to Light Harvesting and Biomass
Generation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2 Light Harvesting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3 Energy Transfer and Reaction Centers . . . . . . . . . . . . . . . . . . 23
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3 Redox Potential and Galvanic Cells . . . . . . . . . . . . . . . . . . . . .... 31
3.1 Functioning Principles . . . . . . . . . . . . . . . . . . . . . . . . . . .... 31
3.2 Conventions on the Electrodes . . . . . . . . . . . . . . . . . . . . .... 35
3.3 Calculation of the Redox Potential at Non-standard
Conditions, the Nernst Law . . . . . . . . . . . . . . . . . . . . . . .... 36
3.4 A Remark on the Artificial Splitting of Water to Produce
Hydrogen, Electrolysis . . . . . . . . . . . . . . . . . . . . . . . . . .... 43
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... 45
4 Overview of the Main Mechanisms of Photosynthesis . . . . ....... 47
4.1 Transport Phenomena in Reaction Centers . . . . . . . . ....... 47
4.2 Light and Dark Reactions of Photosynthesis . . . . . . . ....... 49
4.3 Binding of Carbon Dioxide and Production of Sugar
Molecules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ....... 52
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ....... 55

xv
xvi Contents

5 Work from Light . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57


5.1 Incoming Radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.2 A Simple Derivation of the Stefan–Boltzmann Law . . . . . . . . 59
5.3 The Solid Angle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
5.4 The Photosynthetic Active Radiation . . . . . . . . . . . . . . . . . . . 64
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
6 Some Types of Analyses Conducted on Biomass . . . . . . . . . . . . . .. 71
6.1 Determination of the Total Nitrogen Content, the Kjeldahl
Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 71
6.2 Determination of the Higher Heating Value . . . . . . . . . . . . .. 74
6.3 A Procedure for the HHV Calculation Based on a Particular
Instrument . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 77
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 89
7 Thermogravimetric Analysis (TGA) . . . . . . . . . . . . . . . . . . . . . . . . 91
7.1 Basic Principles of Thermogravimetric Analysis . . . . . . . . . . . 91
7.2 TGA Classification and Types of Equipment . . . . . . . . . . . . . 93
7.3 Procedure and Sample Analysis . . . . . . . . . . . . . . . . . . . . . . . 94
7.4 Positive and Negative Mass Variation . . . . . . . . . . . . . . . . . . 94
7.5 Reporting TGA Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
7.6 Differential Thermal Analysis and Differential Scanning
Calorimetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..... 96
7.7 TGA and the Study of the Kinetics of Diverse Thermal
Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..... 97
7.8 TGA Coupled with Mass Spectrometry, Gas
Chromatography, and Infrared Spectroscopy . . . . . . . . . ..... 98
7.9 Methods and Experimental Conditions Suggested . . . . . . . . . . 100
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
8 Chromatography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
8.1 Principles of Chromatography . . . . . . . . . . . . . . . . . . . . . . . . 103
8.2 The Chromatogram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
8.3 Distribution Constants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
8.4 Retention Time . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
8.5 Volumetric Flow Rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
8.6 Retention Time and Distribution Constant . . . . . . . . . . . . . . . 113
8.7 Retention Factor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
8.8 Parameters Characterizing a Chromatographic Column . . . . . . 114
8.8.1 Selectivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
8.8.2 Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
Contents xvii

8.9 Resolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119


8.10 Parameters Influencing the Resolution . . . . . . . . . . . . . . . . . . 120
8.11 Qualitative Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
8.12 Elemental Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
9 Examples of Quantitative Content Determination
in Chromatography and Elemental Analysis . . . . . . . . . . . . . . . . . . 127
9.1 Instrumentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
9.2 Preparation and Introduction of the Sample . . . . . . . . . . . . . . 129
9.3 Calculation of the Calibration Coefficients . . . . . . . . . . . . . . . 130
9.3.1 Calibration Constant for Hydrogen . . . . . . . . . . . . . . 131
9.3.2 Calibration Constant for Carbon Monoxide . . . . . . . . 131
9.3.3 Calibration Constant for Methane . . . . . . . . . . . . . . . 131
9.3.4 Calibration Constant for Carbon Dioxide . . . . . . . . . . 132
9.3.5 Calibration Constant for Hydrogen Sulfide . . . . . . . . . 132
9.3.6 Calibration Constant for Nitrogen . . . . . . . . . . . . . . . 132
9.3.7 Calibration Constant for Oxygen . . . . . . . . . . . . . . . . 133
9.3.8 Calibration Constant for Ethylene . . . . . . . . . . . . . . . 133
9.3.9 Calibration Constant for Ethane . . . . . . . . . . . . . . . . . 133
9.3.10 Summary of the Results . . . . . . . . . . . . . . . . . . . . . . 134
9.4 An Example of Results from Elemental Analysis . . . . . . . . . . 134
9.4.1 Instrumentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
9.4.2 Calculation of the Calibration Coefficients . . . . . . . . . 136
9.4.3 Summary of the Results . . . . . . . . . . . . . . . . . . . . . . 137
9.5 Analysis of Unknown Substances . . . . . . . . . . . . . . . . . . . . . . 138
Reference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
10 Some Considerations and Statistical Derivations for the
Concentration Profile and Gaussian Curve . . . . . . . . . . . . . . . . . . . 141
Reference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143

Part II Notions of Biochemical Reactors Design and Biofuels


Production
11 Introduction to Part 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
12 Integrated Biorefinery Concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
12.1 An Introduction to Biorefineries and to the Sustainable
Utilization of Bioresources . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
12.2 Industrial Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
xviii Contents

12.2.1 Biodiesel Production from Bio Oils by


Hydrogenation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
12.2.2 Metsä Group’s Äänekoski Plant . . . . . . . . . . . . . . . . . 165
12.3 Production Concepts for Bioethanol Production . . . . . . . . . . . 166
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
13 Electronegativity and Microbial Catalysis . . . . . . . . . . . . . . . . . . . . 173
13.1 Production of Biofuels and Classification of the Related
Microorganisms Needed . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
13.2 Pathways in Biological Processing of Biomass Feedstock . . . . 175
13.3 Main Molecules Used as Energy Carriers in Biological
Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
13.4 Oxidation and Reduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
13.5 An Introduction to Modeling of Digestion Processes . . . . . . . . 180
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
14 Main Reactors Configurations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
14.1 Introduction to Reactor Design and the Different
Configurations Used . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
14.2 Performance of a Batch Stirred Tank Reactor . . . . . . . . . . . . . 194
14.3 Performance of a Continuous Stirred Tank Reactor . . . . . . . . . 196
14.4 First-Order Chemical Reaction in Stirred Tank Reactor
with Changing Reaction Volume . . . . . . . . . . . . . . . . . . . . . . 198
14.5 Plug Flow Reactor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
14.6 The Case of Reversible Reaction . . . . . . . . . . . . . . . . . . . . . . 202
14.7 Performance Equations for Plug Flow Reactors . . . . . . . . . . . . 204
14.8 Autocatalitic Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
14.9 Optimal Choice of Reactors . . . . . . . . . . . . . . . . . . . . . . . . . . 206
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
15 Enzyme Kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
15.1 Product Formation in Enzyme Catalyzed Biological
Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
15.2 Enzime Kinetics. The Michaelis–Menten Model . . . . . . . . . . . 211
15.3 Enzime Kinetics, Briggs–Haldane Approach . . . . . . . . . . . . . . 212
15.4 The Monod Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
15.5 Main Characteristics of the Briggs–Haldane Equation . . . . . . . 216
15.6 Residence Time in a Mixed Flow Fermentor . . . . . . . . . . . . . 218
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
16 Balances on Microbial Fermentation . . . . . . . . . . . ........... . . 221
16.1 Cell Balances . . . . . . . . . . . . . . . . . . . . . . . ........... . . 221
16.2 Models for Microbial Growth Used in Power Generation . . . . 223
16.3 Kinetic Rate Expressions . . . . . . . . . . . . . . . ........... . . 224
16.4 Decay of Organisms . . . . . . . . . . . . . . . . . . ........... . . 226
Contents xix

16.5 Overall Balance in Continuous STR . . . . . . . . . . . . . . . . . . . . 227


16.6 Effluent Concentration of the Substrate . . . . . . . . . . . . . . . . . . 227
16.7 Production of Ethanol by Digestion . . . . . . . . . . . . . . . . . . . . 229
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
17 Processes of Bioethanol Production . . . . . . . . . . . . . . . . . . . . . . . . . 233
17.1 Some Systems Adopted and Product Compositions . . . . . . . . . 233
17.2 Bioethanol from Lignocellulosic Biomass . . . . . . . . . . . . . . . . 239
17.3 Corn Starch as a Feedstock . . . . . . . . . . . . . . . . . . . . . . . . . . 242
17.4 Ethanol Distillation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
17.5 Ethanol Yields from Different Kinds of Feedstock . . . . . . . . . 249
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 250
18 Biodiesel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
18.1 Fundamentals of Biodiesel Production . . . . . . . . . . . . . . . . . . 253
18.2 Ionization Potential and Electron Affinity of an Element . . . . . 257
18.3 Base-Catalyzed Production of Biodiesel . . . . . . . . . . . . . . . . . 259
18.4 Acid-Catalyzed Production of Biodiesel . . . . . . . . . . . . . . . . . 260
18.5 More Detailed Process Description for Alkali-Catalyzed
Biodiesel Production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264
19 Some Chemical Analyses in Biodiesel Production and Biofuel
Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
19.1 Amount of Free Fatty Acid . . . . . . . . . . . . . . . . . . . . . . . . . . 267
19.2 Iodine Number Determination . . . . . . . . . . . . . . . . . . . . . . . . 270
19.3 Saponification Value of Fats and Oils . . . . . . . . . . . . . . . . . . 272
19.4 Biodiesel Characteristics and Properties . . . . . . . . . . . . . . . . . 273
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284
20 Fischer–Tropsch (FT) Synthesis to Biofuels (BtL Process) . . . . . . . 287
20.1 Background Information . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
20.2 Chain Growth Probability (Products Distribution)
for the FT Synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 296
20.3 Syngas Treatment for FT Synthesis . . . . . . . . . . . . . . . . . . . . 298
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
21 Notions of Biomass Gasification . . . . . . . . . . . . . . . . . . . . . . . . . . . 307
21.1 Preliminary Notions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307
21.2 Supercritical Water Gasification . . . . . . . . . . . . . . . . . . . . . . . 319
21.2.1 SCWG Kinetics and Thermodynamics . . . . . . . . . . . . 323
21.2.2 Catalysis in SCWG . . . . . . . . . . . . . . . . . . . . . . . . . . 326
21.2.3 Salts and Inorganics in SCWG of Biomass . . . . . . . . 327
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 331
xx Contents

Part III Review of Some Concepts of Thermodynamics


and Support Material
22 Preliminary Concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 337
22.1 Definition of Mole . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 337
22.2 Conservation of Mass, Closed Systems,
Degree of Advancement of Reaction . . . . . . . . . . . . . . . . . . . 338
22.3 About Internal Energy and Enthalpy . . . . . . . . . . . . . . . . . . . 338
22.4 Relations Between Specific Heat at Constant Volume
and Pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 339
22.5 The First Law of Thermodynamics . . . . . . . . . . . . . . . . . . . . . 340
22.6 Some Relations Between Pressure and Temperature in Gases.
Gas Laws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341
22.7 A Remark on Reversible Processes . . . . . . . . . . . . . . . . . . . . 341
22.8 Pressure and Internal Energy, the Kinetic Model . . . . . . . . . . . 342
22.9 Specific Heat and Universal Gas Constant . . . . . . . . . . . . . . . 343
22.10 Change of State for an Ideal Gas . . . . . . . . . . . . . . . . . . . . . . 345
22.11 Adiabatic Transformation of an Ideal Gas . . . . . . . . . . . . . . . . 348
22.12 Energy Balance for a Given System:
Macroscopic Balances . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 351
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 352
23 Introduction to Entropy and Second Law . . . . . . . . . . . . . . . . . . . . 353
23.1 Mixing of One or Two Gases at the Same T and P . . . . . . . . . 353
23.2 Statistical Derivation of Entropy . . . . . . . . . . . . . . . . . . . . . . 357
23.3 Additional Remarks on Entropy with Variation in Moles,
i.e., Chemical Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 360
23.4 Clausius Statement and Kelvin–Planck Statement
of the Second Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 361
23.5 Entropy Balances . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 363
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 364
24 Thermodynamics in Chemical Reactions Engineering . . . . . . . . . . . 365
24.1 Reaction Rate and Its Dependence on Temperature . . . . . . . . . 365
24.2 Derivation of the van’t Hoff Equation . . . . . . . . . . . . . . . . . . 366
24.3 Derivation of the Arrhenius Equation . . . . . . . . . . . . . . . . . . . 368
24.4 Chemical Potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 370
24.5 Statistical Implications of the Boltzmann Distribution Law . . . 371
25 Some Parameters and Properties of Biomass Fuels . . . . . . . . . . . . . 375
25.1 Some Useful Figures on Biomass Feedstock . . . . . . . . . . . . . . 375
25.2 Overall Energy (Heat) Balances . . . . . . . . . . . . . . . . . . . . . . . 379
Contents xxi

25.2.1 Design Variables Affecting Thermochemical


Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 379
25.2.2 Additional Data on Moisture and Ash Content
of Fuels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 383
25.2.3 Properties of Agrofuels and Typical Air
Requirements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 385
25.3 Stoichiometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 390
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 393
26 A Simple Estimation of the Efficiency for a Biomass
Power Plant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 397
26.1 Mass Balances . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 397
26.2 Thermal Performance and Efficiency . . . . . . . . . . . . . . . . . . . 398
26.3 Incomplete Combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 402
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 403
27 Some Data on Oxidation and Reduction States and Half-Cell
Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 405
27.1 Reduction and Oxidation States, Electron Affinities,
and Ionization Potentials . . . . . . . . . . . . . . . . . . . . . . . . . . . . 405
27.2 Electronegativity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 408
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 410
Acronyms

A0 Chlorophyll
A1[Q] Phylloquinone
AAE Average absolute error
ADH Alcohol dehydrogenase
AFEX Ammonia fiber expansion
AFW Animal fat waste
aq Aqueous (if used as subscript in equations)
ASF Anderson–Schulz–Flory (Distribution)
ASTM American Society for Testing and Materials
ASU Air separation unit
ATP Adenosine triphosphate
ATR Autothermal reforming
BAL Balanced
BOCLE Ball on cylinder lubricity evaluator
BPG 1,3-biphosphoglycerate
BSTR Batch stirred tank reactor
CALB Candida antarctica lipase B
CBP Consolidated bioprocessing
Chl Chlorophyll
CMM Computational molecular modeling
CN Cetane number
CPA Centrally planned Asia and China
CSSC Chlorophyll-sensitized solar cell
Cyt Cytochrome
Cyt bLp and Cyt Cytochrome b low- and high-potential forms
bHp
Cyt f Cytochrome f
DDGS Dried distillers grain solubles
DG Diacylglycerides
DSC Differential scanning calorimetry

xxiii
xxiv Acronyms

DTA Differential thermal analysis


DTG Derivative thermogravimetry
EA Elemental analyzer
EMF Electromotive force
EMP Embden–Meyerhof Pathway
ENO Enolase
EoS Equation of state
FAME Fatty acid methyl ester
FBA Fructose bisphosphate aldolase
Fd Ferredoxin
FFA Free fatty acids
FID Flame ionization detector
FNR Flavo-protein ferredoxin–NADP reductase
FSU Former Soviet Union
FT Fischer–Tropsch
Fx Iron–sulfur center
GC Gas chromatography
GCV Gross calorific value
GDP Gross domestic product
HC Hydrocarbons
HFRR High-frequency reciprocating ring
HHV Higher heating value
HPLC High-pressure liquid chromatography
HRT Hydraulic residence time
HXK Hexokinase
IAPWS International Association for the Properties of Water and
Steam
IEA International Energy Agency
IGCC Integrated gasification combined cycle
IIASA International Institute for Applied Systems Analysis
IR Infrared
LAM Latin America
LHC Light-harvesting complex
LHV Lower heating value
MEA Middle East and North Africa
MG Monoacylglycerides
MS Mass spectroscopy
NADP Nicotinamide adenine dinucleotide phosphate
NADP* Oxidized nicotinamide adenine dinucleotide phosphate
NAM North America
NCV Net calorific value
NEXBTL Next-generation biomass to liquid (By Neste)
Acronyms xxv

NIST National Institute of Standards and Technology


OECD Organisation for Economic Co-operation
and Development
OY Osakeyhtiö (Limited Company)
P680 Primary electron donor of PSII composed
of chlorophyll (Chl)
P700 Primary electron Chl donor of PSI
PAO Pacific OECD
PAR Photosynthetic active radiation
PAS Other Pacific Asia
PC Plastocyanine
PDC Pyruvate decarboxylase
PFK Phosphofructokinase
PFR Plug flow reactor
PGA 3-phosphoglycerate
PGAL Phosphoglyceraldehyde
PGI Phosphoglucose isomerase
PGK 3-phosphoglycerate kinase
PGM Phosphoglycerate mutase
Pheo Pheophytin
PK Pyruvate kinase
POX Partial oxidation
PSI Photosystem I
PSII Photosystem II
Qa Plastoquinone A
Qb Plastoquinone B
QHE Quantum heat engine
QYP Quantum yield of photosynthesis
R&D Research and Development
RC Reaction center
RI Refractive index
RuBP Ribulose 1,5 bi-phosphate
RUE Radiation use efficiency
s Solid (if used as subscript in equations)
SAS South Asia
SBL Strong black liquor
SCWG Supercritical water gasification
SCWO Supercritical water oxidation
SHF Separate hydrolysis and fermentation
SRK Redlich–Kwong
SSF Simultaneous saccharification and fermentation
SSFF Simultaneous saccharification, filtration, and fermentation
STR Stirred tank reactor
TAG Triacylglycerols
TAN Total acid number
xxvi Acronyms

TCD Thermal conductivity detector


TDH Triose phosphate dehydrogenase
TDI Triose phosphate isomerase
TG Triacylglycerides
TGA Thermogravimetric analysis
VFA Volatile fatty acid
WBL Weak black liquor
WCO Waste cooking oils
WEC World Energy Council
WEU Western Europe
WFGD Wet flue gas desulfurization
WGS Water-gas shift
WGSR Water-gas shift reaction
Yo Symmetrically related tyrosine to YZ
YZ Tyrosine
List of Figures

Fig. 1.1 World total population by year. Data from Statista (2018). . .. 4
Fig. 1.2 Alkalinity of oceans and the CO2 equilibrium . . . . . . . . . . . .. 5
Fig. 1.3 A simple schematics on the relation between CO2
concentration, atmospheric temperature, and oceans pH . . . .. 6
Fig. 1.4 World energy consumption, 1990–2040 (Btu1015).
Data from US Energy Information Administration (2013) . . .. 7
Fig. 1.5 World carbon dioxide emissions by region, left: all OECD
countries; right: OECD Europe. Data from US Energy
Information Administration (2013). . . . . . . . . . . . . . . . . . . . .. 8
Fig. 1.6 World gross domestic product (GDP) by region. Data from
US Energy Information Administration (2013) . . . . . . . . . . .. 8
Fig. 1.7 World consumption of renewable energy by region.
Data from US Energy Information Administration (2013) . . .. 9
Fig. 1.8 Potential bio-energy of crop residues by world region,
gigajoules per hectare of cultivated land, NAM,
North America; LAM, Latin America; WEU, Western
Europe; FSU, former Soviet Union; MEA, Middle East
and North Africa; CPA, Centrally planned Asia and
China; SAS, South Asia; PAS, other Pacific Asia; PAO,
Pacific OECD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 10
Fig. 2.1 Visible spectrum among the total wavelength spectrum.
Adapted from Taiz et al. (2014) . . . . . . . . . . . . . . . . . . . . . .. 18
Fig. 2.2 Energy at the earth’s surface. Adapted from
Taiz et al. (2014) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 19
Fig. 2.3 Absorption of chlorophyll-sensitized solar cells (CSSCs).
Chlorophyll, main regions where the chlorophyll absorbs the
most and the less energy in. Data from
Hassan et al. (2016) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 20
Fig. 2.4 Excited states of chlorophyll, a qualitative representation . . .. 20

xxvii
xxviii List of Figures

Fig. 2.5 Chlorophyll and beta-carotene molecules . . . . . . . . . . . . . . . .. 21


Fig. 2.6 Quantum yield of photosynthesis, in the visible spectrum.
Adapted from Zeiger and Taiz (1991) . . . . . . . . . . . . . . . . . .. 23
Fig. 2.7 Chloroplast structure, a qualitative representation . . . . . . . . .. 24
Fig. 2.8 Transfer of energy in light-harvesting antennas . . . . . . . . . . .. 25
Fig. 2.9 Oxidation of water into the reaction center and transported
by tyrosine molecule . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 26
Fig. 2.10 Z-scheme of photosynthesis . . . . . . . . . . . . . . . . . . . . . . . . . .. 27
Fig. 3.1 Functioning of a galvanic cell where a salt bridge
is employed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 32
Fig. 3.2 Galvanic cell with no liquid junction. A representation
of the hydrogen anode; hydrogen is pumped on to the
electrode by using a particular casing . . . . . . . . . . . . . . . . . . . 33
Fig. 3.3 Plus–Right Rule . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
Fig. 3.4 Z-scheme in terms of the redox potential . . . . . . . . . . . . . . . . . 42
Fig. 3.5 Schematic representation of an electrolytic cell . . . . . . . . . . . . 44
Fig. 4.1 Cytochrome b6f complex pumping H+ ions. Adapted from
Taiz et al. (2014) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 48
Fig. 4.2 Summary of the main mechanisms of photosynthesis.
The FNR refers to the ferredoxin—NADP+ reductase.
Adapter from Holtzegel (2016) . . . . . . . . . . . . . . . . . . . . . . .. 49
Fig. 4.3 Oxygenic photosynthesis, the light side. Adapted from US
Department of Energy (2018) . . . . . . . . . . . . . . . . . . . . . . . .. 51
Fig. 4.4 Summary of the oxygenic photosynthesis, light and dark
reactions. Adapted from Ksenzhek and Volkov (1998) . . . . .. 52
Fig. 4.5 Binding the carbon dioxide, the Calvin–Benson cycle.
Adapted from Ksenzhek and Volkov (1998) . . . . . . . . . . . . .. 53
Fig. 4.6 Calvin–Benson cycle, a simplified scheme. The numbers 1–5
refer to the steps of the cycle. PGAL is the
phosphoglyceraldehyde. Adapted from Starr et al. (2016) . . .. 54
Fig. 4.7 Structure of D-glucose and the derived linear combination
to give cellulose . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 55
Fig. 5.1 Solid angle evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 62
Fig. 5.2 Light-harvesting engine, simplified representation . . . . . . . . .. 66
Fig. 5.3 Efficiency of light absorption as a function
of the temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 67
Fig. 6.1 A qualitative description of the turbidity measurement
by using laser/light diffraction technique . . . . . . . . . . . . . . . . . 72
Fig. 6.2 Simplified illustration of the Kjeldahl method . . . . . . . . . . . . . 73
Fig. 6.3 A laboratory setup for calorimetric measurements . . . . . . . . . . 75
Fig. 6.4 Main parts of the overall calorimeter . . . . . . . . . . . . . . . . . . . . 78
Fig. 6.5 Main components of the calorimetric embodiment . . . . . . . . . . 79
Fig. 6.6 Scale used for sample weight measurement . . . . . . . . . . . . . . . 79
Fig. 6.7 Positioning of the capsule inside the embodiment ring . . . . . . 80
List of Figures xxix

Fig. 6.8 Cutting the ignition wire . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 81


Fig. 6.9 Positioning of the ignition wire, to be noticed that the wire
has to be in this kind of shape but of course has to contact
the sample . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 81
Fig. 6.10 Oxygen filling of the calorimetric embodiment . . . . . . . . . . .. 82
Fig. 6.11 Oxygen bottle with pressure meters . . . . . . . . . . . . . . . . . . . .. 83
Fig. 6.12 A top view of the calorimetric embodiment closed
and ready . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
Fig. 6.13 Calorimeter jacket and cover . . . . . . . . . . . . . . . . . . . . . . . . . . 84
Fig. 6.14 Water container . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
Fig. 6.15 Positioning of the water container . . . . . . . . . . . . . . . . . . . . . . 85
Fig. 6.16 Handling of the calorimetric embodiment by suitable
handle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 86
Fig. 6.17 Positioning of the calorimetric embodiment inside
the water container and connection of the cables . . . . . . . . . .. 87
Fig. 6.18 Positioning of the jacket cover and the elastic belt . . . . . . . .. 87
Fig. 6.19 Data logger . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 88
Fig. 7.1 Example of thermogravimetric analyzer. Curtesy of Prof.
Antonio Galvagno and Mauro Prestipino at University of
Messina, Italy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 92
Fig. 7.2 Simple representation of the thermogravimetric method. . . . .. 92
Fig. 7.3 Three of the most common TGA configurations: cantilever
(a), hang down (b), and top-loading column (c) . . . . . . . . . .. 93
Fig. 7.4 TGA (solid line) and DTG (dashed line) signals of biomass
conversion in inert atmosphere . . . . . . . . . . . . . . . . . . . . . . .. 95
Fig. 7.5 TGA/DTA sample holder of a top-loading column
TGA/DTA device . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 97
Fig. 7.6 Conversion versus time curves of different residual chars
tested through a TGA in 50 kPa steam atmosphere
at 750 °C and isothermal conditions. Data from
Prestipino et al. (2018). . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 100
Fig. 8.1 Simple representation of the mains steps
in chromatography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
Fig. 8.2 Draft of the chromatographic column . . . . . . . . . . . . . . . . . . . 104
Fig. 8.3 Functioning principle of refractive index detectors . . . . . . . . . 106
Fig. 8.4 Basic principles of the ultraviolet detectors . . . . . . . . . . . . . . . 107
Fig. 8.5 System adopted for chromatographic measurements . . . . . . . . 108
Fig. 8.6 Main steps involved during the measurement of gas/liquid
concentration in chromatography. Adapted from Rouessac
and Rouessac (2007) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 109
Fig. 8.7 Example of chromatogram with main parameters . . . . . . . . .. 109
Fig. 8.8 Typical response as a function of distance . . . . . . . . . . . . . . .. 116
Fig. 8.9 Typical Gaussian curves for a component
into a chromatographic column . . . . . . . . . . . . . . . . . . . . . . .. 117
xxx List of Figures

Fig. 8.10 Relation between flow and column height . . . . . . . . . . . . . . .. 118


Fig. 8.11 Main parameters and resolution of a chromatogram.
Adapted from Skoog et al. (2006) . . . . . . . . . . . . . . . . . . . . .. 119
Fig. 8.12 Schematics of a CHNS/O analyzer functioning . . . . . . . . . . .. 123
Fig. 9.1 A photo of the gas chromatography instrumentation . . . . . . .. 128
Fig. 9.2 Example of output of the chromatographic analysis.
As shown by the instrument . . . . . . . . . . . . . . . . . . . . . . . . . . 128
Fig. 9.3 Collection of samples of known composition . . . . . . . . . . . . . 129
Fig. 9.4 Injection of the gas into the gas chromatograph. . . . . . . . . . . . 129
Fig. 9.5 Elemental analyzer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
Fig. 9.6 Example of output of the elemental analyzer . . . . . . . . . . . . . . 135
Fig. 9.7 Reference substances . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
Fig. 11.1 Forest growth plotted with the sustainable harvesting
potential and industrial use of wood. Metsä Group.
With permission, the figure is based on Natural
Resource Institute Finland 2015 (Luke) . . . . . . . . . . . . . . . . .. 148
Fig. 11.2 Undernourished population in the world by region.
Data from Statista (2018a) . . . . . . . . . . . . . . . . . . . . . . . . . . .. 149
Fig. 11.3 Carbon cycles and the production of fossil fuels . . . . . . . . . .. 150
Fig. 12.1 Integration concept of regional biorefinery. The Vaasa
region, Finland. Figure made by Lena Karlsson. Reprinted
by permission of Vaasa Water—Public Utility Company . . .. 156
Fig. 12.2 Sectoral integration concept aimed at the supply-chain
network . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 157
Fig. 12.3 Supply-chain structures. The centralized biomass conversion
(Route 1), the distributed biomass conversion (Route 2),
and the centralized–distributed biomass conversion . . . . . . . .. 157
Fig. 12.4 Different approaches in biomass processing . . . . . . . . . . . . . .. 158
Fig. 12.5 LignoForceTM System, WBL weak black liquor,
SBL strong black liquor . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 159
Fig. 12.6 An example of process for regional biomass conversion.
Adapted from Özdenkçi et al. (2017) . . . . . . . . . . . . . . . . . . .. 161
Fig. 12.7 Integration of fermentation, hydrogenation, HTL,
and CHP plant. Adapted from Coma et al. (2017) . . . . . . . . .. 162
Fig. 12.8 NEXBTL process used for biodiesel production . . . . . . . . . .. 164
Fig. 12.9 Production of renewable diesel from tall oil.
Permission to use the figure granted by UPM . . . . . . . . . . . .. 164
Fig. 12.10 A next-generation wood biorefinery proposed by Metsä
Group. Used with permission (von Weymarn 2016) . . . . . . .. 166
Fig. 12.11 Ethanolix process and distribution flow . . . . . . . . . . . . . . . . .. 167
Fig. 13.1 Embden–Meyerhof pathway (EMP). Adapted from
Madigan et al. (2010) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 176
Fig. 13.2 Most important energetic molecules in living organisms.
Adapted from Madigan et al. (2010) . . . . . . . . . . . . . . . . . . .. 177
List of Figures xxxi

Fig. 13.3 Example of electron transport system for purple


photosynthetic and green sulfur bacteria. Adapted from
Rasmussen and Minteer (2014) . . . . . . . . . . . . . . . . . . . . . . .. 178
Fig. 13.4 Main steps of the pathway for the anaerobic digestion
performed by bacteria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 180
Fig. 13.5 Classification of biological processes related to the biofuels
production. Adapted from Levenspiel (1999) . . . . . . . . . . . . .. 182
Fig. 13.6 Commonly used symbol for the
Saccharomyces cerevisiae . . . . . . . . . . . . . . . . . . . . . . . . . . .. 183
Fig. 13.7 Example of ethanol production plant integrated with a
cellulose nanofibers production plant. AFEX, ammonia fiber
expansion. Adapted from Leistritz et al. (2006) . . . . . . . . . . .. 183
Fig. 14.1 Three main reactors configuration for chemical engineering
processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
Fig. 14.2 Example of plain axis in PFR reactor . . . . . . . . . . . . . . . . . . . 190
Fig. 14.3 Illustration of the batch stirred tank reactor . . . . . . . . . . . . . . . 195
Fig. 14.4 Simple schematics of a STR . . . . . . . . . . . . . . . . . . . . . . . . . . 197
Fig. 14.5 Illustration of a PFR and its differential change in conversion
and concentration with respect to the compound A along the
distance from the inlet axes . . . . . . . . . . . . . . . . . . . . . . . . . .. 200
Fig. 14.6 Inverse of the reaction rate as a function of the conversion
(left) and concentration (right) for the component A . . . . . . .. 204
Fig. 14.7 General trend of the reaction rate as a function of the
conversion for autocatalytic reactions . . . . . . . . . . . . . . . . . .. 206
Fig. 14.8 Comparison between reactor volumes of PFR and STR at a
fixed conversion. Optimal choice of the reactors for a given
conversion with autocatalytic reactions . . . . . . . . . . . . . . . . .. 207
Fig. 15.1 Transesterification of castor oil driven with enzymes
as catalysts. Adapted from Andrade et al. (2019) . . . . . . . . .. 210
Fig. 15.2 Simple schematics showing the major components
in fermentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 214
Fig. 15.3 Main phases in the variation of the cell concentration
with time. Adapted from Levenspiel (1999). . . . . . . . . . . . . .. 215
Fig. 15.4 Production rate of product as a function of the substrate
concentration. Adapted from Levenspiel (1999) . . . . . . . . . . .. 216
Fig. 15.5 Feedstock consumption rate function and its main parts . . . .. 217
Fig. 15.6 Linearized function for the concentration of the substrate
as a function of time to find the constant parameters . . . . . . .. 218
Fig. 15.7 Linearized form of the substrate concentration as a function
of the residence time expression at low and high enzyme
concentration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 219
Fig. 16.1 Conversion of the substrate against time for autocatalytic
reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 223
xxxii List of Figures

Fig. 16.2 Glycolytic pathway for ethanol production in S. cerevisiae.


Adapted from Drapcho et al. (2008) . . . . . . . . . . . . . . . . . . .. 230
Fig. 17.1 Schematics for the organosolv process with ethanol.
Adapted from Alriols et al. (2010) . . . . . . . . . . . . . . . . . . . . .. 235
Fig. 17.2 Pretreatment and derivation of clarified juice in bioethanol
production from sugarcane. Adapted from Cortes-Rodríguez
et al. (2018) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 236
Fig. 17.3 Juice concentration and sterilization process in bioethanol
production from sugarcane. Adapted from
Palacios-Bereche et al. (2014) . . . . . . . . . . . . . . . . . . . . . . . .. 236
Fig. 17.4 A scheme of the crystallization process where molasses
and raw sugar is produced. Adapted from
Velasquez et al. (2013) . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 237
Fig. 17.5 Fermentation and distillation steps in bioethanol production
from sugarcane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 238
Fig. 17.6 Vinasse concentration unit (left) and ethanol dehydration
unit (right). Adapted from Kujawski and Zielinski (2006) . . .. 238
Fig. 17.7 One example of process for bioethanol production from
sugarcane. Adapted from Amores et al. (2013) . . . . . . . . . . .. 239
Fig. 17.8 Production of ethanol from lignocellulosic biomass
feedstock. A simplified scheme . . . . . . . . . . . . . . . . . . . . . . .. 241
Fig. 17.9 Schematic representation of the dry-milling process for
ethanol production. Adapted from Vohra et al. (2014) . . . . . .. 243
Fig. 17.10 Schematic representation of the wet-milling process for
ethanol production. Adapted from Vohra et al. (2014) . . . . . .. 243
Fig. 17.11 An example of the animal a-amylase enzyme. The figure is
produced by Jawahar Swaminathan and his colleagues at the
European Bioinformatics Institute and made available for the
public and free to use. The author highly acknowledge
Jawahar and his collaborators . . . . . . . . . . . . . . . . . . . . . . . .. 244
Fig. 17.12 Schematic representation of a dedicated dry-grind ethanol
production plant. Adapted from Vertes et al. (2010) . . . . . . .. 245
Fig. 17.13 Schematic representation of dedicated wet-mill ethanol
production plant. Adapted from Vertes et al. (2010) . . . . . . .. 246
Fig. 17.14 A simple schematics of a distillation column . . . . . . . . . . . . .. 247
Fig. 17.15 Graphical Evaluation of the number of stages
in a distillation column . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 248
Fig. 17.16 A top–down visual of a distillation tray with suitable passes
for the vapor phase (left) and a schematic showing the
passage of the liquid and gas phase over the distillation
tray (right) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 248
Fig. 18.1 Reaction of a vegetable oil with an alcohol. Adapted from
Lotero et al. (2005) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 254
Fig. 18.2 FFA reaction with a base catalyst and formation of soap. . . .. 256
List of Figures xxxiii

Fig. 18.3 Carboxylic acid representation. Adapted from


Jones and Fleming (2014) . . . . . . . . . . . . . . . . . . . . . . . . . . .. 256
Fig. 18.4 Nucleophilic substitution between N–H
and C–OH groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 257
Fig. 18.5 Formation of the tetrahedral intermediate in the esterification
of hydroxyl acids. Adapted from
Jones and Fleming (2014) . . . . . . . . . . . . . . . . . . . . . . . . . . .. 258
Fig. 18.6 All steps in Fischer esterification. Adapted from
Jones and Fleming (2014) . . . . . . . . . . . . . . . . . . . . . . . . . . .. 258
Fig. 18.7 Reaction involving the alkoxide group (−OR) and the
hydroxide group (−OH). Adapted from
Jones and Fleming (2014) . . . . . . . . . . . . . . . . . . . . . . . . . . .. 259
Fig. 18.8 Catalyst alkoxide ion and attach on the carboxylic
group. Adapted from Lotero et al. (2005) . . . . . . . . . . . . . . .. 260
Fig. 18.9 Acid-catalyzed reactions on the triglyceride. R1, R2, and R3
are carbon chains of the fatty acids, while R4 is an alkyl
group of the alcohol. Adapted from Lotero et al. (2005) . . . .. 261
Fig. 18.10 Example of homogeneous acid-catalyzed process for
biodiesel production . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 262
Fig. 18.11 A more detailed flow diagram for biodiesel production
by a homogeneous alkali-catalyzed process. Adapted from
Apostolakou et al. (2009) . . . . . . . . . . . . . . . . . . . . . . . . . . .. 263
Fig. 19.1 Percentage of pollutants reduction as a function
of the fuel blend. B20 means “fuel blend,” in this case, 20%
of biodiesel is mixed with 80% of normal diesel. Data from
Drapcho et al. (2008b) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 268
Fig. 19.2 Structure of the phenolphthalein molecule in its acidic
for and base form. Adapted from
Zumdahl and Zumdahl (2014) . . . . . . . . . . . . . . . . . . . . . . . .. 269
Fig. 19.3 Mechanism of iodine saturation of the double bonds
in unsaturated molecules . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 270
Fig. 19.4 Tristearin structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 272
Fig. 19.5 a Cetane (normal hexadecane) C16H34, cetane number =
100. b Isocetane C16H34, 2,2,4,4,6,8,8-heptamethylnonane
with cetane number = 15 . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 276
Fig. 19.6 Structure of Iso-octane. Adapted from Morrison
and Boyd (1992) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 276
Fig. 20.1 A schematic view of BtL technology . . . . . . . . . . . . . . . . . . .. 288
Fig. 20.2 A diagram showing the steam gasification
of biomass process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 289
Fig. 20.3 SASOL slurry-bed reactor. Adapted from Steynberg et al.
(2004, p. 2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 294
xxxiv List of Figures

Fig. 20.4 Chain growth for compounds with n size on the catalyst.
Adapted from Cheng et al. (2008). The (ad) means adsorbed
on the catalyst, while (g) refers to the release of compounds
in the gas phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 296
Fig. 20.5 The a value and product molar fraction against the chain
length. The dashed lines represent the ideal ASF distribution
while the solid curves are the deviated ASF distribution
values. Adapted from Cheng et al. (2008) . . . . . . . . . . . . . . .. 297
Fig. 20.6 Distribution functions for diverse components and their mass
fraction. Qualitative estimation . . . . . . . . . . . . . . . . . . . . . . .. 298
Fig. 20.7 Wet-cold gas cleaning process. Adapted from Triantafyllidis
et al. (2013) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 299
Fig. 20.8 Dry-hot gas cleaning process for removing tar components
from syngas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 300
Fig. 20.9 Bubbling fluidized bed (BFB) (left), circulating fluidized bed
(CFB) (center), fixed bed reactor (right). Adapted from Kunii
and Levenspiel (1991) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 302
Fig. 20.10 Integrated gasification combined cycle (IGCC) power plant.
Adapted from Shadle and Breault (2012) . . . . . . . . . . . . . . . .. 303
Fig. 20.11 Integration of gasification with FT synthesis and production
of power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 303
Fig. 21.1 Stages of the gasification process. Adapted from
Chen et al. (2018) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 308
Fig. 21.2 A simple schematics of the GE Energy Gasifier.
Adapted from De Agarwal et al. (2018) . . . . . . . . . . . . . . . .. 311
Fig. 21.3 Drawing of the ConocoPhillips E-gas gasifier . . . . . . . . . . . .. 312
Fig. 21.4 Shell gasifier with a description of the constituting parts
of the reactor wall. Adapted from Lisandy et al. (2016) . . . .. 312
Fig. 21.5 Siemens gasification reactor with quench water system . . . . .. 313
Fig. 21.6 Schematics of the quench process and its contribution
to the WGSR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 314
Fig. 21.7 Representation of the KBR Transport Gasifier (TRIGTM).
Adapted from Ariyapadi et al. (2008) . . . . . . . . . . . . . . . . . .. 314
Fig. 21.8 Description of a British Gas Lurgi Gasifier. Adapted from
Krishnamoorthy and Pisupati (2015) . . . . . . . . . . . . . . . . . . .. 315
Fig. 21.9 Lurgi MPG. Adapted from Koss and Schlichting (2005) . . . .. 315
Fig. 21.10 A system developed by Mitsubishi Heavy Industries for
gasification of MSW. Adapted from Mitsubishi Heavy
Industries (2019) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 316
Fig. 21.11 U-Gas method developed by synthesis energy systems.
Adapted from Lau (2009) . . . . . . . . . . . . . . . . . . . . . . . . . . .. 317
Fig. 21.12 A scheme for the High Temperature Winkler Gasifier.
Adapted from Toporov and Abraham (2015) . . . . . . . . . . . . .. 318
List of Figures xxxv

Fig. 21.13 PRENFLO® Gasifier/Boiler. Adapted from ThyssenKrupp


(2019) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 319
Fig. 21.14 Visual example for SCWG of black liquor and liquid
products. Photograph by Cataldo De Blasio. . . . . . . . . . . . . .. 320
Fig. 21.15 Density of water as a function of temperature at 250 bar.
Data from Wagner and Kretzschmar (2008). . . . . . . . . . . . . .. 321
Fig. 21.16 Isothermal variation of ionic product for water.
Left, 374 °C. Right: 650 °C. Data from Wagner
and Kretzschmar (2008) . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 321
Fig. 21.17 Density versus viscosity of water at 386 °C. Data from
Wagner and Kretzschmar (2008) . . . . . . . . . . . . . . . . . . . . . .. 322
Fig. 21.18 Specific heat of SCW against temperature
at 24 MPa, Left, and 36 MPa, Right . . . . . . . . . . . . . . . . . . .. 322
Fig. 21.19 Density of SCW plotted against pressure at 410 °C, left,
and 374 °C, right . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 322
Fig. 21.20 Enthalpy change against temperature at 30 MPa.
Adapted from Kruse (2008) . . . . . . . . . . . . . . . . . . . . . . . . . .. 323
Fig. 21.21 General procedure for evaluating the composition
of a mixture at equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . .. 326
Fig. 21.22 Dielectric constant of water against temperature
at different pressures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 328
Fig. 21.23 Na2CO3 solubility against temperature. Adapted from
Jones et al. (1976) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 329
Fig. 21.24 Pressure drop due to plugging of reactor and sudden release
of the blockage wall . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 330
Fig. 21.25 Schematics of a SCWG setup for laboratory-scale
experiments. Adapted from De Blasio et al. (2019) . . . . . . . .. 330
Fig. 22.1 Simple representation of a space. The box
with a moving object in it . . . . . . . . . . . . . . . . . . . . . . . . . . .. 342
Fig. 22.2 Constant volume system with varying pressure . . . . . . . . . . .. 345
Fig. 22.3 Simple representation of an isobaric process . . . . . . . . . . . . .. 346
Fig. 22.4 Isothermal process. The points 1 and 2 are on the same curve
in this case: the isothermal P-V curve . . . . . . . . . . . . . . . . . . . 347
Fig. 22.5 Adiamatic process of compression of an ideal gas . . . . . . . . . . 350
Fig. 22.6 Volume with its boundaries . . . . . . . . . . . . . . . . . . . . . . . . . . . 351
Fig. 23.1 Adiabatic mixing of two ideal gases . . . . . . . . . . . . . . . . . . . . 354
Fig. 23.2 Object moving under the effect of a force field, gravity. . . . . . 354
Fig. 23.3 Initial and ending state of a variable with two possible routes
to arrive to the final point . . . . . . . . . . . . . . . . . . . . . . . . . . .. 355
Fig. 23.4 Mixing of two gases at the same temperature and adiabatic
conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 358
Fig. 23.5 Simple representation of the Kelvin–Planck statement of the
second law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 362
xxxvi List of Figures

Fig. 23.6 Example of combined thermodynamic systems . . . . . . . . . . . . 362


Fig. 24.1 Probability density function for the velocity values of a
molecule in a box and one direction . . . . . . . . . . . . . . . . . . . . 372
Fig. 25.1 A simple scheme of the gross calorific
value determination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 378
List of Tables

Table 1.1 Some studies concerning biomass supply and its usage in
energy conversion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 9
Table 2.1 Carbon release per GJ for some biomass types and fossil
fuels, data from Ksenzhek and Volkov (1998) . . . . . . . . . .. 14
Table 2.2 Values for the refraction index for selected materials,
with references . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 18
Table 3.1 Evaluation of the standard cell potential . . . . . . . . . . . . . . .. 35
Table 3.2 Diameters of non-hydrated, value on top, and hydrated,
second value, ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 40
Table 3.3 Standard potentials of selected half reactions
of biological importance, 25 °C and pH 7
(Ksenzhek and Volkov 1998; Ross 1991) . . . . . . . . . . . . . . . 41
Table 4.1 Reaction steps in photosynthesis . . . . . . . . . . . . . . . . . . . . . . 50
Table 5.1 Common quantities in electromagnetic transport energy . . . . 60
Table 6.1 Protein factors for selected substances . . . . . . . . . . . . . . . . . . 74
Table 9.1 Substances with known composition used . . . . . . . . . . . . . . . 130
Table 9.2 Calculation of the calibration constant for hydrogen . . . . . . . 131
Table 9.3 Calculation of the calibration constant
of carbon monoxide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 131
Table 9.4 Calculation of the calibration constant of methane . . . . . . .. 132
Table 9.5 Calculation of the calibration constant of carbon dioxide . .. 132
Table 9.6 Calculation of the calibration constant
of hydrogen sulfide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
Table 9.7 Calculation of the calibration constant of nitrogen . . . . . . . . 133
Table 9.8 Calculation of the calibration constant of oxygen . . . . . . . . . 133
Table 9.9 Calculation of the calibration constant of ethylene . . . . . . . . 133
Table 9.10 Calculation of the calibration constant of ethane . . . . . . . . . . 134
Table 9.11 Summary of the results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
Table 9.12 Elemental composition of the reference substances . . . . . . . . 137
Table 9.13 Calibration coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
Table 9.14 Calibration coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

xxxvii
xxxviii List of Tables

Table 9.15 Calibration coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137


Table 9.16 Calibration coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
Table 9.17 Calibration coefficients . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
Table 9.18 Summary of the results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
Table 12.1 Reaction conditions for the process
of regional biomass conversion . . . . . . . . . . . . . . . . . . . . . .. 161
Table 13.1 Moles of ATP produced on the basis of the particular
enzyme which catalyzes the oxidation reactions . . . . . . . . .. 179
Table 13.2 Some references where mathematical models
are given for different processes involved in anaerobic
digestion and where there is substrate inhibition . . . . . . . . .. 182
Table 17.1 Some properties of bioethanol derived from different kinds
of feedstock . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 234
Table 17.2 Concentration (g/100 g) of the main components
in sugarcane juice and sugarcane beet juice . . . . . . . . . . . .. 240
Table 17.3 Composition of produced molasses from cane and beets . . .. 241
Table 17.4 Composition of crops and residues and theoretical ethanol
yield from agricultural residues . . . . . . . . . . . . . . . . . . . . . .. 249
Table 17.5 Theoretical ethanol yield from crop-processing
by-products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 250
Table 17.6 Examples of values for the main parameters in batch
fermentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 250
Table 18.1 Main esters of biodiesel in reference to different kinds of
feedstock used . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 255
Table 18.2 Examples of ionization potential and electron affinities
(in brackets) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 257
Table 18.3 Composition of waste cooking oils (WCO) . . . . . . . . . . . . .. 262
Table 18.4 Properties of waste cooking oils and diesel . . . . . . . . . . . . .. 263
Table 19.1 Biodiesel characteristics (adapted from
Phan and Phan 2008) . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 274
Table 19.2 Properties of diesel and biodiesel . . . . . . . . . . . . . . . . . . . .. 274
Table 19.3 a, b, c coefficients for different groups . . . . . . . . . . . . . . . .. 275
Table 19.4 Diverse kinds of feedstock with values for their level
of unsaturation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 278
Table 19.5 Fuel properties of different feedstock kinds and biofuels . . .. 279
Table 19.6 Examples of microbial-based source of Biodiesel . . . . . . . .. 280
Table 19.7 Properties of diesel, biodiesel, and different
kinds of fuels. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 282
Table 20.1 Main reactor parameters in FT synthesis . . . . . . . . . . . . . . .. 291
Table 20.2 Main reactor parameters in FT synthesis with conversion
and H2/CO ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 293
Table 20.3 Intrinsic kinetic expressions for the reaction between
carbon monoxide and hydrogen . . . . . . . . . . . . . . . . . . . . .. 295
Table 20.4 Adapted from Yates and Satterfield (1991) . . . . . . . . . . . . .. 295
List of Tables xxxix

Table 20.5 Effect of FT process parameters on the chain growth


probability, olefin/paraffin ratio, carbon deposition, and
methane selectivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 298
Table 20.6 Main reactions involved in a wet flue gas desulfurization
(WFGD) process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 300
Table 20.7 FT synthesis gas cleaning requirements for diverse
contaminant and cleaning methods . . . . . . . . . . . . . . . . . . .. 301
Table 21.1 Fuel properties range for the U-Gas gasifier . . . . . . . . . . . .. 317
Table 25.1 Composition of common types of biomass, percentages
in weight . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 376
Table 25.2 Proximate values for wood, paper, peat, and coal.
A comparison . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 377
Table 25.3 Proximate and ultimate analysis of some biomass feedstock
with HHV values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 380
Table 25.4 Properties of biomass and their effect on thermochemical
processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 382
Table 25.5 Moisture, ash, volatile matter, HHV, fixed carbon
and elemental analysis for diverse kind of trees . . . . . . . . .. 384
Table 25.6 Heating values, moisture and ash content of some
examples of biomass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 385
Table 25.7 Different property values for agrofuels . . . . . . . . . . . . . . . .. 385
Table 25.8 Typical air requirements for agrofuel expressed
as kg/kcal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 385
Table 25.9 Main properties of peat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 386
Table 25.10 Ultimate analysis on dry and ash free basis of peat. . . . . . . . 386
Table 25.11 Combustion properties of minor agrofuels. . . . . . . . . . . . . . . 387
Table 25.12 Moisture content, HHV, LHV, density, and calorific value
for wood-related biomass . . . . . . . . . . . . . . . . . . . . . . . . . .. 388
Table 25.13 Proximate and ultimate analyses for corn, straw, husks,
and wood chips . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 389
Table 25.14 HHV and LHV for selected agricultural residues, crops,
forest, and urban residues . . . . . . . . . . . . . . . . . . . . . . . . . .. 390
Table 25.15 Thermodynamic properties for selected alkane
compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 391
Table 25.16 Parameters used in Table 25.15 . . . . . . . . . . . . . . . . . . . . . .. 393
Table 26.1 Enthalpy calculation and related coefficients for Eq. (26.6)
for selected components. Data from Van Loo and
Koppejan (2008) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 399
Table 26.2 Enthalpy calculation and related coefficients for Eq. (26.6)
for selected components . . . . . . . . . . . . . . . . . . . . . . . . . . .. 400
Table 26.3 Coefficients for specific heat calculations related to
methanol, CO, and hydrogen. . . . . . . . . . . . . . . . . . . . . . . .. 401
Table 27.1 Reduced and oxidized states of some elements with
examples (Drapcho et al. 2007) . . . . . . . . . . . . . . . . . . . . . .. 405
xl List of Tables

Table 27.2 Ionization potential (first value) and electron affinities


(second value) in eV (Jones and Fleming 2014) . . . . . . . . .. 406
Table 27.3 Periodic table of the ionization potentials (Jones and
Fleming 2014) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 407
Table 27.4 Quantum representation of atoms, quantum numbers
(Jones and Fleming 2014) . . . . . . . . . . . . . . . . . . . . . . . . . .. 408
Table 27.5 Electronegativity values for the components of the periodic
table (Jones and Fleming 2014) . . . . . . . . . . . . . . . . . . . . .. 409
Table 27.6 Standard electrochemical potentials for biological half-cell
reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 410
Part I
Notions of Biomass Formation
and Development with Some
Analytical Methods
Chapter 1
Introduction

Give me a lever and I will move the Earth.


Sentence generally attributed to Archimedes of Syracuse,
Sicily. 287–212 BC. The author of the: On the Equilibrium of
Planes and father of the first laws of mechanics.

1.1 Biomass as a “Almost” Renewable Energy Source

It is well known that alternative source of energy is required and necessary because
of the increasing pollution and global warming. The world population is reaching
8 billion, and the world energy demand is constantly increasing, as can be seen from
Figs. 1.1 and 1.4.
Up to now, we cannot rely on a sustainable system of energy conversion. This
means that our resources are going to be depleted within decades according to some
forecasts. As a matter of fact, the end of “cheap” oil has been already predicted in the
last decades and inevitably this will happen sooner or later (Campbell and Laherrere
1998).
The increased amount of CO2 in the atmosphere is to be taken seriously into
account since the current situation is that we have 350–450 ppm of CO2 in the
atmosphere and the acceptable levels are at less than 600 ppm. The very dangerous
situation is that the CO2 increases the temperature of our atmosphere and an increase
of temperature reflects on higher solubility of calcium carbonate rocks (De Blasio
et al. 2009). Now the question is, how much CO2 is actually stored within sedimentary
rocks or oceans? The amount is surely scary, considering that a big portion of the
earth rocks is actually constituted by this kind of formation. Also to be noticed that
enormous amounts of CO2 is actually stored within the oceans, and if the temperature
of the oceans will increase, this will cause the release of huge quantities of CO2 , and
in return, cause the dissolution of the sedimentary rocks which are sited at the bottom
of the oceans. This gives terrible results.
Let us look at these facts a little closer. In the literature (Houghton and Woodwell
1989), it is possible to find data on the main carbon flows at global level and some

© Springer Nature Switzerland AG 2019 3


C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_1
4 1 Introduction

Fig. 1.1 World total population by year. Data from Statista (2018)

more data will be also given later on in this manuscript. Carbon is stored in the
world vegetation, living animals and organisms, the soil, the atmosphere, sedimentary
rocks, oceans and of course fossil fuels. Now if we consider that the amount of CO2
stored in biomass would be 500 Gton (we do not need to be precise here), the amount
of CO2 stored within the oceans would be five times more! And we are just looking
at the oceans. On the other hand, if the CO2 released by burning fossil fuels is
around 5 Gton per year, then the amount of CO2 transferred from and to the oceans
will be 20 times more. Plants help in capturing around 50 Gton of carbon per year.
Fortunately at the present times, CO2 is captured by oceans and because of human
destruction of the environment and fossil fuels, the net increase of carbon in this
cycle is around 3 Gton per year. On the basis of these reflections, the carbon dioxide
equilibrium is dependent mostly on solid–liquid reactions systems as demonstrated
here (De Blasio et al. 2013).

MgCO3(s) ↔ Mg2+ + CO2−


3 (1.1)

CaCO3(s) ↔ Ca2+ + CO2−


3 (1.2)


H3 O+ + CO2−
3 ↔ HCO3 + H2 O (1.3)

H3 O+ + HCO−
3 ↔ CO2(aq) + 2H2 O (1.4)

CO2(aq) ↔ CO2(g) (1.5)


1.1 Biomass as a “Almost” Renewable Energy Source 5

In addition to the equilibrium reactions given previously, the dissociation of water


should also be taken into consideration.

2H2 O ↔ H3 O+ + OH− (1.6)

Equilibrium constants and related thermodynamics for the above mentioned reactions
can be found in the literature, Sillén and Martell (1965) or Millero (1979). Here, it
is sufficient to point out that there is indeed huge potential for catastrophic events if
these equilibria are distorted. It should be also acknowledged that indeed the ocean
water is working as a buffer in this sense and it would be not that easy to affect these
equilibrium states, however the increment of temperature could play a major role.
Some compounds which are contributing to the degree of alkalinity in oceans
are shown in Fig. 1.2. Here, the carbon dioxide is in equilibrium with the ocean
water which is absorbing it. The carbon dioxide which is solubilized in water is
also in equilibrium with the carbonic acid, the bicarbonate, and the carbonate ions.
In addition, there are a number of compounds which are contributing further to the
alkalinity of water.
The world is also subject to normal and very regular oscillations; for instance,
every year we have winter and summer and the change of the seasons heavily affects
the biomass production. As a consequence, the CO2 captured and stored by plants
and microorganisms is also subject to cycles. This reflects on the atmospheric con-
centration of carbon dioxide as well; as a matter of fact, it is reported in the literature
that the amplitude of the CO2 oscillation within the year between the north and south
hemisphere is 15 times the accumulation rate (Box 1988; D’Arrigo et al. 1987; Fung
et al. 1987). It is also reported that this amplitude is increasing with time (Bacastow
et al. 1985). Photosynthesis and emissions play a major role within this equilibrium,

Fig. 1.2 Alkalinity of oceans and the CO2 equilibrium


6 1 Introduction

Fig. 1.3 A simple


schematics on the relation
between CO2 concentration,
atmospheric temperature,
and oceans pH

and the alkalinity of water is also something that should be really monitored. A
simple representation of this cycle is given in Fig. 1.3.
To be noticed that a decrease of pH of the oceans within a 0.1 range could give
an enormous increase of the concentration of CO2 in the atmosphere.
Hydrocarbons (including biofuels) are light compared to other kind of fuels. For
instance, it is well known that batteries are quite heavy when we buy them at the
shop and to give a further example, the battery of my electric scooter weights around
12 kg and I mostly can do 10 km when it is fully charged. To be specified that the
battery I am using is not among the best on the market, but still there is no doubt that
they are heavy. Compared to the metals of batteries, the molecular weight of carbon
and hydrogen is very low. Furthermore, the amount of energy in hydrocarbons is
very high compared to their weight and also it has to be noticed that it is very easy
to store liquid fuels and they are mostly not explosive. Hydrogen is very powerful,
but it is not that safe to use it. In fact, if hydrogen comes into contact with oxygen at
high temperatures, it can give some explosions.
For this reason, when hydrogen is used for experiments in laboratory it is a good
idea to have extra precautions concerning safety. For instance, a Plexiglas protection
is needed/suggested to be put as protection prior the reactor area.
Aircraft are well known to consume enormous quantity of fuel when operating;
therefore, also for this process, the utilization of renewable kinds of feedstock has
been proposed. One example is the Neste MY Renewable Jet fuel and, in addition,
there is an increasing will from the European Union to increase the production of
biofuels for the aircraft sector. Concerning the aviation and the related fuels, in this
case, there are not much of alternative choices, since the fuels should be in any case
liquid at the flying conditions and they should still contain a sufficient amount of
energy per unit mass.
Biomass is considered in the literature as a renewable type of energy resource for
the reason that plants, for instance, utilize solar energy and CO2 from the atmosphere
to develop themselves. In this way, the total carbon emissions are considered to be
zero and biomass is said to not contribute to the global warming effect. Nevertheless,
processes like photosynthesis (which will be considered in the following sections)
are not limited to plants, but are extensively used by algae (Thomas et al. 1984).
It is important to keep this in mind, especially when algae are considered to fuel
the third-generation biorefineries of the future and that biodiesel can be produced
by using them. The concepts of microorganisms’ growth and the kinetics involved
1.1 Biomass as a “Almost” Renewable Energy Source 7

will be considered in the second part of this manuscript along with some examples
of algae utilization.
Algae which utilize photosynthesis can produce diverse kinds of oils and ethanol,
and it has also referred that they represent the energy of the future by representatives
of some companies working in this field (Mascarelli 2009). The very good thing
with algae is that they could grow in spaces where they are not competing with other
biomass sources, for instance they could be cultivated off shore. However, at this
moment, the production of algae for these purposes is indeed expensive, and it is
difficult for this technology to compete with fossil fuels considering the actual prices
(Mascarelli 2009).
It is not always feasible to utilize wood-based biomass for energy conversion. As
a matter of fact, it is indeed true that water and the energy of the sun are available
in large quantities (this is true mostly for the second statement); however, plants do
need to have inorganic materials in order to function and grow. For this reason, the
depletion of land and soil should be taken into account when planning of producing
crops with the purpose of energy conversion since there is a diminishing availability
of resources for these scopes (Breure et al. 2018).
Of course, if the public do not have knowledge of how the plant actually works
to grow themselves up, it is very probable that they will consider trees as a very
nice source of energy and propose their utilization for large-scale energy production
(Fig. 1.4). Of course politicians are no exception. Knowing the fundamental concepts
of biomass generation and the related simple mass balances, will give the foundations
for a critical thinking on the matter of feasibility of biomass and “power production.”
Biomass can be produced constantly, even if some types of biomass are seasonal,
and they can offer short-term sequestration of carbon in non-harvested aboveground

600.0

500.0

400.0
Btu.1015

300.0

200.0

100.0

0.0
1980 1990 2000 2010 2020 2030 2040 2050

Non-OECD OECD

Fig. 1.4 World energy consumption, 1990–2040 (Btu·1015 ). Data from US Energy Information
Administration (2013)
8 1 Introduction

Fig. 1.5 World carbon dioxide emissions by region, left: all OECD countries; right: OECD Europe.
Data from US Energy Information Administration (2013)

Fig. 1.6 World gross domestic product (GDP) by region. Data from US Energy Information Admin-
istration (2013)

growth and medium-term sequestration in belowground root biomass. The world


amount of CO2 emissions is reported in Fig. 1.5.
Figure 1.6 shows instead an updated version from 2013, of the gross domestic
product, GDP, for OECD countries divided by continents. Also in this case, it is
possible to notice that the GDP projection to 2040 is steadily increasing; this means
that the production will also increase and for this reason the energy demand.
Following the need for alternative forms of energy, the usage of renewables is also
increasing; this is shown in Fig. 1.7 where the overall consumption of renewables is
reported by regions and for only the European OECD countries.
However, there are more recent studies which show that the extensive and large
use of biomass for industrial purposes of producing energy is not feasible and it
would not give a null net carbon emission balance. One study is reported here as
an example of this theory (Schulze et al. 2012); the study done at the Max Planck
Institute together with other important institutions refers that with an extensive usage
of biomass, we will have younger forest and less forest pools, depleted soil nutrients
together with loss of the ecosystem function of the forests themselves. The depletion
of the soil fertility will make the extensive use of biomass unsustainable.
1.1 Biomass as a “Almost” Renewable Energy Source 9

Fig. 1.7 World consumption of renewable energy by region. Data from US Energy Information
Administration (2013)

Table 1.1 Some studies Study References Study References


concerning biomass supply
and its usage in energy WEC World LESS/BI Williams
conversion Energy (1995)
Council
(1994)
IIASA-WEC Grubler et al. LESS/IMAGE Leemans
(1996) et al. (1996)
FFES Lazarus et al. BATTJES Battjes
(1993) (1994)
EDMONDS Edmonds GLUE Yamamoto
et al. (2003) et al. (1999)
SWISHER Swisher FISHER Fischer and
(1993) Schratten-
holzer
(2001)
USEPA Lashof and DESSUS Dessus et al.
Tirpak (1992)
(1990)
SØRENSEN Soerensen SHELL Shell (1995)
et al. (1999)
HALL Hall (1993) SRES/IMAGE Nakicenovic
et al. (2000)
RIGES Johansson
(1993)

Other than utilizing only biomass, the solution would be to utilize all the possible
alternatives and renewable energy sources and at the minimum possible rate. Meaning
that the overall consumption of energy must be reduced.
The usage and the contribution of biomass supply for energy conversion has been
studied by several authors, and different scenarios have been provided. Table 1.1
shows some of the most important references on this matter. In the literature, there
is information about the approach and the time frame taken by these authors and
organizations.
10 1 Introduction

40
Gigajoules per hectare of culƟvated land

35

30 1990

25 2000
2010
20
2020
15
2030
10
2040
5 2050 2050
0 2030
AFR CPA 2010
EEU FSU LAM MEA NAM 1990
PAO PAS SAS WEU

Fig. 1.8 Potential bio-energy of crop residues by world region, gigajoules per hectare of cultivated
land, NAM, North America; LAM, Latin America; WEU, Western Europe; FSU, former Soviet
Union; MEA, Middle East and North Africa; CPA, Centrally planned Asia and China; SAS, South
Asia; PAS, other Pacific Asia; PAO, Pacific OECD

An estimated bio-energy potential for the crop residues is given in Fig. 1.8; data
were retrieved from Fischer and Schrattenholzer (2001).
From Fig. 1.8, it is possible to notice that there will be an increase of the bio-energy
potential, and this is because there will be agricultural progress, at least according to
the forecasts provided.
In a report given by the IIASA (Nakicenovic and Riahi 2001), the economic
evolution of the world is given following different scenarios according to:
• Scenario A1: Rapid economic growth;
• Scenario A2: A differentiated world with conservation of local entities;
• Scenario B1: Convergent world with a rapid change;
• Scenario B2: The focus here is on local sustainability solutions.
The report gives a detailed analysis and forecast about future energy demand and
economic growth.
In this text book, the focus will be not only dedicated to energy crops, but also
on other types of biomass utilized for power generation; however, since there are
many options and the topic is really vast, it was preferred to privilege only some
types of biomass. For this reason in the first part of the manuscript, the focus will
be on light harvesting and how plants generate work in order to bind CO2 during
photosynthesis. Only the main routes for this process will be covered, and after this,
the main procedures used for the assessment of biomass before utilization will be
treated with particular concentration on proximate and ultimate analysis.
References 11

References

Bacastow, R. B., Keeling, C. D., & Whorf, T. P. (1985). Seasonal amplitude increase in atmospheric
CO2 concentration at Mauna Loa, Hawaii, 1959–1982. Journal of Geophysical Research: Atmo-
spheres, 90(D6), 10529–10540. https://doi.org/10.1029/JD090iD06p10529.
Battjes, J. J. (1994). Global options for biofuels from plantations according to IMAGE simulations
(No. IVEM-SR-77). Rijksuniversiteit Groningen (Netherlands). Interfacultaire Vakgroep Energie
en Milieukunde. Retrieved from http://inis.iaea.org/Search/search.aspx?orig_q=RN:26031212.
Box, E. O. (1988). Estimating the seasonal carbon source-sink geography of a natural, steady-
state terrestrial biosphere. Journal of Applied Meteorology, 27(10), 1109–1124. https://doi.org/
10.1175/1520-0450(1988)027%3c1109:ETSCSS%3e2.0.CO;2.
Breure, A. M., Lijzen, J. P. A., & Maring, L. (2018). Soil and land management in a circular
economy. Science of the Total Environment, 624, 1125–1130. https://doi.org/10.1016/j.scitotenv.
2017.12.137.
Campbell, C. J., & Laherrere, J. (1998). The end of cheap oil. Scientific American, 278(3). Retrieved
from https://www.jstor.org/stable/26057679.
D’Arrigo, R., Jacoby, G. C., & Fung, I. Y. (1987). Boreal forests and atmosphere–biosphere exchange
of carbon dioxide. Nature, 329(6137), 321–323. https://doi.org/10.1038/329321a0.
De Blasio, C., Ahlbeck, J., & Westerlund, T. (2009). Modeling the hydrodynamics and mass-transfer
phenomena for sedimentary rocks used for flue gas desulfurization. The effect of temperature. In
R. M. de Brito Alves, C. A. O. do Nascimento, & E. C. Biscaia (Eds.), Computer aided chemical
engineering (Vol. 27, pp. 411–416). Elsevier. https://doi.org/10.1016/S1570-7946(09)70289-5.
De Blasio, C., Carletti, C., Westerlund, T., & Järvinen, M. (2013). On modeling the dissolution of
sedimentary rocks in acidic environments. An overview of selected mathematical methods with
presentation of a case study. Journal of Mathematical Chemistry, 51(8), 2120–2143. https://doi.
org/10.1007/s10910-013-0202-3.
Dessus, B., Devin, B., & Pharabod, F. (1992). World potential of renewable energies actually
accessible in the nineties and environmental impacts analysis. Houille Blanche, 47(1), 21–70.
Edmonds, J. A., Wise, M. A., Sands, R., Brown, R., & Kheshgi, H. (2003). Agriculture, land use, and
commercial biomass energy: A preliminary integrated analysis of the potential role of biomass
energy for reducing future greenhouse related emissions. In Proceedings of the 6th Greenhouse
Gas Control Technologies Conference (pp. 0-08-044045–2). Oxford UK: Elsevier Inc.
Fischer, G., & Schrattenholzer, L. (2001). Global bioenergy potentials through 2050. Biomass and
Bioenergy, 20(3), 151–159. https://doi.org/10.1016/S0961-9534(00)00074-X.
Fung, I. Y., Tucker, C. J., & Prentice, K. C. (1987). Application of advanced very high
resolution radiometer vegetation index to study atmosphere-biosphere exchange of CO2 .
Journal of Geophysical Research: Atmospheres, 92(D3), 2999–3015. https://doi.org/10.1029/
JD092iD03p02999.
Grubler, A., Jefferson, M., & Nakicenovic, N. (1996). Global energy perspectives: A summary of
the joint study by IIASA and world energy council (Monograph). Retrieved July 9, 2018, from
http://pure.iiasa.ac.at/id/eprint/4860/.
Hall, D. O. (1993). Biomass for energy: Supply prospects. In Renewable energy: Sources for fuels
and electricity (pp. 593–651). Washington D.C.: Island Press.
Houghton, R. A., & Woodwell, G. M. (1989). Global climatic change. Scientific American, 260(4),
36–47.
Johansson, T. B. (1993). A renewables-intensive global energy scenario. In Renewable energy:
Sources for fuels and electricity (pp. 1071–1143). Washington D.C.: Island Press.
Lashof, D. A., & Tirpak, D. A. (1990). Policy options for stabilizing global climate. U.S.: Environ-
mental Protection Agency.
Lazarus, M., Greber, L., Hall, J., Bartels, C., Bernow, S., Hansen, E., … Von Hippel, D. (1993).
Towards a fossil free energy future. The next energy transition. Stockholm Environment Institute
Boston Center.
12 1 Introduction

Leemans, R., van Amstel, A., Battjes, C., Kreileman, E., & Toet, S. (1996). The land cover and
carbon cycle consequences of large-scale utilizations of biomass as an energy source. Global
Environmental Change, 6(4), 335–357. https://doi.org/10.1016/S0959-3780(96)00028-3.
Mascarelli, A. L. (2009). Gold rush for algae. Nature, 461(7263), 460–461. https://doi.org/10.1038/
461460a.
Millero, F. J. (1979). The thermodynamics of the carbonate system in seawater. Geochimica et
Cosmochimica Acta, 43(10), 1651–1661. https://doi.org/10.1016/0016-7037(79)90184-4.
Nakicenovic, N., & Riahi, K. (2001). An assessment of technological change across selected energy
scenarios (Monograph). Retrieved July 9, 2018, from http://pure.iiasa.ac.at/id/eprint/6521/.
Nakicenovic, N., Alcamo, J., Grubler, A., Riahi, K., Roehrl, R. A., Rogner, H.-H., & Victor, N.
(2000). Special report on emissions scenarios (SRES), a special report of working group III of the
intergovernmental panel on climate change. Cambridge: Cambridge University Press. Retrieved
from http://pure.iiasa.ac.at/id/eprint/6101/.
Schulze, E.-D., Körner, C., Law, B. E., Haberl, H., & Luyssaert, S. (2012). Large-scale bioenergy
from additional harvest of forest biomass is neither sustainable nor greenhouse gas neutral. GCB
Bioenergy, 4, 611–616. https://doi.org/10.1111/j.1757-1707.2012.01169.x.
Shell. (1995). Evolution of the world’s energy system 1860–2060. London: Shell Center.
Sillén, L. G., & Martell, A. E. (1965). Stability constants of metal-ion complexes (Sillen, Lars
Gunnar; Martell, Arthur E.). Journal of Chemical Education, 42(9), 521. https://doi.org/10.1021/
ed042p521.1.
Soerensen, B., Meibom, P., & Kuemmel, B. (1999). Long-term scenarios for global energy demand
and supply. Four global greenhouse mitigation scenarios. Final Report (No. IMFUFA-TEKST--
359). Roskilde Universitetscenter (Denmark): Inst. for Studiet af Matematik og Fysik samt deres
Funktioner i Undervisning. Retrieved from http://inis.iaea.org/Search/search.aspx?orig_q=RN:
30029769.
Statista. (2018). World—Total population 2007–2017 | statistic. Retrieved December 31, 2018, from
https://www.statista.com/statistics/805044/total-population-worldwide/.
Swisher, J. (1993). Renewable energy potentials, Chap. 3. Energy, 18(5), 437–459. https://doi.org/
10.1016/0360-5442(93)90022-6.
Thomas, W. H., Seibert, D. L. R., Alden, M., Neori, A., & Eldridge, P. (1984). Yields, photosynthetic
efficiencies and proximate composition of dense marine microalgal cultures. I. Introduction and
Phaeodactylum tricornutum experiments. Biomass, 5(3), 181–209. https://doi.org/10.1016/0144-
4565(84)90022-2.
U.S. Energy Information Administration. (2013). International energy outlook 2013. Retrieved
from http://www.eia.gov/forecasts/ieo/.
Williams, R. H. (1995). Variants of a low CO2 -emitting energy supply system (LESS) for the world.
Prepared for the IPCC Second Assessment Report Working Group IIa, Energy Supply Mitigation
Options.
World Energy Council. (1994). New renewable energy resources. Kogan Page Ltd.
Yamamoto, H., Yamaji, K., & Fujino, J. (1999). Evaluation of bioenergy resources with a global land
use and energy model formulated with SD technique. Applied Energy, 63(2), 101–113. https://
doi.org/10.1016/S0306-2619(99)00020-3.
Chapter 2
Light Harvesting and Biomass
Generation

So when we consider the Photosystem II and their reaction


centers, they are producing oxygen and oxygen in combination
with light is very often very toxic. What happens there is that you
get an oxidative degradation of one protein unit and nature has
developed a way to exchange these damaged proteins every 20
minutes. So every 20 minutes a repair has to be made and I do
not think that we can mimic this process in a technical manner.
Michel Hartmut, Nobel Prize in Chemistry 1988, talking
about renewable energy in 2008.

2.1 Preliminary Data Related to Light Harvesting


and Biomass Generation

The question of how the biomass is produced is central to understand the concept
of feasibility of this feedstock in energy conversion processes. All biological pro-
cesses utilize mechanisms of oxidation of substrate molecules and reduction of final
acceptors, and photosynthesis is one good example which allows for understanding
many useful concepts. In the photosynthetic process, there is oxidation of a substrate
molecule, water, followed by the generation of an electrochemical potential and a
related “current” of electrons. The basic explanation of this process is reported here
for pedagogical purposes to demonstrate the analogy between biological mechanisms
and some technical solutions.
According to Ksenzhek and Volkov (1998), 105 net Gton of carbon is fixed by
photosynthesis on land and into the oceans every year. While Kell (2012) refers that
soil could accommodate 50 ton ha−1 of carbon in the same period. On the other
hand, 260 Gton of oxygen are released into the atmosphere every year (Ksenzhek
and Volkov 1998). We can consider that the amount of oxygen given by plants could
be increasing with time considering that today, plants store 25% of all the carbon
emissions produced by the utilization of fossil fuels (Owen 2017). In the world, there
are more than 390,000 known species of plants (Chapman 2009), and it is reported
in the literature (Ksenzhek and Volkov 1998) that the atmosphere contains 1.2 × 106
Gton oxygen.
© Springer Nature Switzerland AG 2019 13
C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_2
14 2 Light Harvesting and Biomass Generation

Table 2.1 Carbon release per


Fuel kg C/GJ
GJ for some biomass types
and fossil fuels, data from Wood 26.1–29.9
Ksenzhek and Volkov (1998) Peat 30.0
Coal 23.9–25.8
Crude oil 19.0–21.4
Natural gas 13.6–15.4

Furthermore in all OECD countries, the amount of CO2 released by fuel combus-
tion was 11,720 Mton in 2015 (IEA 2017), and it can be noticed that 1 GJ of energy
from oil results in the net release of 73 kg CO2 while natural gas gives 52 kg of CO2
for 1 GJ (Ksenzhek and Volkov 1998). In all the world, the amount of CO2 released
by fuel combustion was 32,294.2 Mton in 2015 (IEA 2017).
Concerning the carbon release from biomass and fossil fuels, Table 2.1 gives some
data related to the most common types of feedstock.
Some estimations for the potential biomass production give that the power which
could be obtained by photosynthetic processes could be between 1500 and 2250 TW,
while the total power coming from the sun is estimated to be 178,000 TW (Pisciotta
et al. 2010). It is also estimated that 220 billion ton of dry biomass can be produced
per year, and this would be equal to 4 × 1018 MJ of energy. In addition, the standing
biomass is estimated to represent 36 × 1018 MJ (Ksenzhek and Volkov 1998).
From the previously presented data, it can be noticed that there is huge potential
concerning the utilization (and the production) of biomass for the purposes of power
generation. There are, however, positive and negative aspects with biomass produc-
tion for the purposes of energy conversion. Among the positive aspects, we have as
follows:

• The environmental impact varies in relation to the source and the applied technol-
ogy; however, there is much less impact than the fossil fuels in particular for the
emissions of greenhouse gases.
• It is possible to evaluate the total reduction of CO2 and CH4 emissions taking into
consideration the analysis of the life cycle for the renewable sources.
• The renewable alternatives have a low environmental impact and offer a possibility
for economic and social development; new work places will be created, and there
will be a positive impact on the territory.

Among the limits of biomass production and with reference to energy crops:

• They are available at intervals of time with no continuity.


• It is necessary to utilize a large surface in order to have a significant amount of
potential energy.
• The climate of course plays a fundamental role in biomass growth.

Some examples of biomass which could be utilized for power conversion:


2.1 Preliminary Data Related to Light Harvesting and Biomass Generation 15

• wood and wood waste;


• municipal organic solid waste (one example of activities which could be done in
this case is the incineration and methane utilization from landfills);
• herbaceous biomass and agricultural residues;
• aquatic biomass;
• industrial solid organic wastes;
• municipal bio-solids;
• products from farm animal manures;
• algae.

Among the lignocellulosic kinds of biomass, we have:

• agricultural residues, residues from forests, paper industry, and wood industry;
• energetic cultivations.

Of course there is also a series of microorganisms to be considered as biomass, and


these microbes can indeed be utilized for energy conversion. For instance, microal-
gae and several microorganisms can drive fermentation and production of biogas.
However, these concepts will be covered within the second part of this manuscript.

2.2 Light Harvesting

In a very general sense, biofuels are derived from the energy coming as radiation
from the sun. The incoming energy is “harvested” by plants and microorganisms, and
by using the existing CO2 of the atmosphere, new biomass can be then produced. In
order to develop biomass (grow), a plant has to perform a series of activities which
are listed as follows:

• The photosynthetic process needs water and light to work. Now in case we consider
a tree, the water should be brought up where the leaves are. For this reason, water
should be taken from the soil.
• The same process in order to work needs minerals and therefore, extraction of
mineral components from soil is also necessary;
• The nitrogen should be bond into the process as well, for instance activities related
to its storage and cycling (Li and Coleman 2018);
• Finally, the carbon dioxide of course should be taken from air. This can be also
directly related to a novel trend of research which includes the artificial photosyn-
thesis (Liu et al. 2016).

In addition to all these operations, there are all a series of other activities to be
performed and these activities have the goal of synthesizing the organic matter:

• The plant should obtain the photosynthesis products from a larger surface (the
leaves).
• Transport of metabolites and metals through the system (Kobayashi et al. 2019).
16 2 Light Harvesting and Biomass Generation

• Intracellular movements (for instance, in plant cells and among which chloroplasts
or nucleus) (Nagai 1993).
• Maintenance of an electrochemical potential difference on cellular membranes
(De Vrieze et al. 2018).
• Building of plant structures on a cellular and macroscopic level.
• Plant tropism (Gilroy and Masson 2008).

In order to do all this work, plants need energy. The source of all this energy is
the sun: Solar radiation is collected and used by plants to perform the photolysis of
water (this is also referred as photochemical water treatment) (Furatian and Mohseni
2018), produce a transmembrane electrochemical potential, and finally bond CO2
molecules with the help of ATP molecules. This process is called photosynthesis.
The incoming radiation is of course measured in terms of energy and power, and
for a determinate incoming radiation, there will be some production of biomass.
Now, not all the plants for instance have the same capability of producing biomass
from incoming radiation from the sun and therefore, they are classified and divided
by their radiation use efficiency, RUE. One preliminary example is given here for the
type of cultivation Nicknejad Cultivar, and the value of the RUE in this case is 0.84 g
(of biomass produced) on MJ of radiation coming on 1 m2 of land (Miranzadeh et al.
2011). Several examples of RUE for diverse kinds of biomass will be given later in
this manuscript.
Coming back to the introduction of the photosynthesis, this process produces
sugar and oxygen from CO2 and water with the help of solar radiation following the
general reaction:

6CO2 + 6H2 O → C6 H12 O6 + 6O2 (2.1)

There are two kinds of photosynthesis: oxygenic and anoxygenic (Blankenship et al.
2007); in oxygenic photosynthesis, there is production of oxygen (Nelson 2011),
while the anoxygenic photosynthesis is carried on by oxygen-intolerant prokaryotic
organisms or phototrophic bacteria (Qi et al. 2018) which utilize sun radiation and
bacteriochlorophyll.
Light harvesting is done only at determinate wavelength or frequencies, and this
is because only a part of the photosynthetic spectrum has the necessary energy to
carry on the required steps for water oxidation (McCree 1981). This can be described
more simply by the following example: Let us assume that we want to break a stone
by letting the stone to fall down from a determinate altitude; we could see that the
altitude has to be necessary in order for the stone to gain the required kinetic energy
from its potential energy. We could see, for instance, that the energy required to break
the stone could not be sufficient if the object does not gain the required energy. In
the same way, the photons should have a sufficient energy to allow for the process
of oxidation.
It is good to remember that oxidation means that we are taking electrons from a
substrate, for instance if we oxidize iron means that electron are taken from it. On the
other hand, reduction indicates that we are actually giving electrons to a substrate, for
2.2 Light Harvesting 17

this reason an object receiving electrons is said to be reduced. The radiations utilized
for the photosynthetic processes are included in the so-called photosynthetic active
radiation, PAR (Sun et al. 2017); this means that radiations with an energy less than
PAR are discarded and not used. An electromagnetic wave, such as light, is produced
by the movement of electrons in the source of the radiation itself. Electrons in atoms
are situated in energy levels (Jones and Fleming 2014), and when they move from
one level to another, they release the difference in energy between the two levels;
this energy is released as an electromagnetic wave. There is much theory about this,
and students interested more on the argument can read further references of physics
and even better, quantum mechanics. Here only some basic information about the
atoms will be given, and more information will be given during the course of the
manuscript when the oxidation state of atoms and molecule will be described. For a
more detailed description of where electrons are located around the nucleus of atoms
and how they are involved in chemical bonds, I invite the readers to refer to textbooks
like Brown et al. (2014) or Jones and Fleming (2014).
The velocity of an electromagnetic wave, c, is directly proportional to its fre-
quency, ν, and its wavelength, λ, and it is given by:

c =ν·λ (2.2)

and the energy associated with it is:

E =h·ν (2.3)

where h is the Planck constant (6.626 × 10–34 J s) and ν is the frequency of the
wave. So as seen from the above equations, the higher the frequency, the lower is the
wavelength, and the more is the value of the frequency the more is the energy carried
by the wave. The constant c is the speed of the wave; now since light is considered
as an electromagnetic wave, c is the speed of light by definition. The wave velocity
is a function of the medium where the wave propagates; in the vacuum, the speed
of light is c = 299.8 × 103 km/s, and in water for instance, the speed of light is
different and equal to c = 224.9 × 103 km/s (Resnick et al. 2001). The speed of the
electromagnetic wave in the medium is a function of the index of refraction of the
same medium in the form:
c
n= (2.4)
v
where n is the index of refraction, c is the speed of light in vacuum, and v is the
velocity of the wave in the medium considered. Table 2.2 gives some values for the
refraction index for selected materials, with references.
Figure 2.1 shows the visible spectrum among a range of wavelengths.
It is possible to notice from Fig. 2.1 that the visible spectrum represents only a
very small region of all the possible values for the frequency and the wavelengths;
the infrared region is at lower frequency (larger wavelength), and the ultraviolet is
18 2 Light Harvesting and Biomass Generation

Table 2.2 Values for the refraction index for selected materials, with references
Material n References Material n References
Vacuum 1 Fused silica 1.45282 Malitson (1965)
glass
Air 1.0003 Ciddor (1996) Sodium chloride 1.52894 Bass et al.
(2009)
Water 1.42433 Daimon and Cellulose 1.4608 Kasarova et al.
Masumura (2007)
(2007)
Ethyl 1.36652 Rheims et al. Gold 0.28 Palik (1985)
alcohol (1997)
Calcite 1.63703 Ghosh (1999) Glycerol 1.4631 Rheims et al.
(1997)

Fig. 2.1 Visible spectrum among the total wavelength spectrum. Adapted from Taiz et al. (2014)

at higher frequency and shorter wavelengths. Energy is coming from a source as


electromagnetic waves, a part of this energy is reaching a determinate object, and a
simple energy balance on the object would give that a part of the incoming energy is
absorbed; a part is transmitted through the object, and a part is reflected. So the total
incoming radiation would then be:

TR = α · (TR) + τ · (TR) + ρ(TR) (2.5)

where TR, α, τ, ρ are, respectively, the total incoming radiation, the absorbance, the
transmittance, and the reflectance coefficient.
For this reason, when the radiation arrives on the earth, it has to go through the
atmosphere in order to reach its surface. Figure 2.2 shows the energy arriving on the
earth surface as a function of the wavelength.
It is possible to notice from Fig. 2.2 that the energy arriving on the surface is lower
at larger wavelengths and it is not a smooth curve; this is due to the fact that at larger
2.2 Light Harvesting 19

Fig. 2.2 Energy at the earth’s surface. Adapted from Taiz et al. (2014)

wavelengths, we have lower energy of the wave (because we have less frequency)
and also because the atmosphere does not absorb in the same way at all wavelengths.
This is mostly due to the absorbance of the steam in the atmosphere.
Like the atmosphere, also plants absorb incoming sun radiation. The incoming
radiation absorbed by chlorophyll is reported in the literature as a function of the
wavelength. It is possible to see that between 550 and 600 nm, chlorophyll does not
absorb and instead it reflects light at that wavelength. This is just the wavelength of
the green color, and this is why chlorophyll appears to be green at our eyes. This is
still valid for chlorophyll-sensitized solar cells (CSSCs) for which chlorophyll shows
a similar behavior; see Fig. 2.3.
Figure 2.4 describes the possible states of chlorophyll. Radiation is arriving at an
energy level, and it is giving energy at the surface of the plant. The incoming energy
at higher level (more energy per photon) is then absorbed, lost as heat or reflected
as, for instance, luminescence.
The most common method used by plants to grow is the photosynthesis carried
out by means of chlorophyll. Chlorophylls are only some of the photosynthetic units;
they constitute the light-harvesting (LH) antenna system. The energy in the antenna
complex (or light-harvesting complex, LHC) is transferred by resonance transfer,
and in the literature, it is also mentioned that energy transfer is done by “jumping”
of energy among discrete energy levels (Blankenship 2013; Mirkovic et al. 2017;
Scholes and Fleming 2005). We could do here the hypothesis that 95–99% of the
photons absorbed by the antennae are transferred to the reaction center. Within the
LHC, the photosystem II is composed by diverse structures; for instance, it is referred
that it is constituted by six types of light-harvesting complexes (Xu et al. 2017).
In the first step of the photosynthetic process, light is absorbed by the pigments
associated with the antenna proteins (the LHCs) which are organized as hetero-
oligomers.
20 2 Light Harvesting and Biomass Generation

The way in which the pigments are organized and linked to the LHC determines
the way in which the energy is transferred. This energy leads to the formation of
chemical reactions, and this phenomenon is called photochemistry. As referred by
Xu and collaborators, the most studied LHC is the one related to the PSII while less
information is available for the PSI structures.

Fig. 2.3 Absorption of chlorophyll-sensitized solar cells (CSSCs). Chlorophyll, main regions
where the chlorophyll absorbs the most and the less energy in. Data from Hassan et al. (2016)

Fig. 2.4 Excited states of chlorophyll, a qualitative representation


2.2 Light Harvesting 21

3 2
CH3
CH3
4 1 17
H 3C CH3
+ CH3 16
N N 5 6
2- H 3C 7
H Mg
+ 18
N N 8
H 3C 9
CH3 CH3
H 19
10
H
11
O O
12
O 13
CH3
20
14

15
CH3
15'

14'
13'
H 3C
20'
12'
CH3
11'

10'
9'
H 3C
19'
8'

7' CH3
H 3C CH3 18'

H 3C 6' 5'
16' 1' 4'
H 3C
17'
2' 3'

Fig. 2.5 Chlorophyll and beta-carotene molecules

The thylakoid membrane contains chlorophyll and the electron transport system
that carries out the initial light energy. Figure 2.5 describes the most common compo-
nents of the light-harvesting complex, basically the most common are chlorophylls
and carotenoids.
A simplification of the main reaction of photosynthesis is given as follows:

CO2 + H2 O → (CH2 O) + O2 (2.6)

In case we have wavelengths at 0.68 µ, the energy absorbed is 1760 kJ/mol (O2 )
and the energy necessary to carry on the reaction given previously is 467 kJ/mol.
22 2 Light Harvesting and Biomass Generation

This is simply evaluated as the heat of combustion of glucose (which is 2801 kJ/mol)
divided by six.
In the reaction center, the most important step of all the process takes place: the
conversion of solar energy to chemical and electrical energy. Within this step, the
solar energy (the one given by the photons) is converted in a stable charge flow of
electrons and this is done by ultrafast electron transfer. However, as also reported in
other parts of this manuscript, this “near unity” quantum efficiency of the process
is still unknown (Romero et al. 2014). The reaction center works also as a quantum
heat engine (QHE) since it converts solar energy into useful work, and this is done
by the charge separation and the formation of the necessary fuels (ATP molecules)
with the related chemical energy.
Nevertheless, it has been studied that quantum coherence could enhance the pho-
tosynthetic efficiency of the charge separation (the oxidation of water) in these QHE,
and it has also been showed that, even if quantum coherence could play a major role
in this process, photosynthesis does not require coherent light to work (Dorfman
et al. 2013). This enhancement suggests to reproduce natural solutions to improve
photovoltaic cells or related technical QHE. To be remembered here that with coher-
ent light, it is meant a beam of light where the photons have all the same frequency
and wavelength (a laser) and it is remarkable that plants have been adapted even to
use not coherent light, like the one coming from the sun. What a great engine is a
leaf of a plant!
On the other hand, it is left to us to define the quantum yield of photosynthesis,
QYF, which is (Sommer et al. 2015):
n mol
QYF = ; (2.7)
nq

where n mol is the number of molecules converted and n q is the number of quanta
absorbed. The amount of molecules converted is a function of the energy carried by
the electromagnetic waves (light). Figure 2.6 describes, qualitatively, the quantum
yield of photosynthesis as a function of the wavelength (energy) of the sun radiation.

Plants and microorganisms which utilize photosynthesis to reproduce or growth


their biomass can be considered as energy-converting machines. Since these
machines could be considered as black boxes, energy balances and thermodynamic
rules apply also here. The next paragraphs describe more in detail the so-called reac-
tion centers of photosynthesis; these are very complex molecular structures and as
it will be mentioned, many of the steps involved during the oxidation of water are
still not well understood. It will be shown in the following chapters that the creation
of transmembrane electrochemical potential is a key step for the generation of the
necessary work to function the ATP synthase enzyme. This functioning is equivalent
to the same principles applied in galvanic cells or batteries (van Rotterdam et al.
2002), and this is explained here since it has pedagogical value, at least according to
this author opinion.
2.2 Light Harvesting 23

Fig. 2.6 Quantum yield of photosynthesis, in the visible spectrum. Adapted from Zeiger and Taiz
(1991)

Attempts to simplify the photosynthetic process and to give a thermodynamic


treatment of the photosynthetic energy-converting devices (the plants) have been
carried on in the literature (Albarrán-Zavala and Angulo-Brown 2007; Brittin and
Gamow 1961). These studies often utilize data given in the literature for the standard
state Gibbs free energy of overall biochemical reactions (Madigan et al. 2000; Dean
1992; Conn and Stumpf 1972). Superior plants and cyanobacteria produce glucose
molecules from CO2 and water; however, alternative methods exist to form glucose
from different sources of hydrogen; purple and green bacteria are some examples.

2.3 Energy Transfer and Reaction Centers

As previously mentioned, the focus is given here to superior plants and their way to
generate the building blocks for the development of the biomass. This is also because,
it has been derived that they are among the most efficient energy-converting living
organisms. The chlorophyll and its reaction centers are responsible for the necessary
oxidation of water. All the chlorophyll is contained within the thylakoid membrane
in particular cell structures called chloroplast. The reactions of carbon reduction,
catalyzed by enzymes which are soluble in water, take place in the stroma. This is
the membrane sited in the chloroplast outside the thylakoids. The stroma lamellae
are membranes which extend then to the stacks of the thylakoids (Yahia et al. 2019).
The light-driven reactions are taking place mostly within the thylakoids; however,
also within the photosystem I there are light-driven reactions and here the energy
produced within the thylakoids in the form of ATP and NADPH molecules is utilized
to reduce and utilize the CO2 which is coming from the pores of the chloroplasts.
24 2 Light Harvesting and Biomass Generation

Fig. 2.7 Chloroplast structure, a qualitative representation

The exact structure of the light-harvesting complexes proteins has not been deter-
mined yet; however, this is not the only topic that still needs research; as it will be
shown within the manuscript, there are many principles yet to be discovered con-
cerning light harvesting. The purpose of this section is to introduce the reader to the
general principles of power (and biomass) generation in plants and how these pro-
cesses can have analogy with the working principles of galvanic cells and batteries.
As a matter of fact, there is an electron flow within the chloroplast structure and
the creation of the transmembrane electrochemical potential. The figures that follow
are made by the author of this book, and they might be not the most beautiful ones;
however, these figures were created with the purpose of being simple and to give a
conceptual description of the subject at hand.
Figure 2.7 describes the structure of the chloroplasts and the inner thylakoids,
grana lamellae and stroma lamellae.
The reaction centers, the antenna pigment–protein complexes, and most of the
electron transport enzymes are all integral membrane proteins (Kumari 2017). The
internal part of the thylakoid is called lumen. More in detail, the photosystem II, PSII,
with the related reaction center and chlorophylls, is situated in the grana lamellae,
formed by the thylakoids. The photosystem I (PSI) instead, with the related reaction
center and the electron transport system, is situated in the stroma lamellae. The
cytochrome b6f is situated in between the stroma and grana (Nicholls and Ferguson
2013). In PSII, the oxidation of two molecules of water produces four protons, four
electrons, and one O2 . Figure 2.8 gives a simple representation of how the energy
2.3 Energy Transfer and Reaction Centers 25

Fig. 2.8 Transfer of energy in light-harvesting antennas

is transferred among the antenna complexes and arrives to the reaction center of the
PSII.
The reaction centers of the two photosystems act as a catalyzer; the water is trans-
ported firstly by the tyrosine molecule to a manganese–oxygen-evolving complex,
and here the reactions of oxidation are taking place. The tyrosine is then coupled to
the Mn cluster (Noguchi et al. 1997).
The Step 3 illustrated in Fig. 2.9 releases oxygen and from this, two atoms of
hydrogen are available together with electrons to create the necessary transmem-
brane electrical potential. The exact working mechanism for the manganese–oxygen-
evolving complex and the release of the oxygen is not yet understood exactly. How-
ever, in this manuscript, the goal is to just give general information related to these
phenomena and for more detailed information, the reader could refer to more spe-
26 2 Light Harvesting and Biomass Generation

Fig. 2.9 Oxidation of water into the reaction center and transported by tyrosine molecule

cialized literature. The goal here is to understand that the water is oxidized (removal
of electrons) and this is because the plant needs a “current” (flow of electrons) and
a transmembrane electrochemical potential to function some of the very important
enzymes like for instance the ATP synthase.
In the Photosystem II, the oxidation of water and the release of electrons together
with H+ ions are taking place by means of a reaction center P680 which is naturally
a strong oxidant (we have to oxidize water and this is the most difficult task). How
photosynthetic reaction centers control the oxidation power in the canters P680, P700,
and P870 is explained more in detail in the literature (Ishikita et al. 2006, p. 870). After
this, the reaction center has some more electrons and therefore its electrochemical
potential is increased. By means of an electron transport chain (imagine this as a
wire), the electrons are transported to a second reaction center which is situated in
the PSI this time, the reaction center P700 where 680 and 700 refer to the wavelength
of the incident electromagnetic wave at which the reaction center absorbs the most of
the energy. At the reaction center P700, the electrochemical potential should increase
further to carry on subsequent actions and therefore the P700 reaction center should
be a strong reductant. This is illustrated generally in Fig. 2.10 where the so-called Z-
scheme of photosynthesis is demonstrated. The Z-scheme represents a fundamental
process if light harvesting is utilized for biomass production and generation of power
within the system, as a matter of fact today’s trend is to develop direct Z-scheme
photocatalysts (Xu et al. 2018). Artificial photosynthesis is naturally a trend as well
(El-Khouly et al. 2017).
References 27

Fig. 2.10 Z-scheme of photosynthesis

References

Albarrán-Zavala, E., & Angulo-Brown, F. (2007). A simple thermodynamic analysis of photosyn-


thesis. Entropy, 9(4), 152–168. https://doi.org/10.3390/e9040152.
Bass, M., DeCusatis, C., Enoch, J., Lakshminarayanan, V., Li, G., MacDonald, C., … Van Stryland,
E. (2009). Handbook of optics, volume IV: Optical properties of materials, nonlinear optics,
quantum optics (3rd ed.). McGraw-Hill Education.
Blankenship, R. E. (2013). Molecular mechanisms of photosynthesis. Wiley.
Blankenship, R. E., Sadekar, S., & Raymond, J. (2007). The evolutionary transition from anoxygenic
to oxygenic photosynthesis, Chap. 3. In P. G. Falkowski & A. H. Knoll (Eds.), Evolution of
primary producers in the sea (pp. 21–35). Burlington: Academic Press. https://doi.org/10.1016/
B978-012370518-1/50004-7.
Brittin, W., & Gamow, G. (1961). Negative entropy and photosynthesis. Proceedings of the National
Academy of Sciences of the United States of America, 47(5), 724–727.
Brown, T. E., LeMay, H. E., Bursten, B., Murphy, C., Woodward, P., & Stoltzfus, M. E. (2014).
Chemistry: The central science (13th ed.). Pearson.
Chapman, A. D. (2009). Numbers of living species in Australia and the world (2nd ed.). Canberra,
AU: Australian Government, Department of the Environment, Water, Heritage and the Arts.
Ciddor, P. E. (1996). Refractive index of air: New equations for the visible and near infrared. Applied
Optics, 35(9), 1566–1573. https://doi.org/10.1364/AO.35.001566.
Conn, E. E., & Stumpf, P. K. (1972). Outlines of biochemistry (3rd ed.). New York, USA: Wiley.
Daimon, M., & Masumura, A. (2007). Measurement of the refractive index of distilled water from
the near-infrared region to the ultraviolet region. Applied Optics, 46(18), 3811–3820. https://doi.
org/10.1364/AO.46.003811.
Dean, J. A. (1992). Lange’s handbook of chemistry. New York, USA: McGraw-Hill.
28 2 Light Harvesting and Biomass Generation

De Vrieze, J., Arends, J. B. A., Verbeeck, K., Gildemyn, S., & Rabaey, K. (2018). Interfacing anaero-
bic digestion with (bio)electrochemical systems: Potentials and challenges. Water Research, 146,
244–255. https://doi.org/10.1016/j.watres.2018.08.045.
Dorfman, K. E., Voronine, D. V., Mukamel, S., & Scully, M. O. (2013). Photosynthetic reaction
center as a quantum heat engine. Proceedings of the National Academy of Sciences of the United
States of America, 110(8), 2746–2751. https://doi.org/10.1073/pnas.1212666110.
El-Khouly, M. E., El-Mohsnawy, E., & Fukuzumi, S. (2017). Solar energy conversion: From natural
to artificial photosynthesis. Journal of Photochemistry and Photobiology C: Photochemistry
Reviews, 31(June), 36–83.
Furatian, L., & Mohseni, M. (2018). Temperature dependence of 185 nm photochemical water
treatment—The photolysis of water. Journal of Photochemistry and Photobiology A: Chemistry,
356, 364–369. https://doi.org/10.1016/j.jphotochem.2017.12.030.
Ghosh, G. (1999). Dispersion-equation coefficients for the refractive index and birefringence of
calcite and quartz crystals. Optics Communications, 163(1), 95–102. https://doi.org/10.1016/
S0030-4018(99)00091-7.
Gilroy, S., & Masson, P. H. (2008). Plant tropism. Blackwell Publishing.
Hassan, H. C., Abidin, Z. H. Z., Chowdhury, F. I., & Arof, A. K. (2016). A high efficiency chlorophyll
sensitized solar cell with quasi solid PVA based electrolyte. International Journal of Photoenergy,
2, 1–9. https://doi.org/10.1155/2016/3685210.
IEA. (2017). Emissions from fuel combustion highlights. International Energy Agency
IEA. Retrieved from https://www.iea.org/publications/freepublications/publication/
CO2EmissionsfromFuelCombustionHighlights2017.pdf.
Ishikita, H., Saenger, W., Biesiadka, J., Loll, B., & Knapp, E.-W. (2006). How photosynthetic reac-
tion centers control oxidation power in chlorophyll pairs P680, P700, and P870. Proceedings of the
National Academy of Sciences, 103(26), 9855–9860. https://doi.org/10.1073/pnas.0601446103.
Jones, M., & Fleming, S. A. (2014). Organic chemistry (5th ed.). W. W. Norton & Company.
Kasarova, S. N., Sultanova, N. G., Ivanov, C. D., & Nikolov, I. D. (2007). Analysis of the dispersion
of optical plastic materials. Optical Materials, 29, 1481–1490. https://doi.org/10.1016/j.optmat.
2006.07.010.
Kell, D. B. (2012). Large-scale sequestration of atmospheric carbon via plant roots in natural
and agricultural ecosystems: Why and how. Philosophical Transactions of the Royal Society B:
Biological Sciences, 367(1595), 1589–1597. https://doi.org/10.1098/rstb.2011.0244.
Kobayashi, T., Nozoye, T., & Nishizawa, N. K. (2019). Iron transport and its regulation in plants.
Free Radical Biology and Medicine, 133, 11–20. https://doi.org/10.1016/j.freeradbiomed.2018.
10.439.
Ksenzhek, O. S., & Volkov, A. G. (1998). Plant energetics. San Diego, California: Academic Press.
Kumari, A. (2017). Electron transport chain, Chap. 3. In Sweet biochemistry remembering structures,
cycles, and pathways by mnemonics (1st ed.). Academic Press.
Li, G., & Coleman, G. D. (2018). Nitrogen storage and cycling in trees. In Advances in botanical
research. Academic Press. https://doi.org/10.1016/bs.abr.2018.11.004.
Liu, S.-Q., Zhou, S.-S., Chen, Z.-G., Liu, C.-B., Chen, F., & Wu, Z.-Y. (2016). An artificial pho-
tosynthesis system based on CeO2 as light harvester and N-doped graphene Cu(II) complex as
artificial metalloenzyme for CO2 reduction to methanol fuel. Catalysis Communications, 73,
7–11. https://doi.org/10.1016/j.catcom.2015.10.004.
Madigan, M. T., Martinko, J. M., & Brock, P. J. (2000). Biology of microorganisms (8th ed.). Prentice
Hall.
Malitson, I. H. (1965). Interspecimen comparison of the refractive index of fused silica*,†. JOSA,
55(10), 1205–1209. https://doi.org/10.1364/JOSA.55.001205.
McCree, K. J. (1981). Photosynthetically active radiation. In O. L. Lange, P. S. Nobel, C. B.
Osmond, & H. Ziegler (Eds.), Physiological plant ecology I: Responses to the physical environ-
ment (pp. 41–55). Berlin, Heidelberg: Springer Berlin Heidelberg. https://doi.org/10.1007/978-
3-642-68090-8_3.
References 29

Miranzadeh, H., Emam, Y., Sayyed, H., & Zare, S. (2011). Productivity and radiation use efficiency
of four dryland wheat cultivars under different levels of nitrogen and chlormequat chloride.
Journal of Agricultural Science and Technology, 13, 339–351.
Mirkovic, T., Ostroumov, E. E., Anna, J. M., van Grondelle, R., Govindjee, & Scholes, G. D. (2017).
Light absorption and energy transfer in the antenna complexes of photosynthetic organisms.
Chemical Reviews, 117(2), 249–293. https://doi.org/10.1021/acs.chemrev.6b00002.
Nagai, R. (1993). Regulation of intracellular movements in plant cells by environmental stimuli. In
K. W. Jeon & J. Jarvik (Eds.), International review of cytology (Vol. 145, pp. 251–310). Academic
Press. https://doi.org/10.1016/S0074-7696(08)60429-5.
Nelson, N. (2011). Photosystems and global effects of oxygenic photosynthesis. Biochimica et Bio-
physica Acta (BBA)—Bioenergetics, 1807(8), 856–863. https://doi.org/10.1016/j.bbabio.2010.
10.011.
Nicholls, D. G., & Ferguson, S. J. (2013). Photosynthetic generators of proton motive force, Chap. 6.
In Bioenergetics (4th ed., pp. 159–196). Academic Press.
Noguchi, T., Inoue, Y., & Tang, X.-S. (1997). Structural coupling between the oxygen-evolving
Mn cluster and a tyrosine residue in photosystem II as revealed by fourier transform infrared
spectroscopy. Biochemistry, 36(48), 14705–14711. https://doi.org/10.1021/bi971760y.
Owen, A. (2017). Plants release up to 30 per cent more CO2 than previously thought, study
says. ABC News. Retrieved from http://www.abc.net.au/news/2017-11-18/plant-respiration-co2-
findings-anu-canberra/9163858.
Palik, E. D. (1985). Handbook of optical constants of solids. Boston: Academic Press.
Pisciotta, J. M., Zou, Y., & Baskakov, I. V. (2010). Light-dependent electrogenic activity of
cyanobacteria. PLoS ONE, 5(5), e10821. https://doi.org/10.1371/journal.pone.0010821.
Qi, X., Ren, Y., Liang, P., & Wang, X. (2018). New insights in photosynthetic microbial fuel cell
using anoxygenic phototrophic bacteria. Bioresource Technology, 258, 310–317. https://doi.org/
10.1016/j.biortech.2018.03.058.
Resnick, R., Halliday, D., & Krane, K. S. (2001). Physics (5th ed.). Wiley.
Rheims, J., Köser, J., & Wriedt, T. (1997). Refractive-index measurements in the near-IR using
an Abbe refractometer. Measurement Science & Technology, 8(6), 601. https://doi.org/10.1088/
0957-0233/8/6/003.
Romero, E., Augulis, R., Novoderezhkin, V. I., Ferretti, M., Thieme, J., Zigmantas, D., et al. (2014).
Quantum coherence in photosynthesis for efficient solar-energy conversion. Nature Physics,
10(9), 676–682. https://doi.org/10.1038/nphys3017.
Scholes, G. D., & Fleming, G. R. (2005). Energy transfer and photosynthetic light harvesting. In
Adventures in chemical physics (pp. 57–129). Wiley Ltd. https://doi.org/10.1002/0471759309.
ch2.
Sommer, M. E., Elgeti, M., Hildebrand, P. W., Szczepek, M., Hofmann, K. P., & Scheerer, P. (2015).
Structure-based biophysical analysis of the interaction of rhodopsin with G protein and arrestin,
Chap. 26. In A. K. Shukla (Ed.), Methods in enzymology (Vol. 556, pp. 563–608). Academic
Press. https://doi.org/10.1016/bs.mie.2014.12.014.
Sun, Z., Liang, H., Liu, J., & Shi, G. (2017). Estimation of photosynthetically active radiation using
solar radiation in the UV–visible spectral band. Solar Energy, 153, 611–622. https://doi.org/10.
1016/j.solener.2017.06.007.
Taiz, L., Zeiger, E., Møller, I. M., & Murphy, A. (2014). Plant physiology and development. Sinauer
Associates, Oxford University Press.
van Rotterdam, B. J., Crielaard, W., van Stokkum, I. H. M., Hellingwerf, K. J., & Westerhoff, H. V.
(2002). Simplicity in complexity: The photosynthetic reaction center performs as a simple 0.2 V
battery. FEBS Letters, 510(1–2), 105–107. https://doi.org/10.1016/S0014-5793(01)03210-0.
Xu, P., Roy, L. M., & Croce, R. (1858). Functional organization of photosystem II antenna com-
plexes: CP29 under the spotlight. Biochimica et Biophysica Acta (BBA)—Bioenergetics, 10,
815–822. https://doi.org/10.1016/j.bbabio.2017.07.003.
30 2 Light Harvesting and Biomass Generation

Xu, Q., Zhang, L., Yu, J., Wageh, S., Al-Ghamdi, A. A., & Jaroniec, M. (2018). Direct Z-scheme
photocatalysts: Principles, synthesis, and applications. Materials Today. https://doi.org/10.1016/
j.mattod.2018.04.008.
Yahia, E. M., Carrillo-López, A., Barrera, G. M., Suzán-Azpiri, H., & Bolaños, M. Q. (2019).
Photosynthesis, Chap. 3. In E. M. Yahia (Ed.), Postharvest physiology and biochemistry of fruits
and vegetables (pp. 47–72). Woodhead Publishing. https://doi.org/10.1016/B978-0-12-813278-
4.00003-8.
Zeiger, E., & Taiz, L. (1991). Plant physiology. Redwood City, CA: Benjamin-Cummings.
Chapter 3
Redox Potential and Galvanic Cells

You know…, they were all things which were born


spontaneously. One did what he could do and according to his
personal attitudes and the possessed abilities. One was doing
one task and another one was doing some other tasks and there
was very much friendship among us. And then, there was
enormous passion in what we were doing.
A suitable description of a successful research group. Freely
translated from the words of Emilio Segrè, Nobel Prize in
Physics, 1959.

3.1 Functioning Principles

The schematics given in Fig. 2.10 can be also described as a galvanic cell
(van Rotterdam et al. 2002). Since, and as we will see later within this text, many
biochemical reactions follow the purpose of oxidizing a substrate and producing a
quantity of electrons which are acting as a current. This will give the power to carry
on several biochemical functions. Later on, tables will be given including biochem-
ical half-cell reactions. To better understand this, in this section, an explanation of
the principles for the electrochemical power generation and the Nernst law is given.
This will provide the reader with the basics and the understanding of how the electro-
chemical potential can be actually estimated for biological redox couples. This is also
a topic that mostly relates to engineers, since they are focused on the quantification
of processes and phenomena.
In this kind of electronic generators of power (the galvanic cell), a reaction is
normally taking place on the anode of the cell. Here, electrons are produced and sent
to flow in a wire. The electrons flow as a function of a difference in electrochemical
potential between the anode and the cathode give the power to do work according to
the relation:

1W = 1V · 1A (3.1)

© Springer Nature Switzerland AG 2019 31


C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_3
32 3 Redox Potential and Galvanic Cells

Fig. 3.1 Functioning of a galvanic cell where a salt bridge is employed

To give one example, in Fig. 3.1, the flow of electrons is given from the right to
the left. At the right side and in the figure presented, a CuSO4 solution is positioned
at the anode where a Cu metallic plate is situated. Here, from the CuSO4 solution,
the SO4 2− ions transfer electrons to the anode side. As a consequence, electrons start
to flow in the direction of the cathode. A salt bridge is positioned between the anode
and the cathode. In Fig. 3.1, a salt bridge composed by KCl solution is given; the
CL− ions produced will compensate for the ones lost in the anode by the CuSO4
solution. In a similar way, at the cathode, there will be an accumulation of silver
anions and therefore this is compensated by the K+ ions from the salt bridge. The
galvanic cell is not working anymore when equilibrium is reached.
To be noticed that it is important to understand the flow of ions and electrons
within the solutions and the salt bridge. The flows follow the driving force given by
the difference in concentrations: This means that anions are going from higher con-
centrations to lower concentrations, and the same is with cations and even electrons
are no exception in this sense since they move from a region at higher electrochemical
potential to a lower one. For a more accurate treatise of the galvanic cells thermo-
dynamics, I invite the reader to consider books of physical chemistry like (Levine
2009).
Nevertheless, it is pointed out here that, when a thermodynamic analysis should
be carried out on these particular systems, a standard should be defined. This will
include parameters like pressure and temperature, and this also concerns the kind of
3.1 Functioning Principles 33

Fig. 3.2 Galvanic cell with no liquid junction. A representation of the hydrogen anode; hydrogen
is pumped on to the electrode by using a particular casing

cell itself. In case we consider the standard electrochemical potential, this is taken
at 25 °C, 1 Atm and when the effective concentration for the ions involved is 1 M.
A standard galvanic cell is drafted in Fig. 3.2 where hydrogen is pumped onto the
electrode by using a particular casing. From the reactions:

H2 (g) ↔ H2 (aq) (3.2)

H2 (aq) ↔ 2H+ + 2e− (3.3)

electrons are released from the hydrogen molecule, collected by a wire, and trans-
ferred to a silver electrode where the following reaction is taking place:

AgCl(s) → Ag+ + Cl−



(3.4)

In this case, the silver cations react with the electrons and precipitate as Ag(s)
according to the equation:

Ag+ + e− ↔ Ag(s) (3.5)

The concept of redox potential needs to be explained further at this point. The redox
potential consists of an electrical measurement. This measurement shows the ten-
34 3 Redox Potential and Galvanic Cells

dency of a object to transfer electrons to a reference electrode or from a reference


electrode.
When Ag is giving electrons, the reaction taking place at the standard electrode
is:

Ag  Ag+ + e− (3.6)

Even in this case the previous equation is written as a reversible reaction. When the
reaction is more directed on the right hand side, the electronegativity of the reference
electrode becomes higher than the point of measure; for this reason, the voltmeter
will measure less electronegativity for the sample measured.
Let us consider another example here; the following is the Zn–Cu cell.

Zn(s) + Cu2+ (aq) → Zn2+ (aq) + Cu(s), (3.7)

where at the cathode the electrons are accepted as

Cu++ + 2e− → Cu; (3.8)

and the anode is given by:

Zn → Zn++ + 2e− . (3.9)

The standard overall electrochemical potential E 0 is calculated by the formula:

E 0 = E 0 (cathode) + E 0 (anode) (3.10)

Cu and Zn are very much used in galvanic cells, and even it is proposed to boost
them by inserting anode–cathode couples within the same electrolyte (Lacina et al.
2018).
Since the standard state is defined as the state at which the system is at 25 °C,
the liquid reactant is at 1 M concentration, and the gaseous reactant is at 1 Atm, the
standard reaction of H+ reduction is taken at that condition.

2H+ (aq) + 2e− → H2 (g) E 0 = 0 (3.11)

For this reason, we chose to evaluate the standard potential for the anode and the
cathode with respect to the hydrogen electrode. The following is the representation
of the overall cell reaction as a sum of the two half-cell reactions taking place,
respectively, at the anode, Eq. (3.12), and cathode, Eq. (3.13).
   
Zn(s)Zn2+ (aq)H+ (aq)H2 (g)Pt E 0 = 0.76 V (3.12)
   
PtH2 (g)H+ (aq)Cu2+ (aq)Cu(s) E 0 = 0.34 V (3.13)
3.1 Functioning Principles 35

Table 3.1 Evaluation of the Electrode Reaction


standard cell potential
Anode Zn(s) → Zn2+ (aq) + 2e− E 0 = +0.76 V
Cathode Cu2+ (aq) + 2e− → Cu(s) E 0 = +0.34 V
Overall Cu2+ (aq) + Zn(s) → Cu(s) + Zn2+ (aq) E 0 =
1.10 V

Equations (3.12) and (3.13) are a typical way of writing the electrodes, in this
case as a reference to the standard electrode of hydrogen. The left side represents the
anode side (where electrons are released) while on the right side there is the cathode
which is receiving electrons. The standard Gibbs free energy is given by (Levine
2009):
 
G 0 = −n(F) E 0 (3.14)

where F is the Faraday constant which is equal to 96,485.33 C mol−1 (NIST Physical
Measurement Laboratory 2018), and n is the number of electron moles moving.
Equation (3.14) will give dimensions of Coulomb times Volt which will be equivalent
to joules, and this is the correct dimension for the energy. The author of this book
suggests the readers to do often this kind of reasoning on the dimensions to check
for the consistency of the formulas at hand.
The overall cell reaction for the Zn(s)–Cu(s) system is given by:
  
Zn(s)Zn2+ (aq)Cu2+ (aq)Cu(s) (3.15)

In Table 3.1, a clear example of how to calculate the overall standard electrochemical
potential of a cell is given.
To understand better the calculations, it is convenient to remember that, when a
half-cell reaction is encountered, the sign of the standard electrochemical potential
determines if the reaction is spontaneous or not. In fact, when the given standard
potential is positive, then the reaction is spontaneous. It is not the same case when
the reaction gives negative values. However, there are some conventions to take into
account when data are retrieved.

3.2 Conventions on the Electrodes

With respect to Table 3.1, it is important to notice that when a reaction is in reality
taking place in the way it is written, then the potential that is read on the reading
instrument is always positive. In the case presented, the Zn(s) is indeed an anode
and the Cu(s) is indeed receiving electrons. However, by convention the electrode
potentials are a measure of the driving force for the two half reactions when the half
36 3 Redox Potential and Galvanic Cells

Fig. 3.3 Plus–Right Rule

reactions are written as reductions; this would mean that the electrons are always
written on the left side of the equation:

2AgCl(s) + 2e− ↔ 2Ag(s) + 2Cl− (3.16)

2H+ + 2e− ↔ H2 (g) (3.17)

This means also that the anode side is written on the left and the cathode side is
written on the right. Figure 3.3 gives a simple description of the convention sides for
the galvanic cell.
Considering the reference electrode as in Fig. 3.3, when the voltage is positive, the
electrons are flowing on the right side, so it means that the right side is the cathode and
the left side is the anode. If written according to the convention (with the electrons
on the left side), the anode will have a negative electrochemical potential and this
would also be in accordance with the fact that (always by convention) electrons are
indicated as having a negative charge. The more electronegative is a site, the more
will be the electrons at that site and therefore the more electrons will move on the
basis of the difference in concentration.

3.3 Calculation of the Redox Potential at Non-standard


Conditions, the Nernst Law

If the reaction is written as:

pP + qQ + · · · + ne− ↔ r R + sS + · · · (3.18)

where the capital letters refer to general components and the non-capital letters
refer to the stoichiometric coefficients. The overall electrochemical potential will be
expressed by the Nernst Law as follows (Levine 2009):
3.3 Calculation of the Redox Potential at Non-standard … 37
 
RT (aR )r · (aS )s · · ·
E=E − 0
· ln   q  (3.19)
nF (aP ) p · aQ · · ·

where E is the overall electrochemical potential, E 0 is the standard electrochemical


y
potential evaluated as previously. The terms ax refer to the activities of the reactants
and products. The standard potential is defined as the potential of a half-cell reaction
when all the reactant and products have activity = 1. In the standard galvanic cell,
the reference hydrogen electrode is on the left side (so if the electrode to be measured
has positive potential, then it will be a cathode accordingly to what was mentioned
previously).
We remember that the total Gibbs free energy will be evaluated as:

G = −n F E (3.20)

where this time E is the overall electrochemical potential or also mentioned as elec-
tromotive force (EMF) under non-standard conditions. As an example, consider the
following reaction, related to the Ag (silver) electrode (Holze 2009):

Ag+ + e− ↔ Ag(s) with E 0 = +0.799 V; (3.21)

this would be a cathode and we want to evaluate the Nernst equation for one molar
concentration of Ag(s). Remember that:

J
R = 8.316 ; (3.22)
mol K
and

C
F = 96,487 ; (3.23)
mol

at 25 °C the quantity RT
F
= 0.0592 for this reason:

0.0592 1
E = 0.799 − log ; (3.24)
1 aAg+

and this is because n = 1 (we have one electron mole transferred). Remember also
that activities are taken as 1 for solids, for dilute solutions the activity coefficient is
equal to 1 (do not confuse the activity coefficient with the activity values themselves),
so the activity can be approximated with the molar concentration. The activity of
saturated solutions is also considered to be 1 (Levine 2009). The equation above is
obtained by changing the base of the logarithm from natural to decimal.
For the sake of clarity, a simple derivation of the change of logarithm base is
reported step by step; we start from the definition of the general logarithm on base
a:
38 3 Redox Potential and Galvanic Cells

x = loga C; (3.25)

which implies:

a x = C; (3.26)

for this reason:

logb a x = logb C; (3.27)

and from one of the properties of the logarithms:

xlogb a = logb C; (3.28)

therefore we have:

logb C
loga C = . (3.29)
logb a

Consider the following cell reaction:

AgCl(s) + e− ↔ Ag(s) + Cl− ; (3.30)

and suppose that the concentration of Cl− ions is known. We want to calculate the
potential of the cell by the Nernst equation rewritten here:
 
RT (aR )r · (aS )s · · ·
E=E − 0
· ln   q  (3.31)
nF (aP ) p · aQ · · ·

The activity of a specie x is: ax = γx cx , and the activity coefficient γx is calculated


from the Debye–Hückel equation (Hückel and Debye 1923) which is fully derived
by (Bockris and Reddy 1998) and a simplified version is given in the literature as
follows (Levine 2009):

Z 2A A μ
−ln γ A = √ (3.32)
1 + Bα A μ
 
−9
where A = 1.1744 kg
mol
and B = 3.285 × 10 kg
mol
· m−1 . Substituting these
values in the previous equation and changing the base of the logarithm to the decimal
one, this will be written as:

0.510Z 2A μ/m 0
−logγ A =   (3.33)
1 + 0.328 α A /Å μ/m 0
3.3 Calculation of the Redox Potential at Non-standard … 39

and it could be further simplified as (not specifying for the dimensions involved):

0.510Z 2A μ
−logγ A = √ (3.34)
1 + 3.28α A μ

where Z A is the charge of the A specie, α A is the effective diameter of the hydrated
specie (this is given in nanometers in Eq. (3.36), μ is the ionic strength of the solution.
The ionic strength is calculated from:

1 
μ= c1 Z 12 + · · · + cn Z n2 (3.35)
2
where the subscript numbers (1,… n) are referring to the species involved. For exam-
ple, if we have an HCl solution:

HCl ↔ H+ + Cl− (3.36)


 
Z would be 1 for both H+ and Cl− and the ionic strength will be 2 × 21 c · Z 2 ,
this is because c (the molar concentration) is equal for both ions and Z is also 1 for
both.
The effective diameter of the hydrated specie in nanometers can be found in the
literature, for instance, the data given in Table 3.2 are taken from Kielland (1937).
Table 3.3 gives standard electrochemical potential for some half-cell reactions; at
this point, the reader is invited to notice that these values are related to the reactants
which are present in biological systems. In fact, we should remember here that the
purpose of this treatise is to give the fundamental notions of how the biomass is
generated.
To give some additional description of how the redox potential relates with the
Z-scheme of photosynthesis given previously, Fig. 3.4 represents the Z-scheme in
relation to the redox potential.
Notice that in Fig. 3.4 the higher is the potential the more negative is its value.
This is because the system will be at a more “electronegative” state. It is good to
remember that it is commonly referred to electrons to have a negative charge and
therefore, with this in mind it will be easier to understand the Z-scheme. The redox
potential is a measure of the affinity of a substance to attract electrons.
When an electron proceeds “downhill”, there is the formation of free energy, G,
which is calculated according to Eq. (3.20) (Acrivos 1988), which is repeated here:

G = −n · F · (E) (3.37)

We remember that n is the moles of electrons moving; F, is the Faraday constant,


23.062 kcal/V per mole, and it is the energy released when a mole of electron is
moved by a potential of 1 V.
40

Table 3.2 Diameters of non-hydrated, value on top, and hydrated, second value, ions
H+ He

900
Li+ Be+2 B C N O F− Ne
80 60 150
600 80 350
Na+ Mg2+ Al3+ Si P Cl− Ar
100 90 80 80 190
450 800 900 900 300
K+ Ca2+ Sc3+ Ti V Cr3+ Mn2+ Fe2+ Co2+ Ni2+ Cu2+ Zn2+ Ga3+ Ge4+ As5+ Se Br− Kr
160 140 140
300 600 900
Rb+ Sr2+ Y3+ Zr4+ Nb Mo Tc Ru Rh Pd Ag+ Cd2+ In3+ Sn2+ Sb3+ Te I− Xe
180 170 160 160 150 140 140 – 220
250 500 900 1100 250 500 900 600 300
Cs+ Ba2+ Lu Hf Ta W Re Os Ir Pt Au Hg2+ TI+1 Pb2+ Bi3+ Po At− Rn
210 210 150 140 –
250 500 500 250 590
NH4 + La3+ Ce3+ Pr3+ Nd3+ – Sm3+ Sn4+ NO3- OH−
– 200 – – – – – – –
250 900 900 900 900 900 1100 300 350
Ce4+ Fe3+ CO3 2− NO2 − SO4 2−
– – – – –
1100 900 450 300 400
Adapted from Kielland (1937). Values are expressed in picometers
3 Redox Potential and Galvanic Cells
3.3 Calculation of the Redox Potential at Non-standard … 41

Table 3.3 Standard Half reaction E 0 (V)


potentials of selected half
reactions of biological 1
2 O2 + 2H+ + 2e− → H2 O 0.816
importance, 25 °C and pH 7    
(Ksenzhek and Volkov 1998; Ferredoxin Fe3+ + e− → ferredoxin Fe2+ −0.432
Ross 1991)
2H+ + 2e− → H2 −0.414
NO−
3 + 2H+ + 2e− → NO−
2 + H2 O 0.421

α − ketoglutarate + CO2 + 2H+ + 2e−


→ isocitrate −0.38
 3+  −   0.365
Cytochrome f Fe +e → cytochrome f Fe2+
Acetoacetate + 2H+ + 2e− → β − hydroxybutyrate −0.38
Fe(CN)3− − 4− 0.36
6 + e → Fe(CN)6
NAD+ + H+ + 2e− → NADH −0.32
NADP+ + H+ + 2e− → NADPH −0.324
2H+ + 2e− → H2 at pH = 0 0.00000
   
cytochrome a Fe3+ + e− → cytochrome a Fe2+ 0.29
   
cytochrome c Fe3+ + e− → cytochrome c Fe2+ 0.254
   
cytochrome b Fe3+ +e− → cytochrome b Fe2+ 0.077

Ubiquinone + 2H+ + 2e− → Ubiquinol + H2 0.045


Fumarate2− + 2H+ + 2e− → succinate2− 0.031
O2 + 2H+ + 2e− → H2 O2 0.295
Crotonyl − CoA + 2H+ + 2e− → butyryl − CoA −0.015
Oxaloacetate2− + 2H+ + 2e− → malate2− −0.166
Pyruvate− + 2H+ + 2e− → lactate− −0.185
Acetaldehyde + 2H+ + 2e− → ethanol −0.197
FAD + 2H+ + 2e− → FADH2 −0.219
Glutathione + 2H+ + 2e− → 2 reduced glutathione −0.23
S + 2H+ + 2e− → H2 S −0.243
Lipoic acid + 2H+ + 2e− → dihydrolipoic acid −0.29

To give one example here, if for one molecule of glucose we need 24 electrons
from oxygen in water and the associated electrochemical potential at non-standard
conditions will be −1.24 V, this will give:

23.062 kcal
G = −24 (−1.24)V (3.38)
V

To be noticed that the value of the Gibbs free energy is positive here because
we have calculated it for the process of gathering the necessary electrons from O2 ;
42 3 Redox Potential and Galvanic Cells

Fig. 3.4 Z-scheme in terms of the redox potential

therefore, we need energy to do that. In the case presented here and referring to
literature data (Sinha 2013):

E oxygen = +0.82 V,

and

E carbon in carbohydrates = −0.42 V

Therefore:

E = −0.42 − 0.82 = −1.24.

Some examples of redox potentials in biological systems are shown in the following
equations which are reported to demonstrate different kinds of electron acceptors in
biological systems:

CO2 + 4H+ + 4e− → CH2 O + H2 O (3.39)

1
NO− + −
3 + 6H + 5e → N2 + 3H2 O (3.40)
2
O2 + 4H+ + 4e− → 2H2 O (3.41)
3.3 Calculation of the Redox Potential at Non-standard … 43

Equation (3.39) shows how the CO2 is actually reduced within the process of
photosynthesis; Eq. (3.40) shows how nitrogen oxides (nitrate) also accept electrons
and the last equation refers to the respiration of living organisms including humans.
This is done also to point out that when an electron flow is generated, there has
to be also a ultimate electron acceptor. In anaerobic respiration, for instance, the
ultimate electron acceptor can be organic3+and also inorganic and more in detail we0
can have nitrate, sulfate SO2− 4 and Fe as ultimate acceptors. Because the E
values are calculated for n = 1 and given in the literature in this way, the reactions’
Eqs. (3.39)–(3.41) should be divided by 4, 5, 4, respectively.

3.4 A Remark on the Artificial Splitting of Water


to Produce Hydrogen, Electrolysis

From the previous sections, it is possible to get an understanding of why biological


systems have many things in common with galvanic cells and the principles of elec-
trochemical potential, electron flow, and the related power obtained. Here, we take
into consideration very briefly the process of water reduction and as a consequence
the production of hydrogen by artificial electrolysis. Let us consider the following
figure.
In the figure, there are two electrodes, one (Electrode 1) is connected to the cathode
of a battery and the second (Electrode 2) is connected to the anode of a battery. When
the battery will be on, there will be the generation of charges on the electrodes and
for this reason, the electrode on the right of Fig. 3.5 will have positive charges on it.
Within the cell, there is a solution of NaOH which will give OH− ions; now, because
of the generated electric field, the OH− ions near the positively charged electrode
will be attracted by the electrode and releasing electrons to it. Because of this, the
section 1 will be the anode (where electrons are taken). The electrons will flow from
the anode to the cathode (section 1) and the section 2 in Fig. 3.5 will be charged
negatively. Water is a polar compound, which means that the molecule has a partial
positive charge on the hydrogen atoms and a partial negative charge on the oxygen
atoms. Because of the electric field generated by the battery, the positive side of the
water molecules will be attracted by the negative electrode when they are near the
electrode itself. At the negative electrode, the water molecule is reduced to give the
formation of hydrogen according to the reaction:

4H2 O + 4e− → 2H2 + 4OH− , E 0 = −0.83 V (3.42)

where this would represent an half-cell reaction with a standard electrochemical


potential equivalent to −0.83 V. This means that this reaction (written from the left
to the right as above) is not spontaneous, but in this case the power required to have
the reaction written in this way is given by the battery.
44 3 Redox Potential and Galvanic Cells

Fig. 3.5 Schematic representation of an electrolytic cell

On the other hand, at the anode we have:

4OH− → O2 + 2H2 O + 4e− , E 0 = −0.4 V (3.43)

To be noticed that a reasonable question would be if the sodium ions would


also react with the negative electrode and give the formation of molecular sodium.
However, as we can notice, the following reaction is difficult to achieve since it is
highly not spontaneous

Na+ + e− → Na, E 0 = −2.7 V (3.44)

Therefore, the reaction (3.42) is the one mostly taking place and which gives
the formation of OH− contributing to the equilibrium between the Na+ and OH− .
Equations (3.42) and (3.43) can be summed to give:

2H2 O → O2 + 2H2 , E 0 = −1.23 V (3.45)

which is the minimum required voltage necessary to carry on the steps above by the
battery.
References 45

References

Acrivos, J. (1988). Physical chemistry, third edition (Levine, Ira N.). Journal of Chemical Education,
65(12), A335. https://doi.org/10.1021/ed065pA335.3.
Bockris, J. O. M., & Reddy, A. K. (1998). Modern electrochemistry (2nd ed.). Berlin: Springer.
Holze, R. (2009). Experimental electrochemistry: A laboratory textbook (1st ed.). Wiley.
Hückel, E., & Debye, P. (1923). The theory of electrolytes: I. lowering of freezing point and related
phenomena. Physikalische Zeitschrift, 24, 185–206.
Kielland, J. (1937). Individual activity coefficients of ions in aqueous solutions. Journal of the
American Chemical Society, 59(9), 1675–1678. https://doi.org/10.1021/ja01288a032.
Ksenzhek, O. S., & Volkov, A. G. (1998). Plant energetics. San Diego, California: Academic Press.
Lacina, K., Sopoušek, J., Skládal, P., & Vanýsek, P. (2018). Boosting of the output voltage of
a galvanic cell. Electrochimica Acta, 282, 331–335. https://doi.org/10.1016/j.electacta.2018.06.
080.
Levine, I. N. (2009). Physical chemistry (6th ed.). McGraw-Hill Education.
National Institute of Standards and Technology. (2018). NIST chemistry webbook. Retrieved October
3, 2018, from https://webbook.nist.gov/chemistry/.
Ross, J. R. (1991). Practical handbook of biochemistry and molecular biology. Biochemical Edu-
cation, 19(2), 95–96. In G. D Fasman (Ed.) (pp. 601). Boca Raton, Florida, USA: CRC Press,
1989. $00 ISBN 0-8493-3705-4. https://doi.org/10.1016/0307-4412(91)90020-9.
Sinha, R. K. (2013). Modern plant physiology (2nd ed.). Alpha Science International Ltd.
van Rotterdam, B. J., Crielaard, W., van Stokkum, I. H. M., Hellingwerf, K. J., & Westerhoff, H. V.
(2002). Simplicity in complexity: the photosynthetic reaction center performs as a simple 0.2 V
battery. FEBS Letters, 510(1–2), 105–107. https://doi.org/10.1016/S0014-5793(01)03210-0.
Chapter 4
Overview of the Main Mechanisms
of Photosynthesis

Lastly, I believe that a man who sees the light could admire the
sun itself and not only its image reflected in the water or on any
other surface.
Freely translated from Plato, The Republic Book VII, 360 b.c.

4.1 Transport Phenomena in Reaction Centers

After a little parenthesis on how the biological mechanisms described previously can
be compared to the functioning of Galvanic cells, an overview of the most important
systems in photosynthesis (aerobic) is given here. More information can be found
also in literature (Nelson and Cox 2012; Hall and Rao 1999; Hoptkins 1999).
We recall that the principal parts needed to carry on the process of photosynthesis
are called photosynthetic systems. The main photosynthetic systems are:
• The photosystem II, PSII, which oxidizes water to O2 in structures called thylakoids
(that is why we need here a strong oxidant) and in this process releases protons
into a region called lumen.
• The cytochrome b6 f receives electrons from the previously named PSII and delivers
them to the photosystem I, PSI. It also transports additional protons into the lumen
from the stroma.
• The photosystem I, PSI, reduces NADP+ to NADPH in the stroma (that is why we
need here a strong reductant) by the action of ferredoxin (Fd) and the flavoprotein
ferredoxin–NADP reductase (FNR) (Fultz and Durst 1982).
• The ATP synthase is responsible for producing ATP which is a highly energetic
molecule. The ATP synthase is driven by the diffusion of protons which diffuse
back through it from the lumen (where they had higher concentration) into the
stroma. To be noticed here that, a lot of work is done to obtain a difference in
concentration concerning the H+ ions and this is with reason.
The cytochrome pumping of H+ ions is shown in Fig. 4.1.
From the figure above, it can be observed that the reduced quinone, Q complex
takes 2H+ from the stroma and completely reduces the quinone itself.
© Springer Nature Switzerland AG 2019 47
C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_4
48 4 Overview of the Main Mechanisms of Photosynthesis

Fig. 4.1 Cytochrome b6 f complex pumping H+ ions. Adapted from Taiz et al. (2014)

From the photosystem II, it can be stated that:


– the pheophytin takes the electrons;
– gives them to the quinones, and completely reduce them (QH2 );
– the quinones are oxidized and two electrons are available;
– one goes to photosystem I;
– one goes to a cycle which takes 2H+ from the stroma to the lumen.
Figure 4.2 shows the overall process of photosynthesis with the main systems
involved. To be noticed here that, in order for the ATP synthase to work, a substantial
amount of hydrogen ions is necessary on the lumen side. The difference in elec-
trochemical potential and the chemical diffusion of hydrogen ions through the ATP
synthase channel drives the ATP synthase enzyme to produce ATP molecules.
In the first stage, light-dependent reactions, or also named as light reactions,
capture the energy of light and use it to make the energy storage molecules (ATP).
It will be repeated here that the ATP molecules represent a font of energy storage
in the sense that when these molecules react to form ADP, an important quantity of
energy is released.
During the second stage of photosynthesis, the process accounts for light-
independent reactions. These reactions use these products to capture and reduce
carbon dioxide by means of NADPH.
As a summary, the main reaction centers (RC) are:
• photosystem II;
• cytochrome b6 f;
4.1 Transport Phenomena in Reaction Centers 49

Fig. 4.2 Summary of the main mechanisms of photosynthesis. The FNR refers to the
ferredoxin—NADP+ reductase. Adapter from Holtzegel (2016)

• photosystem I;
• ferredoxin–NADP reductase (FNR);
• the ATP synthase.

4.2 Light and Dark Reactions of Photosynthesis

Light reactions can be described by the following overall reaction (Arnon, 1971):

2H2 O + 2NADP+ + 3ADP + 3Pi + light → 2NADPH + 2H+ + 3ATP + O2


(4.1)

While dark reactions can be written as follows:

ATP + H2 O → ADP + Pi + n H + G = −34 kJ per mole at p H 7 (4.2)

Table 4.1 gives the main steps of photosynthesis. As can be noticed, we have
mainly six steps while the overall reaction is demonstrated at the end of the table.
It is referred in literature (Klass 1998) that for each of the two light reactions, one
or two photons are required to transfer each electron from the oxidation of the water;
according to this calculation, a total of eight photons is, for this reason, required to
50 4 Overview of the Main Mechanisms of Photosynthesis

Table 4.1 Reaction steps in


Step Characteristic reaction
photosynthesis
1st 2H2 O(l) → 4H+ + 4e + O2 (g)
2nd 4Fd+3 + 4e → 4Fd+2
3rd 3ADP + 3P → 3ATP
4th 4Fd+2 → 4Fd+3 + 4e
5th 2NADP + 4H+ + 4e → 2NADPH2
6th CO2 (g) + 3ATP + 2NADPH2 →
(CH2 O) + 3ADP + 2NADP + 3P + H2 O(l)
Overall CO2 (g) + H2 O(l) → (CH2 O) + O2 (g)
Fd ferredoxin

fix one molecule of CO2 . This is not really important now to argue if it is one photon,
two or many photons required to reduce water and take one electron as this seems
to be difficult to really predict and of course we could do many speculations about
it. It is important now to understand that there are many photos arriving on the plant
with a sufficient energy and within the photosynthetic active radiation required and
these photos are responsible for the oxidation of water.
We assume that CO2 is in the gaseous phase and the initial product is glucose;
consider now the reverse reaction between the water (six molecules) and glucose
to give six molecules of CO2 . In this reaction, 24 atoms of hydrogen are produced.
Now, considering that each of these atoms could be represented as H+ , for each of
these atoms, one electron is then available and therefore, a total of 24 electrons are
produced. Going backward, for each water molecule, four electrons are then derived.
Just to have an idea of the energies involved in this process: the standard Gibbs
free energy change at standard conditions (25 °C) is +0.48 MJ (+114 kcal) per mole
of CO2 assimilated (Klass 1998) and the corresponding enthalpy change is +0.47 MJ
(+112 kcal) (Othmer 2005, p. 12).
The main terms of Fig. 4.3 are explained as follows:
• YZ = tyrosine;
• Fx = Iron–Sulfur center;
• P680 = primary electron donor of PSII composed of chlorophyll (Chl);
• Pheo = pheophytin;
• Qa Qb = plastoquinone A and B (they receive electrons from pheophytin) Qb
receives two protons from the stroma side;
• cytochrome f (Cyt f). cytochrome b low- and high-potential forms (Cyt bLp and
Cyt bHp);
• PC = plastocyanin;
• P700 = primary electron Chl donor of PSI;
• A0 = chlorophyll; A1[Q] = phylloquinone;
• Fd = ferredoxin; FNR = ferredoxin NADP reductase;
• NADP* = oxidized nicotinamide adenine dinucleotide phosphate;
• Yo = symmetrically related tyrosine to YZ.
4.2 Light and Dark Reactions of Photosynthesis 51

Fig. 4.3 Oxygenic photosynthesis, the light side. Adapted from US Department of Energy (2018)

The oxidation of water is described by Eq. (4.3) which is Eq. (3.41) with opposite
direction:

2H2 O → O2 + 4e− + 4H+ (4.3)

Water is oxidized by four states of excitation of manganese ions (Kolling et al.


2012):

S0 → S1 → S2 → S3 → S4

These states are probably four due to the fact that, as mentioned previously, four
electrons are taken for each water molecule and at each of these steps, one electron
will be released. As a summary, four protons are transported across the membrane
for every two electrons delivered to P700.
Concerning the common structure of Type II and Type I reaction center, proteins of
the Type II and Type I reaction centers show a pseudo-twofold axis symmetry (Tsiotis
et al. 1997). The reasons for this particular structure are still not fully understood.
Additionally, it is still not clear how Type II centers are able to differentiate their
active and inactive branches. Figure 4.4 shows a schematic view of how the light
reactions are connected with the dark reactions to form glucose molecules.
Electrons from the light process in the thylakoid membrane are responsible for the
reduction of nicotinamide adenine dinucleotide (NAD*) or its phosphorylated form
(NADP*) (Archer and Barber 2004). We need the electron flow and the production
52 4 Overview of the Main Mechanisms of Photosynthesis

Fig. 4.4 Summary of the oxygenic photosynthesis, light and dark reactions. Adapted from Ksen-
zhek and Volkov (1998)

of electrons to provide the reduced forms NADH or NADPH which in turn will give
the necessary electrons (reducing power) necessary for the CO2 fixation. The ATP is
the molecule which will help in this process by giving additional energy generated
by photosynthetic phosphorylation.
The reaction of water with ATP is shown as follows. This reaction is then used in
the process of binding the CO2 and formation of carbohydrates.

ATP + H2 O → ADP + Pinorganic + nH+ (4.4)

This reaction is reported to have a standard Gibbs free energy of −34.5 kJ/mol at
pH 7.0 (Ksenzhek and Volkov 1998).

4.3 Binding of Carbon Dioxide and Production of Sugar


Molecules

Figure 4.5 shows the Calvin–Benson cycle. This cycle is a common way used by
living organisms to produce sugars from CO2 and therefore, it is very important to
study it and to analyze it in the most accurate way. In literature, many models are given
on this cycle and also quantitative comparisons are made between different models
(Arnold and Nikoloski 2011). From the figure, it should be noted that from ribulose-
4.3 Binding of Carbon Dioxide and Production of Sugar Molecules 53

Fig. 4.5 Binding the carbon dioxide, the Calvin–Benson cycle. Adapted from Ksenzhek and Volkov
(1998)

1,5-biphosphate, it is possible to obtain glyceraldehide 3-phosphate molecules which


condense to form ribulose 5-phosphate. For each five turns of the cycle, we have
six ribulose 5-phosphates. This molecule, in turns, forms again the ribulose-1,5-
biphosphate. It is shown here that ATP and NADP-H molecules are effectively used
during these steps.
54 4 Overview of the Main Mechanisms of Photosynthesis

Fig. 4.6 Calvin–Benson cycle, a simplified scheme. The numbers 1–5 refer to the steps of the
cycle. PGAL is the phosphoglyceraldehyde. Adapted from Starr et al. (2016)

A more simplified diagram is shown in Fig. 4.6. This version is usually more
frequent in literature.
As a summary, it should be remembered here that photosynthesis is the only
biochemical system able to do water oxidation rewritten here as:

2H2 O → O2 + 4H+ + 4e− (4.5)

The chemical–physical mechanisms of photosynthesis are not yet exactly known


even if a large quantity of information is provided in literature. Regarding the
molecule of ATP, this molecule is an extremely fast and unstable compound; it stays
in this form for a very short time and it reacts to give inorganic phosphorus, H+ ions,
and energy (Ingermann et al. 1997). The equation is rewritten here as:

ATP + H2 O ↔ ADP + Pi + nH+ (4.6)

With a G = −34.5 kJ/mol at pH 7.


This process and the production of ATP are done in the thylakoid membrane while
NADPH is used during the reduction of CO2 . This is more clearly written as:
4.3 Binding of Carbon Dioxide and Production of Sugar Molecules 55

Fig. 4.7 Structure of d-glucose and the derived linear combination to give cellulose

NADP+ + H+ + 2e ↔ NADPH (4.7)

The molecules which are storing carbohydrates are d-glucose and the linear com-
bination of d-glucose which forms the cellulose. Their structure is shown here in
Fig. 4.7.
These molecules are split with the normal digestion and completely oxidized
in CO2 and H2 O by means of the process denominated as glycolysis. The energy
released is used to produce 36 molecules of ATP, the energy-carrying molecule in
all living organisms.

6CO2 + 12NADPH + 18ATP → C6 H12 O6 + 12NADP + 18ADP


+ 18Pinorganic + 6H2 O (4.8)

References

Archer, M. D., & Barber, J. (2004). Molecular To global photosynthesis. London: Imperial College
Press.
Arnold, A., & Nikoloski, Z. (2011). A quantitative comparison of Calvin-Benson cycle models.
Trends in Plant Science, 16(12), 676–683. https://doi.org/10.1016/j.tplants.2011.09.004.
Arnon, D. I. (1971). The Light reactions of photosynthesis. Proceedings of the National Academy
of Sciences, 68(11), 2883–2892.
Fultz, M. L., & Durst, R. A. (1982). Mediator compounds for the electrochemical study of biological
redox systems: A compilation. Analytica Chimica Acta, 140(1), 1–18. https://doi.org/10.1016/
S0003-2670(01)95447-9.
Hall, D. O., & Rao, K. (1999). Photosynthesis. Cambridge: Cambridge University Press.
Holtzegel, U. (2016). The Lhc family of Arabidopsis thaliana. Endocytobiosis and Cell Research,
27(2), 71–89.
Hoptkins, W. G. (1999). Introduction to plant physiology (2nd ed.). New York, USA: Wiley Inc.
Ingermann, R., Bencic, D., & Herman, J. (1997). Stability of nucleoside triphosphate levels in the
red cells of the snake. Journal of Experimental Biology, 200(7), 1125–1131.
Klass, D. L. (1998). Biomass for renewable energy, fuels, and chemicals. Academic Press.
Kolling, D. R. J., Cox, N., Ananyev, G. M., Pace, R. J., & Dismukes, G. C. (2012). What are the
oxidation states of manganese required to catalyze photosynthetic water oxidation? Biophysical
Journal, 103(2), 313–322. https://doi.org/10.1016/j.bpj.2012.05.031.
Ksenzhek, O. S., & Volkov, A. G. (1998). Plant energetics. San Diego, California: Academic Press.
Nelson, D. L., & Cox, M. M. (2012). Lehninger principles of biochemistry (6th ed.). W.H: Freeman.
56 4 Overview of the Main Mechanisms of Photosynthesis

Othmer, K. (2005). Kirk-othmer encyclopedia of chemical technology (5th ed., Vol. 12). Wiley-
Interscience.
Starr, C., Evers, C. A., & Starr, L. (2016). Biology: Concepts and applications (10th ed.). Cengage
Learning and National Geograohic Learning.
Taiz, L., Zeiger, E., Møller, I. M., & Murphy, A. (2014). Plant physiology and development. Sinauer
Associates, Oxford University Press.
Tsiotis, G., Hager-Braun, C., Wolpensinger, B., Engel, A., & Hauska, G. (1997). Structural analy-
sis of the photosynthetic reaction center from the green sulfur bacterium Chlorobium tepidum.
Biochimica et Biophysica Acta (BBA)—Bioenergetics, 1322(2), 163–172. https://doi.org/10.1016/
S0005-2728(97)00073-X.
U.S. Department of Energy. (2018). Photosynthesis production of hydrogen from water. U.S. Depart-
ment of Energy Office of Biological and Environmental Research. Retrieved from https://public.
ornl.gov/site/gallery/detail.cfm?id=152&topic=&citation=&general=&restsection.
Chapter 5
Work from Light

Il sole é la lampadina del mondo.


The sun is the light bulb on the world.
Daniel Lucas Ettore De Blasio, 10 October 2018.

5.1 Incoming Radiation

It is clear now that the energy coming from the sun plays a fundamental role in
biomass generation and in how plants and organisms also adapted themselves to the
nature of this energy source. In this part of the manuscript, some notions will be given
on how the incoming radiation from the sun affects the thermodynamics involved.
The sun is our star, and it is well known that life depends on this energy source;
therefore, it is good to know a little about it and some data are given as follows
(Bonanno et al. 2002; Williams 2013):
• The age of the sun is around 4.6 billion years.
• Its inner temperature is around 1.57 × 107 K.
• The inner pressure is enormous and estimated around 2.447 × 1011 bar.
• The inner density is something like 1.622 × 105 kg/m3 .
• The temperature at surface is commonly assumed to be 5700 K.
It is also stated in the literature that at least 4000 years are needed for the radiation
energy to reach the surface of the sun from the center (Odenwald 2018). This is
because, in case we consider the main heat transfer process, heat will be put forward
by radiation mainly. And radiation, as it will be explained later in this manuscript, is
strongly dependent on the direction, which goes from the considered source point to
the receiving unit. For instance, starting from the center, heat will be transferred to
the nearest surfaces (molecules) and forwarded to the nearest facing surfaces. These
surfaces will reflect, absorb, and transmit energy to the next ones and so on until the
surface is reached. Conduction cannot play a major role considering the structure of
the sun, its density, and also the elevated temperatures. The way radiation transferred

© Springer Nature Switzerland AG 2019 57


C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_5
58 5 Work from Light

in the sun can be modeled by considering the so-called drunkard’s walk problem or
“random walk” problem which are mathematical problems (Tijms 2003).
The incoming radiation from the sun is constituted by its photons, and the photons’
density is directly proportional to the temperature of the source. Now let us consider
a general number of particles with a particular energy E, and the overall energy
will be determined as:

n(E)E = g(E) · f (E) · E; (5.1)

where n is the number of photos in this case, E is the energy range considered,
while g(E) is denominated as the number of energy states per unit volume in the
energy range considered, and it is given by Richtmyer et al. (1969):


g(E) = E 2; (5.2)
(h · c)3

while f (E) is the Bose–Einstein distribution:

1
f (E) = E (5.3)
Ae KT −1

Therefore, the original expression, Eq. (5.1) will become:

8π 1
n(E)dE = E2 · · dE (5.4)
(h · c) 3 E
Ae K T − 1

Substituting

E
= x;
KT
this will give

E = K · T · x;

and

dE = K · T dx;

Equation (5.11) will be then:

8π ∞ x2
n(E)dE = · K 3 3
T ∫ dx (5.5)
(h · c)3 0 Ae − 1
x
5.1 Incoming Radiation 59

The integral is solved for A = 1 in the case when photons are considered and it is
equal to the Riemann zeta function which multiplies the Euler beta function:

∞ x2
∫ dx = ζ ( p) · ( p); (5.6)
0 Aex − 1

where p in our case is equal to 3.


Values of the Riemann zeta function and for the Euler beta function are tabulated,
and in this case:

ζ ( p) · ( p) = 2.404 (5.7)

This will mean that the number of photos per unit volume, which is their density, is
equal to:
 3
∗ 8π kT
n = 2.404 3 ; (5.8)
c h

and in terms of number of moles of photons:


 3
8π kT
D = 2.404 (5.9)
NA c3 h

where
D = moles of photons (E)/m3 .
h = 6.626 × 10−34 J s is the Planck constant.
k = 1.38 × 10−23 J/K is the Boltzmann constant.
c = 2.998 × 108 m/s is the speed of light.
NA = 6.02 × 1023 is the Avogadro number (the photons in one photon mole in this
case).

5.2 A Simple Derivation of the Stefan–Boltzmann Law

It is also known that the radiative energy (Joules) is directly proportional to the
temperature of the source, the solid angle between the source and the receiving body,
and a constant value which is the usual proportionality constant. This is expressed
by the Stefan–Boltzmann law (Incropera et al. 2006):

E s = ωs aTs4 ; (5.10)

where the proportionality constant a is expressed by:


60 5 Work from Light

Table 5.1 Common


Name of the quantity Symbol Dimensions
quantities in electromagnetic
transport energy Energy density u ML−1 T−2
Velocity of light c LT−1
Electron mass m M
Charge of electron e M1/2 L3/2 T−1
Absolute temperature θ θ
Gas constant k ML2 T−2 θ−1
With M = mass; L = length; T = time; θ = temperature

2π 4 k 4
a= (5.11)
15c2 h 3
The following is a derivation of the Stefan–Boltzmann law done by dimensional
analysis. This is done here to demonstrate a simpler derivation for the formula and
also to give some consideration on the reasoning on the dimensions of the problem at
hand. As a matter of fact, just by reasoning on the dimensions of a problem, this could
give an important number of information and also influence our decisions before
we start actually to do our complicated balances. Notions of dimensional analysis
are reported in the literature (Bridgman 1963), and here, the analyzed problem is
discussed with additional comments.
Consider the quantities shown in Table 5.1.
We would like to find the function of the energy density from the other variables. At
this point, we have four dimensions and six relations given in the previous table, and
this means that we can obtain two non-dimensional numbers. We choose arbitrarily:

First non-dimensional number = uk α θ β eγ m δ (5.12)

Second non-dimensional number = kθ α1 u β1 m γ 1 cδ1 (5.13)

Since the two numbers have no dimensions, it follows that

1 + α + 0 + 21 γ + δ = 0 condition on M
−1 + 2α + 0 + 23 γ + 0 = 0 condition on L
−2 − 2α + 0 − γ + 0 = 0 condition on T
β −α =0 condition on θ

which gives:

α = −4, β = −4, γ = 6, δ = 0
5.2 A Simple Derivation of the Stefan–Boltzmann Law 61

Solving also for Eq. (5.13), we have that the two non-dimensional numbers are
expressed as:

First non-dimensional number = ue6 k −4 θ −4 (5.14)

Second non-dimensional number = kθ · m −1 c−2 (5.15)

Notice that if we would have put together the terms (u, c, m, e, θ ) in the system
of equation, we would have obtained 1 = 0 which is impossible. This is why by
inspecting we notice that θ and m cannot stay alone, because they have only one
dimension so for instance θ should be with k.
From Eqs. (5.14) and (5.15), it can be stated that
 
u = k 4 e−6 θ 4 f k · θ · m −1 c−2 (5.16)

And this is because, since the two numbers are non-dimensional, they are simply
numbers and therefore any number can be expressed as a function of another number
in the same way the number 3 can be derived as 6/2.
Because the argument of the function is very small (notice the power of the velocity
of light at the denominator), and it is not varying much with the temperature, then
the solution becomes:

u = Const k 4 e−6 θ 4 (5.17)

And this leads to

u = aθ 4 (5.18)

which is a version of the Stefan–Boltzmann law.


In reference to our thermodynamic system, the plant, it was shown previously that
plants utilize energy from the sun to bind CO2 and form organic matter. In case we
consider the plant as an energy converter, a simple energy balance would be:

W = Es − Ec − Q (5.19)

where E c is the energy absorbed by the converter and has a similar expression of E s
(energy of the source) while Q is the heat lost by the converter at the temperature of
the converter.
62 5 Work from Light

5.3 The Solid Angle

The energy coming from an object and transported by radiation is highly dependent
on what is the inclination of the surface source or the inclination of the receiving
surface. This is very well understood if we consider the sun as a point and the earth
surface. As a matter of fact, it is very well known that at the equator it is much warmer
than at the north or south pole. Concerning the evaluation and the understanding of
the solid angle concept, the author will try to give also in this case a simplified
derivation of the solid angle expression.
Let us give a look at Fig. 5.1. The radiation is coming from a distance (let us
consider a large distance; in this way, the source can be considered as a point) on a
surface. Now let us assume that the surface on which the incident radiation is going
is not perpendicular to the direction of the incoming photons. The surface dA where
the photons are going should be multiplied by cos( ) to obtain the perpendicular
surface.
Consider again Fig. 5.1 and the following proportion:

dW s2
= 2 (5.20)
dω σ
And the additional proportion:
s r
= (5.21)
σ 1
From Eqs. (5.20) and (5.21), we have:

dW  r 2
= = r 2; (5.22)
dω 1
and

Fig. 5.1 Solid angle evaluation


5.3 The Solid Angle 63
 
cos
dω = dA; (5.23)
r2

which gives
 
cos
A1
ω= ∫ dA; (5.24)
A r2

where A refers to the differential area shown in the figure. Now, let us suppose
that the incoming radiation is going to plants and consider the plants as an infrared
thermometer working in the photovoltaic mode. In this case, the responsivity of the
detecting body is (DeWitt and Nutter 1988):

dS
R(λ) = (5.25)
d (λ)

where dS is the signal output (Ampere) and d (λ) is the quantity of radiation arriving
on the detector (Watt). Equation (5.25) can be also written as:

i
R(λ) = (5.26)
d (λ)

where i is the current. The relation is valid for linearly responding photodiode. The
energy of the photons is (Liddle 2015):

E = h · ν (J) (5.27)

where h = 6.626176 × 10−34 J s is the Plank constant and ν is the wavelength


frequency ν = cλ0 , c0 is the speed of the wavelength in the considered medium, and
λ is the wavelength at hand

h · c0
E= (5.28)
λ
The radiant flux is given by

n·E
(λ) = ; (5.29)
t
where n is the number of photons and t is the time; therefore, the radiant flux is
expressed by:
64 5 Work from Light

n · h · c0
(λ) = (5.30)
t ·λ
The number of photons absorbed per unit time is:

n · α(λ) n(1 − ρ(λ) − τ (λ))


= ; (5.31)
t t
where α is the absorbance, ρ is the reflectance, and τ represents the transmittance.
These are fractions between 0 and 1 and therefore with no dimensions. The sum of
these three coefficients is equal to one. For this reason:

n · [1 − ρ(λ) − τ (λ)] · q(λ) · e


i= ; (5.32)
t

where q(λ) is number of charges/absorbed photon, and e = 1.6022 × 10−19 C, the


charge of the electron. We obtain finally:

[1 − ρ(λ) − τ (λ)] · q(λ) · λ  


R(λ) = h·c0
A W−1 (5.33)
e

where c0 = 2.9979 × 1014 μms


.
The thermodynamic efficiency of an energy converter machine driven by radiative
energy is often expressed by empirical formulations, for instance (Ksenzhek and
Volkov 1998):
   
W 4 Tc 1 Tc 4
η= =1− + ; (5.34)
Es 3 Ts 3 Ts

where s refers to the source of the radiation while c is referring to the energy converter
(the organism using that radiation), and T is the temperature. The previous equation
shows a straight dependence of the efficiency on the temperatures of the sources and
the converter.

5.4 The Photosynthetic Active Radiation

Plants absorb photons which are above a particular frequency (Aaslyng et al. 1999),
and this could be for instance referred as 430 × 10−12 Hz with an energy of 1.77 eV
(electronvolt) (Archer and Barber 2004). However, there is also a maximum at which
plants absorb energy and this value is 750 THz. The photon energy is expressed also
as:
5.4 The Photosynthetic Active Radiation 65

hν0
0 = (5.35)
e0

in this case:
h = 6.626 × 10−34 J s is the Planck constant.
ν0 is the threshold frequency: 430 × 1012 Hz.
e0 is the electron charge, 1.602 × 10−19 C.
As a matter of fact, the value of 1.77 eV is obtained if we do the following
calculation:

1012 (Hz)
h · 430 × = 1.77 eV (5.36)
charge of electron

It is good to stress here that only a fractional amount of the intensity of solar radiation
is utilized as photosynthetic active radiation, PAR (Knyazikhin et al. 1998). To give
a more practical example, the PAR can be compared to the activation energy needed
for chemical reactions to take place; more simply, a comparison can be made between
two molecules and two stones: Let us assume that we want to break the stones or at
least one of them by means of using their kinetic energy, and the stones will not break
unless we give to them a sufficient amount of energy (in this case kinetic energy). The
same is happening for chemical reactions where the energy in that case is determined
by the temperature, and, as a matter of fact, there is a direct correlation between the
kinetic energy and the temperature for molecules. This is no exception when photons
are considered in this case.
The photosynthetic process can be also simplified to take into account for the
main steps required to convert the incoming energy to work. In Fig. 5.2, a simplified
version of what can be denominated as “the light-harvesting engine” is given.
In Fig. 5.2, E s is the energy coming from the source (sun for instance), E a is the
energy absorbed, E e is the energy required to constitute an excited state, E r is the
part reflected, E q is the part lost under the form of heat, M is the imaginary engine,
and W is the work obtained. T represents the temperature at the various stages.
The four stages for the conversion of energy into work are as follows. According
to the literature (Hall and Rao 1999) 47% of the incoming radiation from the sun is
lost because of being outside the PAR range. This can be expressed as:

∫vv10 I (v)dv
η1 = (5.37)
∫∞
0 I (v)dv

where I (v) is the intensity of the incoming radiation as a function of the frequency,
while v0 and v1 are the frequencies related to the PAR. This will give 53% as the
fraction of energy in the PAR (100 − 47 = 53). Now, about 30% of this 53% is lost
to incomplete absorption (Hall and Rao 1999). This will leave another 15.9% from
the 53% giving the previously mentioned 37%.
66 5 Work from Light

Fig. 5.2 Light-harvesting engine, simplified representation

Ea
η2 = (5.38)
Eu

where E u is the energy which can be utilized and given previously. Considering that
24% of this 37% is lost due to inefficiency in creating the excited state (Hall and Rao
1999), this will leave another 8.9% from the original 100% and therefore (37 − 8.9
= 28.2). The efficiency in generating the excited state will be then:

Ee 28.2
η3 = = = 0.76 (5.39)
Ea 37

where E a is the absorbed energy and E e is the one referred to the excited state. This
could be also expressed as:

Te
η3 = 1 − · r; (5.40)
Ts

where r is a coefficient related to a dilution factor (Ksenzhek and Volkov 1998). The
utilization of the excited state of the working body can be expressed as:
 0.5
Tc
η4 = 1 − ; (5.41)
Te

where c refers to the converter and e refers to the excited state. As a summary: η1
is the fraction of photosynthetic active radiation for different temperatures; η2 is the
5.4 The Photosynthetic Active Radiation 67

Fig. 5.3 Efficiency of light absorption as a function of the temperature

efficiency for the absorption of this energy; η3 is the part of energy transferred in the
excited state; η4 is the fraction of energy in the threshold level.
From Fig. 5.3, it is possible to see that the highest absorption efficiency is situated
around 7000 K (Ksenzhek and Volkov 1998); however, it is known that the sun is at
a temperature of 5762 K at the surface (Albarrán-Zavala and Angulo-Brown 2007)
which suggests that the sun could have been diminished its temperature at the surface
since the evolution of the plants.
It should be mentioned here that there are several values given in the literature for
the overall efficiency calculated for the energy conversion in the process of photo-
synthesis and there are diverse methods. Here, it is reported one definition given by
Ksenzhek and Volkov (1998).
 
rate of enthalpy storage W m−2 in plants as D-glucose
ηPS =   (5.42)
solar irradiance W m−2 acting on the plant

which can be also written as:

JλS αλ λ
∫λλmax
min NA
H dλ
ηPS = ∞ S (5.43)
∫0 E λ dλ

where
NA is the Avogadro number; 
photons
JλS is the solar photon flux m 2 s nm , at a wavelength between λ-min and λ-max;

E λS is the solar irradiance [W/(m2 nm)];


αλ is the fraction of light absorbed by the plant (absorbance);
68 5 Work from Light

λis the quantum yield for a consumption of CO2 , and ideal value is 0.125
photons/(moles of O2 produced);
H is the enthalpy of the photosynthesis (overall reaction).
In the literature, thermodynamic models for the photosynthetic process are avail-
able. In case we consider the overall reaction taking place to form glucose molecules
(Albarrán-Zavala and Angulo-Brown 2007):

6CO2 (g) + 6H2 O(liq) → C6 H12 O6 (s) + 6O2 ; (5.44)

data for the standard enthalpy of formation could be retrieved from databases
(National Institute of Standards and Technology 2018). Considering standard condi-
tions, the enthalpies of formation for the CO2 , H2 O, and glucose are respectively −
393.52 kJ/mol, −285.820 kJ/mol, and −1273.3 kJ/mol. Summing up according to:

0
E Ph = HGlu
0
+ 6Hox
0
− 6HCO
0
2
− 6Hw0 ; (5.45)

where Glu stands for Glucose, ox is the oxygen, and w is the water, this calculation
will give that the energy required to form one mole of glucose would be E Ph 0
=
2802.74 kJ/mol. Additional data are given in terms of standard internal energy at
a particular wavelength and in terms of Gibbs free energy (Albarrán-Zavala and
Angulo-Brown 2007).

kJ
U 0 = 10555.287 at λ = 680 nm
mol
kJ
G 0 = 2880.31 (sugar)
mol

References

Aaslyng, J. M., Rosenqvist, E., & Høgh-Schmidt, K. (1999). A sensor for microclimatic measure-
ment of photosynthetically active radiation in a plant canopy. Agricultural and Forest Meteorology,
96(4), 189–197. https://doi.org/10.1016/S0168-1923(99)00057-X.
Albarrán-Zavala, E., & Angulo-Brown, F. (2007). A simple thermodynamic analysis of photosyn-
thesis. Entropy, 9(4), 152–168. https://doi.org/10.3390/e9040152.
Archer, M. D., & Barber, J. (2004). Molecular to global photosynthesis. London: Imperial College
Press.
Bonanno, A., Schlattl, H., & Paternò, L. (2002). The age of the sun and the relativistic corrections
in the EOS. Astronomy & Astrophysics, 390(3), 1115–1118. https://doi.org/10.1051/0004-6361:
20020749.
Bridgman, P. W. (1963). Dimensional analysis. London: Yale University Press.
DeWitt, D. P., & Nutter, G. (1988). Theory and practice of radiation thermometry. Wiley.
Hall, D. O., & Rao, K. (1999). Photosynthesis. Cambridge University Press.
Incropera, F. P., Bergman, T. L., & Lavine, A. S. (2006). Introduction to heat transfer (5th Edition).
Wiley.
References 69

Knyazikhin, Y., Martonchik, J. V., Myneni, R. B., Diner, D. J., & Running, S. W. (1998). Syner-
gistic algorithm for estimating vegetation canopy leaf area index and fraction of absorbed pho-
tosynthetically active radiation from MODIS and MISR data. Journal of Geophysical Research:
Atmospheres, 103(D24), 32257–32275. https://doi.org/10.1029/98JD02462.
Ksenzhek, O. S., & Volkov, A. G. (1998). Plant energetics. San Diego, California: Academic Press.
Liddle, A. (2015). An introduction to modern cosmology (3rd ed.). Wiley.
National Institute of Standards and Technology. (2018). NIST chemistry webbook. Retrieved October
3, 2018, from https://webbook.nist.gov/chemistry/.
Odenwald, S. (2018). IMAGE Education Center. Imager for magnetopause to aurora global explo-
ration. US: NASA National Aeronautics and Space Administration. Retrieved from https://image.
gsfc.nasa.gov/poetry/ask/a11354.html.
Richtmyer, F. K., Kennard, E. K., & Cooper, J. N. (1969). Introduction to modern physics (6th ed.).
McGraw-Hill.
Tijms, H. C. (2003). A first course in stochastic models. Chichester, West Sussex, England: Wiley.
Williams, D. R. (2013). Sun fact sheet. NASA Goddard Space Flight Center.
Chapter 6
Some Types of Analyses Conducted
on Biomass

The creatures instead …, the creatures are divided in creatures


of love and creatures of freedom.
Adapted translation from the book: Così parlò Bellavista.
Napoli, amore e libertà. Mondadori, 2010.
Luciano De Crescenzo.

6.1 Determination of the Total Nitrogen Content,


the Kjeldahl Method

One of the major problems encountered when microorganisms are utilized for biofuel
production is the determination of how much microorganisms are actually contained
within the slurry at a determinate time. To be noticed is that there is a difference
between the overall amount of biomass contained in the reactor and the overall
amount of biomass that has to do with the actual microbes. For this reason, different
methods have been developed for this purpose. One classification of these methods
involves the measurement of cell mass with instrumental analysis by:
• dry weight technique;
• measurement of the total nitrogen content;
• measurement of turbidity.
Figure 6.1 shows a qualitative description of the turbidity measurement by using
for instance laser diffraction technique; in this case, the more is the obscuration
measured, the higher is the number of cells in the slurry. However, this kind of
analysis is not the most accurate even if it gives a good indication of the amount of
cells present in the slurry sample.
Other examples of techniques used to measure the amount of cells present in the
biochemical reactor are given as follows:

• measurement of cell number (counting);


• weighing;
• separation of organisms from the medium;

© Springer Nature Switzerland AG 2019 71


C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_6
72 6 Some Types of Analyses Conducted on Biomass

Fig. 6.1 A qualitative description of the turbidity measurement by using laser/light diffraction
technique

• drying of biomass;
• light absorbance measurements.
In this manuscript, the evaluation of cell mass by measuring one cell component will
be taken into account. In this case, the component will be the nitrogen, N, and that
will give an approximate amount of protein from which the number of cells or their
concentration can be derived. It is known for instance that 60–65% of the dry weight
of bacteria is protein, or more in detail, there are 3 × 106 proteins
µm3
(the volume refers to
the cell volume) (Milo 2013). Additionally, it is known that the amount of nitrogen
is 8–12% of the dry cell weight (Bradley and Nichols 1918).
The main methods for the determination of the protein content are:
• Biuret [in this case we have a titration, change in color (Keppy and Allen 2018)];
• Lowry (titration, change in color) (Mæhre et al. 2018);
• Kjeldahl (measurement of total nitrogen by chemical treatment, there is release of
ammonia which is then titrated).
In this book, the Kjeldahl method (Kjeldahl 1883) will be explained since it is a quite
suitable method to measure the total organic nitrogen content of the biomass sample.
The main reactions of the method are listed as follows:

Organic N + H2 SO4 → (NH4 )2 SO4 + H2 O + CO2 + by products (6.1)

(NH4 )2 SO4 + 2NaOH → 2NH3 ↑ +Na2 SO4 + 2H2 O (6.2)


6.1 Determination of the Total Nitrogen Content, the Kjeldahl Method 73

2NH3 + H2 SO4 → (NH4 )2 SO4 + Excess (H2 SO4 ) (6.3)

The organic nitrogen is first digested with an excess of sulfuric acid, Eq. (6.1), while
ammonium sulfate is produced; then, the ammonium sulfate is reacting with sodium
hydroxide, Eq. (6.2), to release all the nitrogen as ammonia which is condensed to an
additional flask bottle. The ammonia is then reacting with a known excess amount
of sulfuric acid which will form ammonium sulfate, Eq. (6.3). The excess sulfuric
acid left is then titrated with sodium hydroxide, Eq. (6.4).

Excess (H2 SO4 ) + 2NaOH → Na2 SO4 + 2H2 O (6.4)

Figure 6.2 is a simplified illustration of the process.


The wt% of nitrogen is then calculated from the formula:

[L(H2 SO4 ) · N(H2 SO4 ) − L(NaOH) · N(NaOH)] · 14.0067


wt% N = × 100
sample gram
(6.5)

where
L liters;
N normality.
To be remembered is that the molecular weight of N = 14.0067 g/mol.

Fig. 6.2 Simplified illustration of the Kjeldahl method


74 6 Some Types of Analyses Conducted on Biomass

Table 6.1 Protein factors for Protein factor Protein sources


selected substances
6.38 Milk and dairy
6.25 Grains
5.95 Rice
5.70 Wheat flour

The wt% of N is finally multiplied by the protein factor giving the amount of
protein in weight percentage. The so-called protein factor is calculated for different
substances (Mariotti et al. 2008), and some examples are given in Table 6.1.
The Kjeldahl method can be executed also by using HCL; in this case, the proce-
dure follows the steps:

1. Nitrogenous organic compound + concentrated H2 SO4 → (NH4 )2 SO4 (ammo-


nium sulfate)
2. (NH4 )2 SO4 + 2NaOH → 2NH3 + Na2 SO4 + 2H2 O
3. NH3 + HCl (0.5 N) → NH4 + Cl− + HCl (this quantity is the left back)
4. The left back HCl is titrated with standard NaOH (0.5 N).

The percentage of N is then estimated by the simple relation:

weight of N in grams
% of N = × 100 (6.6)
weight of substance in grams

An additional procedure exists (Martín et al. 2017). In this procedure, the previous
step No. 3 of the HCL procedure is done by using boric acid.

NH3 + H3 BO3 → NH+ −


4 H2 BO3 (ammonium borate) (6.7)

Ammonium borate can be directly titrated by standard acid using methyl red as
indicator.
The advantages with this last method are:
• We could prevent the loss of ammonia by volatilization.
• Boric acid is an acid which is too weak to interfere with further titration of ammo-
nium borate.
• There is no need of back titration as in the case of reaction with HCl.

6.2 Determination of the Higher Heating Value

The determination of the gross calorific value for a biomass sample is of crucial
importance when the biomass itself will be used as fuel for power plants. Let us
assume that we have a biomass-fired power plant, and there are diverse sources from
6.2 Determination of the Higher Heating Value 75

where we could buy the fuel. The first step, in this case, is to assess the fuel itself, and
the first kind of analysis to be performed would be the evaluation of the gross calorific
value. This is done by using the so-called calorimetric bomb. The understanding of
how this measure is made will help the reader to avoid mistakes done commonly in
the calculation of the overall efficiency of a power plant. To be noticed is that, in
case the lower heating value would be taken into consideration for the inlet energy,
the overall efficiency of the power plant could be over 100% if also a condenser is
used. The efficiency of a power plant cannot be over 100% as obvious and also by
referring to the common laws of the thermodynamics.
The functioning of the calorimetric bomb is quite simple (these steps will be
described more in detail later):

1. The biomass is inserted in a chamber.


2. The chamber is closed and inserted in a volume of water.
3. The biomass is combusted completely with excess of oxygen.
4. The heat produced is transferred to the known amount of water.

Figure 6.3 shows a photograph of common laboratory equipment used at Åbo


Akademi University, Faculty of Science and Engineering. The equipment is a Parr
1672 calorimetric bomb which includes a thermometer.
The amount of energy released by the biomass sample during its combustion and
the condensation of the produced water is not straightly related to the specific heat of
water and its variation in temperature. As a matter of fact, diverse factors should be
taken into consideration and some parameters are related to the particular instrument
used. In this case, the released heat is calculated according to Eq. (6.8)

Fig. 6.3 A laboratory setup for calorimetric measurements


76 6 Some Types of Analyses Conducted on Biomass

H = m H2 O · cpH2 O · T + K · T (6.8)

where m H2 O is the mass of water which is present outside the calorimetric bomb and
K is an instrument-related constant. Remember that the value of cp, the specific heat
of water, is not constant with the temperature; however, the amount of water is quite
large compared with the mass of the fuel that we want to analyze; for this reason,
the water temperature will be raised only of some degrees. If the temperature is not
changing much, then it is a good approximation to keep it constant. In any case, there
is the machine-related constant helping us.
The main parts of the calorimeter are:

• the bomb cylinder with sample holder;


• cooling water vessel;
• stirring device;
• ignition system;
• temperature measurement system (temperature increase measurement system).

The following factors affect the measurement results:

• the stirring device operation;


• the heat loss through external walls;
• the increase in temperature of metal components of the system (we take into
consideration of this in the water equivalent coefficient);
• the presence of other combustible materials within the cylinder.

The higher heating value for the considered instrument is evaluated by:

(m w + K )cw · t − qi
HHV = (6.9)
m
where

• mw —mass of water in the vessel;


• K = 0.441 kg—water equivalent value of the calorimeter;
• cw —water specific heat capacity;
• t—increase in water temperature throughout the experiment;
• qi —sum of correction factors for additional thermal effects;
• m—weight of the combusted fuel.

To be specific, some heat is required to combust the wire and thread which are used
during the starting of the process according to:

qi = HHVwi m wi + HHVt m t (6.10)

where
• mwi —mass of burned wire;
• mt —mass of combusted thread;
6.2 Determination of the Higher Heating Value 77

• HHVwi = 6700 kJ/kg—higher heating value of wire;


• HHVt = 14,700 kJ/kg—higher heating value of thread.
Specific heats, enthalpies, specific volumes, internal energies, and entropy values
are tabulated for several compounds; a good reference for obtaining such data is for
instance the Handbook of Chemical Engineers (Green and Perry 2008); additional
data concerning water can be found in The IAPWS Formulation 1995 for the Ther-
modynamic Properties of Ordinary Water Substance for General and Scientific Use
(Wagner and Pruß 2002).
Data for the specific heat can be obtained from enthalpy data, and the specific
heat at 25 °C, for instance, can just be approximated by the enthalpy value at 25 °C
minus the enthalpy at 24 °C. It is possible to derive the cp from these data since it
will be possible to plot them, and with a suitable program, we can evaluate a relation
for cp = f(T).
The relation between the mass, the specific heat, and the temperature is given as
a consequence:

H = mc p T (6.11)

where m is in kg, cp in (kJ/kg K), and T in Kelvin, and H is then the variation in
enthalpy expressed in kJ.
Dividing by the weight of the sample, we have produced enthalpy per unit weight
of sample. The lower heating value, LHV, is evaluated from the HHV as:
 
kJ
LHV = HHV − 2441.8 (9 · H + moisture fraction) (6.12)
kg

In Eq. (6.12) the value 2441.8 (kJ/kg) refers to the heat of vaporization of water,
while H is the fraction of hydrogen in the fuel sample expressed in weight.
We take into account the hydrogen which is present in the fuel since when the
hydrogen reacts with oxygen it releases a big quantity of energy and forms water.
Now, the water is not at the liquid state when the sample is combusted, but it is
present in the form of vapor, and for this reason, some extra calories are required to
keep this state. When the combustion is ended, the vapor will give back this amount
of energy while it is condensing.

6.3 A Procedure for the HHV Calculation Based


on a Particular Instrument

With reference to Fig. 6.4, the main parts of the overall calorimeter are listed as
follows: thermometer (1), thermometer bracket (2), thermometer support washer
(3), thermometer lens (4), thermometer support rod (5), motor (6), motor pulley
(7), stirrer drive belt (8), stirrer pulley (9), stirrer bearing assembly (10), ignition
78 6 Some Types of Analyses Conducted on Biomass

wires (11), stirrer shaft and impeller (12), bucket (13), calorimeter jacket and cover
(14), and oxygen combustion bomb or calorimetric embodiment (15). From now on,
the part of the calorimeter where the combustion of the sample takes place will be
denominated as calorimetric embodiment.
The main components constituting the calorimetric embodiment are shown in
Fig. 6.5.
The assembly of the embodiment itself and the arrangements of the biomass
sample within the calorimeter is made according to following steps:

a. The first step is to weight the capsule (where the sample will be positioned) and
set the tare (Fig. 6.6).
b. Then, the sample should be put in the capsule and a note of the weight on the
display should be taken (hint: it is a good idea to let the value stabilize since this
measure should be quite accurate).
c. The capsule should be arranged in the ring of the calorimeter embodiment. This
ring is pending from the top of the embodiment as shown in Fig. 6.7.
d. The next step is to cut 10 cm length of the ignition wire, using the ruler already
drawn on the reel (Fig. 6.8).
e. The following operation is not that easy to perform and requires a certain manual
capacity. The goal here is to combust the sample pellet, and, in order to do this,
we must heat up the sample in some ways. This is done by means of the resistance
wire which has to be in contact with the sample. At this point, the extremities
of the wire should be inserted in holes which are positioned in top-lateral part

Fig. 6.4 Main parts of the overall calorimeter


6.3 A Procedure for the HHV Calculation Based on a Particular Instrument 79

Fig. 6.5 Main components of the calorimetric embodiment

Fig. 6.6 Scale used for


sample weight measurement
80 6 Some Types of Analyses Conducted on Biomass

of the calorimetric embodiment (see Fig. 6.9) and, as mentioned previously, the
wire should be shaped in such a way that it must be in contact with the pelletized
fuel, but without touching the capsule (otherwise you will create a short circuit
which will burn only the wire).
f. The next step is the closure of the calorimetric bomb. To do this, it has to be kept
in mind that if there is some extra pressure into the container, the vessel could not
close properly. Therefore, we should unscrew the knurled knob which is situated
on the top of the vessel (this valve must be open to properly close the bomb).
The top part of the embodiment should be carefully inserted on the cylinder and
closed. The knob should be closed too (see Figs. 6.10 and 6.12).
g. At this point, the embodiment shown in Fig. 6.10 should be pressurized with
oxygen. The concept here is that the wire will be brought to very high temperature
and become incandescent. When these conditions are achieved, the oxygen will
be combusted powerfully. The oxygen for instance could have a pressure of
30 bar, and in order to do this, a oxygen bottle should be connected to the top of
the embodiment through the suitable one-way valve. Of course, the pressure is
read by means of a pressure meter as the one given in Fig. 6.11.
h. At this stage, the calorimetric embodiment is ready (Fig. 6.12).

Fig. 6.7 Positioning of the


capsule inside the
embodiment ring
6.3 A Procedure for the HHV Calculation Based on a Particular Instrument 81

Fig. 6.8 Cutting the ignition wire

Fig. 6.9 Positioning of the


ignition wire, to be noticed
that the wire has to be in this
kind of shape but of course
has to contact the sample

Figure 6.13 shows the section where the calorimetric embodiment is placed. In this
section, the water will “collect” the calories contained within the combusted biomass
sample and its temperature will be raised accordingly. In this device, the rise in
water temperature is observed, registered, and subsequently processed by the digital
thermometer.
Continuing with the procedure for the utilization of the instrument, the following
steps should be performed:
82 6 Some Types of Analyses Conducted on Biomass

Fig. 6.10 Oxygen filling of


the calorimetric embodiment

i. Figure 6.14 shows the container where the calorimetric embodiment is placed
and where water is filled. In this case, the procedure is to fill the container with
exactly 2 L of distilled water. Of course, a tare should be taken on the empty
container before filling it with water.
j. This container should be placed within the calorimetric jacket, and usually, there
are some signs where the container should be fitted. This signs are also at the
bottom of the jacket (Fig. 6.15).
k. The calorimetric embodiment shown in Fig. 6.16 should be, at this stage, inserted
into the container filled with water. This operation should be done with care by
using a lifting handle to be attached to two suitable holes at the top sides of
the embodiment itself. The embodiment should be transported carefully to not
disturb the sample and its arrangements. The embodiment is placed at the bottom
of the water contained where suitable signs are present to allow for its correct
positioning. The two ignition electrodes should be plugged into the terminal
connections on the top of the calorimetric embodiment. Attention should be paid
to not remove any water from the bucket during the whole procedure (Fig. 6.17).
When the sample is combusted into the calorimetric embodiment, the temperature of
the water is measured. However, it is also true that the temperature should be uniform
within the medium in order to have an accurate measure. For this reason, stirring is
required. The procedure continues with the following actions:
l. The jacket should be covered and the cover should be positioned in the right way
so that we have the stirring in the water. The stirring is done by means of a belt
(see Fig. 6.18), and this belt is connected to an electric motor with a switch.
6.3 A Procedure for the HHV Calculation Based on a Particular Instrument 83

Fig. 6.11 Oxygen bottle


with pressure meters

Data acquisition and elaboration are performed by the data logger, Fig. 6.19, which
eventually provides with the higher heating value of the sample of interest and corrects
it from the influence of nitrogen and sulfur occasionally present in the fuel and from
the amount of heat generated by the ignition wire.
In this case, the procedure continues as follows, to be noticed that this procedure
refers to the particular instrument used:

m. If the assembly has been done in the correct way and depending on the instrument
used, on the display of the data logger we could have a message of “READY”,
or we could confirm that we have finished the procedure.
n. The value of the sample mass previously weighted should be input.
o. At this point, the measurement has started and no actions are required. If every-
thing has been assembled correctly, the process can take about 15 min before
outputting the HHV value. During this time, the sample will pass through three
“periods”: the preperiod, during which the mixing makes the water temperature
in the water container homogeneous; the rise period starts when the sample is
84 6 Some Types of Analyses Conducted on Biomass

Fig. 6.12 A top view of the


calorimetric embodiment
closed and ready

Fig. 6.13 Calorimeter jacket and cover


6.3 A Procedure for the HHV Calculation Based on a Particular Instrument 85

Fig. 6.14 Water container

Fig. 6.15 Positioning of the


water container
86 6 Some Types of Analyses Conducted on Biomass

Fig. 6.16 Handling of the


calorimetric embodiment by
suitable handle

combusted; the post-period starts when the temperature stops rising and will
carry on till the temperature becomes constant.
p. The machine will advise with a message or sound when the measurement is
ended, and the HHV value will be displayed.

The calorimeter could be programmed to perform the calculation of the gross calorific
value in accordance to Eq. (6.13).

W · T − e1 − e2 − e3
HHV = (6.13)
m
where

HHV Higher heating value (J/kg);


W the energy equivalent of the calorimeter, the amount of energy to rise 1 °C
the calorimeter T (J/°C);
e1 the heat produced by burning the nitrogen entrapped in the bomb to form
nitric acid (J);
6.3 A Procedure for the HHV Calculation Based on a Particular Instrument 87

Fig. 6.17 Positioning of the


calorimetric embodiment
inside the water container
and connection of the cables

Fig. 6.18 Positioning of the jacket cover and the elastic belt
88 6 Some Types of Analyses Conducted on Biomass

Fig. 6.19 Data logger

e2 the extra heat produced by burning sulfur to sulfur trioxides and forming
sulfuric acid instead of sulfur dioxide (J);
e3 the heat produced by the burning fuse wire (J);
m the mass of the sample (g).

The influence of the fuse could be estimated by considering the length of the left-
over wire. We assume in this case that the disappeared part was actually combusted,
and the following calculation steps will be done:

E tot = HHVmeas · m sample (6.14)

E fuse = HHVfuse · lburnt fuse · 4.186 × 10−6 (6.15)

E sample = E tot − E fuse (6.16)

E sample
HHVsample = (6.17)
m sample

where
The term 4.186 × 10−6 is a conversion factor.

msample is the mass of the sample (kg);


l burnt fuse is the length of the burnt wire (cm);
E tot is the total energy developed by the combustion (MJ);
6.3 A Procedure for the HHV Calculation Based on a Particular Instrument 89

E fuse is the energy developed by the combustion of the only fuse wire (MJ);
E sample is the energy developed by the combustion of the only sample (MJ);
HHVmeas is the measured higher heating value, that is, the amount of energy gen-
erated by the combustion of all substances inside the bomb, related to
the mass of sample (MJtot /kgsample );
HHVfuse is the higher heating value of the nickel/chromium wire as indicated on
the reel. In this case, this is equal to 2.3 (calfuse /cm);
HHVsample is the higher heating value of the sample only (MJsample /kgsample ).

The chlorine content determination is done in a similar way to the determination


of the higher heating value. The method takes into account the combustion of a test
sample in an oxygen bomb calorimeter (same as the HHV determination), and then:

• The solution is rinsed with distilled water.


• A solution which contains chloride ions is obtained.
• The solution is reacting with silver nitrate to form silver chloride and then recovered
using a precipitation technique.

The reaction is expressed as follows:

Cl− + Ag+ (aq) → AgCl (s) (6.18)

References

Bradley, H. C., & Nichols, M. S. (1918). Nitrogen content of bacterial cells. Journal of Biological
Chemistry, 33, 525. Retrieved from http://www.jbc.org/content/33/3/525.full.pdf.
Green, D. W., & Perry, R. H. (2008). Perry’s chemical engineers’ handbook (8th ed.). McGraw-Hill.
Retrieved from https://doi.org/10.1036/0071422943.
Keppy, N. K., & Allen, M. W. (2018). The biuret method for the determination of total protein using
an evolution array 8-position cell changer (pp. 2). Thermo Fisher Scientific.
Kjeldahl, J. (1883). Neue Methode zur Bestimmung des Stickstoffs in organischen Körpern.
Zeitschrift für Analytische Chemie, 22(1), 366–383.
Mæhre, H. K., Dalheim, L., Edvinsen, G. K., Elvevoll, E. O., & Jensen, I. -J. (2018). Protein
determination—Method matters. Foods, 7(1). https://doi.org/10.3390/foods7010005.
Mariotti, F., Tomé, D., & Mirand, P. P. (2008). Converting nitrogen into protein—Beyond 6.25 and
Jones’ factors. Critical Reviews in Food Science and Nutrition, 48(2), 177–184. https://doi.org/
10.1080/10408390701279749.
Martín, J., Fernandez Sarria, L., & Asuero, A. G. (2017). The Kjeldahl titrimetric finish: On the
ammonia titration trapping in boric acid. In Advances in titration techniques. https://doi.org/10.
5772/intechopen.68826.
Milo, R. (2013). What is the total number of protein molecules per cell volume? A call to rethink
some published values. BioEssays, 35(12), 1050–1055. https://doi.org/10.1002/bies.201300066.
Wagner, W., & Pruß, A. (2002). The IAPWS formulation 1995 for the thermodynamic properties
of ordinary water substance for general and scientific use. Journal of Physical and Chemical
Reference Data, 31(2), 387–535. https://doi.org/10.1063/1.1461829.
Chapter 7
Thermogravimetric Analysis (TGA)

When Alexander the Great (the king of the Greek kingdom


of Macedon) asked to Diogenes of Sinope: “what I can do
for you?”
Diogene answered: “Stand aside so I can get some sun”.
Free translation from: Epictetus, The Discourses, Book 3,
108 AD.

7.1 Basic Principles of Thermogravimetric Analysis

Thermogravimetric analysis (TGA) is often used within studies of thermal decom-


position of biomass. For instance, with this method we could study the pyrolysis
of substances at diverse conditions and heating rates. The instrumental measure in
some cases can give the so-called differential thermal gravimetric (DTG) thermo-
gram from which assumptions are done on the reaction rates and the rates of the
diverse steps in the thermal treatment. To give one example, the pyrolysis of some
organic substances and wastes like manures is studied by this method (Chong et al.
2019) or, for instance, the decomposition of micro-algae (Xu et al. 2017).
The thermogravimetric method is obviously belonging to the so-called thermal
analysis, and this is because the sample is treated at diverse temperatures and with
different atmospheres. The method allows for a simultaneous measurement of tem-
perature, time, and mass of a sample in a controlled dynamic atmosphere (Fig. 7.1).
The foundation on the basis of these kinds of measures is the variation of weight
of the sample as a consequence of the thermal treatment. Because of the process,
some mass will be lost in terms of volatiles or decomposition and this weight loss is
measured normally by means of a microbalance. The sample is inserted in a “sample
pan” or container which is specifically designed for this kind of measure. Concerning
the temperature conditions and its measurement, the temperature variation is set
according to a customized temperature program which may include isothermal and
ramp steps with different heating rates, while the temperature is measured with
thermocouples which are in contact with the sample container (Fig. 7.2).
The pan is usually positioned on a sample holder, which is connected with the
mass sensitive element (the microbalance). The TG system (sample, sample holder,
© Springer Nature Switzerland AG 2019 91
C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_7
92 7 Thermogravimetric Analysis (TGA)

Fig. 7.1 Example of thermogravimetric analyzer. Curtesy of Prof. Antonio Galvagno and Mauro
Prestipino at University of Messina, Italy

Fig. 7.2 Simple representation of the thermogravimetric method


7.1 Basic Principles of Thermogravimetric Analysis 93

pan) is heated using an electric furnace, which in particular cases can reach up to
2000 °C, depending on the specific materials and design of the furnace and other
components. The sample pans that are used in a TGA device can be of different
shapes and materials. They should be able to safely store the sample, to not react
with the sample, and suitable for the desired temperature. The most common sample
pans are made of alumina, platinum, and aluminum.

7.2 TGA Classification and Types of Equipment

The TGA devices can be grouped in three main types, depending on the design of
the sample holder-microbalance system, as shown in Fig. 7.3: cantilever (a), hang
down (b), and top-loading column (c).
The hang down and the top-loading systems are different with respect to the
position of the scale. In the first case, the sample is hanged and the scale is on top of
it, and in the second case, the scale is at the bottom. Both methods take into account
the gravity force. The devices with two sample holders are used for differential
thermal analysis with reference materials, as discussed later in this section.
During a TGA, the wide temperature variation at which the gas may be subjected
involves a variation in the purge gas density that cannot be neglected. This effect is
denominated as buoyancy effect and can be observed as an apparent mass variation,
whose magnitude depends on the kind of the TG apparatus. As a matter of fact,
this phenomenon depends on the volume of the sample and the density of the gas
which is forming the gas atmosphere into the chamber. In the numerical processing
of the TGA curves, this phenomenon is often hidden or not taken into consideration
(Auroux 2013).
For these reasons, some of the trials require a blank test (i.e., with the empty
sample pan) for each TGA program, in order to make the correction curve that will
be subtracted from the sample’s test. A TGA program is the combination of the
different temperature variation steps, characterized by specific heating rates and gas

Fig. 7.3 Three of the most common TGA configurations: cantilever (a), hang down (b), and top-
loading column (c)
94 7 Thermogravimetric Analysis (TGA)

flow. When a variation occurs from a test to another, a new blank test should be
performed.
This method of analysis can be coupled with other methods like differential ther-
mal analysis (DTA) and differential scanning calorimetry (DSC). This is usually
done when the composition of the sample is not known (Van Humbeeck 1998).

7.3 Procedure and Sample Analysis

The TGA apparatus is usually loaded with a sample mass in the range 2–20 mg,
depending on the density of the sample, the sample pan size, and the purpose of
the analysis. Concerning the physical properties of the sample, these can affect the
TG curves. For instance, in the case of an exothermic reaction, higher mass loading
may lead to higher onset temperature because of the higher heat released by the
sample, which involves a sudden local temperature increase. Hence, considering the
thermal inertia of the system and the fact that the thermocouple is positioned in
close contact with the sample, the detected temperature may apparently result higher
than the one observed during tests with lower mass. The magnitude of this effect is
correlated to the physical properties of the sample, the sample pan, and the amount
of heat released or absorbed during the sample transformation. A similar effect is
noticeable with different sample’s sizes and shapes. Powders, flakes, and blocks have
different effects on heat diffusion, which can lead to an apparent retard or anticipation
of the onset temperature or time. Furthermore, in the case of powders, significant
differences in the particle size between two runs may affect the repeatability of the
results. To give one example, the results of thermogravimetric analysis has been
reported to be highly dependent on the sample size for particular samples (Stawski
2009), while also the gas flow and the geometry of the pan can affect the results.

7.4 Positive and Negative Mass Variation

The thermogravimetric analysis is used for characterizing the thermal behavior of


materials during thermal treatment in either inert or reactive atmosphere. In a reactive
atmosphere, it is also possible to study the reactions that occur between the sample
and the purge gas if it involves mass variation (increase or decrease). Hence, it is
possible to classify the thermal processes that occur in a TGA device according to
the positive or negative variation of mass as follows:
Positive mass variation:

– adsorption;
– oxidation;
– reduction.
7.4 Positive and Negative Mass Variation 95

Negative mass variation:


– desorption;
– thermal decomposition (with volatiles formation);
– oxidation (combustion);
– vaporization;
– sublimation.
From the list above, another classification can be made between thermal and
thermochemical processes, depending on the capacity of the purge gas to react with
the sample in the temperature range of the test.

7.5 Reporting TGA Results

As mentioned above, TGA results are plotted as mass versus temperature or time.
Another kind of representation of the results is the derivative thermogravimetry
(DTG), which is the first derivative of the TGA curve. Hence, it describes the mass
change rate. DTG is also used to detect more subtle effects or when kinetic parame-
ters should be investigated (Gallagher 1998). Since the TGA signal is reported in mg
or % (100% is the initial mass at time zero), the DTG is usually plotted as mg/min
or %/min. Figure 7.4 shows the plot of both TGA and DTG curves obtained with a
thermogravimetric test of a biomass sample under nitrogen atmosphere and constant
heating rate up to 750 °C. The TGA curve reported in figure describes the thermal

Fig. 7.4 TGA (solid line) and DTG (dashed line) signals of biomass conversion in inert atmosphere
96 7 Thermogravimetric Analysis (TGA)

decomposition of biomass following different steps that occur at a different temper-


ature. From a TGA plot, a constant slope characterizes the different steps of thermal
decomposition. In the case of Fig. 7.4, these steps are hardly detectable from the
different slopes of the curve, because they occur in overlapped temperature ranges.
The DTG plot helps the detection of transformations that overlap or are associated
with poor mass loss. Indeed, the peaks of DTG indicate a specific transformation,
and the minimum values coincide with the maximum rate of mass loss of the related
decomposition step.

7.6 Differential Thermal Analysis and Differential


Scanning Calorimetry

The TGA can be combined with differential scanning calorimetry (DSC) or differ-
ential thermal analysis (DTA) in order to carry out a complete analysis of the thermal
behavior of materials with one equipment. Indeed, in addition to the mass variation,
it is possible to provide quantitative and qualitative information on the endothermic
and exothermic phenomena. Both DTA and DSC are based on the comparison of the
different thermal behavior between the reference material and the sample (Pope and
Judd 1997).
The DTA measures the temperature difference between the reference and the sam-
ple when they are subjected to the same heat flux. As it is for the TGA, the differential
temperature is plotted against time or temperature. Any change in the sample that
involves heat release (exothermic process) also involves the detection of higher tem-
perature in the sample than the reference. The opposite happens when the changes in
the sample cause heat absorption (endothermic process). An important observation
for the correct interpretation of the DTA and DSC data is that the differences in
heat capacity and thermal conductivity between the sample and the reference may
cause variation of the differential temperature as well, even if no transformations
are occurring. It follows that these thermal analytical techniques allow measuring
the heat capacity of materials. Furthermore, they allow for the detection of those
transitions that do not involve heat release or adsorption but imply the variation of
thermal properties of the sample, as in the case of some solid-phase changes.
The main difference between the DSC and DTA lies in the method used for cal-
culating the different behavior between the inert reference material and the sample.
Indeed, the DSC records the energy that is necessary to provide to the sample or the
reference in order to determine a zero temperature difference between them. A DSC
device has two individual heaters, one for the sample and one for the reference, and
two individual temperature sensors. This method is also called the power compen-
sation principle. Some devices are claimed as DSC even if they have a single heat
source. In this case, they should be defined as false DSC, because the principle is
closer to a DTA since the energy needed for keeping zero temperature is calculated
and not directly recorded by the heating unit, which is shared by both sample and
7.6 Differential Thermal Analysis and Differential Scanning Calorimetry 97

Fig. 7.5 TGA/DTA sample


holder of a top-loading
column TGA/DTA device

reference. The latter method is also called heat flow DSC. It, therefore, follows that,
in many cases, combined TGA-DSC devices do not have real DSC. In principle, the
two DTA and DSC are equivalent for qualitative analysis, with the difference that
DSC gives better performance at a low heating rate.
Furthermore, because of the materials used for accurate measurement, most of
power compensation DSC devices can run in an inert atmosphere and at a lower
temperature than DTA (600–750 °C). However, for accurate quantitative analysis
power compensation DSC should be recommended. Figure 7.5 shows the TG/DTA
sample holder of a top-loading sample holder with both reference and sample pans.
The study of biomass thermal behavior through TGA provides information about
its composition and reactivity. Indeed, several characteristics can be determined,
such as moisture content, fixed carbon, volatile matter, and ash (Basu 2013), making
TGA a reliable and rapid method for performing the proximate analysis of biomass.
Some authors developed methodologies for biomass characterization based on ther-
mogravimetry, which allows calculating the content of the main pseudo-components
of biomass, i.e., hemicellulose, cellulose, and lignin, as well as the proximate anal-
ysis, with a single thermogravimetric test (Saldarriaga et al. 2015). These methods
combine TGA, deconvolution of the DTG signals, and empirical correlations.

7.7 TGA and the Study of the Kinetics of Diverse Thermal


Processes

The correlation between the mass variation and time in TGA signals allows studying
the kinetic of thermochemical conversion under different atmospheres, which means
it is possible to investigate the kinetics of torrefaction, pyrolysis, gasification, and
combustion. The knowledge of the kinetic behavior of biomass conversion is of
primary importance in order to optimize the design of the reactors and the operating
conditions (residence time, temperature, pressure, oxidant medium flows).
The thermogravimetric experiments for carrying out the kinetic studies can be
performed according to isothermal, linear, or stepwise linear heating programs. With
98 7 Thermogravimetric Analysis (TGA)

such analysis, it is possible to determine the evolution of conversion rates as a func-


tion of the conversion’s level, temperature, or time. Also, it is possible to extrapolate
the kinetic parameters of biomass conversion by performing TGA at different par-
tial pressures (only in case of reactive purge gas and by fixing the temperature) and
different temperatures (fixing the partial pressure of the reactive gas), for isothermal
runs, or different heating rates for non-isothermal runs. The details of the differ-
ent approaches that can be applied to calculating the kinetic parameters are widely
described in the specific literature (Vyazovkin et al. 2011). Some authors suggested
alternative methods that allow computing the activating energy or reaction order with
a single run by repeatedly shifting the temperature or the gas partial pressure between
two predetermined levels, respectively (Zimbardi 2000). Additional research works
showed the possibility to use some empirical correlations between the conversion
models and the inorganics embodied in biomass, as well as between the kinetic
constant and the inorganics, in order to predict the conversion profiles and kinetics
of thermogravimetric processes under specific reacting environment (Dupont et al.
2011; Prestipino et al. 2018).

7.8 TGA Coupled with Mass Spectrometry, Gas


Chromatography, and Infrared Spectroscopy

The study of the thermochemical behavior of biomass through TGA can be improved
by coupling the TG devices with the analysis of the evolved gas. This upgrade can
enhance the outcomes of the tests by providing the variation of the gas evolution from
the sample as a function of time and temperature, correlated to the corresponding
mass variation. The most common techniques for the evolution gas analysis coupled
with TGA are mass spectrometry (MS) and infrared spectroscopy (IR). In some
cases, gas chromatography combined with mass spectroscopy (GC/MS) is also used
for a more comprehensive characterization (Cai et al. 2018). TG-FTIR and GC/MS
can be used together to study the mechanisms involved in the thermal treatment of
the sample; for instance, they can be coupled to study the mechanisms of pyrolysis
(Xu et al. 2018). These solutions are usually said to be hyphenated, and they can be
even used when we want to know the composition of an aqueous sample (Garavaglia
2012). Hyphenated techniques are applied to study diverse kinds of samples, and for
different purposes, they usually include two or more analytical techniques which are
coupled together online (Niessen 2017).
The kinetic studies of pyrolysis and torrefaction are usually carried out in nitrogen,
argon, or helium flow in order to simulate endothermic processes in the absence of
oxygen. With the most common TG devices, the maximum heating rate for reliable
results is 50 K/min (Cai et al. 2018), even though most of the manufacturers state
that the devices can work up to 100 K/min. This poses some limitations in scaling up
the results of the TGA pyrolysis to real-scale processes, which are often performed
under fast and flash conditions (Basu 2013).
7.8 TGA Coupled with Mass Spectrometry, Gas Chromatography … 99

During thermochemical gasification, biomass undergoes to various steps, such


as drying, pyrolysis, gasification, and partial combustion, which are characterized
of different temperatures. After pyrolysis, the solid biomass is converted into char.
The latter is the actual solid reactant that is subjected to the partial oxidation and the
different heterogeneous reactions with the gasification mediums and the products of
combustion. Hence, the kinetics of gasification and combustion is studied using the
char of parent biomass. The gasification step is also the slowest one, which involves
its reactions being the controlling ones and the most important in the design of the
whole gasification process reactors. For the kinetics of combustion, different oxy-
gen partial pressures are adopted, while in the study of gasification kinetics, char is
subject to a reacting environment consisting of carbon dioxide or steam at differ-
ent partial pressures, using an inert gas as the complement. When steam is used as
gasification medium in combination with air in a real gasification process, the ther-
mogravimetric characterization with steam is of primary importance to understand
the kinetic behavior. Indeed, in steam or air-steam gasification, the heterogeneous
steam–carbon reaction is the most relevant one. The reaction between carbon diox-
ide and carbon, with the formation of carbon monoxide (reverse Boudouard), should
always be studied, primarily when the air is used as gasification medium since it is
the most limiting one, even in the case of air-steam gasification. However, at temper-
atures lower than 750 °C and high steam partial pressures, the kinetics of steam-char
is still the most relevant for the design of the reactor and the proper modeling of the
process.
The importance of studying the conversion kinetics of biomass is evidenced in
Fig. 7.6. The figure shows the isothermal runs of different residual chars in a steam
environment with a TG device, where the conversion (X) is plotted as a function of
time. In Fig. 7.6, we have results for the olive pomace char (OPChar), citrus peel
char (CPChar), and grape pomace char (GPChar).
The conversion expresses the fraction of mass loss compared to the final theoretical
mass of complete conversion (X = 1), as expressed in Eq. (7.1)

mi − mt
X= (7.1)
mi − mf

where mi , mt, and mf are the initial mass, the mass at time t, and the theoretical final
mass, respectively.
From Fig. 7.6, it is possible to observe that using different parent biomass may
lead to very different degrees of conversion at a fixed time t. Indeed OPChar achieves
only 55% of conversion after 1000 s, at 750 °C and 50 kPa of steam partial pressure,
while CPChar and GPChar are almost entirely converted. The different behavior is
attributed to the different composition of the inorganics, which affect the kinetic
of steam gasification of chars (Dupont et al. 2011; Prestipino et al. 2018). It fol-
lows that performing a thermogravimetric run at proper conditions it is a rapid and
straightforward method that allows understanding the affinity of different feedstock
to be converted at the same conditions, in addition to the possibility of providing
preliminary information on the proper process parameters.
100 7 Thermogravimetric Analysis (TGA)

Fig. 7.6 Conversion versus time curves of different residual chars tested through a TGA in 50 kPa
steam atmosphere at 750 °C and isothermal conditions. Data from Prestipino et al. (2018)

7.9 Methods and Experimental Conditions Suggested

When performing TGA for studies of kinetics related to gasification and combustion
of chars, some precautions should be taken into account in order to avoid diffusion
limitation and keeping kinetic controlled conditions. To this regards, the tempera-
ture should be <900 °C and the particle size <0.1 mm (Di Blasi 2009). As particle
size increases, the diffusion of gas reactants into the char is limited, which leads
to the diffusion-controlled regimes. The heat transfer plays an essential role in the
transition between kinetic and diffusion-controlled regimes. Indeed, higher temper-
ature involves higher heat and mass transfer, moving away from the kinetic regime.
Although using TGA it is possible to get very accurate data for the kinetics of biomass
conversion, the difference between the particles shape and size in the laboratory anal-
ysis and those in a real-scale reactors poses the problem of addressing the mass and
heat transfer of real-scale biomass particles. This limitation can be partially over-
come by using macro-TGA devices that allow for testing biomass samples in the mass
range 30–200 g where heat and mass transfer limitations can be taken into account.
In some cases, the reactor itself can be used as a macro-TG if it is continuously
weighted (Onsree and Tippayawong 2017). The drawbacks of using macro-TGA are
the loss of resolution and precision in the study of the real reactions’ kinetics.
References 101

References

Auroux, A. (Ed.). (2013). Calorimetry and thermal methods in catalysis. Berlin, Heidelberg:
Springer. Retrieved from www.springer.com/gp/book/9783642119538.
Basu, P. (2013). Biomass gasification, pyrolysis and torrefaction—Practical design and theory (2nd
ed.). Oxford, UK: Academic Press. Retrieved from https://doi.org/10.1016/c2011-0-07564-6.
Cai, J., Xu, D., Dong, Z., Yu, X., Yang, Y., Banks, S. W., et al. (2018). Processing thermogravimetric
analysis data for isoconversional kinetic analysis of lignocellulosic biomass pyrolysis: Case study
of corn stalk. Renewable and Sustainable Energy Reviews, 82, 2705–2715. https://doi.org/10.
1016/j.rser.2017.09.113.
Chong, C. T., Mong, G. R., Ng, J. -H., Chong, W. W. F., Ani, F. N., Lam, S. S., et al. (2019).
Pyrolysis characteristics and kinetic studies of horse manure using thermogravimetric analy-
sis. Energy Conversion and Management, 180, 1260–1267. https://doi.org/10.1016/j.enconman.
2018.11.071.
Di Blasi, C. (2009). Combustion and gasification rates of lignocellulosic chars. Progress in Energy
and Combustion Science, 35(2), 121–140. https://doi.org/10.1016/j.pecs.2008.08.001.
Dupont, C., Nocquet, T., Da Costa, J. A., & Verne-Tournon, C. (2011). Kinetic modelling of steam
gasification of various woody biomass chars: Influence of inorganic elements. Bioresource Tech-
nology, 102(20), 9743–9748. https://doi.org/10.1016/j.biortech.2011.07.016.
Gallagher, P. K. (1998). Thermogravimetry and thermomagnetometry, Chap. 4. In M. E. Brown
(Ed.), Handbook of thermal analysis and calorimetry (Vol. 1, pp. 225–278). Elsevier Science
B.V. https://doi.org/10.1016/S1573-4374(98)80007-1.
Garavaglia, M. (2012). Analysis of an unknown aqueous sample by TG-IR-GC/MS. PerkinElmer.
Retrieved from http://photos.labwrench.com/equipmentManuals/14602-5645.pdf.
Niessen, W. M. A. (2017). Hyphenated techniques, applications of in mass spectrometry. In J. C.
Lindon, G. E. Tranter, & D. W. Koppenaal (Eds.), Encyclopedia of spectroscopy and spectrometry
(3rd ed., pp. 174–180). Oxford: Academic Press. https://doi.org/10.1016/B978-0-12-409547-2.
05240-9.
Onsree, T., & Tippayawong, N. (2017). Application of Gaussian smoothing technique in evaluation
of biomass pyrolysis kinetics in macro-TGA. Energy Procedia, 138, 778–783. https://doi.org/10.
1016/j.egypro.2017.10.059.
Pope, M. I., & Judd, M. D. (1997). Differential thermal analysis—A guide to the technique and its
applications. Heyden.
Prestipino, M., Galvagno, A., Karlström, O., & Brink, A. (2018). Energy conversion of agricultural
biomass char: Steam gasification kinetics. Energy, 161, 1055–1063. https://doi.org/10.1016/j.
energy.2018.07.205.
Saldarriaga, J. F., Aguado, R., Pablos, A., Amutio, M., Olazar, M., & Bilbao, J. (2015). Fast char-
acterization of biomass fuels by thermogravimetric analysis (TGA). Fuel, 140, 744–751. https://
doi.org/10.1016/j.fuel.2014.10.024.
Stawski, D. (2009). The effect of sample weight in thermogravimetric analysis of low viscosity
polypropylene in air atmosphere. Polymer Testing, 28(2), 223–225. https://doi.org/10.1016/j.
polymertesting.2008.12.003.
Van Humbeeck, J. (1998). Simultaneous thermal analysis, Chap. 11. In M. E. Brown (Ed.), Hand-
book of thermal analysis and calorimetry (Vol. 1, pp. 497–508). Elsevier Science B.V. https://
doi.org/10.1016/S1573-4374(98)80014-9.
Vyazovkin, S., Burnham, A. K., Criado, J. M., Pérez-Maqueda, L. A., Popescu, C., & Sbirrazzuoli,
N. (2011). ICTAC kinetics committee recommendations for performing kinetic computations on
thermal analysis data. Thermochimica Acta, 520(1), 1–19. https://doi.org/10.1016/j.tca.2011.03.
034.
Xu, F., Wang, B., Yang, D., Ming, X., Jiang, Y., Hao, J., et al. (2018). TG-FTIR and Py-GC/MS
study on pyrolysis mechanism and products distribution of waste bicycle tire. Energy Conversion
and Management, 175, 288–297. https://doi.org/10.1016/j.enconman.2018.09.013.
102 7 Thermogravimetric Analysis (TGA)

Xu, Q., Ling, C., & Li, J. (2017). Microalgae decomposition in CO2 atmospheres by thermogravi-
metric analysis. Energy Procedia, 123, 381–386. https://doi.org/10.1016/j.egypro.2017.07.273.
Zimbardi, F. (2000). Evaluation of reaction order and activation energy of char combustion by
shift technique. Combustion Science and Technology, 156(1), 251–269. https://doi.org/10.1080/
00102200008947305.
Chapter 8
Chromatography

We consider a body to be in motion in case it changes its relative


position with respect to a group of bodies which is considered to
be at rest. But as all bodies, even the ones which appear in the
most perfect rest, may be in motion.
Free translation from Mecanique Celeste, Pierre Simon
Marquis de Laplace. 1799–1825.

8.1 Principles of Chromatography

The basic principles of chromatography are given here with a particular focus on gas
chromatography. Gas and liquid chromatography have many things in common, and
they are widely described in the literature (Rouessac and Rouessac 2007); however,
in this section the focus is given on the main parameters of the derived chromatogram
for pedagogical reasons. Gas chromatography is used for analytical purposes, and
most commonly refers to the concentration measurement of gas mixtures. The gas
to be analyzed is sent to a chromatographic column which is constituted by a sta-
tionary phase and a mobile phase. The substance behaves differently depending on
the different affinities of each substance in the mixture with the phases (Fig. 8.1).
The chromatographic instrument is simply described to be constituted by an oven
which should be thermostable. Within the oven, there is the chromatographic column.
The column is formed commonly by a thin glass capillary tube which can be some
meters long. In the tube, a thin layer of the fixed phase is fixed on the internal walls.
The substance forming the fixed layer on the tubes of the column should have a
certain degree of affinity with the gas mixture, and this is because in this column
we want to separate the single components before arriving to the sensor section.
There is a mobile carrier which is used to transport the mixture into the capillary
column. This gas usually is inert, and hydrogen, H2 ; nitrogen, N2 ; and helium, He
are used (Rouessac and Rouessac 2007; Skoog et al. 2006). The sample is inserted
through an injector. Figure 8.2 shows a simple drawing with the components of the
chromatographic column.
The basic functioning principle is described as follows: There is a certain affinity
between the components of the gas mixture to be analyzed and the stationary phase
© Springer Nature Switzerland AG 2019 103
C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_8
104 8 Chromatography

Fig. 8.1 Simple representation of the mains steps in chromatography

Fig. 8.2 Draft of the chromatographic column

of the column. Because of this, some elements will travel faster than others into the
column. The consequences will be that each of the components will be separated
among each other. Because the column will be long enough (otherwise there is some
serious problems with the design of the apparatus), each of the components will
across the final detector one at the time. In this way, only when all the molecules
of the faster element have crossed that section, the second element will arrive and
so forth the other elements. This is why, as we will see later, each column has a
particular resolution, and the value of the resolution has to be large enough. As an
example, a detector can convert difference in absorbance in diverse current signals
(mA) and plot them against time. This is the so-called chromatogram.
8.1 Principles of Chromatography 105

Concerning the different kinds of detectors used in chromatography, some of


them will be listed here. One of the most used, especially for gas chromatography,
is the thermal conductivity detector (TCD). This type of detectors is based on the
principle of heat conductivity, and the central part of the detector is a Wheatstone
bridge circuit. In this system, the sample gas which is coming from the column has a
different conductivity with respect to the carrier gas. Now, let us assume that we have
a hot filament where current is passing; the filament will be at diverse temperatures
if the gas flowing on it will be of different nature and this is because they have
different conductivity. If the temperature of the filament is higher, this will increase
or decrease the electrical resistance of the filament and therefore, the voltage will
be also different. The sensor gives a continuous comparison of the voltage produced
by the carrier gas and the voltage produced by the sample gas. This is because
two resistances are used in parallel. The concentration of the sample gas is directly
proportional to the voltage output.
The sensitivity of these sensors is directly proportional to the current flowing
into the filament, and it is increased by higher temperature differences between the
filament and the case which encloses the filament itself. However, it has to be noticed
that the temperature of the filament cannot be increased over certain limits to not
burn the wire itself and for this reason the sensitivity of this kind of detectors cannot
be of very high values. The sensitivity is inversely proportional to the flow of gas.
The flame ionization detector (FID) is the most used within gas chromatography;
the principle on the basis of this detector is that the sample is combusted in a flame
produced by hydrogen and air. During the combustion, in case we have molecules
where carbon is present, we have the formation of CH radicals, which in turn can
produce CHO+ ions as follows:

CH + O → CHO+ + e− (8.1)

Now, because a difference in voltage of some hundreds volt is applied between


two electrodes where the electrons are, a current is formed and this current can be
recorded (Stauffer et al. 2008b). The current produced is directly proportional to the
formyl cation produced. This method is used for organic compounds, but it is not
really suitable for compounds like O2 , CO2 , H2 O, and NH3 .
The refractive index detector (RI) is used for instance in high-pressure liquid
chromatography, HPLC, and they generate a signal which is based on the difference
between the RI of the carrier mobile phase and the RI of the sample mixed with the
mobile phase. The sample cell and the reference cell are separated; when the sample
cell contains only the carrier phase, the incoming beam is focused on a photodetector
(which can be a photo-resistor), and when the sample is present, the beam will be
refracted to an angle. The more the concentration of the sample compound, the
more the beam is refracted. A simple description of the functioning principles of RI
measurements is given in Fig. 8.3.
106 8 Chromatography

Fig. 8.3 Functioning principle of refractive index detectors

HPLC is used for studying a variety of liquid products, and it can be used to
analyze the composition of fatty acid methyl esters (FAME) within the production
of biodiesel (Syed 2017).
Ultraviolet detectors, also denominated as ultraviolet photoelectric detectors, are
also widely used in chromatography. Here, UV or visible radiation in the range of
190–800 nm of wavelength is used; the light is then absorbed by the sample within
the flow cell, and chromophoric groups are then detected by this method (Baiocchi
et al. 2002). A simple representation of the functioning principles of this method is
given in Fig. 8.4.
The UV light is directed through a filter and a slit to a first mirror from where the
light is sent to a diffraction grating which divides the light into different wavelengths
and frequencies. This is the key step in this process; the diffracted light is then sent
to a second mirror and divided into two beams; one is going to a reference diode,
and the other one is going to the flow cell with the sample. The light from the two
diodes will produce a signal which is recorded and compared at all times. In this
case, absorbance-time plots are obtained and peak areas are directly proportional to
the sample concentration.
There are mainly three types of UV detectors:

– fixed-wavelength filter-photometers;
– variable-wavelength spectrophotometers;
– photodiode-array detectors.

The positive aspects that have made chromatography used broadly can be summarized
as follows (Skoog et al. 2006):
8.1 Principles of Chromatography 107

Fig. 8.4 Basic principles of the ultraviolet detectors

• very short analysis times (one measure can take 15 min for instance);
• possibility of separating, working in suitable conditions any mixture of substances;
• possibility of carrying out analysis in series, since the same column can be regen-
erated continuously by the carrier gas;
• high sensitivity (amount of substance analyzed can be in the range of 10−5 –10−12
g).
Figure 8.5 shows the basic schematic of a gas chromatograph. The cylinder contains
the gases that constitute the mobile phase (this will be the carrier); this is purified by
the usage of some traps before being inserted into the column. The chromatographic
column containing the stationary phase is placed inside a chamber which is ther-
mostabilized appropriately in order to keep the various constituents in the gaseous
phase of the mixture to be separated (Guiochon and Guillemin 1988). The mixture of
gases to be analyzed is introduced into the column through the injector. This can be
injected in the gaseous phase if the mixture is a gas, or dissolved in a suitable solvent,
if the mixture to be analyzed is liquid or solid. To give one example, if the sample
gases are collected in a gas sampling bag with septum and valve options, then the gas
could be taken directly with a syringe and injected directly into the chromatographic
column.
108 8 Chromatography

Fig. 8.5 System adopted for chromatographic measurements

The parts shown in Fig. 8.5 are:

1. gas carrier tank;


2. injector block;
3. room columns;
4. chromatographic column;
5. electronic control system of the temperature;
6. detector;
7. pure gas tanks in service of the detector;
8. electronic processing of the detector signal;
9. recorder;
10. dedicated computer integrator or PC connected to the gas chromatograph, with
software for GC.

An elementary representation of the main steps involved during the measurement of


gas/liquid concentration in chromatography is shown in Fig. 8.6. The mechanisms
of chromatographic separation are based on adsorption, distribution, ion exchange,
exclusion, and affinity. Figure 8.6 shows the sequence of these steps:
The main steps shown in Fig. 8.6 are described as follows:
(a) the necessary parts of this measurement are column, SP—stationary phase,
MP—mobile phase, and sample;
(b) introduction of the sample;
(c) start of elution;
(d) recovery of the products following the separation.
8.2 The Chromatogram 109

Fig. 8.6 Main steps involved during the measurement of gas/liquid concentration in chromatogra-
phy. Adapted from Rouessac and Rouessac (2007)

8.2 The Chromatogram

As mentioned previously, at the end of the column is placed a detector which is able
to highlight the various substances emerging at different times. They emit a signal,
and the intensity of this signal is proportional to their concentration. The signal is
recorded by an instrument that will thus provide a chromatogram as shown in Fig. 8.7.

Fig. 8.7 Example of chromatogram with main parameters


110 8 Chromatography

The most important parameters in a chromatogram are as follows. In reference to


Fig. 8.7, some of them are shown as:
• w is the peak amplitude.
• w1/2 is the amplitude of half height (in h/2); in practice, it is the amplitude of the
curve where the height is half of the maximum.
• h is the peak height.
• t R is the retention time. This parameter is of outmost importance, and it is the
time needed for the sample to exit from the column. To be noticed that this is
approximately the point at which 50% of the total quantity has passed the detector.
In other words, this would be the value of the abscissa at the point where the
concentration is maximum. The retention time is dependent on the nature of the
component.
• t M is the so-called dead time. This is the time required for the carrier to across all
the column; the carrier is suppose to not spend time on the stationary phase and
therefore this is the shortest time a component will spend in the column.
• QC is the carrier flow; this is the flow rate of the carrier gas through the column.
• V R is the retention volume. This is the volume of the carrier gas used in the transport
of the sample from the starting point (when we inject the sample into the column)
to the exit of the column. This is calculated by:

VR = tR Q C (8.2)

• V M is the so-called dead volume. This is the volume of carrier gas that passes
through the column within the dead time.
• V S is the corrected retention volume, and this is, on the other hand, the actual
volume of carrier necessary for the transport of a certain component. The corrected
retention volume is calculated accordingly to:

VS = VR − VM (8.3)

• j is the compressibility factor. This value is related to the pressure of the gas inside
the chromatographic column. One simple way to assess the compressibility factor
of a real gas is to evaluate it from a simple relation like (Acrivos 1988):

PV
Z= (8.4)
n RT
And for ideal gases, the value of Z would be then equal to 1.
The net volume, V NET , is calculated according to Eq. (8.5). This is given by the
actual volume of carrier, which is first corrected with a suitable factor to give VM and
then multiplied by the compressibility of the gas in the column;
8.2 The Chromatogram 111

VNET = VM · j (8.5)

It is useful to examine in detail all the factors involved in the physics of the system.

8.3 Distribution Constants

During the chromatography process, there is equilibrium between the solute in the
stationary and mobile phases (Dettmer-Wilde and Engewald 2014):

Amobile ⇔ Astationary (8.6)

The equilibrium constant K C for the distribution of species A between the two phases
is called the distribution constant. This is defined as:

(a A )S
KC = (8.7)
(a A )M

where

– (a A )S is the activity of solute A in the stationary phase.


– (a A )M is the activity of solute A in the mobile phase.

Under some conditions which can be low concentrations or when, for instance, non-
ionic species are involved, the activity coefficients are almost equal to 1; then it is
possible to write:
nS
CS VS
KC = = nM (8.8)
CM VM

where

– C S is the molar analytical concentration of the solute in the stationary phase.


– C M is the molar analytical concentration of the solute in the mobile phase.
– nS is the number of moles of analyte in the stationary phase.
– nM is the number of moles of analyte in the mobile phase.

K C is constant over a large range of solute concentrations. Because of this, it is pos-


sible to assume that C S is directly proportional to C M . This type of chromatography
is called linear chromatography and provides the results as symmetric Gaussian-type
peaks and retention times independently of the quantity of analyte in input.
112 8 Chromatography

8.4 Retention Time

This quantity is a function of K C , and it is measurable. For a better understanding, it


is good to observe Fig. 8.7 where the small peak on the left represents a specie that is
not retained by the column. This is happening often since the sample to be analyzed
or the mobile phase contains a specie (or more) which is not retained. When one or
more species are not retained, this could help in the identification of the other species
present in the sample. The larger peak on the right is that of a component, and the
time required to reach the detector after the sample injection is called retention time,
t R , as seen previously. The specie has been retained because it spends a time t S in the
stationary phase. Then the retention time can be defined as (Rouessac and Rouessac
2007; Skoog et al. 2006):

tR = t S + t M (8.9)

In addition to the previous discussion, it is also possible to define the following


parameters:
• v: average linear rate of solute migration through the column (cm/s);

L
v= (8.10)
tR

• u: average linear velocity of the mobile-phase molecules (cm/s);

L
u= (8.11)
tM

where L is the length of the column packing.

8.5 Volumetric Flow Rate

In chromatography, the mobile-phase flow is usually characterized by the volume


flow rate F at the column outlet (cm3 /min). For an open tubular column, we have:

F = u 0 A = u 0 πr 2 (8.12)

where

– u0 , linear velocity at the column outlet;


– A, cross-sectional area of the tube.
8.5 Volumetric Flow Rate 113

When a stationary phase is present, for instance in the case of packed columns, the
volume of the entire conduct is not available to the gas flow and therefore, a correction
term must be introduced as follows:

F = u 0 πr 2 ε (8.13)

where ε is the column porosity, a fraction of total column volume available to the
liquid.

8.6 Retention Time and Distribution Constant

Consider the average linear rate of solute migration through the column, v, Eq. (8.10),
and the average linear velocity of the mobile-phase molecules, u, Eq. (8.11). These
two quantities are related as follows:
tM
v = uf = u (8.14)
tR

where f is the fraction of time the solute spends in the mobile phase.
This fraction is also equal to the average number of mole of solute in the mobile
phase divided by the total number of moles of solute in the column at any instant:

moles of solute in mobile phase


v=u· (8.15)
total moles of solute
Furthermore, we consider that:

moles of solute in mobile phase = C M V M


moles of solute in stationary phase = C S V S

And therefore:

CM VM 1
v=u =u (8.16)
CM VM + CS VS 1 + CCMS VVSM

Substituting in this equation, the equation of the distribution constants is obtained:

1
v=u (8.17)
1+ K C VS
VM
114 8 Chromatography

8.7 Retention Factor

The retention factor, k, is an experimental quantity widely used to compare the


migration rates of solutes in columns. This quantity is very important because k does
not depend on the geometry or on volumetric flow rate of the column. The retention
factor k A for a solute A is defined as:

K A VS
kA = (8.18)
VM

where K A is the distribution constant for the solute A view previously.


Substituting this equation with the one related to the rate of migration of a solute,
Eq. (8.17), the following expression is obtained:

1
v=u (8.19)
1 + kA

To demonstrate how k A can be derived from a chromatogram, it is shown here how


it is possible to apply some substitutions. Giving that:

L L 1
= (8.20)
tR tM 1 + k A

and rearranging, we have

t R − tM tS
kA = = (8.21)
tM tM

In Eq. (8.21), t S is the corrected retention time, which is the actual retention time of
a certain component in the column. Note that t R and t M are readily obtained from the
chromatography. In case we have values of k much less than unity, this means that
the solute passes through the column at a time close to the value of the dead time. On
the other hand, when the retention factor is greater than 20, the elution times become
excessively long. Then the ideal separation is carried out in conditions where the
retention factor solutes in a mixture are comprised between 1 and 10. Optimization
of the separation therefore occurs by optimizing the values of the retention factor for
the components of interest.

8.8 Parameters Characterizing a Chromatographic


Column

There are different parameters by which we are able to characterize a chromato-


graphic column. Among the most important, there are selectivity, efficiency, and
8.8 Parameters Characterizing a Chromatographic Column 115

resolution. The main references used in the following derivations are in this case
(Rouessac and Rouessac 2007; Skoog et al. 2006; Stauffer et al. 2008a).

8.8.1 Selectivity

The selectivity of a column expresses the capability of a certain column to give


well-separated peaks of components. It can be expressed in relation to the factors of
separation, for example, let us consider two generic components A and B, accord-
ing to previous mathematical relations existing between the various quantities, the
following parameter, α (the selectivity), can be obtained:

KB kB (tR ) B − t M
α= = = (8.22)
KA kA (tR ) A − t M

α should be as far as possible diverse from one. According to this definition, α is


always greater than unity. The last equality permits the determination of α from an
experimental chromatogram.

8.8.2 Efficiency

The following correlated terms are widely used as quantitative measures of chro-
matographic column efficiency:

– Plate height, H;
– Plate count or number of theoretical plates, N.

The two are related in this way:

L
N= (8.23)
H
where L is the length of the column packing (cm).
This efficiency increases if the number of plates, N, grows and if the plate height
H is lesser (Stauffer et al. 2008a). The efficiency of a column depends on the column
type and phases.
Considering a chromatographic column, there will be equilibrium between the
liquid phase and the vapor phase if we are taking into consideration liquids. On
the other hand, there will be also a certain equilibrium between the rate at which
molecules are in the mobile phase and on the fixed phase. To be noticed that when a
component passes the detector, it is derived by logical reasoning that the values of
its concentration with time will have a Gaussian-shaped curve.
116 8 Chromatography

The Gaussian curve is described by its standard deviation σ or its variance σ 2 . It


is convenient to define the efficiency of a column in terms of variance per unit length
of column. For this reason, by definition, the plate height H is given by:

σ2
H= (8.24)
L
Figure 8.8 shows a column having a packing L in length. In the same figure, there
is a curve showing the distribution of the number of molecules (or its concentration)
along the length of the column when the component is crossing the detector. This is
a Gaussian curve, and the position of L + σ and L − σ is given.
It is possible to notice that if σ is in cm2 , then H represents a linear distance
in cm. As a matter of fact, the plate height can be imagined as the length of the
column that contains a fraction of the analyte that lies between L and L − σ . Since
by statistical derivation (Montgomery 2003) the area under a normal curve which is
situated between ±σ is approximately 68% of the total area, the plate height contains
about 34% of the component analyzed.
The variance of the solute peak, τ 2 , expressed in seconds (s2 ) is distinguished
from σ 2 which is expressed in length units (cm2 ). The relation given between these
two standard deviations τ and σ is:

Fig. 8.8 Typical response as a function of distance


8.8 Parameters Characterizing a Chromatographic Column 117

σ σ
τ= = (8.25)
L
tR
v

Figure 8.9 also shows a method for approximating these two standard deviations in an
experimental plot. According to the procedure proposed by Skoog and collaborators
(2006), if we take the tangents at the inflection points on both sides of the peak in order
to form a triangle with the baseline of the chromatogram, then this area subtended by
the triangle is about 96% of the total area under the peak in case the peak is Gaussian
curve. Therefore, one can conclude that these intercepts are at approximately (±2τ )
from the maximum point of the curve, and it is also given that the amplitude of the
base of the triangle W is equal to 4τ .
Substituting w = 4τ into Eq. (8.25) and rearranging, it is possible to find:

LW
σ = (8.26)
4tR

Equation (8.24) will be then:

LW2
H= (8.27)
16tR2

From which:
 2
tR
N = 16 (8.28)
W

Fig. 8.9 Typical Gaussian curves for a component into a chromatographic column
118 8 Chromatography

It is possible also to use the empirical relations given previously to connect the height
equivalent to some basic parameters. For instance, the Van Deemter equation (van
Deemter et al. 1995) connects the quantity H with the flow QC of the carrier gas.

B
H = A+ + C QC (8.29)
QC

This is the equation of a hyperbola which is not equilateral. In Eq. (8.29), we have
that the term A is the micro-eddy diffusion coefficient. This coefficient is a function
of the particles diameters average in the filling. It is mentioned (Skoog et al. 2006)
that the coefficient tends to zero in the capillary columns in which the regime is not
swirling flows.
The term B is the molecular diffusion coefficient. This term is a function of
the diffusion coefficients of gases. It is minimized by using low temperatures, high
pressures, and gas high molecular weight.
The term C is the inertia ratio to the passage of molecules from the liquid phase to
the gas phase. This value is depending on the thickness of the stationary phase (for
instance, if the stationary phase is liquid) and its viscosity. It is minimized at high
temperature.
From Fig. 8.10, it can be seen that to minimize the height H of a theoretical plate,
it is necessary to use the optimal flow values between the values Q1 and Q2 .
The coefficients A, B, and C are characteristic of a certain column and are respon-
sible for enlargement of the slots and therefore for enlargement of the peaks.

Fig. 8.10 Relation between flow and column height


8.8 Parameters Characterizing a Chromatographic Column 119

A similar equation can be written also with regard to the temperature; there is also
an optimum temperature T that minimizes the value of H improving the efficiency
of a given column.

B
H = A+ + CT (8.30)
T

8.9 Resolution

The resolution represents one of the most important parameters given for a chro-
matographic column. As a matter of fact, each column behaves in a different way in
case the same mixture should be analyzed. Each column has a different resolution.
The resolution gives a certain value, and this value indicates how far two peaks are
within the chromatogram. Analytically, the resolution of each column is defined in
this way (Fig. 8.11):
 
Z 2Z 2 (tR ) B − (tR ) A
Rs = W A = = (8.31)
2
+ W2B WA + WB WA + WB

It is given that for a resolution of 1.5, the resolution value is the minimal to have a
good understanding of the concentrations for each of the components of the mixture.

Fig. 8.11 Main parameters and resolution of a chromatogram. Adapted from Skoog et al. (2006)
120 8 Chromatography

On the other hand, a resolution of 0.75 does not give a complete separation of A and
B. With a resolution of 1.0, the peak A contains approximately 4% of B and therefore
the peak B contains about 4% of A. Increasing the number of plates and increasing
also the length of the column give a better resolution of course.

8.10 Parameters Influencing the Resolution

There are mathematical relationships that link the resolution of a column and the
retention factors k A and k B for two solutes, the selectivity factor α, and the number
of plates N. Under the hypothesis that two solutes A and B have retention times close
enough between them, it is possible to write:

WA = WB ≈ W (8.32)

Rewriting Eq. (8.31) which is related to the column resolution, it is possible to find:

(tR ) B − (tR ) A
Rs = (8.33)
W
We rearrange this by substituting
 2
tR
N = 16 ; (8.34)
W

and we obtain:

(tR ) B − (tR ) A N
Rs = (8.35)
(tR ) B 4

In case we use the substitution:

tR − tM t
kA = = R; (8.36)
tM tM

the resolution is obtained as:



kB − kA N
Rs = (8.37)
1 + kB 4

In addition, we could also use the definition of selectivity factor:

kB
α= ; (8.38)
kA
8.10 Parameters Influencing the Resolution 121

and obtain
  √
α−1 kB N
Rs = (8.39)
α 1 + kB 4

In Eq. (8.39), k B is the retention factor related to the slower-moving species and α
represents the selectivity factor. The equation can be also rewritten in the following
form to give the number of plates needed to realize a given resolution:
 2  2
α 1 + kB
N= 16Rs2 (8.40)
α−1 kB

Since occasionally it is needed for certain species to divide among each other, it
would be necessary to have a model which takes into account this variable.
During chromatography, we want to achieve the highest possible resolution in a
short length of the column or residence time. Nevertheless, it is necessary to optimize
the two parameters since they go in opposite directions.
The element which is the slowest in the column represents the limiting step, and
this will determine the length of the column. It is possible to recall from what was
written previously:

L L 1
v= , N= , v=u (8.41)
(tR ) B H 1 + kA

From these equations, we have:

N H (1 + k B )
(tR ) B = ; (8.42)
u
where (t R )B is the time required to bring the peak for B component to the end of the
column when the velocity of the mobile phase is u.
Putting together the following two equations, Eqs. (8.43) and (8.44):

N H (1 + k B )
(tR ) B = ; (8.43)
u

 2  2
α 1 + kB
N = 16Rs2 ; (8.44)
α−1 kB

Equation (8.45) is obtained:


 2
16Rs2 α (1 + k B )3
(tR ) B = (8.45)
u α−1 (k B )2
122 8 Chromatography

Since H is a function of u and α is a function of k, such an equation is difficult to


solve and use except when doing comparisons and simplifications.

8.11 Qualitative Analysis

A qualitative analysis of chromatography is performed by studying different condi-


tions and physical parameters used. For instance, we would like to study the mixture
at hand by using different temperatures or different fixed phases. The quality of the
measurements will be a function of these parameters.
In addition to chromatography, there are different technologies providing a greater
quantity of information with highest level of accuracy compared to chromatography.
These could be IR, nuclear magnetic resonance, or mass spectrometry. Neverthe-
less, chromatography is a widely used technique for identify the presence or not of
components of mixtures.
Chromatography is an essential step in qualitative analyses and can be considered
the first step to do in order to have preliminary information on the sample and continue
with the further investigation. As a matter of fact, with complex samples it is difficult
to obtain the wanted information by using only one method.
As a summary, even if this technique is not the most accurate among the avail-
able methods, it can give very valuable information. For instance, we can see if a
compound is present in the sample or it is entirely absent from the mixture. Chro-
matography is a very good tool when we are analyzing gas products coming for
instance from gasification of biomass to produce syngas (De Blasio et al. 2019).

8.12 Elemental Analysis

The elemental composition of a sample in terms of, for instance, oxygen, nitro-
gen, hydrogen, and sulfur is usually measured by elemental analysis. The principle
on the basis of these types of analyses is that the sample is heated to very high
temperatures and, at these temperatures the sample is oxidized or undergoing to a
thermal degradation where known compounds are formed. By means of specifically
designed columns, the produced gas is further treated to remove excess compounds
end/or producing particular reactions.
Some types of elemental analyzers utilize very high temperatures where even the
molecular bonds of the molecules forming the sample are then broken. Therefore, a
gas consisting of only the elements is formed.
The temperatures used for transforming the sample into its elements can be of
thousands of degrees (for instance, 1500–3000 °C), and the power used to do this
can be of some kilowatts (for instance, 8–9 kW). The kinds of detectors used for
the analyses can be based on infrared detection (one example could be a solid-state
detector) and thermal conductivity (TCD). Recall that some of the detectors used in
8.12 Elemental Analysis 123

chromatography are similar. The analysis allows for detecting concentrations of few
ppm, and therefore, it is quite accurate. Figure 8.12 shows a schematic for the CHNS
and oxygen elemental analysis. This example refers to a Thermo Fisher Scientific
Flash series instrument.
The maximum weight of the sample is normally specified by the particular
machine used for the analysis; the sample is weighted accurately and inserted in
a suitable sampling device. The machine will automatically allocate the sample in
the space where it will be processed thermally. In the case the CHNS amount should
be derived, the sample is usually positioned in a tin capsule. After the sample is posi-
tioned in place, oxygen will be automatically pumped to allow for the combustion
of the sample; this is taking place at around 1800 °C in this example.
The products of combustion are sent to a copper oxide column where we have
mainly N2 , nitrogen oxides, CO2 , water, SO2 , and SO3 . Copper oxide is used here
to help the oxidation process. The excess of oxygen (the one which is reacting with
the carrier gas and the one which is oxidizing further the SO2 ) is removed into a
column where electrolytic copper is placed. In this section, we have the reduction of
nitrogen oxides to nitrogen and the excess oxygen is also trapped here (Gentile et al.
2013). In this place, we have mainly N2 , CO2 , SO2 , and water. These compounds
are then sent to the TCD detector through a CHNS column which is maintained at a
specific temperature depending on the machine used. The results will be shown in a
very similar way that was used for the gas chromatography.
In the case of oxygen analysis, the sample should be enclosed within a silver
capsule and introduced into the sampling unit. Also in this occasion, the machine will

Fig. 8.12 Schematics of a CHNS/O analyzer functioning


124 8 Chromatography

automatically send the sample to the chamber where it will be processed thermally at
high temperatures. Temperatures between 1000 and 1100 °C are used in this example
to convert the sample into gas products which are treated into a nickel-plated carbon
column. Here, mainly CO and N2 are formed at these temperatures (Berezkin 1983).
The components are sent to the detectors after passing an adsorption trap with soda
lime and anhydrone and an oxygen separation column which is maintained at 100 °C.
By absorbing the oxidized molecules, soda lime helps to reduce further the products
from the thermal processing and release all the oxygen to be measured (Weimann
et al. 1997).
In the case of elemental analysis, the calibration of the machine is done with
particular compounds of known composition and, like for the gas chromatography,
the machine will derive a diagram showing different peaks. Each peak will be related
to a different component, and the concentrations of C, H, N, S, and oxygen can be
then retrieved. Since the sample to be inserted into the analyzer is a small solid
quantity, the reference materials have to be also solid. In case liquids should be
tested, the liquid should be then absorbed into an inert powder and then inserted into
the sampling section. Examples of quantitative content determination in GC and EA
will be shown in Sect. 9.5.

References

Acrivos, J. (1988). Physical chemistry, third edition (Levine, Ira N.). Journal of Chemical Education,
65(12), A335. https://doi.org/10.1021/ed065pA335.3.
Baiocchi, C., Brussino, M. C., Pramauro, E., Prevot, A. B., Palmisano, L., & Marcı̀, G. (2002). Char-
acterization of methyl orange and its photocatalytic degradation products by HPLC/UV–VIS
diode array and atmospheric pressure ionization quadrupole ion trap mass spectrometry.
International Journal of Mass Spectrometry, 214(2), 247–256. https://doi.org/10.1016/S1387-
3806(01)00590-5.
Berezkin, V. G. (1983). Chemical methods in gas chromatography. Elsevier Science Publishers.
De Blasio, C., De Gisi, S., Molino, A., Simonetti, M., Santarelli, M., & Björklund-Sänkiaho, M.
(2019). Concerning operational aspects in supercritical water gasification of kraft black liquor.
Renewable Energy, 130, 891–901. https://doi.org/10.1016/j.renene.2018.07.004.
Dettmer-Wilde, K., & Engewald, W. (2014). Practical gas chromatography. A comprehensive ref-
erence. Springer.
Gentile, N., Rossi, M. J., Delémont, O., & Siegwolf, R. T. W. (2013). δ15 N measurement of
organic and inorganic substances by EA-IRMS: A speciation-dependent procedure. Analytical
and Bioanalytical Chemistry, 405(1), 159–176. https://doi.org/10.1007/s00216-012-6471-z.
Guiochon, G., & Guillemin, C. L. (Eds.). (1988). Methodology gas chromatographic instru-
mentation. Journal of Chromatography Library, 42, 319–391. https://doi.org/10.1016/S0301-
4770(08)70081-2.
Montgomery, D. C. (2003). Applied statistics and probability for engineers (3rd ed.). New York,
USA: Wiley.
Rouessac, F., & Rouessac, A. (2007). Chemical analysis: Modern instrumentation methods and
techniques (2nd ed.). Wiley.
Skoog, D., Holler, F. J., & Crouch, S. R. (2006). Principles of instrumental analysis (6th ed.).
Brooks Cole.
References 125

Stauffer, E., Dolan, J. A., & Newman, R. (2008a). Gas chromatography and gas chromatography—-
Mass spectrometry, Chap. 8. In E. Stauffer, J. A. Dolan, & R. Newman (Eds.), Fire debris analysis
(pp. 235–293). Burlington: Academic Press. https://doi.org/10.1016/B978-012663971-1.50012-
9.
Stauffer, E., Dolan, J. A., & Newman, R. (2008b). Detection of ignitable liquid residues at fire
scenes. In E. Stauffer, J. A. Dolan, & R. Newman (Eds.), Fire debris analysis (pp. 131–161).
Burlington: Academic Press. https://doi.org/10.1016/B978-012663971-1.50009-9.
Syed, M. B. (2017). Analysis of biodiesel by high performance liquid chromatography using refrac-
tive index detector. MethodsX, 4, 256–259. https://doi.org/10.1016/j.mex.2017.07.002.
van Deemter, J. J., Zuiderweg, F. J., & Klinkenberg, A. (1995). Longitudinal diffusion and resistance
to mass transfer as causes of nonideality in chromatography. Chemical Engineering Science,
50(24), 3869–3882. https://doi.org/10.1016/0009-2509(96)81813-6.
Weimann, J., Hagenah, J. U., & Motsch, J. (1997). Reduction in nitrogen dioxide concentration by
soda lime preparations during simulated nitric oxide inhalation. British Journal of Anaesthesia,
79(5), 641–644. https://doi.org/10.1093/bja/79.5.641.
Chapter 9
Examples of Quantitative Content
Determination in Chromatography
and Elemental Analysis

9.1 Instrumentation

The quantitative analysis of results obtained from chromatography refers to the eval-
uation of the actual concentration of each component of the mixture. The first step
is the evaluation of the peak area, which usually is given by the instrument used. An
example of chromatography analyses done on gas products derived from supercritical
water gasification of Biomass (De Blasio et al. 2015) is given here.
The instrument used in this case is the gas chromatograph (GC) Clarus 500
(Fig. 9.1). The detector utilized in this case is a thermal conductivity detector (TCD).
The gas chromatograph returns as output a plot, which consists of several peaks
(Fig. 9.2). This is what we have seen previously to be the chromatogram.
Each peak corresponds to a compound among those measurable by the machine.
More specifically, each of them has a different retention time which makes it rec-
ognizable to the operator. The area of these peaks, through certain coefficients, is
proportional to the volumetric concentration of the element in the sample.
For a correct analysis, it is useful to calculate these coefficients related to the
measure of each compound through the preliminary study of a known substance.
This is known as the calibration of the machine. After calibrating the instrument, it
will be possible to evaluate the composition of a sample whose composition is not
known.
In few words, since we know that a compound is detected at a particular residence
time and that the concentration of that compound is directly proportional to the area
under the curve, then, by the utilization of a suitable coefficient of proportionality
(coefficient or function), it will be possible to estimate accurately the concentration
of that component in the sample tested.

© Springer Nature Switzerland AG 2019 127


C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_9
128 9 Examples of Quantitative Content Determination …

Fig. 9.1 A photo of the gas chromatography instrumentation

Fig. 9.2 Example of output of the chromatographic analysis. As shown by the instrument
9.2 Preparation and Introduction of the Sample 129

9.2 Preparation and Introduction of the Sample

The samples to be introduced into the chromatograph are obtained from the collection
of the products of gasification or, in the case of calibration, from bottles of known
composition. In the first case, this operation is done by using the gas bags provided
with a membrane that allows for the injection of a needle without loss of gas (gas
sampling bags with valve and septum). In the second case, it is possible to release
the gas directly into a syringe. This is visible from Fig. 9.3.
Figure 9.4 shows the injection of the gas into the gas chromatograph. When the
sample is inserted into the instrument, a small amount of air could be aspirated
too or for instance, some air could have been still present into the chromatograph
chamber where the sample gas is momentarily stored during the sampling step before

Fig. 9.3 Collection of samples of known composition

Fig. 9.4 Injection of the gas


into the gas chromatograph
130 9 Examples of Quantitative Content Determination …

measuring. For this reason, in the analysis of the final composition of the products,
the oxygen and the nitrogen forming the air can be usually subtracted.

9.3 Calculation of the Calibration Coefficients

To calculate the calibration constants (the coefficient of proportionality), diverse sub-


stances with known composition can be taken into consideration. For this particular
example, these substances are summarized in Table 9.1.
The change in the volume of sample introduced should not cause any change in
the area of the peaks of the individual compounds.
For the calculation of the coefficients, it is necessary to divide the volume fraction
of the individual compound by the peak area resulting from the chromatogram related
to the same compound. This is done during the calibration process where the volume
fractions are known.

Volume fractioncompound
K compound = (9.1)
Peak areacompound

To be noticed that this is a simplification of the procedure, since not necessarily


the volume areas and the known volumetric fractions have a linear dependence. The
accurate procedure would be to measure the areas for several different known com-
positions and then determine the proportionality coefficient which can be a number
of a function.

Table 9.1 Substances with Bottle 1 H2 100% Mol


known composition used
Bottle 2 CO 5% Mol
Air BAL
Bottle 3 CH4 60% Mol
CO2 BAL
Bottle 4 H2 10% Mol
N2 BAL
Bottle 5 H2 S 5000 ppm
N2 BAL
Bottle 6 C2 H4 100%
Bottle 7 CO2 1%
C2 H4 0.099%
C2 H6 0.972%
CH4 1.02%
He BAL
BAL stands for balanced
9.3 Calculation of the Calibration Coefficients 131

9.3.1 Calibration Constant for Hydrogen

The procedures which follow take into account the given Eq. (9.1). To find the correct
calibration constant for hydrogen gas, sampling bottles with different compositions
have been used in this example. The data and the results are visible in Table 9.2.
For the sensitivity of the machine, diluted samples are more reliable compared to
those concentrated (bottle 4 compared to 1). This is also visible from the asymmetry
of the Gaussian curve relative to the concentration 100% H2 . Indeed, the Gaussian
of the sample 4 shows a trend very symmetrical and therefore more precise.

9.3.2 Calibration Constant for Carbon Monoxide

To find the correct calibration constant of carbon monoxide, also in this case, a
calibration bottle has been used. The data and the results are visible in Table 9.3.

9.3.3 Calibration Constant for Methane

In this case, the data and the results are visible in Table 9.4.

Table 9.2 Calculation of the calibration constant for hydrogen


Sample volume (mL) Volume fraction (%) Peak area K
Bottle 1
Test 1 11 100 4,020,820.16 2.49 × 10−5
Test 2 10 100 4,057,932.57 2.46 × 10−5
Test 3 10 100 4,232,846.05 2.36 × 10−5
Bottle 4
Test 1 10 10 422,134.15 2.36 × 10−5
Test 2 10 10 441,125.63 2.27 × 10−5
Test 3 9 10 443,963.95 2.25 × 10−5

Table 9.3 Calculation of the calibration constant of carbon monoxide


Bottle 2 Sample volume (mL) Volume fraction (%) Peak area K
Test 1 15 5 422,441.18 1.18 × 10−5
Test 2 10 5 419,887.24 1.19 × 10−5
Test 3 10 5 416,974.83 1.20 × 10−5
132 9 Examples of Quantitative Content Determination …

Table 9.4 Calculation of the calibration constant of methane


Bottle 3 Sample volume (mL) Volume fraction (%) Peak area K
Test 1 10 60 3,941,285.96 1.52 × 10−5
Test 2 10 60 3,922,941.64 1.53 × 10−5
Test 3 15 60 3,911,516.61 1.53 × 10−5

9.3.4 Calibration Constant for Carbon Dioxide

The data and the results for CO2 are visible in Table 9.5.

9.3.5 Calibration Constant for Hydrogen Sulfide

For H2 S, the data and the results are visible in Table 9.6.

9.3.6 Calibration Constant for Nitrogen

For nitrogen, data and results are visible in Table 9.7.

Table 9.5 Calculation of the calibration constant of carbon dioxide


Sample volume (mL) Volume fraction (%) Peak area K
Bottle 3
Test 1 10 40 3,443,657.61 1.16 × 10−5
Test 2 10 40 3,412,977.81 1.17 × 10−5
Test 3 15 40 3,398,244.43 1.18 × 10−5
Bottle 7
Test 1 20 1 99,606.84 1.00 × 10−5
Test 2 11 1 97,003.13 1.03 × 10−5
Test 3 9 1 96,241.89 1.04 × 10−5

Table 9.6 Calculation of the calibration constant of hydrogen sulfide


Bottle 5 Sample volume (mL) Volume fraction (%) Peak area K
Test 1 13 0.5 32,427.53 1.54 × 10−5
Test 2 10 0.5 26,015.42 1.92 × 10−5
Test 3 10 0.5 25,667.64 1.95 × 10−5
9.3 Calculation of the Calibration Coefficients 133

Table 9.7 Calculation of the calibration constant of nitrogen


Sample volume (mL) Volume fraction (%) Peak area K
Bottle 4
Test 1 10 90 7,529,354.09 1.20 × 10−5
Test 2 10 90 7,469,558.98 1.21 × 10−5
Test 3 9 90 7,429,618.32 1.21 × 10−5
Bottle 5
Test 1 13 99.5 8,339,365.39 1.19 × 10−5
Test 2 10 99.5 8,342,276.79 1.19 × 10−5
Test 3 10 99.5 8,337,046.56 1.19 × 10−5

9.3.7 Calibration Constant for Oxygen

For oxygen, the data and the results are visible in Table 9.8.

9.3.8 Calibration Constant for Ethylene

For ethylene, data and results are shown in Table 9.9.

9.3.9 Calibration Constant for Ethane

For ethane, data and results are shown in Table 9.10.

Table 9.8 Calculation of the calibration constant of oxygen


Bottle 2 Sample volume (mL) Volume fraction (%) Peak area K
Test 1 10 19.95 1,597,280.74 1.25 × 10−5
Test 2 10 19.95 1,597,499.02 1.25 × 10−5
Test 3 15 19.95 1,598,675.99 1.25 × 10−5

Table 9.9 Calculation of the calibration constant of ethylene


Bottle 6 Sample volume (mL) Volume fraction (%) Peak area K
Test 1 10 100 9,871,250.99 1.01 × 10−5
Test 2 10 100 9,702,196.38 1.03 × 10−5
Test 3 20 100 9,600,094.28 1.04 × 10−5
134 9 Examples of Quantitative Content Determination …

Table 9.10 Calculation of the calibration constant of ethane


Bottle 7 Sample volume (mL) Volume fraction (%) Peak area K
Test 1 20 0.972 105,370.12 9.23 × 10−6
Test 2 11 0.972 102,544.15 9.48 × 10−6
Test 3 9 0.972 101,783.81 9.55 × 10−6

Table 9.11 Summary of the Retention time K


results
H2 0.73–0.96 2.29 × 10−5
CO 10.38–10.40 1.19 × 10−5
CH4 9.28 1.53 × 10−5
CO2 1.83–2.08 1.10 × 10−5
H2 S 6.64–6.69 1.80 × 10−5
N2 8.17–8.19 1.20 × 10−5
O2 7.82–7.83 1.25 × 10−5
C2 H4 2.06–2.08 1.03 × 10−5
C2 H6 3.30 9.42 × 10−6

9.3.10 Summary of the Results

Table 9.11 shows a summary of the results obtained for each compound and also the
typical retention times of each of them; the retention time allows to recognize the
compound separated when the substances have of unknown composition.

9.4 An Example of Results from Elemental Analysis

9.4.1 Instrumentation

In order to assess the composition of certain liquid and solid samples, feeds and
products, it is necessary an elemental analysis; for this purpose, a typical instrument
used is the elemental analyzer (EA). An example will be given here (like for the
case of GC). In the following text, a series of laboratory results will be given. The
EA used is shown in Fig. 9.5. In particular, this instrument is a Flash EA1100 from
Thermo Quest.
The EA returns as output a plot consists of several peaks (Fig. 9.6).
Each peak corresponds to an element among those measurable by the machine,
more specifically according to the following order N–C–H–S. The peak areas,
through certain coefficients, are proportional to the concentration of the element
in the sample.
9.4 An Example of Results from Elemental Analysis 135

Fig. 9.5 Elemental analyzer

Fig. 9.6 Example of output of the elemental analyzer


136 9 Examples of Quantitative Content Determination …

Fig. 9.7 Reference


substances

Also in this case, it is necessary to calibrate the machine through the preliminary
study of known substances. After this procedure, it will be possible to evaluate the
composition of a sample whose composition is not known.
In the same way as for the GC, in this case, a proportionality exists between
the weight of the element in the sample and its peak area. Here the proportionality
coefficient is evaluated by dividing the weight of the individual elements by the area
shown on the output plot. Also the area obviously is related to the element considered.

Weightelement
K element = (9.2)
Peak areaelement

The analysis processes are similar to the case of GC, while only the final calcula-
tion changes; it is pointed out that in this case, we assume that a linear proportionality
exists between the weight of the element and its measured area. However, this is an
approximation since the proportionality coefficient could be also represented by a
suitable function.

9.4.2 Calculation of the Calibration Coefficients

To calculate the calibration constants, two substances with known composition have
been taken into consideration. These substances are l-cystine and dl-methionine.
Figure 9.7 shows the two reference samples.
Table 9.12 shows the weight percent of the elements in the two reference sub-
stances.
The following are examples of real measures and the determination of calibration
coefficients for the reference substances. The tests are normally repeated to have
averaged values and more accurate evaluation of the coefficients. Measures are as
follows:
9.4 An Example of Results from Elemental Analysis 137

Table 9.12 Elemental l-Cystine(w%) dl-Methionine (w%)


composition of the reference
substances C 29.99 40.25
H 5.03 7.43
N 11.66 9.39
S 26.69 21.49
O 26.63 21.45

Table 9.13 Calibration Weight (mg) Area (mVs) K


coefficients
C 1.889 7,979,878 2.367 × 10−7
H 0.317 4,034,818 7.857 × 10−8
N 0.735 1,395,280 5.268 × 10−7
S 1.682 3,193,019 5.268 × 10−7

Test n. 1: 6.30 mg of l-cystine (Table 9.13).


Test n. 2: 5.98 mg of l-cystine (Table 9.14).
Test n. 3: 5.24 mg of l-cystine (Table 9.15).
Test n. 4: 6.20 mg of dl-methionine (Table 9.16).
Test n. 5: 4.86 mg of dl-methionine (Table 9.17).

9.4.3 Summary of the Results

The calibration coefficients can be derived from tests performed by averaging the
results. The results obtained are reported in Table 9.18.

Table 9.14 Calibration Weight (mg) Area (mVs) K


coefficients
C 1.793 7,609,669 2.356 × 10−7
H 0.301 3,892,033 7.734 × 10−8
N 0.697 1,277,633 5.455 × 10−7
S 1.596 3,040,440 5.249 × 10−7

Table 9.15 Calibration Weight (mg) Area (mVs) K


coefficients
C 1.572 6,745,124 2.331 × 10−7
H 0.264 3,404,930 7.754 × 10−8
N 0.611 1,091,720 5.597 × 10−7
S 1.399 2,676,692 5.227 × 10−7
138 9 Examples of Quantitative Content Determination …

Table 9.16 Calibration Weight (mg) Area (mVs) K


coefficients
C 2.496 10,669,690 2.339 × 10−7
H 0.461 6,102,209 7.555 × 10−8
N 0.582 1,050,285 5.541 × 10−7
S 1.332 2,583,730 5.155 × 10−7

Table 9.17 Calibration Weight (mg) Area (mVs) K


coefficients
C 1.956 8,390,162 2.331 × 10−7
H 0.361 4,756,623 7.589 × 10−8
N 0.456 808,920 5.637 × 10−7
S 1.044 2,060,116 5.068 × 10−7

Table 9.18 Summary of the results


Test 1 Test 2 Test 3 Test 4 Test 5 Average
Cystine Cystine Cystine Methionine Methionine results
KC 2.37 × 2.36 × 2.33 × 2.34 × 2.33 × 2.35 ×
10−7 10−7 10−7 10−7 10−7 10−7
KH 7.86 × 7.73 × 7.75 × 7.56 × 7.59 × 7.70 ×
10−8 10−8 10−8 10−8 10−8 10−8
KN 5.27 × 5.46 × 5.60 × 5.54 × 5.64 × 5.50 ×
10−7 10−7 10−7 10−7 10−7 10−7
KS 5.27 × 5.25 × 5.23 × 5.16 × 5.07 × 5.19 ×
10−7 10−7 10−7 10−7 10−7 10−7

It is also important to notice that the element with the most variation is hydrogen.
This is easily understood from the fact that the elongated shape of its peak is more
affected by the error compared, for example, with the uniform peak of carbon.
In addition, the values obtained are very good approximations because the results
of the various tests are not very different and a slight variation of them does not cause
a significant error in the measurement of the concentration of the elements.

9.5 Analysis of Unknown Substances

GC
Once the calibration constants of the GC are obtained, the volume concentration of
a compound in the sample is calculated by:

Volume fractioncompound = Kcompound · Peak areacompound (9.3)


9.5 Analysis of Unknown Substances 139

EA
The weight of the elements of a sample of unknown composition is easily achieved
with the help of the EA. In this case, it is necessary to multiply the peak area of the
element for its constant (the proportionality coefficients obtained previously during
the calibration).

Weightelement = Kelement · Peak areaelement (9.4)

For the analysis of liquids, it is a good practice to use a fixed inert and perform
the analysis in the same way previously viewed. The inert material used in this study
is Chromosorb (reagent for elemental analysis).

Reference

De Blasio, C., Lucca, G., Özdenkci, K., Mulas, M., Lundqvist, K., Koskinen, J., et al. (2015). A
study on supercritical water gasification of black liquor conducted in stainless steel and nickel-
chromium-molybdenum reactors. Journal of Chemical Technology and Biotechnology, 91(10),
2664–2678. https://doi.org/10.1002/jctb.4871.
Chapter 10
Some Considerations and Statistical
Derivations for the Concentration Profile
and Gaussian Curve

The time to be lost is never lost.


A suggestion which is almost always valid.

In the previous discussion on the principles of chromatography, it was mentioned


that it is convenient to express the chromatograms’ curves with Gaussian functions.
It is obvious that this assumption is reasonable and it reflects very well the reality.
In fact, in case one element is flowing into a conduct after being injected in it, it
will spread the concentration along with the flowing direction. This is because some
molecules of that “packet” are slower and some others are faster. The spread will be
in some way related to the diffusivity of that element in the particular carrier.
A Gaussian function is expressed as follows:
 
1 (x − μ)2
y= √ · exp − (10.1)
σ 2π 2σ 2

where μ is denominated as the variance or the expected value (or mean value) of
the function, while σ is the standard deviation of the function and σ 2 is its variance.
Equation (10.1) is a probability density function which is related to the measures in
experiments and it is really much applied in science and engineering. Therefore, in
the following discussion, we will demonstrate some of the features of this curve.
The Gaussian function is characterized by a symmetrical curve. For the stan-
dard normal distribution (if we consider the axes origin placed in the middle of the
curve, the maximum is at x = 0, y = 03999), the curve has two inflection points
at x = ±1, and for these points, the ordinate value is 0.242 (this is 60.6% of the
maximum value). We know from literature that the mean of this probability density
function should be μ (Montgomery 2003).

© Springer Nature Switzerland AG 2019 141


C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_10
142 10 Some Considerations and Statistical Derivations …

It will be demonstrated here that the mathematical derivation of the mean value
of a function (integral value) will give μ. The mean value of the function y = f (x)
is:

∞
x f (x)dx = μ (10.2)
−∞

In the resolution of Eq. (10.2), we solve for the indefinite integral and then we
consider the boundaries. We substitute:

x − μ = t, → x = t + μ and dx = dt (10.3)

for this reason, there is no change in dx, and dx is just substituted with dt. After this,
we substitute:

(x − μ)2 t2 
2
= y2 → 2
= y 2 → t 2 = y 2 · 2σ 2 → t = y 2 · 2σ 2 (10.4)
2σ 2σ
In Eq. (10.4), the arrow → means “which gives”. From Eq. (10.4) we have
√ √
t= 2 · σ → ydt = 2 · σ · dy (10.5)

For this reason, substituting in our probability density function Eq. (10.1) we have:

1   √
f (x)dx = √ · exp −y 2 · 2 · σ · dy (10.6)
σ 2π

From:
 √
(x − μ)2 = y 2 · 2σ 2 → x = y 2 · 2σ 2 + μ → x = 2 · σ · y + μ; (10.7)

and therefore:
 √ 1  
x f (x)dx = 2 · σ · y + μ · √ · exp −y 2 dy (10.8)
π

We remember the method of integration per parts:

D[ f (x) · g(x)] = f  (x) · g(x) + g  (x) · f (x)



⇒ D[ f (x) · g(x)]
 
= f  (x) · g(x) + g  (x) · f (x) (10.9)
10 Some Considerations and Statistical Derivations … 143

where f (x) and g(x) are two diverse functions of the variable x. And as a conse-
quence:
 
f  (x) · g(x) = [ f (x) · g(x)] − g  (x) · f (x) (10.10)

Integrating per parts Eq. (10.8):


 √
1  
2 · σ · y + μ · √ · exp −y 2 dy
π
√ √  √ √
π π 2· σ
=√ · 2· σ · y+μ − · √ dy
π·2 2 π
(10.11)

and finally we obtain:


√ √
2 μ 2 μ
· σ· y+ − ·σ· y= (10.12)
2 2 2 2
and this result has to be multiplied by 2 because we have integrated only between
zero and infinite. In this way, Eq. (10.2) is proven.

Reference

Montgomery, D. C. (2003). Applied statistics and probability for engineers (3rd ed.). New York,
USA: Wiley.
Part II
Notions of Biochemical Reactors
Design and Biofuels Production
Chapter 11
Introduction to Part 2

The most terrible of misfortunes, the death, means nothing for


us. Since when we are living, there is no death and when the
death is there, we don’t exist anymore.
Epicurus, Letter to Menoeceus, Third century B.C.

Liquid fuels are particularly suitable for our needs. They contain large amount of
energy in terms of calorific value, they are easily transported and most of all, and the
power delivered by this kind of fuels is constantly the same, no matter the amount of
the fuel present in the tank of your car. This is a quite important aspect of liquid fuels
which is not usually mentioned. If we think about batteries, indeed there are also
positive aspects with their usage, and however, it is also to be realized that they are
heavy and the weight of the batteries is carried around at all times no matter if there
is no power anymore left in them. Additionally, the power delivered by fully charged
batteries is different from when the battery is almost empty. As a matter of fact, the
differential voltage is indeed correlated to the discharge of the battery (Kato et al.
2018). This can be soon realized after buying a simple electric scooter and using it
a bit.
Liquid fuels can be produced by using biomass feedstock, and therefore, this
would make them renewable energy sources. The kinds of feedstock that can be used
are also different; for instance, algae can be utilized for producing biodiesel (Khan
et al. 2017), and also cellulose can be utilized to produce ethanol (Singh et al. 2018).
In order to reduce oil dependency, one of the strategies proposed usually is to
utilize biomass for the production of fuels and also for the production of diverse
chemicals and materials (Klass 1998). To support this sentence, we see from the
following graph (Fig. 11.1) that the use of domestic wood can be increased sustainably
in the future.
An industrial plant where operations are done on biomass to produce energy, fuels,
chemicals, or materials is usually defined as a biorefinery. The goal in a biorefinery
is that all the fractions and sidestreams are efficiently converted into products and
that these products could be reused and sold on the market; this is directly related to

© Springer Nature Switzerland AG 2019 147


C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_11
148 11 Introduction to Part 2

Fig. 11.1 Forest growth plotted with the sustainable harvesting potential and industrial use of
wood. Metsä Group. With permission, the figure is based on Natural Resource Institute Finland
2015 (Luke)

the sustainability of the activities carried on in a biorefinery, and this is the basis for
the development of a bio-economy (Palmeros Parada et al. 2017). In this way, the
waste is minimized, and the production should be more sustainable. The definition
given by the International Energy Agency, IEA, for a biorefinery is: “Biorefinery
is the sustainable processing of biomass into a spectrum of marketable products
and energy” (Liguori and Faraco 2016). Or, as stated by Valmet “Biorefining means
creating more value from biomass” (Valmet 2014). Biorefineries can be classified,
for example, based on the used feedstock or technology and this kind of processes are
utilizing biomass in a more sustainable way and also in a more productive way. As
a matter of fact, three aspects should be considered when biorefinery is considered
(Pandey et al. 2015). The first point is that the process should be economically
attractive, and this means that all the products have a market, and they can be sold
at competitive price. The profitability of the process can be improved by increasing
the efficiency of the biomass utilization. The biorefinery gives also less pollution and
less emission, and therefore, this is the meaning for sustainability. In the literature,
novel biorefinery processes are given (Özdenkçi et al. 2017), and possible options
could be also represented by the integration of technologies like supercritical water
gasification of biomass and hydrothermal liquefaction (De Blasio et al. 2015) even
if some kinds of feedstock would give problems when this kind of technologies are
adopted (De Blasio et al. 2019).
Biofuels can be produced from diverse kinds of feedstock, and this would be the
basis for the classification of biofuels in first generation, second generation, and third
generation. For instance, the first-generation biofuels are produced from vegetable
oil, sugars, starch, or animal fats; the second generation is produced from feedstock
which is not related to food crops, while the third-generation biofuels are those
produced from algae.
Concerning the first-generation biofuels, there are concerns that this kind of bio-
fuels’ production is not even ethical since producing crops not for food purposes but
11 Introduction to Part 2 149

Fig. 11.2 Undernourished population in the world by region. Data from Statista (2018a)

for driving cars is not the best idea thinking that in the world there are millions of
children actually dying because of food scarcity.
In fact, the children dying every year for this reason are millions, and they represent
almost 50% of all the children’s deaths (UNICEF, World Health Organization, and
The World Bank 2018).
Because of the food we do not actually need (the one in excess), we are actually
occupying precious land which could be utilized for other purposes. This is because
we consume extra energy in our activities, for instance, walking.
A simple example will be given here. In the literature, it is possible to find esti-
mations of the consumption of energy of a human being at running or walking, but
many of this data are not congruent, and therefore, in this case some assumptions
will be done.
Let us assume here that the consumption of energy at walking is 0.456 kcal for
each kilogram of carried extra weight and for each kilometer of walking distance.
We assume also that a person normally walks only 3 km per day. Another assumption
is that in a very nice season the production of potatoes could be 22.7 kg for square
meter, and this is very high over-estimation since this normally does not happen.
Now, in the world there are 7.53 billion people (Statista 2018b), and unfortunately,
796 million people are undernourished as can be derived from Fig. 11.2.
Subtracting the undernourished population from the total population, we assume
here that half of the remaining fraction has at least 3 kg of overweight, and this would
be 3.37 billion people. For the energy content of potatoes, it was chosen here the
value of 0.81 kcal/kg. Also in this case, there are many different values given in the
literature. The potato peel, for instance, has been reported to have energy content
very similar to coal (Uzun et al. 2017). In short, it would be needed half hectare just
to provide the energy required to transport some extra weight.
It is clear that the utilization of crops for replacing oil decreases the land area
available for food production. Therefore, this would be a non-sustainable solution.
150 11 Introduction to Part 2

Fig. 11.3 Carbon cycles and the production of fossil fuels

The usage of lignin and residues from industrial processes have been considered to
be much more positive for biofuels production (Liguori and Faraco 2016).
The environmental issues play a major role in the production of biofuels. Consid-
ering forest harvesting, the rate of harvesting should not exceed the annual growth in
the long term. On the other hand, water resources are also very important to preserve,
and the local ecosystems should not be compromised. Furthermore, the production
concept should be socially acceptable, meaning that the social acceptability should
also be considered here. A commonly accepted criterion is that the utilization and
production of the raw materials should not compete with food production, as stated
previously. And this is one reason why the first-generation biofuel production should
not be significantly increased.
Since the production of biofuels involves the transformation of organic molecules
(mainly the ones containing carbon), it is beneficial to remember that carbon is
present in many forms in nature and it can be subject to cycles. We remember from
the first part of this book that the carbon cycle, and especially the equilibrium between
CO2 in the atmosphere, the oceans, and the carbonate rocks, plays a very important
role since a shift in this equilibrium can end up with disastrous consequences. In
Fig. 11.3, a very simple schematics is depicted with respect to the carbon cycles and
the production of fossil fuels, while for a more detailed information on the water–CO2
equilibria, the author invites the reader to consult the introductory section of this book.
From these equilibria, it could be even argued that oil and coal could be considered as
“renewable,” as a matter of fact it takes “only” millions of years for the organic matter
to deposit under the ground and then, by a series of geological events, starting with
the diagenesis, reach the proper conditions to form oil. This is called catagenesis,
and this process is not fast either.
11 Introduction to Part 2 151

Even rocks are formed by enormous quantities of skeletal shells; in fact, the
limestone minerals can be divided into four groups: non-skeletal grains, skeletal
grains, micrite, and cement, while, on the other hand, the diagenesis of carbonates
involves the processes of cementation, microbial, recrystallization, micritization,
neomorphism, dissolution, compaction, and dolomitization (De Blasio 2010). For
this reason, limestone can be under the form of calcite, and it can be also in the form
of dolomite rocks, CaMg(CO3 )2 .
To give some examples of what kind of organisms are forming limestone rocks,
we can have radiolarians, diatoms, sponges, and, also in this case, algae. Among
the geochemical processes, we have that these processes have importance since the
degree of compaction and crystallization of minerals affects the limestone dissolution
itself (De Blasio et al. 2018).
Coming back to our catagenesis and the formation of crude oil, if we take into
account angiosperms plants living in the Paleogene and Neogene periods (this means
from 65 million years to around 3 million years ago), then it would take all this time
for this organic deposits to form kaurane (Kashirtsev et al. 2012).
Living cells, or enzymes, will become more common in the future for the purpose
of biofuels production. Just to give one example, biofuels can be produced by the
fermentation of soybeans (Loman and Ju 2016). Therefore, the fermentation process
will become more and more important in the coming years, and that is why the
biochemical engineering concepts should be carefully understood by future engineers
working in the field of biorefineries. From this need comes one of the goals of
this manuscript: the understanding of biological processes in power production and
biochemical reaction engineering.
There is a series of companies around the world where biomass is converted and
biofuels are produced or converted. Finnish forest and oil companies have devel-
oped novel concepts to refine biomass into bioproducts and biofuels. Some of these
concepts are Neste’s biodiesel and St1’s bio-ethanol production concepts. The raw
materials that have been mainly utilized in biorefineries include maize, soybean,
sugarcane, i.e., the crops which are cultivated on arable land (Cherubini 2010).
Biorefineries can also produce chemicals, biomaterials, and biofuels. In addition,
in biorefineries there is the possibility to produce power, heat, or steam as a side
product. Considering the actual economical situation and the fact that resources are
depleting, the marked for bio-based fuels and chemical will increase in the coming
years.
Today, there is more and more regulations on environmental issues and incen-
tives to substitute oil with biomass-based fuels and alternative fuels. Therefore,
process integration has become the main drivers for companies to develop sustain-
able production processes. Especially in the case of biofuels and power, companies
are encouraged to deliver eco-friendly solutions by governmental financial support.
Furthermore, there is more acceptance from the public toward companies that are
utilizing bio-based feedstock. Companies can get advantage when investing in renew-
able bio-based products and chemicals. One example is the collaboration between
the companies Lego and Ikea who have declared the will to replace products with
bio-based alternatives (Neste Corporation 2016).
152 11 Introduction to Part 2

Biodiesel is produced in large quantities by diverse companies, and one example


is the Finnish company Neste. The feedstock in this case is vegetable oil or waste oil.
Concerning the utilization of materials containing high amount of lignin, enzymes
are used in large quantities during the fermentation process for the production of bio-
ethanol (Kubicek and Kubicek 2016). Techniques like the CRISPR–cas9 gene editing
system developed by the University of California (Schwartz et al. 2016) together with
large amounts of gene sequencing data will greatly improve the efficient production
of biofuels (Kubicek and Kubicek 2016). This will be done also because of the always
better molecular biology technologies.
Genetics will play a sure role in the development of microorganisms, for instance
algae, with the purpose of improving their growth rate, or even their capability of
transforming the harvested energy. Even if it will be difficult to improve the almost
perfect way, the photosynthesis operates. In fact, artificial photosynthesis will also
continue to be researched in the future along with other light-harvesting machines.
As it was provided in the first part of this manuscript, the amount of energy coming
from the sun is indeed enormous and we should improve the ways this energy is stored.
The sun is driving many processes including the carbon cycles reported previously.
New ways of utilizing CO2 will be developed, for instance, by means of methanation,
since CO2 can be processed with hydrogen to give methane. Catalysis, of course, will
be central in the production of biofuels, and new technologies will be implemented
to produce structures where catalysts are allocated.

References

Cherubini, F. (2010). The biorefinery concept: Using biomass instead of oil for producing energy
and chemicals. Energy Conversion and Management, 51, 1412–1421. https://doi.org/10.1016/j.
enconman.2010.01.015.
De Blasio, C. (2010). Reactive dissolution of sedimentary rocks in flue gas desulfurization. Painos-
alama OY: Modeling and Experimental Investigation.
De Blasio, C., Carletti, C., Salonen, J., & Björklund-Sänkiaho, M. (2018). Ultrasonic power
to enhance limestone dissolution in the wet flue gas desulfurization process. Modeling and
results from stepwise titration experiments. ChemEngineering, 2, 53. https://doi.org/10.3390/
chemengineering2040053.
De Blasio, C., De Gisi, S., Molino, A., Simonetti, M., Santarelli, M., & Björklund-Sänkiaho, M.
(2019). Concerning operational aspects in supercritical water gasification of kraft black liquor.
Renewable Energy, 130, 891–901. https://doi.org/10.1016/j.renene.2018.07.004.
De Blasio, C., Lucca, G., Özdenkci, K., Mulas, M., Lundqvist, K., Koskinen, J., et al. (2015).
A study on supercritical water gasification of black liquor conducted in stainless steel and
nickel-chromium-molybdenum reactors. Journal of Chemical Technology and Biotechnology,
91, 2664–2678. https://doi.org/10.1002/jctb.4871.
Kashirtsev, V. A., Gaiduk, V. V., Chalaya, O. N., & Zueva, I. N. (2012). Geochemistry of biomarkers
and catagenesis of organic matter of Cretaceous and Cenozoic deposits in the Indigirka-Zyryanka
basin (northeastern Yakutia). Russian Geology and Geophysics, 53, 787–797. https://doi.org/10.
1016/j.rgg.2012.06.006.
References 153

Kato, H., Kobayashi, Y., & Miyashiro, H. (2018). Differential voltage curve analysis of a lithium-
ion battery during discharge. Journal of Power Sources, 398, 49–54. https://doi.org/10.1016/j.
jpowsour.2018.07.043.
Khan, S., Siddique, R., Sajjad, W., Nabi, G., Hayat, K. M., Duan, P., et al. (2017). Biodiesel
production from algae to overcome the energy crisis. Hayati Journal of Bioscience, 24, 163–167.
https://doi.org/10.1016/j.hjb.2017.10.003.
Klass, D. L. (1998). Biomass for renewable energy, fuels, and chemicals. Academic Press.
Kubicek, C. P., & Kubicek, E. M. (2016). Enzymatic deconstruction of plant biomass by fungal
enzymes. Current Opinion in Chemical Biology Energy Mechanistic Biology, 35, 51–57. https://
doi.org/10.1016/j.cbpa.2016.08.028.
Liguori, R., & Faraco, V. (2016). Biological processes for advancing lignocellulosic waste biore-
finery by advocating circular economy. Bioresource Technology Waste Biorefinery—Advocating
Circular Economy, 215, 13–20. https://doi.org/10.1016/j.biortech.2016.04.054.
Loman, A. A., & Ju, L. -K. (2016). Soybean carbohydrate as fermentation feedstock for production
of biofuels and value-added chemicals. Process Biochemistry, 51, 1046–1057. https://doi.org/10.
1016/j.procbio.2016.04.011.
Neste Corporation. (2016). Neste ja IKEA Sweden kumppaneiksi uusiutuvan biomuovin tuotan-
nossa.
Özdenkçi, K., De Blasio, C., Muddassar, H. R., Melin, K., Oinas, P., Koskinen, J., et al. (2017).
A novel biorefinery integration concept for lignocellulosic biomass. Energy Conversion and
Management, 149, 974–987. https://doi.org/10.1016/j.enconman.2017.04.034.
Palmeros Parada, M., Osseweijer, P., & Posada Duque, J. A. (2017). Sustainable biorefineries,
an analysis of practices for incorporating sustainability in biorefinery design. Industrial Crops
Products Challenges in Building a Sustainable Biobased Economy, 106, 105–123. https://doi.
org/10.1016/j.indcrop.2016.08.052.
Pandey, A., Hofer, R., Larroche, C., Taherzadeh, M., & Nampoothiri, M. (2015). Industrial biore-
fineries and white biotechnology. Elsevier.
Schwartz, C. M., Hussain, M. S., Blenner, M., & Wheeldon, I. (2016). Synthetic RNA polymerase III
promoters facilitate high-efficiency CRISPR-Cas9-mediated genome editing in Yarrowia lipoly-
tica. ACS Synthetic Biology, 5, 356–359. https://doi.org/10.1021/acssynbio.5b00162.
Singh, R., Kemausuor, F., & Wooldridge, M. (2018). Locational analysis of cellulosic ethanol
production and distribution infrastructure for the transportation sector in Ghana. Renewable and
Sustainable Energy Reviews, 98, 393–406. https://doi.org/10.1016/j.rser.2018.09.017.
Statista. (2018a). Forecast: Global undernourished population 1990–2030, by region | Statis-
tic [WWW Document]. Statista. https://www.statista.com/statistics/678742/underunourished-
population-worldwide-by-region/. Accessed December 31, 2018.
Statista. (2018b). World—Total population 2007–2017 | Statistic [WWW Document]. Statista.
https://www.statista.com/statistics/805044/total-population-worldwide/. Accessed December 31,
2018.
UNICEF, World Health Organization, & The World Bank. (2018). Levels and trends in child mal-
nutrition. UNICEF / WHO / World Bank Group Joint Child Malnutrition Estimates Key findings
of the 2018 edition.
Uzun, H., Yıldız, Z., Goldfarb, J. L., & Ceylan, S. (2017). Improved prediction of higher heating
value of biomass using an artificial neural network model based on proximate analysis. Biore-
source Technology, 234, 122–130. https://doi.org/10.1016/j.biortech.2017.03.015.
Valmet. (2014). Valmet technical paper series: Bio-refining.
Chapter 12
Integrated Biorefinery Concepts

Of course, when nonsense comes out of a computer people have


a lot of respect for it, and that is exactly the problem.
Gianni Astarita, Chem. Eng. Sci. 52, 24, 4681–4698.

12.1 An Introduction to Biorefineries


and to the Sustainable Utilization of Bioresources

Different solutions have been proposed as biorefinery concepts and therefore we


have different technologies available with different products.
One simple classification related to biorefineries divides them in “single product”
processes and “integrated process” (Svensson et al. 2015). In an integrated biorefin-
ery, it is possible to get both the bio-based products, for instance pulp or biochemical,
and energy products such as electricity, heat, or biofuels (Pandey et al. 2015). It is
reported that process integration can have a significant influence to the profitabil-
ity of the biorefinery and a suitable product diversification can help companies to
manage periods of low demand and crisis (Pandey et al. 2015). The integration of
processes allows for the reduction of transportation and energy consumption, and in
the case there is a surplus of energy produced, this can be sold or utilized for other
purposes. In Finland, the pulp and paper industry has a leading role on the economy
of the country; in addition, the principal by-product coming from this process, black
liquor, is considered as the most important fuel and also the most abundantly used.
To give one example, 11% of the total heat and power produced in the nation is
derived by this fuel (Helynen et al. 2002). As a matter of fact, black liquor is used in
recovery boilers where this fuel is combusted and the chemicals contained in it are
recovered by suitable reduction–oxidation processes.
A “multi-feed, multi-purpose” biorefinery concept which involves the processing
of black liquor has been proposed by Özdenkci and collaborators (2017). Process
integration of the biorefinery concept and improvements of the costs with reduction
of the power used can be started from the supply. The supply-chain network structure

© Springer Nature Switzerland AG 2019 155


C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_12
156 12 Integrated Biorefinery Concepts

Fig. 12.1 Integration concept of regional biorefinery. The Vaasa region, Finland. Figure made by
Lena Karlsson. Reprinted by permission of Vaasa Water—Public Utility Company

and even the position of the biorefinery have to be planned carefully with respect
to the resources and the companies present on the territory. To give one example of
suitable conditions for the concept of process integration of biorefinery, the Vaasa
region in Finland shown good potential since there is a discreet number of feedstock
producers, natural resources, and biogas facilities as shown in Fig. 12.1.
Sectorial integration network can be taken into account together with new tech-
nologies; this will include pretreatment of the feedstock where it is available and
regional distribution of converted biomass from various sectors. This is followed by
the upgrading of the produced biofuels and the separation of the pollutants which will
be recovered further. A sectoral integration concept example aimed at a supply-chain
network is shown in the Fig. 12.2.
This could reflect the structures of the supply chain; in fact, there are mainly three
types of supply-chain networks: the centralized, distributed, and also a hybrid of
them known as distributed–centralized. This is described by the following Fig. 12.3.
On the other hand, we can have different approaches concerning the way we
utilize biomass. For instance, we can partly decompose the biomass and obtain side
streams and products which will be separated end/or upgraded. In alternative, we
could proceed to the complete destruction of the biomass and obtain syngas which
could, in turn, give further products or it could be utilized in an engine. The third
option would be to fractionate the biomass feedstock to give, for instance, cellulose,
lignin, and hemicellulose. These compounds could be treated to have chemicals,
12.1 An Introduction to BioRefineries and … 157

Fig. 12.2 Sectoral integration concept aimed at the supply-chain network

Fig. 12.3 Supply-chain structures. The centralized biomass conversion (Route 1), the distributed
biomass conversion (Route 2), and the centralized–distributed biomass conversion
158 12 Integrated Biorefinery Concepts

Fig. 12.4 Different approaches in biomass processing

fibers, or they could be processed with a thermal treatment to give syngas or in


addition, sent to a chemical processing unit to obtain biomaterials (Fig. 12.4).
Among the biological processing of biomass, also in this case, we could distin-
guish diverse kinds of approach (Ishola et al. 2015, 2013; Kim et al. 2016; Li et al.
2016; Zheng et al. 2009):
– saccharification and fermentation. Done at the same time (SSF);
– saccharification and co-fermentation performed simultaneously;
– consolidated bioprocessing (CBP);
– saccharification, followed by filtration and fermentation (SSFF). The steps are
done simultaneously;
– genetically engineered yeast (Kwak et al. 2018);
– separate hydrolysis and fermentation (SHF);
while the fractionation of the biomass can be done by means of diverse methods
like:
– alkaline treatment (Kim et al. 2016);
– organosolv treatment with alcohol or acid (Choi et al. 2019)
Biomass can be processed by thermal treatment and this could include:
– Pyrolysis. This includes fast pyrolysis (580–980 °C, 10–200 K/s, residence time
of 10–200 s) and slow pyrolysis (280–680 °C, 0.1–10 K/s, residence time of
450–550 s) (Balat et al. 2009a);
– thermal gasification (Balat et al. 2009b);
– hydrothermal processes (Özdenkçi et al. 2017)
• aqueous phase reforming;
12.1 An Introduction to BioRefineries and … 159

Fig. 12.5 LignoForceTM System, WBL weak black liquor, SBL strong black liquor

• carbonization;
• hydrothermal liquefaction;
• oxidation and gasification (SCWO and SCWG);

– partial wet oxidation (130–270 °C, 30 min residence time, partial oxygen pressure
of 0.5 MPa) (Muddassar et al. 2015, 2014; Özdenkci et al. 2014; Sipponen et al.
2016).
Regarding the PWO of biomass, this is a quite interesting process where the
partial oxidation of the biomass fraction takes place. It is often suggested that this
treatment could be coupled with other processes since this is an exothermic step
where heat could be produced to run side processes which could treat the biomass
further. The PWO gives also the possibility to transform difficult feedstock like lignin
into carboxylic acids, for instance, and these products will be more easily processed
by other methods like SCWG or HTL.
In industry, the LignoForce system is a novel process developed by FPinnovations
in Canada and this process is starting with the PWO of the black liquor fraction to
oxidize it. The sulfur is also oxidized by this method and, as a result, this is improving
the lignin recovery. When the black liquor is acidified, there is the formation of
sulfur compounds which could be volatile. By this method, the sulfur compounds
are transformed into non-volatile species, and the lignin is separated into two forms,
acid form and sodium form. The latter one has been discovered to be particularly
suitable as a substitute of phenol–formaldehyde. In addition, lignin could be utilized
instead of polyols and polyurethane and in thermoplastics. Lignin could also be used
as dispersant and to substitute flocculants and textile industry (Schreiber et al. 2014).
By means of the LignoForce system, lignin can be precipitated by reducing the
pH of the partial wet oxidized black liquor. This can be achieved by pumping CO2
160 12 Integrated Biorefinery Concepts

or adding sulfuric acid to the solution. The pH value should be reduced to 9–10 from
an original value of 13–14 (Kouisni et al. 2016). The process should be followed
by a coagulation of the lignin molecules, a filtration process and a washing process
which is using sulfuric acid and water.
Compounds like dimethyl sulfide and disulfide, methyl mercaptans, and hydro-
gen sulfide are totally reduced forms of sulfur (TRS) which are very dangerous,
and therefore, in case the acidification of the lignin, slurry is done in presence of
pressurized oxygen, this compounds would be oxidized to non-volatile species. The
schematics of the process is shown in Fig. 12.5.
In HTL of biomass, we have temperatures between 280 and 380 °C and pressures
between 70 and 300 bar. At these conditions, the self-dissociation of water (see the
description of hydrothermal gasification explained later in this part of the manuscript)
increases significantly and this will mean that there will be more concentration of
H+ and OH− ions allowing for subcritical water to catalyze acid–base reactions. It is
referred in literature that this process produces bio-oil with less amount of oxygen
(Tekin et al. 2014) and this means that we would need less hydrogen to upgrade it to
liquid fuels. In addition, homogeneous and heterogeneous catalysts can be applied
also in this case.
Alkali salts are used as catalysts and they are reported to accelerate the water gas
shift reaction; on the other hand, potassium salts minimize the formation of char
and improve the oil yield (Christensen et al. 2014). As heterogeneous catalysts, we
recognize zirconium oxides and nickel alloys.
HTL can lead to a yield of 30–40% of oil (dry basis on the feedstock) if we consider
lingo-cellulosic feedstock. The energy recovery is usually around 80% (Toor et al.
2011). However, there is also literature which is not that encouraging with respect
to this process. In fact, it is referred that if we take into account a techno-economic
analysis of the process, this will be not really competitive with respect to other
processes like the ones of petroleum industry (Zhu et al. 2014).
HTL, SCWG, and PWO could be integrated to have a regional system flexible
to different kinds of biomass feedstock. An example of an integrated system which
takes black liquor as incoming feedstock is shown in Fig. 12.6. The system foresees
the recovery of chemicals and energy with the production of syngas and biofuels
(Fig. 12.6).
The process conditions for the system proposed in Fig. 12.6 are written in
Table 12.1. The conditions suggested by Özdenkci and the author of this book are
according to the suggestions given in literature and on the experience gained in
research laboratories with particular reference to the case of PWO and SCWG of
biomass.
Another example is given here in regard to the utilization of waste organic streams.
Waste organic streams could feed algae which, in turn could be used for HTL. The
anaerobic fermentation could give volatile fatty acids, VFA, which could be processed
by esterification and hydrogenation to give ethanol, butanol, and propanol. The VFAs
could be further processed by yeast fermentation and give bio-oil (Fig. 12.7).
According to the cited reference, integration could include also the introduction of
plastics into the HTL. However, this option is not taken into account here, considering
12.1 An Introduction to BioRefineries and … 161

Fig. 12.6 An example of process for regional biomass conversion. Adapted from Özdenkçi et al.
(2017)

Table 12.1 Reaction conditions for the process of regional biomass conversion
Unit Conditions Comments
Acidification T: 80 °C Lignin precipitation at pH 9–10
P: 1 Atm according to literature data
PWO T: 170–240 °C Heating of the biomass feedstock
PO2 : 0.5–1 MPa due to the exothermic conditions
t: 10–20 min
The conversion reactor T: 250–350 °C HTL
P: 4–22 MPa
t: 1–12 min
The conversion reactor T: 600–700 °C SCWG
P: 25 MPa or above
t: 1–5 min
HP G/L separator T: 80–250 °C H2 -rich gas separation, in case of
P: 25 MPa or above SCWG
LP G/L separator T: 80 °C or less CO2 -rich gas separation, in case
P: 1 Atm of SCWG
or
Separation of aqueous valuable
compounds and bio-oil if we
have HTL
Adapted from Özdenkçi et al. (2017)
162 12 Integrated Biorefinery Concepts

Fig. 12.7 Integration of fermentation, hydrogenation, HTL, and CHP plant. Adapted from Coma
et al. (2017)

that HTL of plastics is not effectively working when, for instance, polyethylene and
polypropylene are tested alone, as referred by the same researchers (Coma et al.
2017). Additionally, plastics are a well-known problematic feedstock because it can
lead to toxic and corrosive streams and therefore, some other methods could be
applied instead like recycling.
In the framework of thermochemical-based biorefineries, biomass gasification
plays an important role considering the flexibility of the products quality that can
be obtained. The main product of gasification is a gas-phase mixture of hydrogen,
carbon monoxide, carbon dioxide, methane, and traces of higher hydrocarbons. This
gas-phase product is commonly named as syngas. Depending on the gasification
medium, nitrogen may cover a relevant fraction of the gas mixture. For instance,
nitrogen content can reach values higher than 60% of the total gas volume when air
is used as gasification medium (Pérez et al. 2012). In addition to air, other gasification
mediums are steam, pure oxygen, carbon dioxide, or their mixes. Especially in the
case of power production, the most common gasification mediums are air or mixtures
of air and steam because of economic reasons. Indeed, while pure oxygen reduced
the diluting effect of nitrogen, its costs refrains from its utilization. The addition of
steam has a low impact on the operating costs of gasifiers if a proper heat recovery
system is integrated in the gasification process. Furthermore, it allows increasing the
syngas yield and its heating values. An important advantage of using steam is the
possibility to produce a hydrogen reach syngas (Prestipino et al. 2017), reducing the
efforts for downstream upgrading. Using steam alone as gasification medium leads to
allothermal gasification, with additional costs for external heat supply. Hence, pure
12.1 An Introduction to BioRefineries and … 163

steam gasification is used only in rare cases for the production of products of high
added value (Chiodo et al. 2017; Ning et al. 2018).
Concerning the kinds of feedstock which could be utilized for gasification pur-
poses, wood accounts for the largest source of biomass. However, several studies are
under development for enhancing the use of alternative feedstock. Municipal, indus-
trial, and agricultural wastes and residues are promising alternatives that include
biorefineries in the circular economy systems (G˛adek et al. 2016; Galvagno et al.
2017). Gasification of residues can be also applied and integrated in biochemi-
cal biorefineries processes. Indeed, the lignin-rich residues of second generation
bioethanol production are good candidates for being used as feedstock for energy
and chemicals recovery by thermochemical gasification (Cerone et al. 2017).
The syngas produced from biomass gasification can be used by different kind of
downstream processes. Indeed, the gas composition can be also tuned in order to fit
the requirements for producing hydrogen, combined cooling heating and power, and
liquid biofuels (Chiodini et al. 2017; Kraussler et al. 2018; Palomba et al. 2017).
Different power production units, such as high-temperature fuel cells, internal
combustion engines, and gas turbines, can be fed by syngas for delocalized produc-
tion of energy at different scales (Al-attab and Zainal 2018; Galvagno et al. 2016;
Prestipino et al. 2016).
Renewable gas can be produced by utilizing microorganisms growing at anaerobic
conditions (Croce et al. 2016). Additionally, biogas, which contains large amounts
of carbon dioxide and methane, can be produced by municipal waste. One example
of this kind of production is the company Stormossen situated in the Vaasa region in
Finland. An additional method is to convert fermentable sugars into products which
can be bioethanol or organic acids. This is also done by fermentation.
Furthermore, biofuels like biodiesel can be produced by processing waste veg-
etable oils or waste in general. The main reactions are esterification and transesteri-
fication or for instance, hydrogenation reactions. Later on in this manuscript, these
concepts will be covered.

12.2 Industrial Examples

12.2.1 Biodiesel Production from Bio Oils by Hydrogenation

As it will be discussed later on in this manuscript, the commonly produced fatty acid
methyl ester (FAME) biodiesel is prepared from vegetable oils by transesterification
reaction. However, it is reported in literature that the properties of FAME are not
optimal for the transport industry and this is because of the high oxygen content and
its high cloud point (Ramos et al. 2014).
The hydrogenation of bio-oils to produce biocrude oil and biodiesel will give
a product similar to the common diesel and this represents an alternative solution
adopted by companies like Neste and UPM (Figs. 12.8 and 12.9).
164 12 Integrated Biorefinery Concepts

Fig. 12.8 NEXBTL process used for biodiesel production

Fig. 12.9 Production of renewable diesel from tall oil. Permission to use the figure granted by
UPM

Neste’s next-generation biomass-to-liquid (NExBTL) technology is considered to


be a promising technology (Neste Oil 2018). The biodiesel is produced from oils and
fats which are processed to remove any of the impurities. Hydrogenation is followed
to convert the oils and fats into hydrocarbons and alkenes are obtained together
with alkanes; this offers the possibility of refining them to biodiesel or if possible
to other products. This process offers also the possibility of producing biogasoline
and biopropane. Neste has also investigated the possibility of producing renewable
chemicals like bioplastics, always by using the NEXBTL technology (Bezergianni
2013).
12.2 Industrial Examples 165

Neste is one of the biggest producers of biodiesel; as a matter of fact, the production
of biodiesel by Neste is around two million tons per year and the production plants are
located in Porvoo (Finland), Rotterdam, and Singapore. The biodiesel is produced
from diverse feedstock kinds among which there are vegetable oils and also animal
fat from the waste food industry. The vegetable oils are mainly rapeseed, soya, and
palm oil (Engman et al. 2016).
UPM, another Finnish company, utilizes similar hydrogenation technology for
the production of biodiesel. In this case, the biodiesel is called BioVerno (UPM-
Kymmene 2012). The BioVerno diesel is produced in Lappeenranta, Finland with
an annual production capacity of 120 mL. For this process, tall oil is utilized as raw
material (Vásquez et al. 2017). Tall oil is considered as residual stream in the pulp and
paper industry. This feedstock is firstly purified and then hydrogenated. Gases like
H2 S and other non-condensable gases are removed during the process. The biodiesel
is then separated from the other fractions by distillation. This particular biodiesel is
also suitable for driving in winter time since it has the characteristics of the normal
diesel (Vauhkonen 2015).

12.2.2 Metsä Group’s Äänekoski Plant

One example of next generation biorefinery concept is given by Metsä Group. As


written on the company Web pages, the biorefinery is built in Äänekoski in Finland.
The overall investment has been 1.2 Billion euro, it produces 1.3 million ton per
year and it utilizes 6.5 M m3 of wood annually (Engström 2017). The bio-electricity
production is 1.8 TWh/year (Nousiainen 2018). The additional goal of this company
is to utilize side streams to produce alternative compounds as it can be observed from
Fig. 12.10.
Figure 12.10 shows the additional products which could be obtained from the
biorefinery and some of these products could even be produced with partners com-
panies. In addition, the company claims to produce a surplus of electricity which can
be either sold or utilized in other processes. The claimed aim of the Metsä Group is
to achieve a situation in which no waste is produced.
In the Äänekoski plant, also biogas as biofuel will be utilized, the development
of the biogas process for Äänekoski plant is done by Envor Protech (Envor Protech
OY 2016). The technology involves the fermentation of biomass with high lignin
content. Large content of lignin is always problematic when digestion or any other
treatment process is required; this is because lignin is a very stable compound formed
by aromatic components. We remember from the organic chemistry that an aromatic
ring presents more stability since there are resonant forms of the same component.
166 12 Integrated Biorefinery Concepts

Fig. 12.10 A next-generation wood biorefinery proposed by Metsä Group. Used with permission
(von Weymarn 2016)

12.3 Production Concepts for Bioethanol Production

Bioethanol is produced by a fermentation process in which sugars like glucose is


converted into ethanol and carbon dioxide. The process conditions are anaerobic.
Bioethanol can be mixed to the normal gasoline which is then blended. To give one
example of ethanol utilization and mixing in gas fuels, in Finland, there are two
possible blended options available: the 5% blend and the 10% blend, while in other
European countries, the so-called E10 is not available everywhere (ePURE 2018).
Most of the produced bioethanol is considered as a first-generation biofuel because
the raw material for the process is starch or glucose coming mainly from the arable
land. To give an example of company producing ethanol in Finland, the Finnish com-
pany St1 has developed also new concepts for producing bioethanol. The technology
called Ethanolix® for instance is based on the bakeries and breweries waste products
and the side products can be utilized for feeding animals (Nair et al. 2017).
St1 has four Etanolix plants in Finland and one in Sweden. The production capacity
is, however, quite small since usually this technology is positioned near companies
producing food or for instance beer (Fig. 12.11). The production is accounted to
be between one to seven million liters per year and for instance, the plant sited in
Gothenburg gives five million liters of high-quality ethanol (St1 Nordic Oy 2018,
p. 1).
12.3 Production Concepts for Bioethanol Production 167

Fig. 12.11 Ethanolix process and distribution flow

It is possible to see from the figure above that the facilities producing waste
are near the ethanol plant. The by-products from the process are utilized as animal
feeding and fertilizers; the water is removed in the dehydration section. Finally, the
bioethanol is mixed to the gasoline and goes to the gasoline stations.
Another concept developed by St1 for bioethanol production is the Bionolix. The
raw materials for the process in this case are municipal and industrial biowaste and
also meat products. St1 has one Bionolix plant in Finland. As a side product, it
produces electricity, district heat, and fertilizers.
The Cellunolix is an additional technology proposed by the same company (St1
Nordic Oy 2010). The idea behind this process is to produce bioethanol from biomass
which contains lignin. As the reader can notice, there is more and more motivation to
develop technologies which are able to process lignin for biofuels production. This
is because lingo-cellulosic biomass is largely available in some regions of the world
like Finland.
In this concept, the lingo-cellulosic material is firstly treated with enzymes in
order to release fermentable sugars. Then yeast is utilized to ferment the feedstock in
order to produce ethanol. Distillation and water removal are following this step. Using
lingo-cellulosic material in bioethanol production is a new technology. Therefore, St1
does not yet have any operating Cellunolix plants. However, the first industrial plant
is currently under construction in Kajaani, Finland. The plant will have a capacity
of 10,000 m3 per year. The raw material for the process is sawdust and it will come
from the local sawmills.

References

Al-attab, K. A., & Zainal, Z. A. (2018). Micro gas turbine running on naturally aspirated syngas: An
experimental investigation. Renewable Energy, 119, 210–216. https://doi.org/10.1016/j.renene.
2017.12.008.
168 12 Integrated Biorefinery Concepts

Balat, M., Balat, M., Kırtay, E., & Balat, H. (2009a). Main routes for the thermo-conversion of
biomass into fuels and chemicals. Part 1: Pyrolysis systems. Energy Conversion and Management,
50, 3147–3157. https://doi.org/10.1016/j.enconman.2009.08.014.
Balat, M., Balat, M., Kırtay, E., & Balat, H. (2009b). Main routes for the thermo-conversion of
biomass into fuels and chemicals. Part 2: Gasification systems. Energy Conversion and Manage-
ment, 50, 3158–3168. https://doi.org/10.1016/j.enconman.2009.08.013.
Bezergianni, S. (2013). Catalytic hydroprocessing of liquid biomass for biofuels production.
InTECH.
Cerone, N., Zimbardi, F., Contuzzi, L., Prestipino, M., Carnevale, M. O., & Valerio, V. (2017).
Air-steam and oxy-steam gasification of hydrolytic residues from biorefinery. Fuel Processing
Technology, 167, 451–461. https://doi.org/10.1016/j.fuproc.2017.07.027.
Chiodini, A., Bua, L., Carnelli, L., Zwart, R., Vreugdenhil, B., & Vocciante, M. (2017). Enhance-
ments in Biomass-to-Liquid processes: Gasification aiming at high hydrogen/carbon monoxide
ratios for direct Fischer-Tropsch synthesis applications. Biomass and Bioenergy, 106, 104–114.
https://doi.org/10.1016/j.biombioe.2017.08.022.
Chiodo, V., Urbani, F., Zafarana, G., Prestipino, M., Galvagno, A., & Maisano, S. (2017). Syngas
production by catalytic steam gasification of citrus residues. International Journal of Hydrogen
Energy; Advanced Materials for Fuel Cells and Electrolyzers: The E-MRS 2016 Fall Meeting
Symposium Q (Vol. 42, pp. 28048–28055), September 19–22, 2016, Warsaw, Poland. https://doi.
org/10.1016/j.ijhydene.2017.08.085.
Choi, J.-H., Jang, S.-K., Kim, J.-H., Park, S.-Y., Kim, J.-C., Jeong, H., et al. (2019). Simultaneous
production of glucose, furfural, and ethanol organosolv lignin for total utilization of high recalci-
trant biomass by organosolv pretreatment. Renewable Energy, 130, 952–960. https://doi.org/10.
1016/j.renene.2018.05.052.
Christensen, P. S., Peng, G., Vogel, F., & Iversen, B. B. (2014). Hydrothermal Liquefaction of the
microalgae phaeodactylum tricornutum: Impact of reaction conditions on product and elemental
distribution. Energy and Fuels, 28, 5792–5803. https://doi.org/10.1021/ef5012808.
Coma, M., Martinez-Hernandez, E., Abeln, F., Raikova, S., Donnelly, J., …. Chuck, C. (2017).
Organic waste as a sustainable feedstock for platform chemicals. Faraday Discuss, 202, 175–195.
https://doi.org/10.1039/C7FD00070G.
Croce, S., Wei, Q., D’Imporzano, G., Dong, R., & Adani, F. (2016). Anaerobic digestion of straw and
corn stover: The effect of biological process optimization and pre-treatment on total bio-methane
yield and energy performance. Biotechnology Advances, 34, 1289–1304. https://doi.org/10.1016/
j.biotechadv.2016.09.004.
Engman, A., Hartikka, T., Honkanen, M., Kiiski, U., Kuronen, M., Lehto, K., et al. (2016). Neste
renewable diesel handbook. Espoo: Neste Corp.
Engström, J. (2017). Äänekoski a Mill for the 21th Century (No. 36), Spectrum. Andritz.
Envor Protech OY. (2016). Envor Protech Oy is developing biogas technology to forest industry
[WWW Document]. http://www.envorprotech.fi/en/?newspage=2. Accessed July 14,18.
ePURE. (2018). Fuel blends | ePURE—European renewable ethanol [WWW Document]. https://
epure.org/about-ethanol/fuel-market/fuel-blends/. Accessed July 14, 18.
G˛adek, W., Mlonka-M˛edrala, M., Prestipino, M., Evangelopoulos, P., Kalisz, S., & Yang, W. (2016).
Gasification and pyrolysis of different biomasses in lab scale system: A comparative study. E3S
Web Conference, 10, 00024. https://doi.org/10.1051/e3sconf/20161000024.
Galvagno, A., Prestipino, M., Chiodo, V., Maisano, S., Brusca, S., & Lanzafame, R. (2017). Energy
performance of CHP system integrated with citrus peel air-steam gasification: A comparative
study. In Energy Procedia, ATI 2017—72nd Conference of the Italian Thermal Machines Engi-
neering Association (Vol. 126, pp. 485–492). https://doi.org/10.1016/j.egypro.2017.08.233.
Galvagno, A., Prestipino, M., Zafarana, G., & Chiodo, V. (2016). Analysis of an integrated agro-
waste gasification and 120 kW SOFC CHP system: modeling and experimental investigation.
In Energy Procedia, ATI 2016—71st Conference of the Italian Thermal Machines Engineering
Association (Vol. 101, pp. 528–535). https://doi.org/10.1016/j.egypro.2016.11.067.
References 169

Helynen, S., Sipila, K., Peltola, E., & Holttinen, H. (2002). Renewable energy sources in Finland
by 2030. Helsinki: Eduskunnan kanslia.
Ishola, M. M., Brandberg, T., & Taherzadeh, M. J. (2015). Simultaneous glucose and xylose uti-
lization for improved ethanol production from lignocellulosic biomass through SSFF with encap-
sulated yeast. Biomass and Bioenergy, 77, 192–199. https://doi.org/10.1016/j.biombioe.2015.03.
021.
Ishola, M. M., Jahandideh, A., Haidarian, B., Brandberg, T., & Taherzadeh, M. J. (2013). Simulta-
neous saccharification, filtration and fermentation (SSFF): A novel method for bioethanol pro-
duction from lignocellulosic biomass. Bioresource Technology, 133, 68–73. https://doi.org/10.
1016/j.biortech.2013.01.130.
Kim, J. S., Lee, Y. Y., & Kim, T. H. (2016). A review on alkaline pretreatment technology for
bioconversion of lignocellulosic biomass. Bioresource Technology, 199, 42–48. https://doi.org/
10.1016/j.biortech.2015.08.085.
Kouisni, L., Gagné, A., Maki, K., Holt-Hindle, P., & Paleologou, M. (2016). Ligno force system
for the recovery of lignin from black liquor: Feedstock options, odor profile, and product charac-
terization. ACS Sustainable Chemistry and Engineering, 4, 5152–5159. https://doi.org/10.1021/
acssuschemeng.6b00907.
Kraussler, M., Binder, M., Schindler, P., & Hofbauer, H. (2018). Hydrogen production within a
polygeneration concept based on dual fluidized bed biomass steam gasification. Biomass and
Bioenergy, 111, 320–329. https://doi.org/10.1016/j.biombioe.2016.12.008.
Kwak, S., Jo, J. H., Yun, E. J., Jin, Y. -S., & Seo, J. -H. (2018). Production of biofuels and chemicals
from xylose using native and engineered yeast strains. Biotechnology Advances. https://doi.org/
10.1016/j.biotechadv.2018.12.003.
Li, M.-F., Yang, S., & Sun, R.-C. (2016). Recent advances in alcohol and organic acid fractiona-
tion of lignocellulosic biomass. Bioresource Technology, 200, 971–980. https://doi.org/10.1016/
j.biortech.2015.10.004.
Muddassar, H. R., Melin, K., de Villalba Kokkonen, D., Riera, G. V., Golam, S., & Koskinen,
J. (2015). Green chemicals from pulp production black liquor by partial wet oxidation. Waste
Management and Research Journal of the International Solid Wastes Public Cleaning Association
ISWA, 33, 1015–1021. https://doi.org/10.1177/0734242X15602807.
Muddassar, H. R., Melin, K., & Koskinen, J. (2014). Production of carboxylic acids from alkaline
pretreatment byproduct of softwood. Cellulose Chemistry and Technology, 48, 835–842.
Nair, R. B., Lennartsson, P. R., & Taherzadeh, M. J. (2017). Bioethanol production from agricultural
and municipal wastes. In: J. W. –C. Wong, R. D. Tyagi, A. Pandey (Eds.), Current developments
in biotechnology and bioengineering (pp. 157–190). Elsevier. https://doi.org/10.1016/B978-0-
444-63664-5.00008-3.
Neste Oil. (2018). Neste oil’s nexbtl diesel.
Ning, S., Jia, S., Ying, H., Sun, Y., Xu, W., & Yin, H. (2018). Hydrogen-rich syngas produced by
catalytic steam gasification of corncob char. Biomass and Bioenergy, 117, 131–136. https://doi.
org/10.1016/j.biombioe.2018.07.003.
Nousiainen, I. (2018). Metsä Group’s bioproduct mill has produced more than 1 million tonnes
of pulp, reaching its full production capacity [WWW Document]. https://www.metsagroup.
com/en/media/all-news/Pages/News.aspx?EncryptedId=96853DAD830688F5&Title=Metsa
Groupsbioproductmillhasproducedmorethan1milliontonnesofpulp,reachingitsfullproduction
capacity. Accessed January 6, 19.
Özdenkçi, K., De Blasio, C., Muddassar, H. R., Melin, K., Oinas, P., Koskinen, J., et al. (2017).
A novel biorefinery integration concept for lignocellulosic biomass. Energy Conversion and
Management, 149, 974–987. https://doi.org/10.1016/j.enconman.2017.04.034.
Özdenkci, K., Koskinen, J., & Sarwar, G. (2014). Recovery of sodium organic salts from partially
wet oxidized black liquor. Cellulose Chemistry and Technology, 48, 825–833.
170 12 Integrated Biorefinery Concepts

Palomba, V., Prestipino, M., & Galvagno, A. (2017). Tri-generation for industrial applications:
Development of a simulation model for a gasification-SOFC based system. International Journal
of Hydrogen Energy; Advanced Materials for Fuel Cells and Electrolyzers: The E-MRS 2016 Fall
Meeting Symposium Q (Vol. 42, pp. 27866–27883), September 19–22, 2016, Warsaw, Poland.
https://doi.org/10.1016/j.ijhydene.2017.06.206.
Pandey, A., Hofer, R., Larroche, C., Taherzadeh, M., & Nampoothiri, M. (2015). Industrial biore-
fineries and white biotechnology. Elsevier.
Pérez, J. F., Melgar, A., & Benjumea, P. N. (2012). Effect of operating and design parameters
on the gasification/combustion process of waste biomass in fixed bed downdraft reactors: An
experimental study. Fuel, 96, 487–496. https://doi.org/10.1016/j.fuel.2012.01.064.
Prestipino, M., Chiodo, V., Maisano, S., Zafarana, G., Urbani, F., & Galvagno, A. (2017). Hydrogen
rich syngas production by air-steam gasification of citrus peel residues from citrus juice manu-
facturing: Experimental and simulation activities. International Journal of Hydrogen Energy, 42,
26816–26827. https://doi.org/10.1016/j.ijhydene.2017.05.173.
Prestipino, M., Palomba, V., Vasta, S., Freni, A., & Galvagno, A. (2016). A Simulation tool to
evaluate the feasibility of a gasification-I.C.E. System to produce heat and power for industrial
applications. In Energy Procedia, ATI 2016—71st Conference of the Italian Thermal Machines
Engineering Association (Vol. 101, pp. 1256–1263). https://doi.org/10.1016/j.egypro.2016.11.
141.
Ramos, L. P., Cordeiro, C. S., Cesar-Oliveira, M. A. F., Wypych, F., & Nakagaki, S. (2014). Appli-
cations of heterogeneous catalysts in the production of biodiesel by esterification and transester-
ification, Chap. 16. In Bioenergy research: Advances and applications (pp. 255–276). Elsevier,
Amsterdam.
Schreiber, M., Vivekanandhan, S., Cooke, P., Mohanty, A. K., & Misra, M. (2014). Electrospun
green fibres from lignin and chitosan: a novel polycomplexation process for the production of
lignin-based fibres. Journal Materials Science, 49, 7949–7958. https://doi.org/10.1007/s10853-
014-8481-z.
Sipponen, M. H., Özdenkci, K., Muddassar, H. R., Melin, K., Golam, S., & Oinas, P. (2016).
Hydrothermal liquefaction of softwood: Selective chemical production under oxidative con-
ditions. ACS Sustainable Chemistry and Engineering, 4, 3978–3984. https://doi.org/10.1021/
acssuschemeng.6b00846.
St1 Nordic Oy. (2018). St1 built a waste-based Etanolix® ethanol production plant in Gothenburg
- St1 [WWW Document]. English. https://www.st1.eu/st1-built-a-waste-based-etanolix-ethanol-
production-plant-in-gothenburg. Accessed July 14, 18.
St1 Nordic Oy. (2010). st1 produces bioethanol from straw and cellulose. Focus on Catalysis, 4–5.
https://doi.org/10.1016/S1351-4180(10)70207-5.
Svensson, E., Eriksson, K., & Wik, T. (2015). Reasons to apply operability analysis in the design
of integrated biorefineries. Biofuels, Bioproducts and Biorefining, 9, 147–157. https://doi.org/10.
1002/bbb.1530.
Tekin, K., Karagöz, S., & Bektaş, S. (2014). A review of hydrothermal biomass processing. Renew-
able and Sustainable Energy Reviews, 40, 673–687.
Toor, S. S., Rosendahl, L., & Rudolf, A. (2011). Hydrothermal liquefaction of biomass: A review
of subcritical water technologies. Energy, 36, 2328–2342. https://doi.org/10.1016/j.energy.2011.
03.013.
UPM-Kymmene. (2012). UPM-Kymmene annual report 2011: UPM will build the world’s first
biorefinery producing wood-based biodiesel. Focus on Catalyst, 6. https://doi.org/10.1016/S1351-
4180(12)70177-0.
Vásquez, M. C., Silva, E. E., & Castillo, E. F. (2017). Hydrotreatment of vegetable oils: A review
of the technologies and its developments for jet biofuel production. Biomass and Bioenergy, 105,
197–206. https://doi.org/10.1016/j.biombioe.2017.07.008.
Vauhkonen, V. (2015). Renewable diesel from tall oil. UPM The Biofore Company.
von Weymarn, N. (2016). Metsä group’s bioproduct mill: A next generation wood biorefinery in
Äänekoski. Metsä Fibre: Finland.
References 171

Zheng, Y., Pan, Z., & Zhang, R. (2009). Overview of biomass pretreatment for cellulosic ethanol
production. International Journal of Agricultural and Biological Engineering, 2, 51–68. https://
doi.org/10.25165/ijabe.v2i3.168.
Zhu, Y., Biddy, M. J., Jones, S. B., Elliott, D. C., & Schmidt, A. J. (2014). Techno-economic
analysis of liquid fuel production from woody biomass via hydrothermal liquefaction (HTL) and
upgrading. Applied Energy, 129, 384–394. https://doi.org/10.1016/j.apenergy.2014.03.053.
Chapter 13
Electronegativity and Microbial Catalysis

Deus sive Natura.


God, in other words, the Nature.
Benedict de Spinoza, 1677, Of human slavery, that is, of the
forces of affections. In Part 4 of his book: Ethica Ordine
Geometrico Demonstrata.

13.1 Production of Biofuels and Classification


of the Related Microorganisms Needed

We have seen previously some examples of different technologies utilized for biofuels
productions, and it was possible to see that many of the processes involved are
dealing with fermentation, or more generally biomass feedstock is processed by
microorganisms to get biofuels. In some of the processes mentioned, the fermentation
process is carried out with the aim of enzymes which determine the product yield
and rate (Zhao et al. 2018). Enzymes represent a large expense in the process, and
therefore, it is of outmost importance to optimize the process in order to have less
cost and less energy consumption.
A common fuel produced by these methods is hydrogen. Here, hydrogen is con-
sidered first because this particular fuel has the simplest molecular composition and
very high enthalpy value, for instance, its enthalpy of formation at the standard
conditions is reported as 221.67 kJ/mol (Wagman et al. 1982).
The types of microorganisms utilized for hydrogen production can be classified
on the basis of the substrate (food) they are going to oxidize (remembering that a
substrate is said to be oxidized when electrons are actually taken from it):
• chemoorganotrophs (the ones that oxidize organic compounds);
• chemolithotrophs (the ones that oxidize non-organic compounds).
The two principal steps in the formation of organic matter are indicated usually
as:
• Catabolism (production of energy and ATP);
• Anabolism (use of energy and production of organic matter).
© Springer Nature Switzerland AG 2019 173
C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_13
174 13 Electronegativity and Microbial Catalysis

Catabolism is the process of production of energy and ATP molecules from a


substrate; this process is common to humans and animals in general which do not
produce organic matter (except when they have to grow to the adult size). On the
other hand, anabolism is more familiar to plants and organisms in general, as a matter
of fact, they produce organic matter.
Continuing with the main microorganisms’ classification, the main biological
processes for the conversion of organic substrates are classified commonly as follows:
• aerobic respiration (the electron acceptor in this case is oxygen, and on the other,
the product is CO2 );
• anaerobic respiration (forms reduced compounds like H2 and CH4 or ethanol)
(Tian et al. 2017);
• fermentation (it is commonly referred to have no electron transport chain, and in
this case, there is an endogenous electron acceptor which normally is an organic
compound (Wong et al. 2016). One good example of fermentation carried on by
microorganisms is the one done by cyanobacteria (Stal and Moezelaar 1997);
Each of these biological processes has a relative environment where they can take
place. This environment can be:
• aerobic;
• anoxic (provides electron acceptors different than O2 );
• anaerobic (absence of O2 and inorganic electron acceptors).
In the process of anaerobic respiration, the electron transport system is similar
to the one in aerobic respiration. The difference is that oxygen is not the terminal
electron acceptor.
Hydrogen can be produced by a number of microorganisms. The most common
microorganisms utilized for the production of hydrogen are reported here with their
classification which is done on the basis of the environmental temperature:
• Caldanaerobacter subterraneus, thermophilic (Sokolova et al. 2001);
• Carboxydocella thermautotrophica, thermophilic (Sokolova et al. 2002);
• Carboxydothermus hydrogenoformans, thermophilic (Gerhardt et al. 1991);
• Citrobacter, mesophilic (Oh et al. 2003);
• Rhodopseudomonas palustris, mesophilic (Jung et al. 1999);
• Rhodospirillum rubrum, mesophilic (Zürrer and Bachofen 1979);
• Rubrivivax gelatinosus, mesophilic (Uffen 1976);
• Thermincola carboxydiphila, thermophilic (Sokolova et al. 2005);
• Thermococcus strain, thermophilic (Sokolova et al. 2004b);
• Thermolithobacter carboxydivorans, thermophilic (Sokolova et al. 2007);
• Thermosinus carboxydivorans, thermophilic (Sokolova et al. 2004a).
Particular attention has to be put on a microorganism called Pyrococcus furiosus,
discovered in 1986 by Karl Stetter in Italy (Keller et al. 2013). The organism is the
subject of recent research because it can be modified genetically to produce hydrogen
from CO2 and water. The organism is able to survive in superheated waters or volcano
mud. Therefore, it is classified as a thermophile. Furthermore, the P. furiosus has been
13.1 Production of Biofuels and Classification of the Related … 175

also modified to produce industrial chemicals from CO2 and hydrogen (Keller et al.
2013).
Microorganisms can also produce other kinds of fuels other than hydrogen, for
instance, there is a numerous number of microorganisms able to give ethanol. The
most common of these organisms are given here with the related scientific reference:
• Acetobacterium woodii, mesophilic (BALCH et al. 1977);
• Butyribacterium methylotrophicum, mesophilic (Zeikus et al. 1980);
• Clostridium autoethanogenum, mesophilic (Abrini et al. 1994);
• Clostridium glycolicum, mesophilic (Küsel et al. 2001);
• Clostridium ljungdahlii, mesophilic (Barik et al. 1988);
• Clostridium carboxidivorans, mesophilic (Liou et al. 2005);
• Moorella HUC22-1, thermophilic (Sakai et al. 2004).

13.2 Pathways in Biological Processing of Biomass


Feedstock

In case we consider glucose as the main nutrient, ethanol is produced by both fer-
mentation and respiration method of digestion. This pattern is described by the
Embden–Meyerhof pathway (EMP) (Golder 1979), also called glycolysis, which
is represented by the following simplified chemical reaction:
 
C6 H12 O6 +2ADP + 2NAD+ → 2CH3 COCOOH + 2ATP + 2NADH (13.1)

In the reaction above, 2CH3 COCOOH refers to the pyruvic acid, and this is the
most important part: the production of ATP, the energy carrier. This pathway can
take place in all environments. Good representations of this pathway are given in the
literature (Madigan et al. 2010).
With reference to Fig. 13.1, the intermediate compounds are: (A) glucose-6-P; (B)
fructose-6-P; (C) fructose-1,6-P; (D) dihydroxyacetone-P; (E) glyceraldehyde-3-P;
(F) 1,3-bisphosphoglycerate; (G) 3-P-glycerate; (H) 2-P-glycerate; and (I) Phospho-
enolpyruvate.
The main enzymes in the same figure are: (1) hexokinase; (2) isomerase; (3) phos-
phofructokinase; (4) aldolase; (5) triosephosphate isomerase; (6) glyceraldehyde-3-P
dehydrogenase; (7) phosphoglycerokinase; (8) phosphoglyceromutase; (9) enolase;
(10) pyruvate kinase; (11) lactate dehydrogenase; (12) pyruvate decarboxylase; and
(13) alcohol dehydrogenase.
In the fermentation of yeast, glucose gives finally two molecules of ethanol and
two CO2 with a release of energy of −239 kJ/(mol of glucose). While for lactic
acid bacteria, the glucose gives two molecules of lactate and 196 kJ/mole of glucose
of energy. Yeast fermentation is of interest not only for ethanol production, and
therefore, for producing alcoholic products, but also these kinds of processes have
176 13 Electronegativity and Microbial Catalysis

Fig. 13.1 Embden–Meyerhof pathway (EMP). Adapted from Madigan et al. (2010)

produced a growing interest toward the biowaste utilization and processing for power
and fuel generation (Ntaikou et al. 2018).
In Fig. 13.1, the first stage concerns reactions which do not involve oxidation–re-
duction, while the second stage is the one with this kind of reactions.
The degradation of organic compounds by microorganisms is given in a very
detailed way by the Kyoto Encyclopedia of Genes and Genomes. The database KEGG
(Kanehisa Laboratories 2018) is available on the web and contains a large number
of biological reactions with the related enzymes.

13.3 Main Molecules Used as Energy Carriers in Biological


Processes

In the first part of this manuscript, the Z-scheme of photosynthesis was described to
give an understanding of the importance of the electrochemical potential generation.
The substrate (in that case water) is oxidized, and electrons are taken from it. The
electrons are then transported through an electron transport chain. This is the needed
current which gives power to the molecular functioning. As an introductory concept
here, it is beneficial to remember that a substance which has less reduction potential
than a reference substance (has a more negative value) is reduced less easily than
it is the reference substance. This is quite obvious since this can be imagined as
a substance which has “enough” negative electric charges and therefore does not
accept easily more electrons.
On the other hand, a substance whose reduction potential is positive (is greater than
that of a reference substance) is more easily reduced than the reference. Hydrogen is
commonly used as a reference for these purposes. Here, the main molecules utilized
13.3 Main Molecules Used as Energy Carriers in Biological Processes 177

Fig. 13.2 Most important energetic molecules in living organisms. Adapted from Madigan et al.
(2010)

as energy carriers are described in Fig. 13.2. The Gibbs energy values at standard
conditions, G 0 molkJ
, for the main molecules utilized as energy carriers are as fol-
lows (Madigan et al. 2010): phosphoenolpyruvate: −51.6; 1,3-bisphosphoglycerate:
−52.0; acetyl phosphate: −44.8; ATP: −31.8; ADT: −31.8; acetyl-CoA: −35.7;
AMP: −14.2; and glucose 6-phosphate: −13.8. Some of these molecules are shown
also in Fig. 13.2.
The energy shown, G 0 , is the energy released when the bonds indicated by the
arrows in Fig. 13.2 are hydrolyzed.
On the other hand, an example of electron transport system for purple photosyn-
thetic and green sulfur bacteria is shown in Fig. 13.3.
In reference to this figure, the P870 and P840 are referring to the reaction
center, and their excited forms (see first part of this book). We have different
types of quinones, Q, cytochromes, Cyt, chlorophyll molecules, BChl, pheophytin,
BPh, quinones, A, ferredoxin, FD, plastoquinone, QP, ferredoxin-NADP+ reductase
(FNR), clusters, Fe-S, and menaquinone, MQ.
178 13 Electronegativity and Microbial Catalysis

Fig. 13.3 Example of electron transport system for purple photosynthetic and green sulfur bacteria.
Adapted from Rasmussen and Minteer (2014)

13.4 Oxidation and Reduction

We can have configurations, for the electron transport chains, where there is a reaction
of reduction and NADH is produced. The NADH contains energy because, as we
have seen, when some determinate bonds are broken, this will release energy. Here,
it is also important to notice that the electrochemical potential falls from values at
which is more negative (electronegative) to values where it is at more positive values,
this was also mentioned in the first part of this book as “downhill” potential in the
Z-scheme of photosynthesis. As a matter of fact, this is in agreement with the theory
and the convention adopted: In case we consider electrons possessing a negative
charge, then, the more electrons are present in a determinate region, and the more
negative will be their electrochemical potential.
The electrons are subtracted from a substrate and sent to the electron transport
chain, and finally, we need electron acceptors. This is obvious, since in very sim-
ple words, the electrons have to end up somewhere. As examples here, the nitrate
and sulfate acceptance of electrons in non-oxygenic processes (Camacho 2009) is
described by the following reactions:

1
5e− + NO− +
3 + 6H → N2 + 3H2 O (13.2)
2
8e− + SO2− +
4 + 10H → H2 S + 4H2 O (13.3)

Sulfur-based compounds are not normally utilized by the living organisms to


produce organic matter, but instead these types of molecules are often utilized for their
13.4 Oxidation and Reduction 179

ability to release energy and to give or accept electrons. Sulfur is usually associated
with very strong pollutants which are very famous within the power conversion of
problematic fuels like coal. As a matter of fact, coal can have very high sulfur content
and it is well-known that when combusted, SO2 can give the formation of acidic rains.
Now, when wet flue gas desulfurization is applied to reduce the SO2 emissions, the
key step is the limestone dissolution and the formation of calcium ions which, in
turn, react with the sulfate ions to form gypsum (De Blasio et al. 2016). Indeed,
the sulfite oxidation which takes in processes like flue gas desulfurization could be
coupled with the oxidative degradation of organic compounds, and this idea was
already reported in 1987 (Lee and Rochelle 1987).
This was mentioned to show that sulfate ions are indeed present in other very
different processes. Coming back to our biological processes, the electrochemical
potentials for the equations written previously should be positive for reactions written
in this way if phenomena are spontaneous.
The oxidation process produces ATP molecules (our fuel) and, on the basis of
the particular enzyme (the catalyst) used to perform this operation, a certain amount
of moles or number of NADH or FADH2 and ATP will be given. If we consider
palmitic oil, for instance, from this oil the palmitate is formed and transformed to
palmitoyl-CoA which is then completely oxidized to the final products. The table
which follows shows the number of fuel molecules produced by complete oxidation
of palmitoyl-CoA to CO2 and H2 O (Kamel and Halperin 2017, p. 5) (Table 13.1).

Table 13.1 Moles of ATP Enzyme NADH FADH2 ATP (number)


produced on the basis of the (number)
particular enzyme which
catalyzes the oxidation A-CoA 7 FADH2 14
reactions β-H-CoA 7 NADH 21
Iso-D 8 NADH 24
α-K-D 8 NADH 24
S-CoA 8
synthetase
S-D 8 FADH2 16
M-D 8 NADH 24
A-CoA is the acyl-CoA dehydrogenase, β-H-CoA is the β-
hydroxyacyl-CoA dehydrogenase, Iso-D is the isocitrate dehydro-
genase, S-CoA is the succinyl-CoA synthetase, S-D is the succinate
dehydrogenase, α-K-D is the α-ketoglutarate dehydrogenase, and
M-D is the malate dehydrogenase
Adapted from Nelson and Cox (2012)
180 13 Electronegativity and Microbial Catalysis

13.5 An Introduction to Modeling of Digestion Processes

An organic substrate can be digested by microorganisms which will grow in number


depending on the availability of the “food” and the environmental conditions adopted.
Other than analytical models (which will be described more in details later in this
manuscript); in the literature also, empirical models can be found for the digestion
of a substrate. To give one example, a model for biomass (cells) formation in yeasts
can be described by Eq. (13.4) (van Dijken and Scheffers 1986):

737C6 H12 O6 + 680NH3 + 6H2 SO4 + 931NADPH


+ 1349NAD+ → 1000C4 H7.32 O2.24 N0.68 S0.006 (100 g cells)
+ 424CO2 + 931NADP+ + 1349NADH + 1358H2 O + 418H+ (13.4)

Figure 13.4 shows the main steps of the pathway for the anaerobic digestion
performed by bacteria, from the hydrolysis of complex organic compounds, a series
of shorter chain molecules, are obtained. The fermentation is achieved after the
hydrolysis process and it is followed by acetogenesis where volatile acids and fatty
acids are converted to acetic acid H2 and CO2 , while the methanogenesis allows
for the release of methane. The two processes can be also carried out at the same
time by enriched microbiological cultures. As a matter of fact, the hydrogenotrophic
methanogens are indeed able to carry on the reaction:

4H2 + CO2 → CH4 + 2H2 O; (13.5)

Fig. 13.4 Main steps of the pathway for the anaerobic digestion performed by bacteria
13.5 An Introduction to Modeling of Digestion Processes 181

with a Gibbs free energy of −131 kJ/mol at standard conditions (Ma et al. 2019).
While the homoacetogens are able to finalize the reaction which leads to the formation
of acetate:

4H2 + 2CO2 → CH3 COO− + 2H2 O + H+ (13.6)

with a Gibbs free energy of −95 kJ/mol at standard conditions. By coupling these
two kinds of microbes, we can achieve the carbon reduction and energy conservation
at the same time (Siriwongrungson et al. 2007).
To carry on these processes, since reduction is involved, we need to provide
electrons to the converting species and this is a fundamental requirement. As reducing
species, zero-valent metals are utilized, for instance, zero-valent iron (Kondratenko
et al. 2013).
The anaerobic digestion process is commonly very unstable (Lyberatos and Ski-
adas 1999), and for these reasons, good mathematical models are then required in
order to have a suitable control of the process. As we will see later in this manuscript,
the mathematical models for the substrate depletion and product formation can be
different on the basis of which parameter is the limiting step during the process. For
instance, the product yield can be inhibited by the same substrate or the product, like
in the case of ethanol production from Saccharomyces cerevisiae. Table 13.2 demon-
strates provided mathematical models in which it is assumed that the methanogen
organisms are inhibited by the substrate concentration. All the mathematical models
provided in this table follow a Monod kinetics which will be analyzed later in this
manuscript.
More specifically and for the purposes of this book, it is also possible to individuate
and classify the microorganisms’ species on the basis of the fuel produced. The
microbial modeling of biofuel production is done as follows:
• fermentative bacteria and archaea (which produce H2 and CH4 );
• fungi and algae.
We can have also a classification of biological processes related to the biofuels
production as shown in Fig. 13.5, for instance, alcoholic fermentation:

C6 H12 O6 → 2C2 H5 OH + 2CO2 (13.7)

As preliminary information, it will be mentioned here that the main parameters


used for biomass growth are commonly referred to as:
• ν, number specific growth rate [1/h]
• μ, biomass specific growth rate [1/h]
and this will be analyzed further during the following chapters.
On the other hand, the most known fermenting organism is the S. cerevisiae also
represented by some biologist by the picturesque symbol given in Fig. 13.6.
Also in this case, alcoholic fermentation is studied and simulated to be integrated
with biorefineries. In the literature, alcoholic fermentation is sometimes coupled with
182 13 Electronegativity and Microbial Catalysis

Table 13.2 Some references where mathematical models are given for different processes involved
in anaerobic digestion and where there is substrate inhibition
Organism type and Process Model proposed by Inhibited by
(substrate)
Acetoclastic Methanogenesis Graef and Andrews Unionized VFA or
methanogens (1974) external inhibitor
(unionized VFA as
acetate)
Acid formers Hydrolysis of Hill and Barth (1977) Unionized VTA
(glucose) insoluble organics, Unionized VTA and
Producing methane acidogenesis and unionized NH3
(unionized VTA as methanogenesis
acetate)
Acid formers (soluble Acidogenesis Kleinstreuer and Unionized acetate,
organics) Methanogenesis Poweigha (1982) toxic substances for
Producing methane both processes
(acetate)
Acid formers Acidogenesis Moletta et al. (1986) Unionized acetate
(glucose) Methanogenesis Unionized acetate
Producing methane
(acetate)
Acid formers (soluble Acidogenesis First order Total VFA
organics) Methanogenesis Andrews (1969) Unionized VFAs
Producing methane
(unionized VFAs)
VFA indicated the volatile fatty acids
Adapted from Lyberatos and Skiadas (1999)

Fig. 13.5 Classification of biological processes related to the biofuels production. Adapted from
Levenspiel (1999)
13.5 An Introduction to Modeling of Digestion Processes 183

Fig. 13.6 Commonly used


symbol for the
Saccharomyces cerevisiae

Fig. 13.7 Example of ethanol production plant integrated with a cellulose nanofibers production
plant. AFEX, ammonia fiber expansion. Adapted from Leistritz et al. (2006)
184 13 Electronegativity and Microbial Catalysis

hydro-processing, pyrolysis, gasification, electrolysis, and even synthesis of transport


fuels and chemicals. An example of process integration for a cellulose nanofibers
production plant with a bio-ethanol plant is demonstrated in Fig. 13.7.
To better understand how biofuels production is taking place, for instance, the
microbial fermentation to produce ethanol or biogas, the author will focus the dis-
cussion here on the basics of the bioreactions engineering approach. This kind of
knowledge is necessary when we need a design of the reactors to be used; further-
more, this is the expertise the energy engineer should have in order perform the
above-mentioned design; and with “design,” it is meant here the optimization of the
working features of the reactors. Therefore, the discussion will start here with the
explanation of the main difference between the reactors configurations and then more
in details the kinetics of microbial growth and fuel yield will be treated.

References

Abrini, J., Naveau, H., & Nyns, E.-J. (1994). Clostridium autoethanogenum, sp. nov., an anaerobic
bacterium that produces ethanol from carbon monoxide. Archives of Microbiology, 161, 345–351.
https://doi.org/10.1007/BF00303591.
Andrews, J. F. (1969). Dynamic model of the anaerobic digestion model. Journal of the Sanitary
Engineering Division American Society Civil Engineers SA, 1, 95–116.
Balch, W. E., Schoberth, S., Tanner, R. S., & Wolfe, R. S. (1977). Acetobacterium, a new genus
of hydrogen-oxidizing, carbon dioxide-reducing, anaerobic bacteria. International Journal of
Systematic and Evolutionary Microbiology, 27, 355–361. https://doi.org/10.1099/00207713-27-
4-355.
Barik, S., Prieto, S., Harrison, S. B., Clausen, E. C., & Gaddy, J. L. (1988). Biological production
of alcohols from coal through indirect liquefaction. Applied Biochemistry and Biotechnology, 18,
363–378. https://doi.org/10.1007/BF02930840.
Camacho, A. (2009). Sulfur bacteria. In G. E. Likens (Ed.), Encyclopedia of inland waters (pp
261–278). Oxford: Academic Press. https://doi.org/10.1016/B978-012370626-3.00128-9.
De Blasio, C., Carletti, C., Lundell, A., Visuri, V.-V., Kokkonen, T., Westerlund, T., et al. (2016).
Employing a step-wise titration method under semi-slow reaction regime for evaluating the reac-
tivity of limestone and dolomite in acidic environment. Minerals Engineering, 86, 43–58. https://
doi.org/10.1016/j.mineng.2015.11.011.
Gerhardt, M., Svetlichny, V., Sokolova, T. G., Zavarzin, G. A., & Ringpfeil, M. (1991). Bacterial CO
utilization with H2 production by the strictly anaerobic lithoautotrophic thermophilic bacterium
Carboxydothermus hydrogenus DSM 6008 isolated from a hot swamp. FEMS Microbiology
Letters, 83, 267–271.
Golder, R. H. (1979). Patterns in the Embden-Meyerhof pathway of glycolysis. Biochemical Edu-
cation, 7, 7–8. https://doi.org/10.1016/0307-4412(79)90007-4.
Graef, S. P., & Andrews, J. F. (1974). Stability and control of anaerobic digestion. Journal of Water
Pollution Control Federation, 46, 666–683.
Hill, D. T., & Barth, C. L. (1977). A Dynamic model for simulation of animal waste digestion.
Journal of Water Pollution Control Federation, 49, 2129–2143.
Jung, G. Y., Jung, H. O., Kim, J. R., Ahn, Y., & Park, S. (1999). Isolation and characterization of
Rhodopseudomonas palustris P4 which utilizes CO with the production of H2 . Biotechnology
Letters, 21, 525–529. https://doi.org/10.1023/A:1005560630351.
References 185

Kamel, K. S., Halperin, M. L. (2017). Ketoacidosis, Chap. 5. In K. S. Kamel & M. L. Halperin


(Eds.), Fluid, electrolyte and acid-base physiology (5th ed., pp. 99–139). Philadelphia: Elsevier.
https://doi.org/10.1016/B978-0-323-35515-5.00005-1.
Kanehisa Laboratories. (2018). KEGG: Kyoto encyclopedia of genes and genomes [WWW Docu-
ment]. https://www.kegg.jp/. Accessed July 15, 18.
Keller, M. W., Schut, G. J., Lipscomb, G. L., Menon, A. L., Iwuchukwu, I. J., Leuko, T. T., et al.
(2013). Exploiting microbial hyperthermophilicity to produce an industrial chemical, using hydro-
gen and carbon dioxide. Proceedings of National Academy of Sciences, 110, 5840–5845. https://
doi.org/10.1073/pnas.1222607110.
Kleinstreuer, C., & Poweigha, T. (1982). Dynamic simulator for anaerobic digestion processes.
Biotechnology and Bioengineering, 24, 1941–1951. https://doi.org/10.1002/bit.260240903.
Kondratenko, E. V., Mul, G., Baltrusaitis, J., Larrazábal, G. O., & Pérez-Ramírez, J. (2013). Status
and perspectives of CO2 conversion into fuels and chemicals by catalytic, photocatalytic and
electrocatalytic processes. Energy & Environmental Science, 6, 3112–3135. https://doi.org/10.
1039/C3EE41272E.
Küsel, K., Karnholz, A., Trinkwalter, T., Devereux, R., Acker, G., & Drake, H. L. (2001). Physiolog-
ical ecology of Clostridium glycolicum RD-1, an Aerotolerant acetogen isolated from sea grass
roots. Applied and Environment Microbiology, 67, 4734–4741. https://doi.org/10.1128/AEM.67.
10.4734-4741.2001.
Lee, Y. J., & Rochelle, G. T. (1987). Oxidative degradation of organic acid conjugated with sulfite
oxidation in flue gas desulfurization: Products, kinetics, and mechanism. Environmental Science
and Technology, 21, 266–272. https://doi.org/10.1021/es00157a007.
Leistritz, F. L., Senechal, D. M., Stowers, M. D., McDonald, W. F., Saffron, C. M., & Hodur, N.M.
(2006). Preliminary feasibility analysis for an integrated biomaterials and ethanol biorefinery
using wheat straw feedstock (No. 23500). Agribusiness & Applied Economics Report. North
Dakota State University, Department of Agribusiness and Applied Economics.
Levenspiel, O. (1999). Chemical reaction engineering. New York, US: Wiley.
Liou, J. S.-C., Balkwill, D. L., Drake, G. R., & Tanner, R. S. (2005). Clostridium carboxidivo-
rans sp. nov., a solvent-producing clostridium isolated from an agricultural settling lagoon, and
reclassification of the acetogen Clostridium scatologenes strain SL1 as Clostridium drakei sp.
nov. International Journal of Systematic and Evolutionary Microbiology, 55, 2085–2091. https://
doi.org/10.1099/ijs.0.63482-0.
Lyberatos, G., & Skiadas, I. V. (1999). Modeling of anaerobic digestion—A review. Global NEST
International Journal, 1, 63–76.
Ma, L., Zhou, L., Ruan, M.-Y., Gu, J.-D., & Mu, B.-Z. (2019). Simultaneous methanogenesis
and acetogenesis from the greenhouse carbon dioxide by an enrichment culture supplemented
with zero-valent iron. Renewable Energy, 132, 861–870. https://doi.org/10.1016/j.renene.2018.
08.059.
Madigan, M. T., Martinko, J. M., Stahl, D. A., & Clark, D. P. (2010). Brock biology of microor-
ganisms (13th ed.,). Benjamin Cummings.
Moletta, R., Verrier, D., & Albagnac, G. (1986). Dynamic modelling of anaerobic digestion. Water
Research, 20, 427–434. https://doi.org/10.1016/0043-1354(86)90189-2.
Nelson, D. L., & Cox, M. M. (2012). Lehninger principles of biochemistry (6th ed.). W.H: Freeman.
Ntaikou, I., Menis, N., Alexandropoulou, M., Antonopoulou, G., & Lyberatos, G. (2018). Valoriza-
tion of kitchen biowaste for ethanol production via simultaneous saccharification and fermen-
tation using co-cultures of the yeasts Saccharomyces cerevisiae and Pichia stipitis. Bioresource
Technology, 263, 75–83. https://doi.org/10.1016/j.biortech.2018.04.109.
Oh, Y.-K., Seol, E.-H., Kim, J. R., & Park, S. (2003). Fermentative biohydrogen production by
a new chemoheterotrophic bacterium Citrobacter sp. Y19. International Journal of Hydrogen
Energy, 28, 1353–1359. https://doi.org/10.1016/S0360-3199(03)00024-7.
Rasmussen, M., & Minteer, S. D. (2014). Photobioelectrochemistry: Solar Energy conversion and
biofuel production with photosynthetic catalysts. Journal of the Electrochemical Society, 161,
H647–H655. https://doi.org/10.1149/2.0651410jes.
186 13 Electronegativity and Microbial Catalysis

Sakai, S., Nakashimada, Y., Yoshimoto, H., Watanabe, S., Okada, H., & Nishio, N. (2004). Ethanol
production from H2 and CO2 by a newly isolated thermophilic bacterium, Moorella sp. HUC22-1.
Biotechnology Letters, 26, 1607–1612. https://doi.org/10.1023/B:BILE.0000045661.03366.f2.
Siriwongrungson, V., Zeng, R. J., & Angelidaki, I. (2007). Homoacetogenesis as the alternative
pathway for H2 sink during thermophilic anaerobic degradation of butyrate under suppressed
methanogenesis. Water Research, 41, 4204–4210. https://doi.org/10.1016/j.watres.2007.05.037.
Sokolova, T. G., González, J. M., Kostrikina, N. A., Chernyh, N. A., Slepova, T. V., Bonch-
Osmolovskaya, E. A., et al. (2004a). Thermosinus carboxydivorans gen. nov., sp. nov., a new
anaerobic, thermophilic, carbon-monoxide-oxidizing, hydrogenogenic bacterium from a hot pool
of Yellowstone National Park. International Journal of Systematic and Evolutionary Microbiol-
ogy, 54, 2353–2359. https://doi.org/10.1099/ijs.0.63186-0.
Sokolova, T. G., Jeanthon, C., Kostrikina, N. A., Chernyh, N. A., Lebedinsky, A. V., Stackebrandt,
E., et al. (2004b). The first evidence of anaerobic CO oxidation coupled with H2 production by
a hyperthermophilic archaeon isolated from a deep-sea hydrothermal vent Extrem. Life Extreme
Cond., 8, 317–323. https://doi.org/10.1007/s00792-004-0389-0.
Sokolova, T. G., González, J. M., Kostrikina, N. A., Chernyh, N. A., Tourova, T. P., Kato, C.,
et al. (2001). Carboxydobrachium pacificum gen. nov., sp. nov., a new anaerobic, thermophilic,
CO-utilizing marine bacterium from Okinawa Trough. International Journal of Systematic and
Evolutionary Microbiology, 51, 141–149. https://doi.org/10.1099/00207713-51-1-141.
Sokolova, T., Hanel, J., Onyenwoke, R. U., Reysenbach, A.-L., Banta, A., Geyer, R., et al. (2007).
Novel chemolithotrophic, thermophilic, anaerobic bacteria Thermolithobacter ferrireducens gen.
nov., sp. nov. and Thermolithobacter carboxydivorans sp. nov Extrem. Life Extreme Condition,
11, 145–157. https://doi.org/10.1007/s00792-006-0022-5.
Sokolova, T. G., Kostrikina, N. A., Chernyh, N. A., Kolganova, T. V., Tourova, T. P., & Bonch-
Osmolovskaya, E. A. (2005). Thermincola carboxydiphila gen. nov., sp. nov., a novel anaerobic,
carboxydotrophic, hydrogenogenic bacterium from a hot spring of the Lake Baikal area. Inter-
national Journal of Systematic and Evolutionary Microbiology, 55, 2069–2073. https://doi.org/
10.1099/ijs.0.63299-0.
Sokolova, T. G., Kostrikina, N. A., Chernyh, N. A., Tourova, T. P., Kolganova, T. V., & Bonch-
Osmolovskaya, E. A. (2002). Carboxydocella thermautotrophica gen. nov., sp. nov., a novel
anaerobic, CO-utilizing thermophile from a Kamchatkan hot spring. International Journal of
Systematic and Evolutionary Microbiology, 52, 1961–1967. https://doi.org/10.1099/00207713-
52-6-1961.
Stal, L. J., & Moezelaar, R. (1997). Fermentation in cyanobacteria. FEMS Microbiology Reviews,
21, 179–211. https://doi.org/10.1016/S0168-6445(97)00056-9.
Tian, T., Qiao, S., Yu, C., Tian, Y., Yang, Y., & Zhou, J. (2017). Distinct and diverse anaerobic
respiration of methanogenic community in response to MnO2 nanoparticles in anaerobic digester
sludge. Water Research, 123, 206–215. https://doi.org/10.1016/j.watres.2017.06.066.
Uffen, R. L. (1976). Anaerobic growth of a Rhodopseudomonas species in the dark with carbon
monoxide as sole carbon and energy substrate. Proceedings of the National Academy of Sciences
of the United States of America, 73, 3298–3302.
van Dijken, J. P., & Scheffers, W. A. (1986). Redox balances in the metabolism of sugars by yeasts.
FEMS Microbiology Letters, 32, 199–224. https://doi.org/10.1111/j.1574-6968.1986.tb01194.x.
Wagman, D. D., Evans, W. H., Parker, V. B., Schumm, R. H., Halow, I., Bailey, S. M., et al. (1982).
The NBS tables of chemical thermodynamic properties. Selected values for inorganic and C1 and
C2 organic substances. Journal of Physical and Chemical Reference Data.
Wong, M. T., Cheng, D., Wang, R., & Hsing, I.-M. (2016). Modifying the endogenous electron fluxes
of Rhodobacter sphaeroides 2.4.1 for improved electricity generation. Enyzme and Microbial
Technology, 86, 45–51. https://doi.org/10.1016/j.enzmictec.2016.01.009.
Zeikus, J. G., Lynd, L. H., Thompson, T. E., Krzycki, J. A., Weimer, P. J., & Hegge, P. W. (1980).
Isolation and characterization of a new, methylotrophic, acidogenic anaerobe, the marburg strain.
Current Microbiology, 3, 381–386. https://doi.org/10.1007/BF02601907.
References 187

Zhao, T., Tashiro, Y., Zheng, J., Sakai, K., & Sonomoto, K. (2018). Semi-hydrolysis with low
enzyme loading leads to highly effective butanol fermentation. Bioresource Technology, 264,
335–342. https://doi.org/10.1016/j.biortech.2018.05.056.
Zürrer, H., & Bachofen, R. (1979). Hydrogen production by the photosynthetic bacterium Rho-
dospirillum rubrum. Applied and Environment Microbiology, 37, 789–793.
Chapter 14
Main Reactors Configurations

Who is aiming high, eventually, will more differentiate himself.


That is when addressing at the great book of Nature, which is
the own object of the philosophy; this is the way to rise our view.
In the same book, although all that is read, it is as done by an
Omnipotent Architect, and therefore the most proportionated.
Galileo Galilei, 1632, Dialogue Concerning the Two Chief
World Systems.

14.1 Introduction to Reactor Design and the Different


Configurations Used

This section focuses on what are the main reactors’ configurations and the modeling
of the residence time needed and therefore the size for the reactor necessary to
complete a wanted conversion of substrate to give the target products. It is good to
remember here that the volume of the reactor is represented by the volume of the
actual mass of reagents, products, and catalysts in the reactor.
To use a simple metaphor, the working principles of the chemical reaction engi-
neering could be applied also to normal daily life activities like cooking: you need a
suitable pot (the reactor) and the size of the pot is chosen on the basis of how much
food we would like to prepare, some good ingredients, and salts are also welcome.
You need oil or water and some time is needed before you can serve a nice dish.
The concepts of depletion of a substrate and production of a product can be applied
to any chemical engineering process. Among the general view of the chemical engi-
neering processes, the biochemical reaction engineering represents our particular
case here. The three main types of reactors used in chemical engineering are illus-
trated in Fig. 14.1.
Starting from the left of Fig. 14.1, in the batch reactor, there is no steady-state
conditions: the reagents and the following products are all processed in the reactor and
after a determinate time, the products are then obtained from it. The batch reactor
is also stirred, meaning that the concentrations of reagents and products are yes
changing with time, but in all areas of the reactor, the concentrations are the same.

© Springer Nature Switzerland AG 2019 189


C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_14
190 14 Main Reactors Configurations

Fig. 14.1 Three main reactors configuration for chemical engineering processes

Fig. 14.2 Example of plain axis in PFR reactor

Therefore, the concentrations of reagents and products have a time dependency, but
not a space dependency.
The plug flow reactor (Fig. 14.1—center), PFR, and the Stirred Tank Reactor
(Fig. 14.1—right), STR, are steady-state reactors; this means that reagents and prod-
ucts are entering and exiting the reactors with the same concentrations at all time
(when the steady state is reached). Nevertheless in the case of plug flow reactors
(also called tubular reactors), the concentrations of reagents and products are vary-
ing along the axes in the direction of the flow. While for the stirred tank reactor, there
is no dependency on the position into the reactor.
It has to be acknowledged that in the case of PFR, it is usually assumed that the
concentrations of the reagents and products are not varying along the direction which
is perpendicular to the flow direction (the one also perpendicular to the walls of the
tubular reactor). This is more clear considering Fig. 14.2 where the x-axes are the
one in the direction of the flow and the y-axes are the one perpendicular to it.
In this manuscript, we focus on how microorganisms are utilized in the production
of biofuels from a reactor design point of view. With reactor design, it is meant here
that we want to know the volume of the reactor (and its type) necessary to have a
particular mass flow and desired concentrations of the products.
14.1 Introduction to Reactor Design and the Different Configurations Used 191

The treatment of this theory is started here with the representation of microbial and
enzymatic fermentation. An organic feed is inserted in our reactor to give products.
An enzyme is used as a catalyst:

(organic feed, A) −→ enzime E(product chemical, R); (14.1)

(organic feed, A) −→ microbe C(product chemical, R + more cells, C). (14.2)

From Eqs. (14.1) and (14.2), we see that from an organic feed A, for instance
sugar, with the help of an enzyme, we produce a fuel R. In the first case, the enzyme
takes part to the formation of the fuel but the overall amount of it remains unchanged;
therefore, the enzyme is very well assumed to be a catalyst in this case. While in the
second case, we have that microorganisms’ cells are reproducing to form the fuel
product and also more cells.
For systems which are not with a constant density (the volume is varying), for
instance, when the number of moles in the system is changing by time (notice that
we are talking about moles here and not mass), we have by definition (Coker and
Kayode 2001):

Vx A =1 − Vx A =0
εA = ; (14.3)
Vx A =0

where ε A is the fractional change in volume between no conversion and complete


conversion of the reactant A. In case, the volume is linearly a function of the conver-
sion we have:

V = V0 (1 + ε A X A ); (14.4)

however, this information will be given by experiments and in theory could also have
another function. Considering Eqs. (14.3) and (14.4), we obtain:

CA 1 − XA
= . (14.5)
C A0 1 + εA X A

The demonstration of Eq. (14.4) is coming from the definitions of C A and C A0 ,


with Eq. (14.5) and it is as follows:
Remember that the conversion of an element A at some time t is defined as:
X A = 1 − nnA0A where nA and nA0 are the moles of A at time t and the moles of A
at time zero. This becomes X A = 1 − CCA0A when the volume (of reaction) remains
constant.
As a matter of fact:
nA
CA v
= n A0 ; (14.6)
C A0 v0
192 14 Main Reactors Configurations

and from Eq. (14.4)


v
v0 = ; (14.7)
1 + εA x A

Therefore, we obtain:
n A v0 nA v
· = · ; (14.8)
v n A0 v (n + ε A x A )n A0

which gives

CA (1 − x A )
= . (14.9)
C A0 1 − εA x A

From Eq. (14.9), we have

C A0 − C A
XA = ; (14.10)
C A0 + ε A C A

and therefore by differentiating X A with respect to C A , the following relations are


then obtained:

−C A0 (1 − ε A )
dx A = dC A ; (14.11)
(C A0 + ε A C A )2
(1 − ε A )
dC A = −C A0 · dx A . (14.12)
(1 + ε A X A )2

A little premise should be done here to explain in a very simple way, what is the
reaction rate order. In the following treatment of this theory, it will be referred at the
reaction rate as equal to:

Disappearance = (−r A )(Reactor Volume) = (moles of A reacting)/time.


(14.13)

This will be repeated later on when the derivation of the residence time will be
performed for different configurations. Now, it is important to notice that the reaction
rate is a function of the concentration of the reagents and in other words, it is directly
proportional to the concentration of the reagents, or products, or both. Let us assume
now that the moles of reagent A reacting over time is directly proportional to the
concentration of A. That will be:

(−r A ) = kC 1A ; (14.14)

Where the exponent 1 refers to the reaction order. In this way, for reactions of
zero order, we have:
14.1 Introduction to Reactor Design and the Different Configurations Used 193

(−r A ) = k; (14.15)

And so on. Notice also that the concentration of the substance is actually equal to
the number of moles on the volume of the reactor (which is the volume of substance
into the reactor, not the physical volume of the container) and this volume can change
significantly depending on the number of moles present in the system.
As an example of utilization of the fractional change in volume in our calculations,
one case is given here. The example is reported here with the complete solution for
the residence time calculation.
Consider that the reaction A → 3R is taking place. Concerning the feed: 50%
is formed by the A component and 50% is given by an inert compound. However,
let us assume that the 50% of the moles is inert, from the reaction we have that the
fractional change in volume is:

4−2
εA = = 1. (14.16)
2
Suppose that we want a conversion of 80% of the component A, therefore, for a
reactor working in the plug flow mode, PFR, the residence time is expressed as:

x A f x A f 1/2 0.8 1/2


dx A dX A C A0 1 + xA
τ = C A0 = C A0  0.5 = · dX A .
−Γ A 1/2 1−x A K 1 − XA
0 0 K C A0 1+ε A x A 0
(14.17)

The complete analytical solution of the integral is given here:


Let us do the transformation:

(1 − x A )1/2 = t. (14.18)

This will give:

1 − x A = t 2 → x A = 1 − t 2 → dx A = −2t dt; (14.19)

we substitute the expressions obtained in Eq. (14.17) and now the integral on the
right hand side of Eq. (14.17) looks as follows:

0.8  
(1 + 1 − t 2 )0.5
· (−2t) dt = −2 · 2 − t 2 dt; (14.20)
t
0

b
integrating by parts, which means, f  (x)g(x) dx = [ f (x)g(x)]|ab −
b 
a

a f (x)g (x) dx
194 14 Main Reactors Configurations

This gives:
     
t2
−2 t · 2− t2 + √ dt = −2 2 − t 2 dt. (14.21)
2 − t2
dArcsenx
Notice that the derivative of dx
is:

dArcsenx 1
=√ ; (14.22)
dx 1 − x2

therefore:
    
t2 2 − E2 − 2
2− t 2 dt =t · 2−t + √
2 dt → − √ ; (14.23)
2 − t2 2 − t2

and
    
√ 1
2 − t z dt = t · 2 − t2 − 2 − t2 + 2 √ dt; (14.24)
2 − t2

from which we have:


   
1
2 2 − t 2 dt = t · 2 − t 2 + 2 √ dt; (14.25)
2 − t2

this will give:


 
1  t 0.8
2 − t 2 dt = t · 2 − t 2 + Arcsin √ . (14.26)
2 2 0

Considering again that we have done some transformations in Eqs.


(14.18)–(14.19), the final formula for the residence time at these conditions will
be:
0.8
C
1/2
1  (1 − x A )1/2
τ = A0 (1 − x A )2 · 2 − (1 − x A ) + Arcsin √ . (14.27)
K 2 2 0

14.2 Performance of a Batch Stirred Tank Reactor

In the batch stirred tank reactor, BSTR, as previously mentioned, there is no inlet
and outlet during the reaction process. The feedstock is inserted together with the
reagents in the reactor and then only after some time has passed, the products are
14.2 Performance of a Batch Stirred Tank Reactor 195

Fig. 14.3 Illustration of the


batch stirred tank reactor

then recovered from the reactor. The illustration of this kind of reactors is given in
Fig. 14.3.
In this case, the general balance of moles for our system is:

In − Out − (or +)disappearance = accumulation. (14.28)

Notice that in Eq. (14.28), the term “disappearance” can have a minus or plus sign
depending if the number of moles is increasing or decreasing. I would like to stress
here that this is valid in case we are talking about number of moles and not mass. As
it is well known from the thermodynamics principles, the mass remains unchanged
during the process. However, the number of moles of the system can indeed change.
For instance, if we consider a simple reaction like C+O2 = CO2 we notice that from
two moles (one of carbon and one of CO2 ), we obtain one mole. This is particularly
important when the volume of the system is taken into account: we remember that in
gas phase, for instance, the volume of the gas is directly proportional to the number
of moles, i.e., if we have half moles, the volume occupied will be much decreased.
Let us look at all the terms given in the balance:
 
moles
Input of A = FA0 ; (14.29)
time
Out(A) = FA0 (1 − X a ); (14.30)

where X a is the mole fraction of the A component. The disappearance will be:

mol A reacting
Disappearance = (−r A )V = Volumereactor . (14.31)
time
In the case of batch mode of operation, we do not have inlet and outlet of material;
therefore, in this case, the disappearance is equal to the accumulation of moles:
196 14 Main Reactors Configurations

dn A
(−r A )V = ; (14.32)
dt
and:

dC A
= −r A ; (14.33)
dt
Therefore,

dC A
dt = . (14.34)
−r A

At this point, it is good to remember the definition of conversion:

(C A0 − C A )
xA = . (14.35)
C A0

For this reason, the expression for the residence time in this case is:

tfinal X final
dx A
τ= t = C A0 . (14.36)
−r A
0 x A0

It has to be noticed that, in order to solve the integral, the expression for the
term (−r A ) is needed and (−r A ) is a function of the kinetic expression for what it
was called before “disappearance.” Later, this will become more clear reading the
following pages.

14.3 Performance of a Continuous Stirred Tank Reactor

In a continuous stirred tank reactor, there is inlet of feedstock and outlet of products
(Fig. 14.4) and in this case, the inlet and the outlet are proceeding within steady-state
conditions.
As done previously, we start from the general mass balance:

In − Out − disappearance = accumulation; (14.37)

and in the case of a continuous stirred tank reactor, we have:

Input of A (moles/time) = FA0 ; (14.38)

Out (A) = FA = FA0 (1 − X A ); (14.39)


14.3 Performance of a Continuous Stirred Tank Reactor 197

Fig. 14.4 Simple


schematics of a STR

mol A reacting Vreactor


Disappearance = (−r A )V = . (14.40)
time Volume fluid
In this case, the term related to the accumulation in the reactor is missing because
there are steady-state conditions. This means that at any time, we have the same
inlet flow, the same outlet (at the exit), and the concentrations in the reactor are not
changing with time. Therefore, the balance will give:

V XA
FA0 X A = (−r A )V → = (14.41)
FA0 −r A

And the residence time to have a final conversion X A will be:

V V C A0 C A0 X A
τ= = = ; (14.42)
v0 FA0 −r A

where
V 0 volumetric inlet
V reactor volume, (notice that, here we intend the volume of the substance into
the reactor)
C A0 −C A
For first-order chemical reaction with no change in volume: τ = kC A
and this
is because −r A = kC A .
198 14 Main Reactors Configurations

14.4 First-Order Chemical Reaction in Stirred Tank


Reactor with Changing Reaction Volume

In this section, the case of a first-order chemical reaction is taken into account for STR
in the case of change in volume over time. The volume of substance into the reactor
is directly proportional to the amount of moles within the reactor and therefore, since
the amount of moles can vary significantly, the volume of the reactor (the substance
into the reactor) can vary also.
We start from the definition of the expansion factor (Davis and Davis 2003):

Vx A=1 − Vx A=0
ε= ; (14.43)
Vx A=0

where x A = 1 indicates that the conversion of the component A (sugar for instance)
is one, while for x A = 0, we have a conversion of zero and this takes place at the
reactor inlet. It will be shown here that for first-order reaction and the case of STR,
the residence time is expressed as follows:

X A (1 + ε A X A )
τ= for any ε A . (14.44)
k(1 − X A )

This is demonstrated as follows, the mass balance done previously gives:

V V C A0 C A0 X A
τ= = = ; (14.45)
v0 FA0 −r A

remember that X A is the conversion at the exit and also the conversion measured
hypothetically in the reactor volume. This is because the concentration at the exit
of the reactor is the one which is formed in the reactor after the residence time has
passed. In case we assume that the reactor slurry volume has a linear dependence
from the initial volume we have:

V = V0 (1 + ε A X A ). (14.46)

Substituting Eq. (14.46) in Eq. (14.45), will give:

CA 1 − XA
= ; (14.47)
C A0 1 + εA X A

for first-order reaction and the case of continuous stirred reactor, we have previously
written:

V C A0 X A
τ= = ; (14.48)
v0 kC A
14.4 First-Order Chemical Reaction in Stirred Tank Reactor … 199

this gives:

C A0 X A (1 + ε A X A )
kτ = XA = . (14.49)
CA 1 − XA

Now, let us consider the more complicated case in which the reaction is of second
order:

−r A = kC 2A ; (14.50)

in this case:

V C A0 X A C A0 X A
τ= = = ; (14.51)
v0 −r A kC 2A

from which

C A0 X A (1 + ε A X A ) X A
kτ = = . (14.52)
C AC A (1 − X A ) C A

Remember that:

CA 1 − XA
= ; (14.53)
C A0 1 + εA X A

and therefore,

(1 + ε A X A ) XA (1 + ε A X A )2
kτ = · C A0 (1−X A )
= . (14.54)
(1 − X A ) (1+ε A X A )
(1 − X A )2 C A0

14.5 Plug Flow Reactor

As mentioned previously, in the PFR model, we have steady-state conditions for the
inlet and the outlet. This means that the concentrations and the mass flow at the inlet
and outlet are always the same as regards to time. However, even if it is true that
the concentrations are not varying with time for any of the sections of the reactor,
the conversion and the concentrations are changing along with the axes in the flow
direction (see Fig. 14.2).
Notice here that we have a similitude with the chromatographic columns treated
in the first part of this book, in that case also, a residence time in the chromatographic
column was associated to an equivalent length of the column itself. It is good to keep
in mind that the more will be the reactor length, the more will be the residence time
200 14 Main Reactors Configurations

Fig. 14.5 Illustration of a PFR and its differential change in conversion and concentration with
respect to the compound A along the distance from the inlet axes

and the conversion in the reactor; therefore, length and time go hand to hand in this
case (Fig. 14.5).
Also in this case, we do the general mass balance but this time for a differential
section in the tubular reactor. The concentration of each component is varying with
the length of the reactor: considering two sections along with the flow direction,
x1 and x2 , if we consider one fixed reaction section, the concentrations will not vary
with time, nevertheless if we focus on the next section then the concentrations will
be different with respect to the previous section.
As it was done before:
In – Out + Generation = 0 (the accumulation is 0 in steady state) and:

FA = (FA + dFA ) + (−r A )dV. (14.55)

Also in this case, there is no accumulation in the differential volume, which means
that at any fixed section of the PFR, the concentrations are the same at any time.
Considering:

dFA = d[FA0 (1 − X A )] = −FA0 dX A ; (14.56)

we obtain:

FA0 dX A = (−r A )dV ; (14.57)


14.5 Plug Flow Reactor 201

therefore:

X A f
V V C A0 dX A
τ= = = C A0 . (14.58)
v0 FA0 −r A
0

This is valid for constant density systems. For first-order irreversible homogeneous
reactions:

−r A = kC A ; (14.59)

this gives:

X A X A
dX A C A0 dX A
τ = C A0 · = ; (14.60)
kC A kC A
0 0

considering Eq. (14.53), we have:

CA 1 − XA
= ; (14.61)
C A0 1 + εA X A

and this will help us by giving:

X A
(1 + ε A X A )
kτ = dX A . (14.62)
(1 − X A )
0

Solving the integral:

X A X A X A
1 εA X A 1
kτ = dX A + dX A = dX A
1 − XA (1 − X A ) 1 − XA
0 0 0
X A
−X A
− εA dX A (14.63)
(1 − X A )
0

Which gives:

X A X A X A
1 1 − XA − 1 1
kτ = dX A − ε A dX A = dX A
1 − XA (1 − X A ) 1 − XA
0 0 0
202 14 Main Reactors Configurations

X A X A
1
− εA dX A + ε A dX A (14.64)
1 − XA
0 0

Equation (14.65) is derived:

kτ = − ln(1 − X A ) − ε A X A − ε A ln(1 − X A ) = −(1 + ε A ) ln(1 − X A ) − ε A X A .


(14.65)

14.6 The Case of Reversible Reaction

Consider the following case:

A←
k2
→k r R;
1
(14.66)

where r is the stoichiometric coefficient for the product R, k 1 the forward chemical
reaction constant, and k 2 the backward chemical reaction constant. Then for first-
order chemical reactions we have:

−r A = k1 C A − k2 C R . (14.67)

For convenience, it is possible to define the following variable M:

C R0
= M; (14.68)
C A0

consider also that x A is the conversion at equilibrium and as usual remember the
relation:

CA 1 − XA
= . (14.69)
C A0 1 + εA X A

Keeping this in mind:

r (C A0 − C A ) = (C R − C R0 ) → C R = r [C A0 − C A ] + C R0 ; (14.70)

the following result is obtained

CR CA C R0 CR (1 − x A )
=r 1− + → =r 1− + M; (14.71)
C A0 C A0 C A0 C A0 (1 + ε A x A )

and then:
14.6 The Case of Reversible Reaction 203

1 1
rC A = rC A0 − C R + C R0 → C A = C A0 − C R + C R0 . (14.72)
r r
We substitute the expression for C R obtained, Eq. (14.71), and we have:
 
1 (1 − X A ) 1
C A = C A0 − r · C A0 1 − + C A0 · M + · C R0 ; (14.73)
r (1 + ε A X A ) r

at the equilibrium:

k1 CR
k 1 C A = k 2 Cr → =K = ; (14.74)
k2 CA

where K is the equilibrium constant. After this, Eq. (14.75) is obtained:


    
(1−x A )
CA C A0 · 1
r
rC A0 1 − (1+ε A A )
X
+ C A 0
· M + 1
r
C R 0
k2 = k1 · = k1 ·     .
CR C · r 1 − (1−x A ) + C · M
R0 (1+ε A x A ) A0

(14.75)

In these equations previously written, the x A refers to the fraction at equilibrium.


At this point, we have that k2 = f (k1 ), where f means “function of”. For first-order
forward and backward chemical reactions we have:

dC A
= k1 C A − k2 C R → k2 = k1 · f (X A ). (14.76)
dt
Remember, as usual, the relation:

CA 1 − XA
= ; (14.77)
C A0 1 + εA X A

Therefore, integrating the resulting expression, it will be possible to obtain k 1 as


a function of the fractional change in volume, the fraction of the feedstock, etc.

k1 = f (ε A X A , . . .); (14.78)

in the same way, from the relation:

k1 CR
= ; (14.79)
k2 CA

we have that k2 is a function of the other variables and parameters:

k2 = f (ε A X A , . . .); (14.80)
204 14 Main Reactors Configurations

Fig. 14.6 Inverse of the reaction rate as a function of the conversion (left) and concentration (right)
for the component A

and it will be possible to evaluate τ .

14.7 Performance Equations for Plug Flow Reactors

For PFR, the mass balances can be illustrated by plotting the conversion against the
reaction rate. In Fig. 14.6, it has to be noticed that when we consider the concentra-
tions (plot on the right), the volume of the reactor (with reactor volume it is meant
what is inside the reactor) is considered to remain constant, in the drawing on the
right, the concentration of the feedstock should be read from the right to the left
(C A0 –C A ) while in the drawing on the left, the conversion goes from zero at the inlet
and goes to X A at a generic time t.
Before proceeding further, it is suggested to remember that: x A = 1 − nnA0A this
can be considered as: x A = 1 − CCA0A when the volume (of reaction) remains constant.
The residence time for PFR is expressed as follows:

X A f
V V C A0 dX A
τ= = = C A0 . (14.81)
v0 FA0 −r A
0

For zero-order homogeneous reactions in a plug flow, we have:

kC A0 V
τ= = C A0 X A . (14.82)
FA0

This is because for reactions of order zero, −r A = k. the reaction rate is a constant.
14.8 Autocatalitic Reactions 205

14.8 Autocatalitic Reactions

Microorganisms, which are used to produce biofuels like hydrogen for instance, uti-
lize a determinate feedstock to reproduce and to discard a product which is our biofuel
of interest. This category of processes can be represented by so-called autocatalytic
reactions of this kind:

A + C → C + C. (14.83)

where C is the concentration of microorganisms in the system and A is as usual


our feedstock, for instance sugar. The general case of reaction rate for this kind of
reactions is described by:

−r A = kC aA Ccc . (14.84)

Consider now the general equation for the residence time of A in PFR, Eq. (14.81)
and the definition of conversion when there is no volume change in the reactor:
x A = 1 − CCA0A . Equation (14.35), when the stoichiometric coefficients are both equal
to one becomes:

−r A = k · C A0 (1 − X A ) · C A0 (1 + X A ). (14.85)

This would be reasonable since, in the assumptions taken here, when there is total
conversion of moles for the nutrient (sugar for instance), the overall number of moles
for the C component (the microorganism) will be doubled. Therefore, Eq. (14.81)
becomes now:

X A f X A f
V V C A0 dX A dX A
τ= = = C A0 = C A0
v0 FA0 −r A k· C 2A0 · (1 − X A )(1 + X A )
0 0

X A f
1 dX A
= (14.86)
k · C A0 (1 − X A )(1 + X A )
0

At this point, the integral on the right side of Eq. (14.86) will be solved considering:

1 1 − XA + XA 1 2 · XA
= = + ;
(1 − X A )(1 + X A ) (1 − X A )(1 + X A ) (1 + X A ) 2(1 − X A )(1 + X A )
(14.87)

where
(1 − X A )(1 + X A ) = X 2A + 1.
Now Eq. (14.86) will be given by Eq. (14.88):
206 14 Main Reactors Configurations

Fig. 14.7 General trend of the reaction rate as a function of the conversion for autocatalytic reactions

⎡ ⎤
X A f X A f
1 ⎢ 1 1 2X A ⎥
τ= ⎣ + ⎦; (14.88)
k · C A0 (1 + X A ) 2 X 2A + 1
0 0

which will give:

1   
τ= · ln 1 + X A f X 2A f + 1 . (14.89)
k · C A0

The general trend of the reaction rate for the sugar depletion is described in
Fig. 14.7.

14.9 Optimal Choice of Reactors

At this stage, it is good to remember that for stirred tank reactors, the residence time
is:

V V C A0 C A0 X A
τ= = = ; (14.90)
v0 FA0 −r A

while for plug flow reactors, we have:

X A f
V V C A0 dX A
τ= = = C A0 ; (14.91)
v0 FA0 −r A
0

therefore, we see that the bigger is the residence time, the bigger is the reactor volume.
For stirred reactors, Eq. (14.92) is valid:
14.9 Optimal Choice of Reactors 207

Fig. 14.8 Comparison between reactor volumes of PFR and STR at a fixed conversion. Optimal
choice of the reactors for a given conversion with autocatalytic reactions

V XA
= . (14.92)
FA0 −r A

From what was derived previously, it is possible to see very clearly that the ratio
V
FA0
is proportional to the integral area of the curve described by −r1 A in the first case,
while it is the area of the rectangle given by −r1 A · X A in the second case (Fig. 14.8).
Therefore, for auto-catalytic reactions, it is better to use a stirred tank reactor for
low conversions and a plug flow reactor for high conversions.

References

Coker, A. K., Kayode, C. A. (Eds.). (2001). Introduction to reactor design fundamentals for ideal
systems (Chap. 5). In Modeling of chemical kinetics and reactor design (pp. 260–423). Woburn:
Gulf Professional Publishing. https://doi.org/10.1016/B978-088415481-5/50007-3.
Davis, M. E., Davis, R. J. (2003). Fundamentals of chemical reaction engineering. McGraw-Hill
Higher Education.
Chapter 15
Enzyme Kinetics

Pavia, 12. August. 1895.


Very Illustrious Professor!.
Sir Eng. Vitali coming back from Turin referred to me that he
had the precious occasion of communicate with Your Illustrious
Lordship and that my name was mentioned there. Seen such
benevolence from you, it makes me bold to prey you to write me
a little recommendation letter to be presented to Professor
Weber of Zurich, who I would like to visit before the study
courses will be opened at that university. This is to overcome the
difficulties of attending that program due to my young age.
Thanking you in advance and with the highest consideration of
Your Illustrious Lordship.
Devotedly, Alberto Einstein.
A letter written by Albert Einstein to Galileo Ferraris (known
as the Father of three-phase current) when he was only sixteen.
The letter was published by the Italian journal of finance, Il Sole
24 Ore on the 27th of Nov. 2005.

15.1 Product Formation in Enzyme Catalyzed Biological


Processes

The production of biofuels largely involves the utilization of enzymes and therefore,
from this point of view, it would be beneficial to know more on the way they actually
work. Even if the scope of this book is not to focus on the biological aspects related
to the formation, production, and conservation of such entities, it will be mentioned
here that they could be utilized in a variety of processes including the production
of biodiesel where they are used as alternative catalysts. As a preliminary example
and to underline the importance of the enzymes usage in the production of biofuels,
Fig. 15.1 describes the production of free fatty acids, FFA, and fatty acid methyl
esters, FAME from castor oil by means of enzymes catalyzed route.
The process of biodiesel production will be described later in this manuscript;
however, it is good here to have an idea of the importance of reactor design and
process integration in this kind of productions. Enzymes represent also a big part

© Springer Nature Switzerland AG 2019 209


C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_15
210 15 Enzyme kinetics

Fig. 15.1 Transesterification of castor oil driven with enzymes as catalysts. Adapted from Andrade
et al. (2019)

of the expense related to the biofuel generation; therefore, it is a big interest for
the companies working in this field to have an optimized system where the reactors
volumes are kept to a minimum.
Enzymes are indeed catalysts for the reactions in which they are involved. When
modeling the reactions yields and rates, their total concentration is assumed to be
constant in the sense that their total number will be preserved as it will be shown later.
Of course, this is valid in the theoretical treatment of the process and in practice, a
loss of enzymes is always taking place.
It is reasonable to consider the concentration of enzymes constant, because the
enzymes, like catalysts, are imagined to interact with the substrate, facilitate the
reactions involved, and then they liberate the products. After this, they can be used
for one additional reaction. The enzyme reacts with the substrate and gives the
products P, according to:

k1 ,k−1 k2
E + S ←→ C → E + P; (15.1)

where
E enzyme concentration
S substrate concentration
K 1 , −1, 2 reaction rate constant [1/time for first order]
P product concentration
C [E·S] complex
In this way, if we consider the last reaction as a first-order chemical reaction:
15.1 Product Formation in Enzyme Catalyzed Biological Processes 211

dP
= k2 C (15.2)
dt
and considering the same assumption we have:

dC
= k1 E · S − k−1 C − k2 C. (15.3)
dt
Assuming that the complex C is soon converted in P and E we obtain:

dC
∼ 0. (15.4)
dt
To be noticed that this does not mean that the concentration of C is zero but
just that the concentration of the complex remains constant. To be noticed also
that the equation above written are simply stating that the variation of P is directly
proportional to the concentration of C and that the concentration of C is proportional
to E, S, and P taking into account the amount of this terms that is actually disappearing
and what is the element which is produced instead.

15.2 Enzime Kinetics. The Michaelis–Menten Model

Within this model, the enzyme is reacting with a substrate giving an enzyme–substrate
intermediate. This gives again the free form of the enzyme and the product according
to:

k1 ,k2 k3
E + S ←→ E S → E + P; (15.5)

where
E enzyme
S substrate
ES enzyme–substrate
The rate of reaction written for the products is:

dC p
r= = k3 C E S . (15.6)
dt
Let us assume that the enzyme–substrate is in equilibrium with both: the enzyme
and the substrate according to:

k1 C s C E = k2 C E S . (15.7)

In addition, assume that the total concentration of enzymes is conserved


212 15 Enzyme kinetics

C E0 = C E + C E S ; (15.8)

therefore

C E0 C S
CE S = ; (15.9)
k2
k1
+ Cs

where
k2
k1
= K m is the dissociation constant (equilibrium constant).
Substituting these into the rate of formation of the product, Eq. (15.6), we obtain:

dC p C E0 C S Vmax C S
= V = k3 = . (15.10)
dt Km + CS Km + CS

It is understandable that in the beginning of the reaction, we have the max produc-
tion of P over time, this equals k3 C E0 . This is because at the start (at time zero), we
have that the enzyme is utilizing all the substrate (for instance sugar) at its disposition
and therefore, the ES complex is at its maximum concentration.

15.3 Enzime Kinetics, Briggs–Haldane Approach

Also in this case, the enzyme is reacting with a substrate giving an enzyme–substrate
intermediate which in turn gives again the free form of the enzyme and the product
according to:

k1 ,k2 k3
E + S ←→ E S → E + P. (15.11)

The rate of formation for the products will be:

dC p
r= = k3 C E S . (15.12)
dt
In this case, we assume that the change of C E S with time is really small compared
to the change in the concentrations of the product and substrate, C p and C s .

dC E S
= k1 C S C E − k2 C E S − k3 C E S = 0. (15.13)
dt
It is assumed, like previously, that the total concentration of enzymes is conserved

C E0 = C E + C E S ; (15.14)

therefore:
15.3 Enzime Kinetics, Briggs–Haldane Approach 213

C E0 C S
CE S = k2 +k3
; (15.15)
k1
+ CS

and we obtain the rate of formation for the products as a function of the concentration
of the substrate (sugar) and the initial concentration of enzyme:

dC p C E0 C S Vmax C S
= V = k3 = ; (15.16)
dt Km + CS Km + CS

with

K2 + K3
Km = ; (15.17)
K1

and

Vmax = k3 C E0 ; (15.18)

If k2  k3, then the limiting reaction is the last one and K m becomes the same as
the Michaelis–Menten constant.
Consider a mixed flow reactor where fermentation occurs and let us focus, in this
case, on the development of the microorganisms. The production rate of microbes
is:

kC A CC
rC = ; (15.19)
C A + CM

where
C A concentration of food
C C concentration of cells
C M concentration of A when the cells reproduce at ½ of their maximum rate
with the cell growth rate as:

dC
rc = ; (15.20)
dt
a simplified drawing of this case is shown in Fig. 15.2.
214 15 Enzyme kinetics

Fig. 15.2 Simple


schematics showing the
major components in
fermentation

15.4 The Monod Model

From here on, we will take into consideration how the reaction (conversion) rate
is related to other variables of our system. The variation of the concentration of
microorganisms is affected by the same concentration of microbes (c = cells) and by
the concentration of sugar. With a good probability the most used model for the cell
growth rate is the Monod model. The model is named after Jacques Monod who at
the time of his research was working at the Pasteur institute in Paris, France (Monod
1949). In this model, the cell growth rate is expressed as follows:

kC A CC
rC = ; (15.21)
C A + CM

where
rc = dCc
dt
cell growth
CA concentration of food
CC concentration of cells
CM concentration of A where the cells reproduce at ½ of their maximum rate.
In case, the concentration of cells would be proportional to the amount of food
(which is actually reasonable), C A = k1 Cc , the solution for Eq. (15.21) would be
obtained as follows:

dC k · k1 C c CC
= ; (15.22)
dt k1 C c + C M

with a = k · k1 , and b = C M , Cc = C, Eq. (15.22) becomes:

dC aC 2
= . (15.23)
dt k1 C + b

Rearranging we have:
15.4 The Monod Model 215

Fig. 15.3 Main phases in


the variation of the cell
concentration with time.
Adapted from Levenspiel
(1999)

 
k1 C + B
dC = A · dt; (15.24)
C2

Which will give:

C C t
k1 1
+B = A· dt; (15.25)
C C2
C0 C0 0

and finally:
 
C 1 1
k1 ln +B − = A · t. (15.26)
C0 C0 C

Figure 15.3 describes in a general way the variation of the cell concentration
with time, it is possible to notice three phases, a period in which time is needed to
start reproducing (called lag time), an exponential growth, in which the cells utilize
effectively the food given, a stationary phase, and a phase where the cells start dying
(called decay phase).
While in Fig. 15.4, it is possible to notice that the production rate of product
follows different curves on the basis of if we have high food concentration or low
food concentration.
We consider:
A substrate, (also indicated as S in some references)
E enzyme
R product (also indicates as P in some references)
216 15 Enzyme kinetics

Fig. 15.4 Production rate of product as a function of the substrate concentration. Adapted from
Levenspiel (1999)

At high concentrations of the substrate A, we have that the rate is not dependent on
the substrate concentration and this is because if the substrate is a large quantity, then
it will not matter, it will remain a large quantity anyway. At low concentration of the
substrate, the rate of conversion is proportional to the concentration of the substrate
and therefore, the order of reaction can be considered to be of the first order.

15.5 Main Characteristics of the Briggs–Haldane Equation

The Briggs–Haldane approach (Briggs and Haldane 1925), also known as the quasi-
steady-state approach to enzyme kinetics, is also used to model the product yield for
biochemical processes; this approach is very much similar to the Monod reasoning
or the Michaelis–Menten one (Michaelis and Menten 1913).
To be noticed that in some references, the products are indicated with the letters
R or P while r max has the same meaning of our previous V max, and sometimes K m
has the same meaning of C M , the substrate is written as C s sometimes instead of C A.
With this in mind, we have:

dR C E0 C A Vmax C A
= V = k3 = (15.27)
dt Cm + C A Cm + C A

Consider that:

dR
rR = (15.28)
dt
15.5 Main Characteristics of the Briggs–Haldane Equation 217

Fig. 15.5 Feedstock consumption rate function and its main parts

a detailed representation of the function for the feed consumption rate is represented
in Fig. 15.5.
In this manuscript, the detailed explanation of the terms shown in figure is derived.
At C A = 0, the initial slope is calculated by the following limit:
  
d k3 C E0 C A
lim (15.29)
C A →0 dC A C m + C A

dC
while at C A = C M, we substitute this value in the expression for dt p above to
obtain r. When C A tends to infinite we have an indeterminate form of the −r A
rate since we have an infinite value which divides an infinite value. Applying the
theorem of Guillaume de l’Hôpital, we readily obtain:

−r A will be rMax = k3 C E0 . (15.30)

Considering the expression:

dA E0 A Vmax · A
− = V = k3 = ; (15.31)
dt Cm + A Cm + A

we can solve easily the differential equation between an initial time and a general
time t, therefore:

C A0
C M ln + (C A0 − C A ) = k3 C E0 t (15.32)
Ca

Rearranging, we obtain a form which can be plotted. As shown in Fig. 15.6.


218 15 Enzyme kinetics

Fig. 15.6 Linearized function for the concentration of the substrate as a function of time to find
the constant parameters

C A0 − C A t
= −C M + k3 C E0 (15.33)
ln CCA0A ln CCA0A

Equation (15.33) allows us to have a linear form where it is more easy to find the
key constants C M and k3 .
To better understand the terms in the drawing above, when the concentration of the
substrate is equal to C A0 , we are facing again an indeterminate form at the ordinate
value, by solving this limit we obtain:
 
C A0 − C A
lim = C A0 . (15.34)
C A →C A0 ln CCA0A

After deriving this, it will be possible to evaluate the abscissa from:

C A0 − C A t
= C A0 = −C M + k3 C E0 . (15.35)
ln CCA0A ln CCA0A

15.6 Residence Time in a Mixed Flow Fermentor

From the original form of the balance on the substrate depletion,

dA E0 A Vmax A
− = V = k3 = ; (15.36)
dt Cm + A Cm + A
15.6 Residence Time in a Mixed Flow Fermentor 219

Fig. 15.7 Linearized form of the substrate concentration as a function of the residence time expres-
sion at low and high enzyme concentration

we remember that the residence time in a continuous stirred tank reactor (with no
volume change between t 0 and t) is:

C A0 − C A (C A0 − C A )(C M + C A )
τ= → k3 C E0 τ = ; (15.37)
−r A CA

where C M is the Michaelis–Menten constant. Rearranging we have:


 
C E0 C A τ
C A = −C M + k3 ; (15.38)
C A0 − C A

From Fig. 15.7, it is possible to see how C M can be calculated from linearized
forms of the C A concentration. The slope of the straight line will change according
to the initial enzyme concentration since this depends on that concentration. In the

same figure, notice that the C A concentration is plotted against C A0C −C A
and therefore,
it is quite simple to derive the terms shown on the abscissa axis: for C A = C A0 , we
obtain one value of the abscissa, for C A = 0 another value is obtained and the slope
of the curve is obtained by simply applying the property of the straight lines.

References

Andrade, T. A., Martín, M., Errico, M., & Christensen, K. V. (2019). Biodiesel production catalyzed
by liquid and immobilized enzymes: Optimization and economic analysis. Chemical Engineering
Research and Design, 141, 1–14. https://doi.org/10.1016/j.cherd.2018.10.026.
220 15 Enzyme kinetics

Briggs, G. E., & Haldane, J. B. (1925). A note on the kinetics of enzyme action. Biochemical
Journal, 19, 338–339.
Levenspiel, O. (1999). Chemical reaction engineering. New York, US: Wiley.
Michaelis, L., & Menten, M. (1913). Die kinetik der invertinwirkung. Biochemistry Zeitung, 49,
333–369.
Monod, J. (1949). The growth of bacterial cultures. Annual Review of Microbiology, 3, 371–394.
Chapter 16
Balances on Microbial Fermentation

And then you see that sometimes, to undermine the false


authority of an absurd proposition which is against any
reasoning, even a laughter can be a right instrument. Often the
laughter also serves to confuse evil persons and make their
foolishness shine.
Umberto Eco, 1980. The Name of the Rose, Fabbri publisher.

16.1 Cell Balances

The microbial fermentation is a complicated series of several processes, and each


of them could be represented by some form of reactions, and however, in our case
the scope of this textbook is not to explain all the biological steps involved but
simply model the entire process to have the minimum level of accuracy in order to
dimension the reactors used. In a fermentation process, a substrate A, for instance
sugar, is consumed by microorganisms who will reproduce (until some point) and
produce the product R as follows:

C
A → C + R; (16.1)

where we recall

A = substrate
C = the living organism
R = product

This can be rewritten also as:

A + C → C + R; (16.2)

and we could consider this as an autocatalytic reaction (in case we do not make
a difference between C and R). In other words, a simplification for Eq. (16.1) is
possible if we consider no difference between the microorganisms and the products

© Springer Nature Switzerland AG 2019 221


C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_16
222 16 Balances on Microbial Fermentation

A + R → R + R. (16.3)

Considering the mass balance on the substrate:

dC A
−r A = − = kC A C R ; (16.4)
dt
the sum of the concentration of C A and C R is constant. This is an assumption: It
means that is the substrate A is decreasing, the product R is actually increasing, and
this is most likely true. Therefore:

C0 = C A + C R = C A0 + C R0 = constant; (16.5)

and

dC A
− = kC A (C0 − C A ). (16.6)
dt
Integrating between a time zero and a general time t:

C R /C R0
ln = C0 kt = (C A0 + C R0 )kt. (16.7)
C A /C A0

In terms of fractional conversion X A , the equation obtained can be written as:

M + XA
ln = C A0 (M + 1)kt = (C A0 + C R0 )kt; (16.8)
M(1 − X A )

with

C R0
M= ;
C A0

and

(C A0 − C A )
xA = .
C A0

Figure 16.1 is a more detailed version of Fig. 5.3 and reports the conversion of the
substrate against time for autocatalytic reactions of this kind, it is possible to notice
that at the beginning of the process, the conversion remains quite low with time, and
this is because the microorganisms utilize in the beginning only a fraction of the feed
at their disposal. In other words, it is like they have to “realize” that there is food for
them. After this phase, microorganisms utilize the food in a very effective way also
because they reproduce at this time. Finally, the consumption rate decreases and this
16.1 Cell Balances 223

Fig. 16.1 Conversion of the substrate against time for autocatalytic reactions

could be because of two reasons: One is that the food is decreasing, or on the other
hand that the microorganisms are actually dying.

16.2 Models for Microbial Growth Used in Power


Generation

Concerning the rate of biomass formation, as we have seen previously, for first-order
chemical reaction we have that the rate of formation is equal to a reaction rate constant
multiplied by a concentration; that is:

rC = k · C; (16.9)

where C is the concentration of biomass expressed as g/L.


In case we follow the Monod model, this can be expressed as:

kmax Cs
k= ; (16.10)
K s + Cs

with

k specific growth rate coefficient (check the dimensions always on the basis of
the reaction order);
k max maximum specific growth rate (the equivalent of V max cited previously);
Cs substrate concentration (this refers usually to the substrate that is soluble and
it is the equivalent of A mentioned previously);
Ks half saturation constant.

Inhibition can be provoked also by the same substrate (Edwards 1970), for instance
when the concentration is too high. In case we have substrate inhibition, the specific
growth rate coefficient can be expressed as:
224 16 Balances on Microbial Fermentation

kmax Cs
k= Cs2
(16.11)
K s + Cs + KI

where K I is one inhibition coefficient.


The substrate actually can inhibit the process and slow down the cell growth, and
therefore, the inhibition coefficient is taken into account.
Product inhibition is quite common, for instance in the case of ethanol production,
the ethanol itself is inhibiting the further production and the concentration of ethanol
in these kinds of processes cannot go further than fixed values (Luong 1985). In this
case, the reaction rate constant is given by:
 
kmax Cp n
k=  1− (16.12)
1 + CKss Cpcrit

Here, C p is the product concentration, while the C pcrit is the concentration of products
which will cause inhibition of the growth. The constant n is called inhibition factor.
It is also possible to have inhibition from other compounds, and in this case, the
modeling is as follows:

kmax Cs
k=   (16.13)
Cs 1 + CKII + K s

with C I as the concentration of the inhibitor, while K I is a constant related to the


inhibition. In this textbook, it is preferred not to give explicitly the dimensions of C I
or K I because they could vary depending on the mathematical model adopted. For
instance, they could have the dimensions of mg/L, but not necessarily this can be
always the case.
In addition to the cases presented above, it is also possible to have inhibition from
complementary substrates (Shijie 2016); in this case, the microorganism is eating
two kinds of food and the modeling is done as follows:
  
Cs1 Cs2
k = kmax (16.14)
K s1 + Cs1 K s2 + Cs2

where the number 1 is referring to the substrate 1, and the 2 is related to the substrate
number 2.

16.3 Kinetic Rate Expressions

A simple mass balance on the biomass done in case organisms is consuming two
kinds of substrate can be expressed as follows. From the above discussion we have:
16.3 Kinetic Rate Expressions 225

rC B = kC B (16.15)

now taking into consideration, the following relation


  
Cs1 Cs2
k = kmax (16.16)
K s1 + Cs1 K s2 + Cs2

It is possible then to obtain a complete relationship for the rate of production of


biomass.
In case we consider the rate of growth-associated product formation, the following
equation should also be taken into account:

rC p(g) = dC p(g) /dt = K C p(g)rC B = K C p(g) kC B (16.17)

in this case, C p refers to the concentration of the product, while (g) indicates that the
growth-associated product formation is taken into account.
K C p(g) is the growth-associated product formation ratio (sometimes referred to as
mass of product/mass of biomass).
The growth-associated product formation ratio is obtained from the literature. For
example, for the fermentation of sugar done by Kluyveromyces Marxianus, the yield
coefficient was found to be 0.51 g of ethanol on grams of sugar when the hydraulic
residence time, HRT, was 64.4 h (Ozmihci and Kargi 2008). Sometimes in the liter-
ature, the yield coefficients for different microorganisms at different conditions are
plotted against the dilution rate, D, which is defined as the medium flow rate divided
by the culture volume.
In the same way, the rate of non-growth-associated product formation (Biazar
et al. 2003) can be derived as follows:

rC p(n) = K C p(n) C B (16.18)

Here, the (n) indicates that the non-growth-associated product formation is taken
into account. The K C p(n) is the non-growth-associated product formation constant.
Notice that the dimensions of K C p(n) are a bit different from the previous case
because there is no k in this case.
The meaning is that this rate is associated with the amount of biomass present,
but it is not directly proportional to the variation of the concentration, dC/dt, or, in
other words, the rate of biomass organisms produced.
Taking into account the growth and non-growth-associated product formation,
this yields:

rC p = K C p(g) kC B + K C p(n) C B (16.19)

The biomass yield and product yield are defined as: γC B/Cs and γC p/Cs , and they are
expressed as a function of the substrate depletion
226 16 Balances on Microbial Fermentation

rC B kC B
rCs = = (16.20)
γC B/Cs γC B/Cs

with

γC p/Cs = rC p /rCs

Biomass yield and product yield have been calculated for a series of organisms, for
instance, γx/s of yeast is 0.084 C-mole-biomass/C-mole-sucrose and γC p/Cs = 0.58
C-mole-ethanol/C-mole-sucrose (Krzystek and Ledakowicz 1998). Here, C-mole
means that the comparison is done on the moles of carbon which, in some cases,
would give a more accurate assessment of the fermentation performances.
It is also possible to calculate the rate of O2 utilization with suitable models like:
  1 − Y CCsB
rCo = 1 − Y CCsB rCs = kC B (16.21)
Y CCsB
 
Certainly 1 − Y CCsB is positive since the rate of substrate consumption has to be
larger than the biomass formation. Some strategies of mixed substrate utilization in
microorganisms are given in (Harder et al. 1982).

16.4 Decay of Organisms

Organisms are growing in different kinds of environment, from seawater to sand,


and naturally, they also decay. The decay kinetics can be modeled mathematically,
and also in this case, diverse models can be provided (Zhang et al. 2015). During the
process of biofuels production, the decay of the microorganisms has been studied in
detail by several authors. One common model can be written as:

rC B(d) = −bC B (16.22)

in this case, rC B(d) refers to the rate of decay for the biomass concentration. Remember
that with biomass concentration, we refer here to the concentration of cells into the
system. The term b represents the decay constant.
The overall balance in batch reactors would be then:

dC B
kC B − bCb = (16.23)
dt
16.5 Overall Balance in Continuous STR 227

16.5 Overall Balance in Continuous STR

Previously, we have seen that in the case of continuous STR, we can consider the
simplified case of steady state for the substrate and product (that is still valid). How-
ever, not necessarily this has to be applied for the microorganisms. In this case, the
mass balance would be described as follows:

V̇ V̇ dC B
C Bi − C B + kC B − bC B = (16.24)
V V dt
where C Bi is the concentration of organisms which is inserted into the reactor from
the inlet stream.
The term V̇V is the ratio of the volumetric flow and the reactor volume. With reactor
volume, it is always good to stress that we indicate actually the volume of what is
inside the reactor and not the physical reactor volume.
If the biomass inlet is zero, meaning that we just add food in the reactor, and if
we have steady-state conditions:

1
k= +b (16.25)
τ
Remember also the definition of hydraulic retention time

V
τ= (16.26)

16.6 Effluent Concentration of the Substrate

With reference to the following mass balances, the substrate (food) will be indicated
with the letter s. From the Monod relation:

kmax Cs
k= (16.27)
K s + Cs

And considering Eq. (16.25), we obtain:



1
Ks+b τ
Cs =   (16.28)
kmax − τ1 + b

And therefore:

K s + Cs
τ= (16.29)
Cs (kmax − b) − K s b
228 16 Balances on Microbial Fermentation

The effluent concentration is, of course, the concentration inside the stirred tank
reactor. The balance of substrate (sugar for instance) in continuous steady-state STR
is done as follows. Consider:

V̇ V̇ kC B dCs
Csi − Cs − = (16.30)
V V Y CCsB dt

In this case, we just add food to the reactor, and therefore, the balance given is correct.
Remember also the relations previously written:

rC B kC B
rCs = = (16.31)
γC B/Cs γC B/Cs

where

γC p/Cs = rC p /rCs ;

and

1
k= + b;
τ
In consideration of the previously written relations. The following expression is
obtained:

YC B/Cs (Csi − Cs )
CB = (16.32)
1 + τb

Concerning the balance of products (H2 , ethanol, etc.). In a continuous steady-state


STR, we have:

V̇ V̇ dC p
C pi − C p + kC p(g) kC B + kC p(n) C B = (16.33)
V V dt
Since there is no product inserted into the reactor, the first term on the left side of
the equation above can be neglected. We remember the relations:

rC p(n) = K C p(n) C B ; (16.34)

rC p(g) = K C p(g)rC B = K C p(g) kC B ; (16.35)

and the non-growth-associated (NGA) product formation p:

C p(NGA) = kC p(n) C B τ, (16.36)


16.6 Effluent Concentration of the Substrate 229

expressed as g/L, and the growth-associated (GA) p formation will be:

C p(GA) = kC p(g) kC B τ (16.37)

16.7 Production of Ethanol by Digestion

Ethanol is produced normally by digestion of sugars which are derived from different
kinds of feedstock. The most common organism utilized for the ethanol production
is the Saccharomyces cerevisiae. These organisms can process several number of
sugars (Rodicio and Heinisch 2009) among which we remember maltotriose, sucrose,
galactose, mannose, maltose, glucose, and fructose. The common way of producing
ethanol by this organism is the Embden–Meyerhof–Parnas or EMP pathway (Peretó
2014) also known as the glycolytic pathway. A simplified overall reaction is expressed
as:

C6 H12 O6 + 2Pi + 2ADP → 2C2 H5 OH + 2CO2 + 2ATP + 2H2 O (16.38)

From Eq. (16.38), it is possible to calculate the related yield for the overall process.
The glycolytic pathway for ethanol production in S. cerevisiae (Drapcho et al. 2008)
is demonstrated in Fig. 16.2.
The same ethanol is inhibiting the microorganism. This means that the microbe
is able to grow until a certain concentration of ethanol.
   n
Cs
μ Sg = μ Sog 1 − (16.39)
Cs max

   n∗
Cs
f S f = f So f 1 − (16.40)
C S∗max

In the equations presented, the fermentation rates, f, and the growth rates, μ, are
directly proportional to their value considered when there is no by-product (the
alcohol) and also proportional to a term (in parenthesis) which slows down the same
rate (Nabais et al. 1988). With the term “max,” we indicate here the maximum
concentration allowable for the alcohol product.
230 16 Balances on Microbial Fermentation

Fig. 16.2 Glycolytic pathway for ethanol production in S. cerevisiae. Adapted from Drapcho et al.
(2008)
References 231

References

Biazar, J., Tango, M., Babolian, E., & Islam, R. (2003). Solution of the kinetic modeling of lactic acid
fermentation using Adomian decomposition method. Applied Mathematics and Computation,
144, 433–439. https://doi.org/10.1016/S0096-3003(02)00418-6.
Drapcho, C. M., Nghim, N. P., & Walker, T. (2008). Biofuels, bioproducts and biorefining. McGraw-
Hill.
Edwards, V. H. (1970). The influence of high substrate concentrations on microbial kinetics. Biotech-
nology and Bioengineering, 12, 679–712. https://doi.org/10.1002/bit.260120504.
Harder, W., Dijkhuizen, L., & Postgate, J. R. (1982). Strategies of mixed substrate utilization in
microorganisms [and discussion]. Philosophical Transactions of the Royal Society of London.
Series B, Biological Sciences, 297, 459–480.
Krzystek, L., & Ledakowicz, S. (1998). Yield and maintenance coefficients in S. cerevisiae cul-
tures. Journal of Chemical Technology and Biotechnology, 71, 197–208. https://doi.org/10.1002/
(SICI)1097-4660(199803)71:3%3c197:AID-JCTB825%3e3.0.CO;2-H.
Luong, J. H. (1985). Kinetics of ethanol inhibition in alcohol fermentation. Biotechnology and
Bioengineering, 27, 280–285. https://doi.org/10.1002/bit.260270311.
Nabais, R. C., Sá-Correia, I., Viegas, C. A., & Novais, J. M. (1988). Influence of calcium ion on
ethanol tolerance of Saccharomyces bayanus and alcoholic fermentation by yeasts. Applied and
Environment Microbiology, 54, 2439–2446.
Ozmihci, S., & Kargi, F. (2008). Ethanol production from cheese whey powder solution in a packed
column bioreactor at different hydraulic residence times. Biochemical Engineering Journal, 42,
180–185. https://doi.org/10.1016/j.bej.2008.06.017.
Peretó, J. (2014). Embden-Meyerhof-Parnas pathway. In: Amils R. et al. (Eds.), Encyclopedia of
Astrobiology (pp. 1–1). Berlin, Heidelberg: Springer. https://doi.org/10.1007/978-3-642-27833-
4_503-2.
Rodicio, R., & Heinisch, J. J. (2009). Sugar metabolism by Saccharomyces and non-Saccharomyces
yeasts. In Biology of microorganisms on grapes, in must and in wine (pp. 113–134). Berlin,
Heidelberg: Springer. https://doi.org/10.1007/978-3-540-85463-0_6.
Shijie, L. (2016). Bioprocess engineering. Kinetics, sustainability, and reactor design (2nd ed.).
Elsevier.
Zhang, Q., He, X., & Yan, T. (2015). Differential decay of wastewater bacteria and change of
microbial communities in beach sand and seawater microcosms. Environmental Science and
Technology, 49, 8531–8540. https://doi.org/10.1021/acs.est.5b01879.
Chapter 17
Processes of Bioethanol Production

Creativity is strictly a human characteristics and it belongs to


the neocortical component. Innovation is a man’s advantage to
be able to make new technological developments and uses the
creativity.
Rita Levi Montalcini, Nobel laureate in medicine, an
interview in 2009 for her first 100 years.

17.1 Some Systems Adopted and Product Compositions

One of the most produced biofuels worldwide is bioethanol. The fuel can be produced
from biomass feedstock like sugarcane but also from other sources like corn (Kumar
and Singh 2019), microorganisms (Onay 2018), yeast (Mohd Azhar et al. 2017),
and even marine biomass (Greetham et al. 2018). Some fuel properties are given in
Table 17.1.
To give some ideas of the volumes implied in this production: as an example,
Brazil is surely among the biggest producers of ethanol in the world. The total cost
of ethanol production in Brazil is estimated to be between 0.25 US dollars and
0.30 US dollars per L (including all inputs and factors) (Watanabe 2009). In Brazil,
10% of the total cultivated land (this accounts for 13.8 million acres) is utilized
for sugarcane production. From this land, about 7.4 million of these acres are used
to produce around 30 billion L of ethanol per year. The liquid effluent (which is
called vinasse) is used as a fertilizer and irrigation supply to the cane fields, thereby
eliminating costs for wastewater treatment. In 2007 worldwide, 11.6 BG (billion
gallons) of ethanol were produced from 2 billion ton of corn (Drapcho et al. 2008a).
When we want to produce bioethanol from sugarcane and sugar beets, the sugar
is first isolated by crushing the stalks with specialized rollers to release the juice.
The first operation is the cleaning of the sugarcane, and this can be operated by
dry cleaning and removal of solid external entities like rocks. These operations are
carried by suitable rolls which include pneumatic separation systems. For instance,
these operations are done to remove the presence of leaves and tops. Continuing with
the operations, the extraction of sugarcane juice is done by roll mills and these can
be designed for large-scale operations and small-scale operations.
© Springer Nature Switzerland AG 2019 233
C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_17
234 17 Processes of Bioethanol Production

Table 17.1 Some properties of bioethanol derived from different kinds of feedstock
Feedstock Density Flash point pH value Heating Viscosity
(g/cm3 ) (°C) (mg/L) value (cSt)
(kcal/kg)
Bioethanol 0.789 12 6.5–9.0 6380 1.53
standard
Mucasea 0.824 12 5.8 5524 2.15
Maranta 0.796 14 4.6 5969 3.05
arundinacea
Linn.
Solanum 0.817 13 5.3 5696 2.84
lyycopersicum
Amorphophallus 0.823 19 4.7 5892 2.38
campanulatus
Amorphophallus 0.858 15 2.5 6786 2.76
variabilis
Saccharum 0.801 13 8.5 8756 1.64
officinarum Linn.
Alocasia 0.836 16 3.8 6445 1.96
macrorrhizos
Adapted from Sutjahjo (2018)

The pretreatment of the sugarcane bagasse can be done by operations like steam
explosion or delignification. The cellulose is also hydrolyzed by using acids, while
among the mechanical treatment we can recall here the comminution process which
reduces the bagasse size and therefore, increases the surface of these particles. In
addition, also the cellulose crystallinity is reduced by this method.
The principle behind the steam explosion process is that steam usually is used to
increase the temperature of the feedstock so that the water contained in it becomes
hotter. The pressure will be quickly dropped and, because of this, the steam will
basically go to the steam phase very rapidly like an explosion. This method is also
used in other processes and in the food industry. For instance, in the tomato processing
industry when the tomatoes are peeled.
The delignification is operated by chemical treatment, for example, Organosolv
process. The method includes ethanol and in some cases sulfuric acid which are used
as a pretreatment media (Fig. 17.1).
The juice obtained can be cleaned by physical operations, like filtration or by using
a hydrocyclone. The other possibility is to use chemical treatment with the usage
of phosphoric acid. This acid helps the removal of impurities during the settlement
operation. This processing can include steps like sulfitation, liming, and clarification
with the consequent separation of the mud. The sulfitation is adopted to purify the
cane juice produced, and this is done by using sulfur dioxide gas. In practice, the
juice is heated up at certain temperatures and with a particular pH (for instance, a
temperature around 70 °C and a pH of around 5) and it is processed with SO2 and lime
17.1 Some Systems Adopted and Product Compositions 235

Fig. 17.1 Schematics for the organosolv process with ethanol. Adapted from Alriols et al. (2010)

until the sugar is then separated from the other compounds. At the end of the process,
the clarification separates the mud at certain conditions. For instance, one method
could be to heat the juice at a near boiling temperature and operate a juice clarifier
with a flash tank. The stirring conditions and the design of the clarification unit are
also important to optimize the consumption of the plant and the air entrainment in
the flash unit should also be minimized since the air could prevent the settling of the
mud.
A more detailed example of the preliminary operations performed to obtain the
clarified juice is shown in Fig. 17.2.
In regard to the juice produced after the preliminary treatment of the sugarcane
feedstock, the juice is concentrated (this is called evaporation) and it is also sterilized
to remove germs and bacteria possibly present into the biomass (Fig. 17.3).
In reference to Fig. 17.3, the syrup stream after the cooking process is sent to the
evaporation section. In this section, the crystallization occurs, and this is described
in more detail in Fig. 17.4.
A remark should be introduced here on the sterilization process. This is one of the
most important processes in the food industry and one of the most energy consuming.
Concepts related to condensation and heat transfer are particularly important in this
sense. There are many types of sterilization processes; for instance, they can be
classified in thermal and not thermal. In addition, the thermal processes can be
classified on the basis of if the medium used for the sterilization is directly in contact
with the food or it is in non-direct contact with the medium to be sterilized.
Let us be a bit more specific here: if the juice is a ready-made product and should
be sterilized, then the procedure is usually to put the product into a suitable container
and then proceed on the sterilization of the product by indirect thermal sterilization.
236 17 Processes of Bioethanol Production

Fig. 17.2 Pretreatment and derivation of clarified juice in bioethanol production from sugarcane.
Adapted from Cortes-Rodríguez et al. (2018)

Fig. 17.3 Juice concentration and sterilization process in bioethanol production from sugarcane.
Adapted from Palacios-Bereche et al. (2014)

For example, aluminum cans are commonly used in the food and beverage industry,
and this shape is used for a precise reason. The cylindrical shape of the can offers the
highest volume/surface area with respect to the possible commercial shapes available.
In fact, the sphere is even better but it is not a commercially suitable shape for obvious
reasons. In this way, we spare on the investments related to the metal of the container.
Another choice is the aluminum metal; aluminum offers a heat transfer rate
between the external medium (the steam) and the internal juice which is much higher
than steel. Without going into details with the proof of this; the difference between
aluminum and steel is readily evident when we use the different metals for cooking.
As an example, if two moka pots of the same volumetric capacity are used, the one
made in stainless steel will require more time to make the coffee ready. This has to
do also with the different heat capacities of the two metals and even with the internal
heat conductivity.
17.1 Some Systems Adopted and Product Compositions 237

Fig. 17.4 A scheme of the crystallization process where molasses and raw sugar is produced.
Adapted from Velasquez et al. (2013)

When sterilized, the cans are sent to rolls where steam is injected on their surfaces;
steam offers the possibility of “injecting” calories into the can at higher temperature
compared to water since it can be over-heated. After the sterilization, the containers
should be cooled down to ambient temperature. In some cases, this should be done
as fast as possible since when particular products are produced, there could be the
development of some bacteria which grow at certain high temperatures. This is the
case of the tomato industry products like peeled tomato and related.
The cooling is done also on rolls, where droplets of water are used in this case.
This is also done with a precise reason: droplets allow for the maximization of the
contact surface between the water and the can and the minimization of the water
consumption. There is also another reason why droplets are used and this is related
to the fact that when droplets are falling into air, they will be cooled even further
eventually, and this is because part of the water from the droplet is evaporating and
for this process, some energy is needed and lost by the droplet itself which will be
brought at even lower temperature.
To have a quantitative determination of this process, the reader is suggested to
consult some of the psychrometric charts and the related theory available in liter-
ature. The reader is also suggested to look further on gathering the basis for the
understanding of the heat transfer concepts which are explained in details elsewhere
in literature (Bird et al. 2006).
Coming back to the overall process for the production of bioethanol from sugar-
cane, after the sterilization and the cooling of the produced must, the fermentation
step takes place followed by the distillation process (Fig. 17.5).
In regard to the fermentation process, it is possible to have a distinction between
batch, fed-batch (Veloso et al. 2019), and continuous processes. The distillation
shown in this case is consisting of diverse steps which include reboiler columns and
238 17 Processes of Bioethanol Production

Fig. 17.5 Fermentation and distillation steps in bioethanol production from sugarcane

condenser columns. The distillation process for the ethanol production is described
in more detail at the end of this chapter with some theoretical notions.
The vinasse, produced by the fermentation and distillation process, is sent to a
concentration unit while the produced ethanol is sent to a dehydration unit for further
concentration to almost purity (Fig. 17.6).

Fig. 17.6 Vinasse concentration unit (left) and ethanol dehydration unit (right). Adapted from
Kujawski and Zielinski (2006)
17.1 Some Systems Adopted and Product Compositions 239

Fig. 17.7 One example of process for bioethanol production from sugarcane. Adapted from Amores
et al. (2013)

In reference to the more general representation of the process, Fig. 17.7, and with
respect to the production of sugar; after the process of clarification and the evaporation
of the resulted filtrate, the juice is treated further as was illustrated previously. The
steps that follow are denominated as cooking, crystallization, and centrifugation. In
Fig. 17.3, the evaporator depicted before the separation section is the equivalent of
the crystallization shown in Figs. 17.4 and 17.7. The sugar starts to crystallize in
a sugar-syrup juice, and the syrup traces are washed away from the sugar crystals
by water. The crystals are then separated by centrifugation. The fermentation of the
sugar allows for the production of ethanol which is then distilled and separated at
the end of the process.

17.2 Bioethanol from Lignocellulosic Biomass

In general, the bioethanol can be produced also by feedstock which has a composition
in cellulose, hemicellulose, and lignin. In this case, a pretreatment is necessary to
allow for the separation of these components. The pretreatment includes enzymes
and also chemicals at certain temperatures and pressures. More in detail, the ligno-
cellulosic biomass is processed with water, acid, and steam and after a separation
process (which can include centrifugation) a base is added to the produced mixture
to allow for the enzymes to be used. As a matter of fact, enzymes cannot operate if
the conditions are too acid. Also, the temperature conditions have to be adjusted for
the enzymes. The hydrolysis step is also operated by the enzymes, these powerful
240 17 Processes of Bioethanol Production

Table 17.2 Concentration (g/100 g) of the main components in sugarcane juice and sugarcane beet
juice
Components Sugarcane juice Sugarcane beet juice
Solid 13.7 17.3
Sucrose 12 16.5
Monosaccharides 0.63 0.15
Polysaccharides 0.028 0.019
Lactate 0.016
Acetate 0.035
Sulfate 0.039 0.02
Phosphate 0.033 0.047
Nitrate 0.015
Nitrite 0.005
Aconitate 0.09
K 0.11 0.125
Na 0.005 0.015
Cl 0.003
Ca 0.04
Mg 0.028
Total-N 0.105
Betaine-N 0.046
Amino acid-N 0.026
Ammonia-N 0.006
Amide-N 0.011
Adapted from Drapcho et al. (2008b)

catalysts break the cellulose into glucose (which is the building block of cellulose),
and the hemicellulose chains are then converted to xylose, and this sugars can be
readily fermented into bioethanol. Yeast or other bacteria are then introduced into
the fermentation reactor. The residence time for the ethanol to be produced is quite
long, 3–5 days, and this is a little drawback for the bioethanol production since
bigger reactor volumes are required to ensure a continue production. Sugars could
also be converted to longer chain molecules and other kinds of biofuels. In this case,
we should add hydrogen to these molecules in order to have renewable gasoline and
biodiesel. The so-called stillage and the lignin residues are then separated by suitable
methods (Fig. 17.8).
17.2 Bioethanol from Lignocellulosic Biomass 241

Fig. 17.8 Production of ethanol from lignocellulosic biomass feedstock. A simplified scheme

Table 17.3 Composition of produced molasses from cane and beets


Item Cane molasses Beet molasses
Brix 79.5 79.5
Total solids (wt%) 75.0 77.0
Specific gravity 1.41 1.41
Total sugars (wt%) 46.0 48.0
Crude protein (wt%) 3.0 6.0
Nitrogen free extract (wt%) 63.0 62.0
Ash (wt%) 8.1 8.7
Ca (wt%) 0.8 0.2
P (wt%) 0.08 0.03
K (wt%) 2.4 4.7
Na (wt%) 0.2 1.0
Cl (wt%) 1.4 0.9
S (wt%) 0.5 0.5
Trace minerals
Cu (mg/kg) 36 13
Fe (mg/kg) 249 117
Mn (mg/kg) 35 10
Zn (mg/kg) 13 40
(continued)
242 17 Processes of Bioethanol Production

Table 17.3 (continued)


Item Cane molasses Beet molasses
Vitamins
Biotin (mg/kg) 0.36 0.46
Choline (mg/kg) 745 716
Pantothenic acid (mg/kg) 21.0 7.0
Riboflavin (mg/kg) 1.8 1.4
Thiamine (mg/kg) 0.9 –
Adapted from Drapcho et al. (2008b)

Considering ethanol production by fermentation and from sugarcane and sugar


beets, the composition of the juice produced from sugarcane and sugar beets is
described in the previous table (Table 17.2).
While the produced molasses from cane and beets have the composition shown
in Table 17.3.
The maximal growth of Saccharomyces cerevisiae, i.e., the maximal specific
growth rate μmax , is obtained at a temperature of 30–35 °C (Zakhartsev et al. 2015),
while at temperatures above 44 °C the ethanol production is reduced and the residual
sugar content is increased.

17.3 Corn Starch as a Feedstock

When corn starch is used as a feedstock for ethanol production, the polymers forming
the starch should be broken into the constituents of those polymers (sugars) and this
process is called “saccharification.” Starch refineries use one of the following three
methods for the initial processing of corn grain. More information on these processes
can be found in literature (Schwietzke et al. 2009) and also, with focus on biodiesel
production, some of these processes are described (Veljković et al. 2018):

• A dry-grind process, in this case, the so-called kernels are grinded into a fine
powder.
• A dry-milling process, in this case, there is the utilization of roller mills where
the kernels are processed. This is followed by a separation of microorganisms and
fibers by screening.
• A wet-milling process (Bothast and Schlicher 2005). In this case, sulfuric acid
might be used to pretreat the kernels. In this phase, the so-called swelling of the
kernels takes place and then the kernels go to the milling process. After this, the
separation of microorganisms, fibers, protein, and starch is taking place by density
separation and washing.
17.3 Corn Starch as a Feedstock 243

Fig. 17.9 Schematic representation of the dry-milling process for ethanol production. Adapted
from Vohra et al. (2014)

Fig. 17.10 Schematic representation of the wet-milling process for ethanol production. Adapted
from Vohra et al. (2014)

We can use the wet process or the dry-milling process for ethanol production from
Fig. 17.9.
While the wet-milling process is shown in Fig. 17.10.
The reader is advised that in these representations, the connections to the dis-
tillation columns or the shown reactors are qualitative and they do not refer to the
exact position within the column. In other words, if a link to a distillation column is
244 17 Processes of Bioethanol Production

positioned on the top of the column, this does not mean necessarily that in the real
case the feeding is done at the last trays. This would require more investigation.
To be noticed from the previous figures that the utilization of enzymes in these pro-
cess steps is large. To give one example, in the liquefaction of water and ground corn
slurry, the α-amylase enzyme is widely used (Kumar and Singh 2016). Figure 17.11
represents the structure of the α-amylase enzyme.
Concerning the differences between the two methods, within the wet process, the
starch is separated from the rest of the components and sent to the fermentation;
while in the dry process, all the feedstock is processed first for its reduction of size
and then utilized into the fermentation. In terms of heat production, approximately
29.7 MJ/kg ethanol is generated (De Jong and Van Ommen 2014). Figure 17.12 is a
schematic representation of a dedicated dry-grind ethanol production plant.
Some information will be given here on the utilization of the enzymes and their
process parameters. The glucoamylase enzymes exhibit maximum activity at tem-
peratures of 60–65 °C and at a pH of about 4.5 (Hyun and Zeikus 1985). It is common
that the glucoamylase is injected directly into the fermentation reactor using a process
called “simultaneous saccharification and fermentation” (SSF). Large-scale indus-
trial fermentation processes typically result in ethanol concentrations of 12–15% on
a volume basis (Zhang et al. 2006). The fermentation is typically conducted in a
batch process that lasts for 48–72 h (Vertes et al. 2010).
A schematic representation of a production plant, where the wet-mill ethanol
production method is adopted, is demonstrated in Fig. 17.13.
In this process, the so-called steep water containing 1–2% of sulfur dioxide is
added; the mixture is then held at 49–53 °C for between 22 and 50 h. After this process

Fig. 17.11 An example of the animal α-amylase enzyme. The figure is produced by Jawahar
Swaminathan and his colleagues at the European Bioinformatics Institute and made available for
the public and free to use. The author highly acknowledge Jawahar and his collaborators
17.3 Corn Starch as a Feedstock 245

Fig. 17.12 Schematic representation of a dedicated dry-grind ethanol production plant. Adapted
from Vertes et al. (2010)

of “steeping,” there is evaporation and formation of corn steep liquor and corn is
converted by bacteria to lactic acid. Corn steep liquor is a particular product/feedstock
which could even be utilized as fuel in microbial fuel cells, for instance (Phansroy
et al. 2018). The lactic acid and sulfur dioxide cause the kernel to soften in such
a way that the disulfide bonds are broken, breaking the key linkages between the
starch, germ, and fiber. Finally, the fibers are separated by screens and the gluten is
separated by centrifugation.

17.4 Ethanol Distillation

When ethanol is produced, it will be necessary to increase its concentration among


the products and therefore, the distillation process is applied in this case. In this
small section, only some general information is given on this process and for an
246 17 Processes of Bioethanol Production

Fig. 17.13 Schematic representation of dedicated wet-mill ethanol production plant. Adapted from
Vertes et al. (2010)

accurate study of the distillation columns, the author of this manuscript suggest
to read specialized textbooks available in literature, for instance (Treybal 1980).
The distillation process allows to obtain a higher concentration of a component
in a mixture; this is done by taking advantage of several equilibrium stages where
ethanol in vapor phase is in equilibrium with its liquid phase at each of the distillation
trays. If ethanol and water are mixed together, for instance, then ethanol will have
a higher volatility (and lower boiling point) with respect to water and therefore the
vapor produced by this mixture will contain a higher concentration of ethanol with
respect to the liquid phase. Going upward, this vapor will increase the concentration
of ethanol in the liquid phase of the next tray and with an higher concentration of
ethanol in the liquid phase, the concentration of ethanol at this stage will be even
higher than the previous stage. And, this is repeated for the successive stages in
the distillation column. One of the available references on this argument is Madson
(2003). A simple representation of the process is given in Fig. 17.14.
The so-called stillage is a liquid waste remaining after ethanol distillation
(Choonut et al. 2015). Typical values for the ethanol in the stillage are 0.02 wt%.
The feedstock to the distillation column is called Beer, and this feed has a content
of 10% ethanol in volume typically. The typical number of stripping stages is eight
while the stages in the rectification section are typically 14 (Madson 2003). A sim-
17.4 Ethanol Distillation 247

Fig. 17.14 A simple schematics of a distillation column

ple representation of a graphical evaluation for the distillation stages is given in


Fig. 17.15.
Stripping is another form of distillation. This concerns the first stage of the process
when lighter components are said to be “stripped” from the heavier ones. Usually,
the feed during stripping is entering at the top of the column while the heat source is
at the bottom.
The distillation of two or more substances is somewhat complicated by the fact
that an azeotrope can occur. An azeotrope is found when two or more components
have the same concentrations in the vapor phase and the liquid phase. For this reason,
this mixture cannot be altered by simple distillation. An azeotropic distillation should
be performed in this case.
In a homogeneous azeotropic distillation, a solvent is used to change the volatility
of the components and in this case, we are talking about extractive distillation. While
in heterogeneous azeotropic distillation, we can have (Madson 2003):

• distillation using salt effects;


• pressure-swing distillation (at different pressures the equilibrium curves are trans-
lating).
Some constituents of the distillation tray are shown in Fig. 17.16.
248 17 Processes of Bioethanol Production

Fig. 17.15 Graphical Evaluation of the number of stages in a distillation column

Fig. 17.16 A top–down visual of a distillation tray with suitable passes for the vapor phase (left)
and a schematic showing the passage of the liquid and gas phase over the distillation tray (right)
17.5 Ethanol Yields from Different Kinds of Feedstock 249

17.5 Ethanol Yields from Different Kinds of Feedstock

In literature, several data can be found on the possible ethanol yield from diverse
kinds of feedstock and grains. In this last part of the chapter, some data are given as
examples of yields for ethanol production. However, it should be pointed out that in
literature, there are many data available and sometimes the values provided are in
conflict among them.
Some values of ethanol yield are given in Table 17.4 for agricultural residues.
Biomass residues can come from different sources like corn stover, barley straw,
oat straw, rice straw, sorghum and bagasse, and the provenience can be from different
parts of the world. From the crop-processing by-products, some examples of the
theoretical ethanol yield can be found in Table 17.5.
As for the case of ethanol yield from different feedstock kinds, the same reasoning
can be done in regard to the values of the main parameters used for the production
of ethanol. The literature in this field is large and a complete review of the reactors’
conditions is outside the scope of the first edition of this manuscript. Some examples
of values used for the main parameters in batch fermentation are given in Table 17.6.

Table 17.4 Composition of crops and residues and theoretical ethanol yield from agricultural
residues
Dry matter (wt%) Carbohydrates Lignin (wt%) Ethanol yield
(wt%) (L/kg of dry
biomass)
Barley 88.7 67.10 2.90 0.41
Barley straw 81.0 70.00 9.00 0.31
Corn 86.2 73.70 0.60 0.46
Corn stover 78.5 58.29 18.69 0.29
Oat 89.1 65.60 4.00 0.41
Oat straw 90.1 59.10 13.75 0.26
Rice 88.6 87.50 0.48
Rice straw 88.0 49.33 7.13 0.28
Sorghum 89.0 71.60 1.40 0.44
Sorghum straw 88.0 61.00 15.00 0.27
Wheat 89.1 35.85 0.40
Wheat straw 90.1 54.00 16.00 0.29
Sugarcane 26.0 67.00 0.50
Bagasse 71.0 67.15 14.50 0.28
Adapted from Kim and Dale (2004)
250 17 Processes of Bioethanol Production

Table 17.5 Theoretical ethanol yield from crop-processing by-products


By- Starch Cellulose Lignin HemicelluloseEthanol References
products (wt%) (wt%) (wt%) (wt%) yield
from crop
processing
Corn fiber 20 14 40 0.46 (L/kg Kurambhatti
dry) et al.
(2018),
Kálmán
et al.
(2006)
Rice husk 50 25–30 14–15% Onwuakor
(L/L) of Chijioke
fermented et al.
hydrolysate (2017),
distilled Syarif
et al.
(2016)

Table 17.6 Examples of values for the main parameters in batch fermentation
Process conditions Range References
Cell density 10–54 g of dry weight/L Bunch (1994)
Temperature 33–35 °C (36) Bajpai and Margaritis (1987)
Ethanol concentration Up to 14% (Vol/Vol) Ghareib et al. (1988)
Residence time Up to 10 days Unrean and Srienc (2010)

As we can see from the previous table, the reactors can operate at a very wide
range of parameters.

References

Alriols, M. G., Garcia, A., Chouaibi, F., Hajji, N., Labidi, J., Klemes, J., et al. (2010). Lignocellu-
losic biorefinery processes energy integration. Chemical Engineering Transactions, 21, 553–558.
https://doi.org/10.3303/CET1021093.
Amores, M. J., Mele, F. D., Jiménez, L., & Castells, F. (2013). Life cycle assessment of fuel ethanol
from sugarcane in Argentina. International Journal of Life Cycle Assessment, 18, 1344–1357.
https://doi.org/10.1007/s11367.013-0584-2.
Bajpai, P., & Margaritis, A. (1987). The effect of temperature and pH on ethanol production by
free and immobilized cells of Kluyveromyces marxianus grown on Jerusalem artichoke extract.
Biotechnology and Bioengineering, 30, 306–313. https://doi.org/10.1002/bit.260300222.
Bird, R. B., Stewart, W. E., & Lightfoot, E. N. (2006). Transport phenomena (2nd ed.). Wiley.
Bothast, R. J., & Schlicher, M. A. (2005). Biotechnological processes for conversion of corn into
ethanol. Applied Microbiology and Biotechnology, 67, 19–25. https://doi.org/10.1007/s00253-
004-1819-8.
References 251

Bunch, A. W. (1994). High cell density growth of micro-organisms. Biotechnology and Genetic
Engineering Reviews, 12, 535–562. https://doi.org/10.1080/02648725.1994.10647921.
Choonut, A., Yunu, T., Pichid, N., & Sangkharak, K. (2015). Ethanol production from reused liquid
stillage. In 2015 International Conference on Alternative Energy in Developing Countries and
Emerging Economies. Energy Procedia, 79, 808–814. https://doi.org/10.1016/j.egypro.2015.11.
570.
Cortes-Rodríguez, E. F., Fukushima, N. A., Palacios-Bereche, R., Ensinas, A. V., & Nebra, S.
A. (2018). Vinasse concentration and juice evaporation system integrated to the conventional
ethanol production process from sugarcane—Heat integration and impacts in cogeneration sys-
tem. Renewable Energy, 115, 474–488. https://doi.org/10.1016/j.renene.2017.08.036.
De Jong, W., & Van Ommen, R. (2014). Biomass as a sustainable energy source for the future:
Fundamentals of conversion processes. Wiley.
Drapcho, C. M., Nghiem, J., & Walker, T. (2008a). Biofuels engineering process technology.
McGraw-Hill Education.
Drapcho, C. M., Nghim, N. P., & Walker, T. (2008b). Biofuels, bioproducts and biorefining.
McGraw-Hill.
Ghareib, M., Youssef, K. A., & Khalil, A. A. (1988). Ethanol tolerance of Saccharomyces cerevisiae
and its relationship to lipid content and composition. Folia Microbiologica (Praha), 33, 447–452.
https://doi.org/10.1007/BF02925769.
Greetham, D., Zaky, A., Makanjuola, O., & Du, C. (2018). A brief review on bioethanol production
using marine biomass, marine microorganism and seawater. Current Opinion in Green Sustainable
Chemistry Bioresources and Biochemicals/Biofuels and Bioenergy, 14, 53–59. https://doi.org/10.
1016/j.cogsc.2018.06.008.
Hyun, H. H., & Zeikus, J. G. (1985). General biochemical characterization of thermostable pul-
lulanase and glucoamylase from Clostridium thermohydrosulfuricum. Applied and Environment
Microbiology, 49, 1168–1173.
Kálmán, G., Recseg, K., Gáspár, M., & Réczey, K. (2006). Novel approach of corn fiber utilization.
Applied Biochemistry and Biotechnology, 131, 738–750. https://doi.org/10.1385/ABAB:131:1:
738.
Kim, S., & Dale, B. E. (2004). Global potential bioethanol production from wasted crops and crop
residues. Biomass and Bioenergy, 26, 361–375. https://doi.org/10.1016/j.biombioe.2003.08.002.
Kujawski, W., & Zielinski, L. (2006). Bioethanol—One of the renewable energy sources. Environ-
ment Protection Engineering, 32, 143–149.
Kumar, D., & Singh, V. (2016). Dry-grind processing using amylase corn and superior yeast to
reduce the exogenous enzyme requirements in bioethanol production. Biotechnology for Biofuels,
9. https://doi.org/10.1186/s13068-016-0648-1.
Kumar, D., & Singh, V. (2019). Bioethanol production from corn, Chap. 22. In S. O. Serna-Saldivar
(Ed.), Corn (3rd ed., pp. 615–631). Oxford: AACC International Press. https://doi.org/10.1016/
B978-0-12-811971-6.00022-X.
Kurambhatti, C. V., Kumar, D., Rausch, K. D., Tumbleson, M. E., & Singh, V. (2018). Increasing
ethanol yield through fiber conversion in corn dry grind process. Bioresource Technology, 270,
742–745. https://doi.org/10.1016/j.biortech.2018.09.120.
Madson, P. W. (2003). Ethanol distillation: The fundamentals. In The alcohol textbook. The Not-
tingham University Press.
Mohd Azhar, S. H., Abdulla, R., Jambo, S. A., Marbawi, H., Gansau, J. A., Mohd Faik, A. A.,
et al. (2017). Yeasts in sustainable bioethanol production: A review. Biochemistry and Biophysics
Reports, 10, 52–61. https://doi.org/10.1016/j.bbrep.2017.03.003.
Onay, M. (2018). Bioethanol production from Nannochloropsis gaditana in municipal wastewater.
In 5th International Conference on Energy and Environment Research, ICEER 2018, Prague,
Czech Republic, July 23–27, 2018. Energy Procedia, 153, 253–257. https://doi.org/10.1016/j.
egypro.2018.10.032.
Onwuakor Chijioke, E., Hans-Anukam Uzunma., & Uzokwe Munachi, J. (2017). Production of
ethanol and biomass from rice husk using cultures of Aspergillus flavus, Aspergillus eamarii and
252 17 Processes of Bioethanol Production

Saccharomyces cerevisiae. American Journal of Microbiological Research, 5, 86–90. https://doi.


org/10.12691/ajmr-5-4-3.
Palacios-Bereche, R., Ensinas, A. V., Modesto, M., & Nebra, S. A. (2014). Mechanical vapour
recompression incorporated to the ethanol production from sugarcane and thermal integration to
the overall process applying pinch analysis. Chemical Engineering Transactions, 39, 397–402.
Phansroy, N., Khawdas, W., Watanabe, K., Aso, Y., & Ohara, H. (2018). Microbial fuel cells
equipped with an iron-plated carbon-felt anode and Shewanella oneidensis MR-1 with corn steep
liquor as a fuel. Journal of Bioscience and Bioengineering, 126, 514–521. https://doi.org/10.
1016/j.jbiosc.2018.04.011.
Schwietzke, S., Kim, Y., Ximenes, E., Mosier, N. S., & Ladisch, M. (2009). Ethanol production
from maize. In Molecular genetic approaches to maize improvement. New York: Springer.
Sutjahjo, D. H. (2018). The characteristics of bioethanol fuel made of vegetable raw materials. In
Materials science and engineering. IOP Publishing. https://doi.org/10.1088/1757.899X/296/1/
012019.
Syarif, H. U., Suriamihardja, D. A., Selintung, M., & Wahab, A. W. (2016). Analysis SEM the
chemical and physics composition of used rice husks as an absorber plate. International Journal
of Engineering Science and Applications, 2, 25–30.
Treybal, R. E. (1980). Mass transfer operations (3rd ed.). McGraw-Hill.
Unrean, P., & Srienc, F. (2010). Continuous production of ethanol from hexoses and pentoses using
immobilized mixed cultures of Escherichia coli strains. Journal of Biotechnology, 150, 215–223.
https://doi.org/10.1016/j.jbiotec.2010.08.002.
Velasquez, H. I., Mesa, C., & Giraldo, S. A. (2013). Energy and exergy analysis of the combined
production process of sugar and ethanol from sugarcane (Colombia case study). In International
Conference on Efficiency, Cost, Optimisation, Simulation and Environmental Impact of Energy
Systems. Presented at the ECOS 2013.
Veljković, V. B., Biberdžić, M. O., Banković-Ilić, I. B., Djalović, I. G., Tasić, M. B., Nježić, Z. B.,
et al. (2018). Biodiesel production from corn oil: A review. Renewable and Sustainable Energy
Reviews, 91, 531–548. https://doi.org/10.1016/j.rser.2018.04.024.
Veloso, I. I. K., Rodrigues, K. C. S., Sonego, J. L. S., Cruz, A. J. G., & Badino, A. C. (2019). Fed-
batch ethanol fermentation at low temperature as a way to obtain highly concentrated alcoholic
wines: Modeling and optimization. Biochemical Engineering Journal, 141, 60–70. https://doi.
org/10.1016/j.bej.2018.10.005.
Vertes, A. A., Qureshi, N., Blaschek, H. P., & Yukawa, H. (Eds.). (2010). Biomass to biofuels:
Strategies for global industries. Oxford, UK: Wiley.
Vohra, M., Manwar, J., Manmode, R., Padgilwar, S., & Patil, S. (2014). Bioethanol production: Feed-
stock and current technologies. Journal of Environmental Chemical Engineering, 2, 573–584.
https://doi.org/10.1016/j.jece.2013.10.013.
Watanabe, M. (2009). Ethanol production in Brazil: Bridging its economic and environmental
aspects. International Association for Energy Economics.
Zakhartsev, M., Yang, X., Reuss, M., & Pörtner, H. O. (2015). Metabolic efficiency in yeast Sac-
charomyces cerevisiae in relation to temperature dependent growth and biomass yield. Journal
of Thermal Biology, 52, 117–129. https://doi.org/10.1016/j.jtherbio.2015.05.008.
Zhang, C. -H., Yang, Y., Teng, B. -T., Li, T. -Z., Zheng, H. -Y., Xiang, H. -W., et al. (2006). Study
of an iron-manganese -Tropsch synthesis catalyst promoted with copper. Journal of Catalysis,
237, 405–415. https://doi.org/10.1016/j.jcat.2005.11.004.
Chapter 18
Biodiesel

The energy in biomass doesn’t come for free, you have to use
fertilizers. You have the plowing of the field and even by plowing
the field you will need power. I have read that for 1 hectare 100
liters of diesel will be needed. If the scope of the land is the
production of biodiesel, 1/10 will be lost just for the plowing of
the field.
Michel Hartmut, Nobel Prize in Chemistry in 1988.

18.1 Fundamentals of Biodiesel Production

Biodiesel is sometimes considered as a renewable transportation fuel, and it is nor-


mally produced from vegetable oils and animal fats even if it can be produced from
other sources such as waste cooking oils (Phan and Phan 2008). Biodiesel is consist-
ing of fatty acid methyl esters (FAME) (Hoekman et al. 2012) and because of this
composition required, it will be more understood why they are produced by vegetable
oils. In the following explanations, the author of this book encourages the reader in
considering also more advanced textbooks of organic chemistry (Jones and Fleming
2014) in order to understand the terminology used here.
Vegetable oils present a series of positive aspects that make them suitable for
biodiesel production. More in particular:
• are rich in neutral triacylglycerols (TAG);
• can be rich in free fatty acids (FFA).
Non-conventional sources of these two categories of compounds are, for instance,
non-edible vegetable oils (for instance, Jatropha oil and Polanga oil) or animal fat
wastes (AFWs) (Pollardo et al. 2018) (for instance, chicken fat and white grease).
These feedstocks are characterized with high free fatty acids and water.
As mentioned previously, biodiesel is constituted mainly by FAME, however, the
composition can still vary quite largely when considering the process conditions and
the feedstock utilized. In addition, catalysis plays a fundamental role in biodiesel
production.

© Springer Nature Switzerland AG 2019 253


C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_18
254 18 Biodiesel

In regard to the biodiesel composition, Table 18.1 is derived with data from Knothe
(2005) and Hoekman et al. (2012) and only the main esters are reported here. This
is done to demonstrate that, even if the composition of biodiesel is restricted to only
some compounds, there is still a variety of them (Table 18.1).
The main reactions taking place during the biodiesel production aim at the for-
mation of fatty acid esters (characterized by the group –OOR). If the feedstock is
rich in neutral triacylglycerols (TAG), we have the following reaction:
Catalyst
Triacylglycerol (TAG) + 3R OH ←→ 3R COOR + C3 H5 (OH)3 (18.1)

the component 3R COOR is our fatty acid ester and in addition, glycerol is produced
by this process. Notice that the terms R and R are referred here to carbon chains
(CH3 –CH2 –CH2 – …).
The catalysts can be KOH, NaOH, or H2 SO4 . If vegetable oils are rich in free
fatty acids (Gebremariam and Marchetti 2018), FFA, we have:
Acid Catalyst
RCOOH (FFA) + R OH ←→ RCOOR + H2 O (18.2)

where the catalyst is commonly HCl or H2 SO4 and RCOOR is an alkyl ester. In
transesterification, the biodiesel is obtained by the reaction of a vegetable oil with
an alcohol, in the presence of a catalyst, to give monoalkyl esters.
The overall reaction is represented in Fig. 18.1.
Free fatty acids are formed when oils are treated at high temperatures where
they undergo an oxidation process. Side reactions in the production of biodiesel are
saponification and hydrolysis, and they are represented, respectively, by:
 
RCOOH + NaOH (or KOH) ↔ R COONa+ or R COOK+ + H2 O (18.3)

Base catalyst
RCOOR + H2 O ←→ RCOOH + R OH (18.4)

Fig. 18.1 Reaction of a vegetable oil with an alcohol. Adapted from Lotero et al. (2005)
Table 18.1 Main esters of biodiesel in reference to different kinds of feedstock used
Ester Camelina Canola Coconut Corn Jatropha Palm Rapeseed Safflower Soy Sunflower Tallow Yellow
grease
Lauric 47.7
(dode-
canoic);
12:0
Myristic 18.5
(tetrade-
canoic);
14:0
Palmitic 11.5 14.9 42.5 11.6 24.3
(hexade-
canoic);
16:0
Stearic 18.2
18.1 Fundamentals of Biodiesel Production

(octade-
canoic);
18:0
Methyl 16.8 60.4 26.6 40.4 41.3 59.5 14.2 23.7 21.7 42.2 44.6
octade-
cenoate
(methyl
oleate);
18:1
Methyl 17.0 21.2 58.7 36.2 21.5 74.3 53.8 66.3 25.1
octadeca-
dienoate
(linoleate);
18:2
Methyl 35.6 9.6 8.4
octadeca-
trienoate
(linole-
nate);
18:3
Values in % of FAME composition
255
256 18 Biodiesel

It is referred in literature that when acid-catalyzed reactions occur, the saponification


is greatly reduced (Gebremariam and Marchetti 2018). On the other hand, when free
fatty acids (FFAs) (ROOH is the acidic group) are present in the feedstock, then
there could be the formation of soaps, and this is due to a possible excess of the
catalyst itself. The reaction of saponification due to an excess of base catalyst is also
described in Fig. 18.2.
The formation of soaps is very negative for separation processes, the viscos-
ity will increase for example. In waste oil, the amount of FFA suggested is not
very high and around 1 wt% (Kumar Tiwari et al. 2007). While the FFA reduces
the quality of the feedstock, water is eliminated by centrifugation to an acceptable
level of 0.5% (wt).
Fatty acids directly affect the main parameters of the biodiesel which are:

• the cloud point (the minimum temperature at which small solid crystals can be
observed when the temperature of the sample is decreased);
• the cold filter plugging point (the temperature at which the fuel filter plugs due to
crystals and solid components);
• the pour point (the minimum temperature at which the fuel is actually moving
sufficiently, it becomes a gel).

For the sake of comprehension, we remember here that the structure of a carboxylic
acid is represented in Fig. 18.3.
In Fig. 18.3, it is possible to notice that this molecule has some parts which are
nucleophile (attract protons) while, on the other hand, there are also electrophile
parts (which attract electrons). It is quite important to reason on these sites of the
molecules to understand and design possible reaction routes. For this reason, more
information on the meaning of ionization potential and electron affinity will be given
in the following sections.

Fig. 18.2 FFA reaction with a base catalyst and formation of soap

Fig. 18.3 Carboxylic acid


representation. Adapted from
Jones and Fleming (2014)
18.2 Ionization Potential and Electron Affinity of an Element 257

18.2 Ionization Potential and Electron Affinity


of an Element

The ionization potential is the energy required to subtract one electron from the atom
to form a cation (the easiest electron is the one situated at the upper orbital level),
this is measured in electron Volt, eV, which is equivalent to 1 V·1.6 × 10−19 C, and
this has the dimensions of Joule. Remember also that 1 V = 1 J/1 C and that the
Coulomb is actually the measure of the charge (the electrons).
The electron affinity is, on the other hand, the energy released when the electron
is added to an atom. Some values for the ionization potential of some elements are
given in Table 18.2 adapted from Jones and Fleming (2014).
While an example of nucleophilic substitution is given in Fig. 18.4.
To be noticed that in Fig. 18.4, the N–H group has more negative charge on the N
atom, and therefore the H atom will have a more positive charge. On the other hand,

Table 18.2 Examples of ionization potential and electron affinities (in brackets)
H He
13.60 24.59
(0.75)
Li Be B C N O F Ne
5.39 0.32 8.30 11.26 14.53 13.62 17.42 21.56
(0.62) (0.24) (1.27) (1.47) (3.34)
Na Mg Al Si P S Cl Ar
5.14 7.65 5.99 8.15 10.49 10.36 12.97 15.75
(0.55) (0.46) (1.24) (0.77) (2.08) (3.61
Values are in eV

Fig. 18.4 Nucleophilic substitution between N–H and C–OH groups


258 18 Biodiesel

Fig. 18.5 Formation of the tetrahedral intermediate in the esterification of hydroxyl acids. Adapted
from Jones and Fleming (2014)

Fig. 18.6 All steps in Fischer esterification. Adapted from Jones and Fleming (2014)

the group C–OH has a more negative charge on the OH part and therefore this part
will have more possibility to react with the H atom of the N–H group.
The formation of the tetrahedral intermediate in the esterification of hydroxyl
acids proceeds as shown in Fig. 18.5.
Consider the periodic table with the electron affinities, note that oxygen has an
electron affinity higher than carbon; for this reason, the alcohol will bound on the
carbon and not on the oxygen.
All steps in Fischer esterification are demonstrated in Fig. 18.6.
The ion RH2 O+ goes to R–OH which is a stable compound. The reaction with the
alkoxide group (− OR) is reversible and catalytic, while the reaction of the hydroxide
group (− OH) is not reversible and not catalytic (Fig. 18.7).
18.2 Ionization Potential and Electron Affinity of an Element 259

Fig. 18.7 Reaction involving the alkoxide group (− OR) and the hydroxide group (− OH). Adapted
from Jones and Fleming (2014)

The hydrogen from the hydroxyl acid cannot just be replaced by the R of the alko-
hyde; because, an acid when loses the hydrogen is actually stabilized by resonance.
Remember that R is referring to the carbon chain (CH3 –CH2 …).
As a summary and as referred previously, the catalysts used in the process of
biodiesel production are mainly alkaline:

• NaOH
• KOH.

And, they are involved in the reactions:

CH3 O− Na+ → CH3 O− + Na+ (18.5)

NaOH + CH3 OH ↔ CH3 O− + H2 O + Na+ (18.6)


 
Here, we notice the formation of the methoxide ions CH3 O− , and they are strong
nucleophiles and attach the glyceride molecules.

18.3 Base-Catalyzed Production of Biodiesel

When base catalysts are used in the production of biodiesel, the reaction mechanisms
are illustrated as follows. The base catalysts create an alkoxide ion, BH+ . In this way,
the alkyl ion, RO− , will attack directly the double bond of the carboxylic group.

ROH + B ↔ RO− + BH+ (18.7)

The alkyl ion attack on the double bond of the triglyceride is demonstrated in
Fig. 18.8.
Notice that the RO– ion is a nucleophile. As an example, here, the formation of
the methoxide ion from a base catalyst and an alcohol is demonstrated:
260 18 Biodiesel

CH3 O− Na+ → CH3 O− + Na+ (18.8)

NaOH + CH3 OH ↔ CH3 O− + H2 O + Na+ (18.9)

In particular, the methoxide ion (the one with the methyl group) is a strong nucle-
ophile, which means that it is able to link at lower electrophile carbon.

18.4 Acid-Catalyzed Production of Biodiesel

When an acid will be adopted as catalyzer, the attack on the original triglyceride is
shown in Fig. 18.9.
Acid catalysts do not show measurable susceptibility to FFA as previously men-
tioned; however, remember that, there has to be a limit in FFA content. In case, we

Fig. 18.8 Catalyst alkoxide ion and attach on the carboxylic group. Adapted from Lotero et al.
(2005)
18.4 Acid-Catalyzed Production of Biodiesel 261

consider here an alcohol described by R4 OH (see Fig. 18.9), this would be a nucle-
ophile because it can give some electrons to the cation. Nucleophiles are also called
Lewis bases. If R4 is the methoxide ion, then this will be the methyl ester we want,
the carbonyl oxygen becomes more electrophilic, and this is because obtains a plus
charge, and this makes more easy to host the OH group of the alcohol.
An example of homogeneous acid-catalyzed process for biodiesel production is
given in Fig. 18.10.
As an example of biodiesel production from waste cooking oils (WCO), it will
be reported here the alkali-catalyzed transesterification of waste cooking oils. The
composition of waste cooking oils is demonstrated in Table 18.3.
In this particular study, the temperature of the experiments was 30, 50, or 70 °C.
While the residence time was 20, 40, 60, 90, or 120 min, the molar ratio of methanol
and WCO was from 5:1 to 12:1. The amount of KOH catalyst was 0.5–1.5 wt% of
the WCO.
The transesterification is commonly carried out with an extra amount of alcohol
in order to shift the equilibrium to the proposed product, methyl ester.

Fig. 18.9 Acid-catalyzed reactions on the triglyceride. R1, R2, and R3 are carbon chains of the
fatty acids, while R4 is an alkyl group of the alcohol. Adapted from Lotero et al. (2005)
262 18 Biodiesel

Fig. 18.10 Example of homogeneous acid-catalyzed process for biodiesel production

Table 18.3 Composition of Component g/100 g of total fatty acid methyl ester
waste cooking oils (WCO)
C6:0 0
C8:0 8.82
C10:0 6.21
C12:0 44.65
C14:0 16.31
C15:0 0.00
C16:0 10.59
C16:1 0
C17:0 0
C18:0 3.29
C18:1 8.17
C18:2 1.96
C18:3 0
C20:0 0
C20:1 0
Adapted from Phan and Phan (2008)

Table 18.4 reports a comparison between waste cooking oil and diesel. It is possi-
ble to notice that obviously waste cooking oils are not suitable for an engine, however,
they do have good technical properties.
Methanol concentration has an effect on the conversion of biodiesel; as a matter
of fact, the conversion to biodiesel can be up to 20% in case the concentration of
methanol goes from five to twelve times the concentration of waste cooking oils.
The temperature also affects the conversion of waste cooking oils to biodiesel;
it is reported in the literature that if the concentration of KOH is 0.75 and the
methanol/WCO ratio is 8:1, then the highest conversion has been noticed at 50 °C
(Lotero et al. 2005).
18.5 More Detailed Process Description for Alkali-Catalyzed … 263

Table 18.4 Properties of Parameter WCO Biodiesel


waste cooking oils and diesel
Acid number (mg KOH/g) 2.36 0.11
Iodine number (mg KOH/g) 13.20 –
Saponification number (mg KOH/g) 268.22 –
Density (g/cm3 ) 0.92 0.83
Flash point (°C) 269 69
Cloud point (°C) 21 0
Pour point (°C) 18 <12
Viscosity at 40 °C (mm2 /s) 30.05 3.53
Adapted from Phan and Phan (2008)

Fig. 18.11 A more detailed flow diagram for biodiesel production by a homogeneous alkali-
catalyzed process. Adapted from Apostolakou et al. (2009)

18.5 More Detailed Process Description


for Alkali-Catalyzed Biodiesel Production

Previous studies on biodiesel production and the different alternatives available have
been performed. It has been concluded that the alkali-catalyzed process for biodiesel
production is the most favorable in terms of its economy (Marchetti et al. 2008) and
that heterogeneous-catalyzed processes are also favorable. For this reason, the case
of alkali-catalyzed biodiesel production will be described in more details here. The
process is demonstrated in Fig. 18.11.
The demonstrated process is mostly suitable for oils that have a low content in
FFA, like rapeseed oil. The oil and the related catalyst are sent to a transesterification
264 18 Biodiesel

reactor together with methanol. As mentioned, diverse catalysts could be used in


this stage and to give one example, sodium methoxide could be used (KoohiKamali
et al. 2012); however, also other types of catalysts are utilized as well. The effluent
from the transesterification reactor is sent to a centrifuge where the esterificated
fraction is sent to a second reactor. The glycerol is separated from this fraction
and sent to a mixing tank. The same operation is basically achieved into the second
transesterification reactor, and the obtained product is then sent to the washing process
where the biodiesel fraction is processed with water and hydrochloric acid. The
fraction containing soaps or salts, methanol, water, glycerol, and FFA is sent to a
mixing tank, while the remaining biodiesel (which contains a small fraction of water)
is sent to a flash tank unit where the refined biodiesel is then separated.
The fraction with the highest amount of glycerol and coming from the transesteri-
fication reactors is sent, together with the fraction coming from the washing process,
to be mixed and the separation of FFA occurs by centrifugation with the help of
HCl solution. The derived glycerol and methanol stream is then adjusted for its pH
values with NaOH and sent to a glycerol recovery column where the glycerol is
finally separated. The effluent stream with the highest amount of methanol is sent to
a methanol recovery column where it is separated.

References

Apostolakou, A. A., Kookos, I. K., Marazioti, C., & Angelopoulos, K. C. (2009). Techno-economic
analysis of a biodiesel production process from vegetable oils. Fuel Processing Technology, 90,
1023–1031. https://doi.org/10.1016/j.fuproc.2009.04.017.
Gebremariam, S. N., & Marchetti, J. M. (2018). Biodiesel production through sulfuric acid catalyzed
transesterification of acidic oil: Techno economic feasibility of different process alternatives.
Energy Conversion and Management, 174, 639–648. https://doi.org/10.1016/j.enconman.2018.
08.078.
Hoekman, S. K., Broch, A., Robbins, C., Ceniceros, E., & Natarajan, M. (2012). Review of biodiesel
composition, properties, and specifications. Renewable and Sustainable Energy Reviews, 16,
143–169. https://doi.org/10.1016/j.rser.2011.07.143.
Jones, M., & Fleming, S. A. (2014). Organic chemistry (5th ed.). W. W. Norton & Company.
Knothe, G. (2005). Dependence of biodiesel fuel properties on the structure of fatty acid alkyl esters.
Fuel Processing Technology Biodiesel Processing and Production, 86, 1059–1070. https://doi.
org/10.1016/j.fuproc.2004.11.002.
KoohiKamali, S., Tan, C. P., & Ling, T. C. (2012). Optimization of sunflower oil transesterification
process using sodium methoxide. The Scientific World Journal, 2012, 8 pages. https://doi.org/10.
1100/2012/475027.
Kumar Tiwari, A., Kumar, A., & Raheman, H. (2007). Biodiesel production from jatropha oil
(Jatropha curcas) with high free fatty acids: An optimized process. Biomass and Bioenergy, 31,
569–575. https://doi.org/10.1016/j.biombioe.2007.03.003.
Lotero, E., Liu, Y., Lopez, D. E., Suwannakarn, K., Bruce, D. A., & Goodwin, J. G. (2005). Synthesis
of biodiesel via acid catalysis. Industrial and Engineering Chemistry Research, 44, 5353–5363.
Marchetti, J. M., Miguel, V. U., & Errazu, A. F. (2008). Techno-economic study of different alter-
natives for biodiesel production. Fuel Processing Technology, 89, 740–748. https://doi.org/10.
1016/j.fuproc.2008.01.007.
References 265

Phan, A. N., & Phan, T. M. (2008). Biodiesel production from waste cooking oils. Fuel, 87,
3490–3496. https://doi.org/10.1016/j.fuel.2008.07.008.
Pollardo, A. A., Lee, H., Lee, D., Kim, S., & Kim, J. (2018). Solvent effect on the enzymatic
production of biodiesel from waste animal fat. Journal of Cleaner Production, 185, 382–388.
https://doi.org/10.1016/j.jclepro.2018.02.210.
Chapter 19
Some Chemical Analyses in Biodiesel
Production and Biofuel Characteristics

If I don’t see it, I don’t believe it.


Sentence normally attributed to St. Thomas.

19.1 Amount of Free Fatty Acid

Biodiesel has a series of positive and negative aspects if compared with normal diesel.
The main positive aspects of biodiesel production are listed as (Lin and Lin 2012):
• higher biodegradability;
• pollutant reduction;
• higher flash point;
• biodiesel increases lubricity;
• biodiesel is considered to not contribute to global warming since it is produced
by renewable biomass. However, this is not a suitable conclusion for large-scale
productions, as it was referred in other parts of this book.
As an example of reduction in emissions, Fig. 19.1 shows that when biodiesel is
blended to normal diesel, we have actually a reduction of pollution.
On the other hand, negative aspects with biodiesel production are listed as:
• higher NOx production;
• higher viscosity;
• lower volatility;
• presence of unreacted glycerols.
All the properties listed are highly influenced by the composition of the fuel itself;
therefore, it is beneficial to have some knowledge of some analyses done on the fuel
and the feedstock.
In the previous chapter, it was mentioned that the production of biodiesel is influ-
enced by the feedstock composition and, because of this composition, also the pro-
duction method will be influenced. It is important to understand, ones the product
© Springer Nature Switzerland AG 2019 267
C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_19
268 19 Some Chemical Analyses in Biodiesel Production …

120

100

80
% REDUCTION

60

40

20

0
TH CO2 CO PM Nox Sulfate PAH NPAH
-20

B100 B20

Fig. 19.1 Percentage of pollutants reduction as a function of the fuel blend. B20 means “fuel
blend,” in this case, 20% of biodiesel is mixed with 80% of normal diesel. Data from Drapcho et al.
(2008b)

is obtained, what are the physical and chemical analyses necessary to assess our
product. Among the main kind of analyses, we have the determination of the amount
of free fatty cid (FFA).
The procedures for the FFA or in general the total acid number (TAN) determina-
tion follow diverse standards. For instance, we can recall here the ASTM D664-11a,
the ASTM D974-14, the ASTM D3339-12, and the AOCS Cd 3d 63 (Park et al.
2017). And, there are also several variations of these procedures. The basic princi-
ple behind these types of analyses is that we measure the sample at hand, then the
sample is dissolved into a solution and a particular solvent is used to do this. For
instance, as solvents, we have mixtures of toluene, 2-propanol, and water (Baig and
Ng 2011) or a mixture of ethanol and water (Aricetti and Tubino 2012); or only
propanol can be used. Isopropyl alcohol or isopropanol is also very common (Patel
and Shah 2015, p. 11). The dissolved sample is then titrated with a suitable base,
for instance, KOH and the change in pH values will determine the typical change in
color. The change in color at a precise pH value is determined by particular chemical
compounds, for example, phenolphthalein. When the molecule is in its acidic form,
there is no detectable change in color; however, when the solution is titrated with a
base, KOH, for instance, the pH of the solution will change and at some point, the
phenolphthalein will be converted in its pink base form. The structure of the molecule
in the two states is depicted in the following Fig. 19.2.
A complete procedure will be given in this case with respect to the product
obtained. We assume here that we have produced biodiesel and we want to know if
in our biodiesel there is still a quantity of FFA, however, this kind of analyses can
19.1 Amount of Free Fatty Acid 269

Fig. 19.2 Structure of the phenolphthalein molecule in its acidic for and base form. Adapted from
Zumdahl and Zumdahl (2014)

be performed for any type of feedstock oil or produced bio-oil. The procedure is as
follows:
• The first step is to weigh a quantity of sample (for instance, biodiesel). Let us
assume that we have 15–20 g. Make sure that the scale used is a quite precise one
and reports the exact weight up to the second decimals.
• At this point, we need a suitable solvent to solubilize the sample and one of the
possible solvents available is the isopropyl alcohol. Take a sufficient amount of
isopropyl alcohol (for instance, 120–140 mL) with some drops of phenolphthalein
(for instance, 7–9 drops).
• The sample should be mixed at this point with the titration solvent solution made
previously.
• At this stage, you can titrate with KOH solution (with known concentration of the
basis, for instance, 0.1 N).
• Titrate until the color is purple and the color is stable for a sufficient time (for
instance 20–30 s).
The acid number is calculated by the following:

[(A − B) · N · M W ]
Acid number = (19.1)
Ws

where

A volume (L) of KOH solution necessary for the titration;


270 19 Some Chemical Analyses in Biodiesel Production …

Fig. 19.3 Mechanism of iodine saturation of the double bonds in unsaturated molecules

B volume (L) of titrant used in the blank trial (no biodiesel used);
N normality of titrant (0.1 in this case);
MW molecular weight of the base used for the titration, in this case KOH, 56.1;
Ws weight of sample in grams.

19.2 Iodine Number Determination

The iodine number is commonly measured to have an indication of the saturation


degree of oils or in this case biofuels like biodiesel. The degree of saturation reflects
the amount of double bonds present within the sample molecule. There is a high
correlation between the degree of saturation and the iodine number; for instance, if
we consider the average number of double bonds in the sample, the iodine number
is higher if the amount of double bonds is larger (Giakoumis 2018). Unsaturated
oils’ examples are the canola oil, corn oil, linseed oil, and safflower oil. The result
from this kind of analysis will, in practice, give the amount of iodine required for
performing the saturation of one hundred grams of the sample.
The principle on the basis of this kind of analysis is that the molecules of iodine
monochloride will eventually saturate the double bonds present in the molecules. This
is possible because the iodine monochloride molecule is quite polar with a positive
and negative part, Fig. 19.3, and the double bond (where there is an overcharge of
electrons) will be attacked by the positive part of the iodine monochloride molecule.
The next step would be the formation of a positive site within the oil molecule where
the negatively charged chlorine will attach the remaining “free” space.
19.2 Iodine Number Determination 271

The titration, in this case, is concerning the presence of the remaining iodine
monochloride left into the solution. The method is also denominated as Andrews’
titration after Launcelot W. Andrews in 1903 (Andrews 1903).
The commercially available solution for the iodine monochloride molecules is
called Wij’s solution. For the determination of the iodine number of a particular
sample, the procedure is as follows:
• It is necessary to measure the mass of the sample (for instance, oil or biodiesel),
in this case, a suitable amount would be 0.1–0.2 g.
• At this point, we need a sufficient amount of iodine monochloride. This is available
as Wij’s solution. Add a sufficient amount, for instance, 25–30 mL so we are sure
that this amount is enough to titrate the sample.
• At this stage, we need to add the solvent which, in this case, is the trichloroethene.
A quantity of 10–20 mL will be sufficient in this case.
• The procedure continues by mixing the test flask. Then, some waiting time is
required for the dissolution to take place; therefore, it is good to close the flask
and take it in the dark for half hour at room temperature.
• It is also required to do the same procedure as for the steps written previously, but
in this case, there should be no sample (oil or biodiesel) in the flask. This is called
Blank, we need this titration as a reference.
• After keeping the flask in the dark, it is necessary to add to the solution potassium
iodide. For instance, 15–20 mL of 10% potassium iodide solution (this is an excess
to ensure that all the I-Cl is used). In this way, the leftover of the Wij solution will
oxidize the Iodine in the KI solution, 2I− = I2.
• At this point, it is necessary to titrate the I2 with sodium thiosulphate solution.
This is done in the presence of fresh starch.

– A suitable titration medium would be sodium thiosulphate solution with a known


concentration (for instance, 0.1 M). In this step, the I2 is reduced back to iodide,
I− ions. The titration is continuing until a yellow color appears (straw yellow).
– Then, fresh starch is added (some drops of starch solution), and the color of the
solution turns blue.
– Finally, the titration is continued with sodium thiosulphate until the color turns
white.

This procedure is similar to other procedures found in literature (Yasuda 1931).


The iodine value is calculated as follows:
 g 
(Vblank − Voil ) · [Thio] · M Wt mol 100 g
Iodine value = (19.2)
grams of oil 1000 cc/L

where
V (blank) and V (oil) is the volume of sodium thiosulphate used in titration for the
sample and the blank;
[Thio] is the concentration of the thiosulphate;
M Wt , is the molecular weight of thiosulphate, 126.9 g/mol.
The titration of the I2 with sodium thiosulphate is shown as:
272 19 Some Chemical Analyses in Biodiesel Production …

2Na2 S2 O3 + I2 = Na2 S2 O4 + 2NaI (19.3)

19.3 Saponification Value of Fats and Oils

As we have seen previously within the explanation of the production process for
biodiesel, soaps can be formed by the reaction of the fatty acids and an excess of
homogeneous catalyst, like KOH. The evaluation of the saponification value utilizes
just this reaction. By definition, with saponification value, we consider the milligrams
of KOH needed to neutralize the fatty acids formed by the complete hydrolysis of
one gram of fat or oil. The main reactions in this case are as follows:

Fat + water + KOH = fatty acids + alcohol (19.4)

Fatty acid + KOH = Salt of Fatty acid + alcohol (19.5)

In these reactions, a certain quantity of KOH is consumed from a known quantity


put into the solution and the principle is to titrate the remaining quantity of KOH.
A suitable example for the second reaction reported above is the reaction between
tristearin and KOH:

Tristearin + KOH → glycerol + potassium stearate (3KOOCC17 H35 ) (19.6)

The structure of the tristearin is reported in Fig. 19.4.

Fig. 19.4 Tristearin


structure
19.3 Saponification Value of Fats and Oils 273

The procedure for the estimation of the saponification value is as follows:


• Also, in this case, it is necessary to weight a sufficient amount of sample. For
instance, 0.5 g or 1 g of oil or biodiesel.
• We have then to dissolve the sample in KOH, and this is normally done by boiling
under reflux with excess KOH (for example, 40 mL or 50 mL of KOH 0.1 M).
• We find the moles of KOH left by titration with HCl (for instance, 0.1 M HCl and
some drops of phenolphthalein). The reaction in this case is:

KOH + HCl = KCl + H2 O (19.7)

• At the end of the titration, the color goes from pink to white.

The calculations to be performed are:


• Number of moles used (left) = (concentration) − (volume used);
• Moles of KOH reacted = (initial moles) − (moles left);
• Ml of KOH = 1000 − (moles reacted) − (molecular mass, 56);
• Divide this amount by the grams of oil.

19.4 Biodiesel Characteristics and Properties

When biodiesel is produced, it is necessary to evaluate also some physical charac-


teristics of the fuel. This is a very important assessment since, in order to be utilized
in an engine, biodiesel has to behave in a good manner with respect to:
• The cloud point, which is the temperature at which the crystals of solid biodiesel
first become visible, in case we decrease the temperature.
• The pour point also called gel point, this is the temperature where biodiesel
becomes solid and can no longer be pumped.
• The flash point, this is the minimum temperature at which the vapors of the fuel
ignite using an ignition source.
• The autoignition temperature, this is the temperature at which the vapors ignite
with no source of ignition.
• The fire-point, this is the minimum temperature at which the flame is still burning
with no source of ignition.
Table 19.1 shows some characteristics of biodiesel. In regard to the table, the
nomenclature is as follows: viscosity (vis), acid value (AV), cloud point (CP), pour
point (PP), flash point (FP), density (Den), and water content (WC).
While as a comparison, some properties of diesel and biodiesel are shown in
Table 19.2.
Equation (19.8) (Satyanarayana and Rao 1992) is an example of flash point cal-
culation. Notice that there are also other methods to calculate this value.
274 19 Some Chemical Analyses in Biodiesel Production …

Table 19.1 Biodiesel characteristics (adapted from Phan and Phan 2008)
Parameter EN 14214 Cetinkaya Georgogianni Encinar et al. Phan and
and Karaos- et al. (2007) (2005) Phan (2008)
manoglu
(2004)
Vis 40 °C, 3.5–5.0 5.29–6.46 4.76–4.45 4.8 4.89
mm2 /s
AV, max mg 0.5 0.289 0.8–0.5 – 0.43
KOH/g
CP, °C – 9 (−3) to (−4) 4.7 3
PP, °C – −3 (−4) to (−5) −3.9 0
FP, °C >1.1 176 67–83 177 120
Den (g/mL) 0.86–0.90 0.88–0.89 0.85–0.82 0.89 0.88
WC, max, 500 480.07 – – Trace
ppm
The original references are also reported within the table

Table 19.2 Properties of


Property Diesel Biodiesel
diesel and biodiesel
Oxygen (wt%) 0 11
Composition HC (10–C21) FAME
(C12–C22)
Kinetic viscosity 1.9–4.1 1.9–6.0
(mm2 /s) at 40 °C
Water (vol.%) 0.05 0.05
Specific gravity 0.85 0.88
(g/mL)
Flash point, °C 60–80 100–170
Cloud point, °C (−15) to (5) (−3) to (12)
Pour point, °C (−35) to (−15) (−15) to (16)
Carbon (wt%) 87 77
Hydrogen (wt%) 13 12
BOCLE scuff (g) 3600 >7000
Sulfur (wt%) 0.05 0.05
Cetane number 40–55 48–60
HFRR (µm) 685 314
Standard ASTM D975 ASTM D6751
HC hydrocarbons, FAME fatty acid methyl esters, HFRR high-
frequency reciprocating ring, BOCLE ball on cylinder lubricity
evaluator
Adapted from Lotero et al. (2005)
19.4 Biodiesel Characteristics and Properties 275

Table 19.3 a, b, c coefficients for different groups


Group Number of compounds A B C AAE%
Hydrocarbons 230 225.10 537.60 2217.00 0.410
Alcohols 150 230.80 390.50 1780.00 0.186
Amines 70 222.40 416.60 1900.00 0.2
Acids 40 323.20 600.10 2970.00 0.531
Ethers 80 275.90 700.00 2879.00 0.343
Sulfur 40 238.00 577.90 2297.00 0.484
Esters 120 260.80 449.20 2217.00 0.186
Ketones 80 260.50 296.00 1908.00 0.722
Halogens 200 262.10 414.00 2154.00 0.672
Aldehydes 45 264.50 293.00 1970.00 0.771
Phosphorus 20 201.70 416.10 1666.00 0.815
Nitrogens 125 185.70 432.00 1645.00 0.622
Petroleum fractions 21 237.90 334.40 1807.00 0.294
Data taken from Dean (1985), Aldrich Chemical Company, Inc (1990), Weiss (1980)
AAE% is the percentage of average absolute error in predicted flash point

 2 c

b Tcb e Tb
Tf = a +  2 (19.8)
− c
1 − e Tb

where T f and T b , are the flash and the normal boiling point in Kelvin, while the
parameters a, b, c are given in Table 19.3.
The letters a, b, c refer to the formula for the calculation of the flash point given
previously.
Additional methods and more accurate explanation of the procedures utilized are
available in literature (Rand 2010); however, some more examples will be given here.
For instance, the cetane number (CN) is among the most important parameters for
biodiesel assessment, this value is calculated according to Date (2012):

CN = % by volume of n_hexadecane + 0.15 · (% by volume of HMN) (19.9)

where HMN indicates the 2,2,4,4,6,8,8-heptamethylnonane.


The CN is a measure of the lowest compression ratio that would produce autoigni-
tion.
In the cetane test, we measure the time when the first detectable increase of
pressure is taking in place. When this is happening, it means that the fuel is ignited.
Figure 19.5 describes the structures of the cetane and the isocetane. In the figure,
the white circles represent the hydrogen atoms, while the line crossings have carbon
276 19 Some Chemical Analyses in Biodiesel Production …

Fig. 19.5 a Cetane (normal hexadecane) C16H34, cetane number = 100. b Isocetane C16H34,
2,2,4,4,6,8,8-heptamethylnonane with cetane number = 15

atoms, therefore each of the crossings represents a carbon atom (Mollenhauer and
Tschöke 2010) (Fig. 19.5).
In biodiesel, it is also possible to add additives. For instance, a quantity of ethanol
can be commonly added to biodiesel. Ethanol has 60% of the energy of diesel, but
the octane number is higher. Therefore, the energy amount can be very similar to the
one of diesel.
Concerning the definition of octane number: an engine which has the same self-
ignition properties of a mixture of 2.2,4-trimethylpentane, the so-called iso-octane,
has the octane number equal to the percentage of iso-octane in an iso-octane–heptane
mixture.
The heptane formula is: C7 H16 , while the iso-octane is represented in Fig. 19.6.

Fig. 19.6 Structure of


Iso-octane. Adapted from
Morrison and Boyd (1992)
19.4 Biodiesel Characteristics and Properties 277

In diesel engines, a high cetane number and low octane number is favorable, this
is because the injection of the fuel is done after the compression of the air; therefore,
the fuel should ignite soon after an increase of pressure.
On the other hand, in gasoline engines, the fuel is compressed together with the
air and when the piston reaches the highest pressure, a spark is released to ignite the
fuel and the air mixture. In gasoline engines, we do not want to have ignition before
the spark and therefore we want high octane number and low cetane number.
Table 19.4 shows diverse kinds of feedstock with values for their level of unsat-
uration. As described elsewhere in this manuscript, the iodine number represents
the degree of unsaturation, while the cetane number is a measure of the facility of
ignition of the diesel. The more the cetane number the less time takes for the fuel to
ignite.
In Table 19.4, the MUFA refers to the monounsaturated fatty acids, PUFA stands
for the polyunsaturated acids. While the other abbreviations refer to safflower (SAF),
grape (GRP), silybum marianum (SIL), hemp (HMP), sunflower (SFL), wheat germ
(WHG), pumpkin seed (PMS), sesame (SES), rice bran (RB), almond (ALM), rape-
seed (RPS), peanut (PNT), olive (OL), and coconut oil (COC).
Data on the characteristics of different kinds of feedstock and fuels are reported
in Table 19.5. Among the reported properties, we have the density, the kinematic
viscosity (Kin-VIS), the cetane number (CN), the higher heating value (HHV), the
flash point (FP), the carbon residue (CR), the iodine value (IV), and the pour point
(PP).
Biodiesel has been also produced by utilizing microorganisms, and there is large
amount of research on this topic at the moment. Some microbial-based source of
biodiesel is given in Table 19.6.
As additional data, in Table 19.7, some properties of diesel, biodiesel, and different
kinds of fuels are given. It is useful to do this kind of comparisons when assessing
the suitability of fuels and feedstock. In regard to the table: diesel (D), biodiesel
(BD), sunflower oil (SO), rapeseed oil (RO), soy (S), sunflower oil (SFO), gasoline
(G), ethanol (E), methanol (M), viscosity (Vis), density (Den), flash point (FP),
boiling point (BP), cloud point (CD), pour point (PP), combustion point (ComP),
autoignition temperature (AT), distillation (Dis), cetane index (CI), octane number
(ONum), lubricity (Lub), and heat of vaporization (HOV).
Among the latest trends in biodiesel production, enzyme-catalyzed enzymatic
transesterification is one of the techniques that could be applied as novel technology.
The fatty acid methyl ester (FAME) can be produced from waste choice white
grease (CWG) biocatalyzed by immobilized lipase (Adewale et al. 2015). In this
case, the enzyme could be, for instance, the Candida Antarctica lipase B (CALB).
The yield of FAME is given by:

FAME
Yield(FAME, %) =    (19.10)
TG + DG + MG + FFA + FAME

TG, DG, MG, and FFA are triacylglycerides, diacylglycerides, monoacylglyc-


erides, and free fatty acids, respectively.
278 19 Some Chemical Analyses in Biodiesel Production …

Table 19.4 Diverse kinds of feedstock with values for their level of unsaturation
FA SAF GRP SIL HMP SFL WHG PMS SES RB ALM RPS PNT OL COC
[%]
C6:0 0.52
C8:0 0.01 7.6
C10:0 0.01 5.5
C12:0 0.01 0.01 0.02 0.07 0.09 47.7
C14:0 0.10 0.05 0.09 0.07 0.09 0.17 0.39 0.07 0.04 19.9
C15:0 0.01 0.02 0.04 0.02
C16:0 6.7 6.6 7.9 6.4 6.2 17.4 13.1 9.7 20.0 6.8 4.6 7.5 16.5
C17:0 0.04 0.06 0.06 0.05 0.02 0.03 0.13 0.05 0.04 0.07
C18:0 2.4 3.5 4.5 2.6 2.8 0.7 5.7 6.5 2.1 2.3 1.7 2.1 2.3 2.7
C20:0 0.16 2.6 0.21 0.47 0.63 0.09 1.01 0.43
C22:0 0.14 0.15
C16:1 0.08 0.08 0.05 0.11 0.12 0.21 0.12 0.11 0.19 0.53 0.21 0.07 1.8
(n −
7)
C17:1 0.03
(n −
7)
C18:1cis 11.5 14.3 20.4 11.5 28.0 12.7 24.9 41.5 42.7 67.2 63.3 71.1 66.4 6.2
(n −
9)
C18:1trans 0.14
(n −
9)
C20:1 0.40 0.15 16.5 0.18 7.91 1.08 0.32 1.11 0.16 9.1 0.30
(n −
9)
C18:2cis 79.0 74.7 63.3 59.4 62.2 59.7 54.2 40.9 33.1 22.8 19.6 18.2 16.4 1.6
(n −
6)
C18:3 0.15 0.15 0.88 0.36 0.16 1.2 0.12 0.21 0.45 1.2 1.6
(n −
3)
C18:3 3.0
(n −
6)
SFAs 9.3 10.4 15.1 9.2 9.4 18.2 19.6 16.9 22.5 9.3 6.3 10.7 19.4 92.1
MUFAs 11.6 14.8 20.7 28.1 28.3 20.9 26.1 42.0 44.0 67.9 72.8 71.1 68.2 6.2
PUFAs 79.1 74.9 64.2 62.8 62.4 61.0 54.3 41.2 33.6 22.8 20.9 18.2 18.0 1.6
n−3 0.2 0.2 0.9 0.4 0.2 1.2 0.1 0.2 0.5 0.0 1.2 0.0 1.6 0.0
PUFAs
n−6 79.0 74.7 63.3 62.4 62.2 59.7 54.2 40.9 33.1 22.8 19.6 18.2 16.4 1.6
PUFAs

Adapted from Orsavova et al. (2015)


Table 19.5 Fuel properties of different feedstock kinds and biofuels
Feed Density (kg/m3 ) Kin-VIS (cS) CN HHV (MJ/kg) FP (°C) CR (wt%) IV (gl2/g) PP (°C) S (wt%)
Diesel 820–890 1.6–5.9 >50 >45 >61 <0.15 <1.0
Fuel-oil 920–990 <180 >30 >43 >66 <12 <15 <3
Biodiesel 860–900 3.5–5 >51 120 0.3 120 1
Babassu 946 30 38 150 16
Castor 955 251 42 37.4 83–86
Coconut 918 27 40–42 37.1 8–11 0.01
Corn 910 31–35 38 39.5 277 0.24 103–128 −40 0.01
Cottonseed 915 34 42 38.7–39.5 234 0.24 103–115 −15 to 12 0.01
19.4 Biodiesel Characteristics and Properties

Crambe 905 54 45 40.5 274 0.23


Jatropha 940 34 39 38.8 225 82–98 −15 0.01
Linseed 924 26–27 35 39.3–39.5 241 0.22 180
Mahua 960 25 36 232 58–70
Neem 919 50 65–80
Palm 918 40–45 42 39.5 267 0.23 48–58 −7 0.01
Peanut 903 40 42 39.8 271 0.24 84–100 −32 0.01
Rapeseed 912 35–37 41 39.7 246 0.3 105 −9 0.01
Sesame 913 36 40–42 39.4 260 0.24 103–106 −12 0.01
Soybean 914 29–33 38 39.6 254 0.25 128–143 −7 0.01
Safflower 914 31 41 39.5 260 0.25 145 −15
Sunflower 916 34–36 37 39.6 274 0.27 125–140 0.01
Adapted from Blin et al. (2013)
279
Table 19.6 Examples of microbial-based source of Biodiesel
280

Microbial-based Oil content (% Lauric Palmitic 16:1 18:0 Olein 18:1 Linoleic Linolenic 18:3 ArA 20:4 ω − 6 EpA 20:5 ω − 3 DhA 22:6 ω − 3
oils dw) 14:0 16:0 18:2
Algae
Amphidinium – 8.0 15.0 5.0 – 5.0 6.0 17.0 – 4.0 2.0
carterae
Botryococcus 50 – 15.4 10.6 28.2 13.3 22.12 – – – –
braunii
Chlorella spp 30 – 25.0 2.0 0.9 5.0 20 19.0 – – –
Chlorella 45 – 22.0 3.0 0.9 6.5 18 27.0 – – –
pyrenoidosa
Chlorella 38 – 26.0 2.0 0.8 16.0 24 20.0 – – –
vulgaris
Chlorella 55 1.3 12.9 – 2.7 60.8 17.3 – – – –
protothecoides
Crypthecodinium 15 17.0 17.0 1.0 3.0 10.0 – – – – 44.0
cohnii
Cylindrotheca 14 – 25.1 27.0 1.6 3.8 0.5 0.5 9.1 12 0.7
fusiformis
Dunaliella 8 – 1.7 1.3 – 2.5 2.5 0.5 1.5 1.02 –
bardawil
Isochrysis 29 12.0 10.0 11.0 0.7 3.0 2.0 – – 25.0 11.0
galbana
Nannochloropsis 50 – 27.9 32.4 2.1 10.4 1.9 – 2.1 20.1 0.5
spp
(continued)
19 Some Chemical Analyses in Biodiesel Production …
Table 19.6 (continued)
Microbial-based Oil content (% Lauric Palmitic 16:1 18:0 Olein 18:1 Linoleic Linolenic 18:3 ArA 20:4 ω − 6 EpA 20:5 ω − 3 DhA 22:6 ω − 3
oils dw) 14:0 16:0 18:2
Neochloris 45 – – – – – – – – – –
oleoabundans
Mortiella alpine 56 2.4 15.0 – 2.3 10.0 7.2 4.0 40.3 – –
1S-4
Mortiella 43 – 9.4 – 3.5 50.9 8.2 3.5 16.5 – –
elongate
Mucor 20 – 23 1 – 40 11 16 – – –
circinelloides
Penicillium 64 – 18 – 12 12 43 21 – – –
spinulosum
Pythium 43 16.8 18.6 4.1 2.8 17.3 16.0 1.2 8.2 10.5 –
irregular
Rhizopus 57 – – – 6 – – – – – –
19.4 Biodiesel Characteristics and Properties

arrhizus
Yeast
Cryptococcus 65 – 12.0 1.0 3 73 12 – – – –
albidus
Lipomyces 63 – 34 6 5 51 3 – – – –
starkeyi
Rhodotorula 72 – 37 1 3 47 8 – – – –
glutinis
Trichosporon 65 – 15 – 2 57 24 1 – – –
pullulans

Adapted from Drapcho et al. (2008b)


281
Table 19.7 Properties of diesel, biodiesel, and different kinds of fuels
282

Property Units D BD G SO RO S SFO S E M


Vis (cS at 40 °C) V (40 °C). (cSt) 1.9–4.1 1.9–6 0.8 4.2–4.6 4.8 4.1–4.5 4.5 4.4 1.5 0.75
Den (g/ml) g/mL 0.85 0.88 0.72–0.78 0.86 0.89 0.89 0.87 0.88 0.8 0.8
FP (°C) °C 60–80 100–170 −43 164–183 153 141–188 187 195 13 11
BP (°C) °C 210–235 – 30–225 – – 339 – – 78 65
CP (°C) °C −19 to 5 −3 to 12 – −3 −3 2 −1 1 – –
PP (°C) (°C) −35 to −15 – 15 to – – 6 – – 2 – 8 – 4 – –
16
ComP (°C) °C 92 – – 183 – 171 192 – – –
AT (°C) °C 254 – 370 – – – – – 423 464
Dis (°C) 50% (°C) – – – – – – 353 344 – –
CI [–] [–] 40–55 48–60 <15 49 52 45–47 49.4 48.2 <15 <15
ONum [–] [–] – – 82–92 – – – – – 90–102 89–102
Lub (HFRR µm) HFRR (µm) 685 314 – – – – – – – –
Lub (BOCLE g) BOCLE (g) 3600 >7000 – – – – – – – –
HHV (MJ/kg) 45.2 – 43.5 40.1 40.0 38.8 38.9 40.0 27 20.1
HOV (kJ/kg) kJ/kg 375 – 380 – – – – – 920 1185
H2 O (%vol) %vol 0.05 0.05 – – – – – – – –
Elemental analysis
C (wt%) wt% 87 77 – – – – – – – –
H (wt%) wt% 13 12 – – – – – – – –
O (wt%) wt% 0 11 – – – – – – –
S (wt%) wt% 0.05 0.05 – 0.01 <0.01 0.01 <0.01 0.1 – –
Adapted from Drapcho et al. (2008a)
19 Some Chemical Analyses in Biodiesel Production …
19.4 Biodiesel Characteristics and Properties 283

In enzyme-catalyzed enzymatic transesterification for biodiesel production, the


main parameters are:
• the type of biocatalysts used;
• the different solvents used;
• the effect of speed of agitation;
• the biocatalyst amount;
• effects of different alcohols;
• temperature.
The enzymatic transesterification follows the so-called Ping Pong Bi Bi, mecha-
nism (El Rassy et al. 2004). The main reactions in this mechanism are:
E
TG + M ←→ F + G (19.11)

k1,k2 k3
E + TG ←→ [ETG] → E + F (19.12)

k4,k5   k3
E + M ←→ E M → E + G (19.13)

where k 1 and k 4 are the forward reaction rate constants, while k 2 and k 5 are the
backward reaction rate constants. Triacylglycerides (TG) and methanol (M) are the
substrates, while FAME (F) and glycerol (G) are the products. In this mechanism, the
enzyme (E) must bind the first substrate (TG) to form an enzyme–substrate complex
[ETG], then release the first product (F), and transform intermediate enzyme (E ).
This is followed by binding of a second substrate (M) to E and the breakdown of
the transitory complex [E M] to free enzyme (E) and the second product (G).
The production/consumption rates for the main compounds in enzyme transester-
ification are expressed as:

d(TG)
= −k1 [TG][E] + k2 [ETG] (19.14)
dt
d(M)    
= −k4 [M] E + k5 E M (19.15)
dt
d(F)
= k3 [ETG] (19.16)
dt
d(G)  
= k6 E M (19.17)
dt
d(E)  
= −k1 [TG][E] + k2 [ETG] + k6 E M (19.18)
dt
 
d E    
= −k4 [M] E + k5 E M + k3 [ETG] (19.19)
dt
d(ETG)
= k1 [TG][E] − k2 [ETG] − k3 [ETG] (19.20)
dt
284 19 Some Chemical Analyses in Biodiesel Production …

 
d E M      
= k4 [M] E − k5 E M − k6 E M (19.21)
dt
where (TG) are triacylglycerides, (M) is methanol, (F) is FAME, (G) is glycerol, (E)
is enzyme, (TG) is the first substrate, [ETG] is the enzyme–substrate complex, (F)
is the first product, (E ) is the intermediate enzyme, (M) is the second substrate, and
[E M] is the transitory complex.

References

Adewale, P., Dumont, M.-J., & Ngadi, M. (2015). Enzyme-catalyzed synthesis and kinetics of
ultrasonic-assisted methanolysis of waste choice white grease for fatty acid methyl ester produc-
tion [WWW Document]. https://doi.org/10.1021/acs.energyfuels.5b00849.
Aldrich Chemical Company, Inc. (1990). Aldrich catalogue handbook of fine chemicals. Aldrich
Chemical Company, Inc.
Andrews, L. W. (1903). Titrations with potassium iodate. Journal of the American Chemical Society,
25, 756–761. https://doi.org/10.1021/ja02009a012.
Aricetti, J. A., & Tubino, M. (2012). A green and simple visual method for the determination of the
acid-number of biodiesel. Fuel, 95, 659–661. https://doi.org/10.1016/j.fuel.2011.10.058.
Baig, A., & Ng, F. T. T. (2011). Determination of acid number of biodiesel and biodiesel blends.
Journal of the American Oil Chemists Society, 88, 243–253. https://doi.org/10.1007/s11746-010-
1667-x.
Blin, J., Brunschwig, C., Chapuis, A., Changotade, O., Sidibe, S. S., Noumi, E. S., et al. (2013).
Characteristics of vegetable oils for use as fuel in stationary diesel engines—Towards specifica-
tions for a standard in West Africa. Renewable and Sustainable Energy Reviews, 22, 580–597.
https://doi.org/10.1016/j.rser.2013.02.018.
Cetinkaya, M., & Karaosmanoglu, F. (2004). Optimisation of base-catalysed transesterification
reaction of used cooking oil. Energy & Fuels, 18, 1888–1895.
Date, A. W. (2012). Analytic combustion with thermodynamics, chemical kinetics and mass transfer.
Cambridge: Cambridge University Press. https://doi.org/10.1017/CBO9780511976759.
Dean, J. A. (1985). Lange’s handbook of chemistry (13th ed.). McGraw-Hill.
Drapcho, C. M., Nghiem, J., & Walker, T. (2008a). Biofuels engineering process technology.
McGraw-Hill Education.
Drapcho, C. M., Nghim, N. P., & Walker, T. (2008b). Biofuels, bioproducts and biorefining.
McGraw-Hill.
El Rassy, H., Perrard, A., & Pierre, A. C. (2004). Application of lipase encapsulated in silica
aerogels to a transesterification reaction in hydrophobic and hydrophilic solvents: Bi-Bi Ping-
Pong kinetics. Journal of Molecular Catalysis. B, Enzymatic, 30, 137–150. https://doi.org/10.
1016/j.molcatb.2004.03.013.
Encinar, J. M., González, J. F., & Rodríguez-Reinares, A. (2005). Biodiesel from used frying oil.
Variables affecting the yields and characteristics of the biodiesel. Industrial and Engineering
Chemistry Research, 44, 5491–5499. https://doi.org/10.1021/ie040214f.
Georgogianni, K. G., Kontominas, M. G., Tegou, E., Avlonitis, D., & Vergis, V. (2007). Biodiesel
production: Reaction and process parameters of alkali-catalysed transesterification of waste
frying-oils. Energy & Fuels, 21, 3023–3027.
Giakoumis, E. G. (2018). Analysis of 22 vegetable oils’ physico-chemical properties and fatty acid
composition on a statistical basis, and correlation with the degree of unsaturation. Renewable
Energy, 126, 403–419. https://doi.org/10.1016/j.renene.2018.03.057.
References 285

Lin, C.-Y., & Lin, Y.-W. (2012). Fuel characteristics of biodiesel produced from a high-acid oil
from soybean soapstock by supercritical-methanol transesterification. Energies, 5, 2370–2380.
https://doi.org/10.3390/en5072370.
Lotero, E., Liu, Y., Lopez, D. E., Suwannakarn, K., Bruce, D. A., & Goodwin, J. G. (2005). Synthesis
of biodiesel via acid catalysis. Industrial and Engineering Chemistry Research, 44, 5353–5363.
Mollenhauer, K., & Tschöke, H. (2010). Handbook of diesel engines. Berlin: Springer.
Morrison, R. T., & Boyd, R. N. (1992). Organic chemistry (6th ed.). Prentice Hall.
Orsavova, J., Misurcova, L., Vavra Ambrozova, J., Vicha, R., & Mlcek, J. (2015). Fatty acids
composition of vegetable oils and its contribution to dietary energy intake and dependence of
cardiovascular mortality on dietary intake of fatty acids. International Journal of Molecular
Sciences, 16, 12871–12890. https://doi.org/10.3390/ijms160612871.
Park, L. K.-E., Liu, J., Yiacoumi, S., Borole, A. P., & Tsouris, C. (2017). Contribution of acidic
components to the total acid number (TAN) of bio-oil. Fuel, 200, 171–181. https://doi.org/10.
1016/j.fuel.2017.03.022.
Patel, N. K., & Shah, S. N. (2015). Biodiesel from plant oils, Chap. 11. In S. Ahuja (Ed.), Food,
energy, and water (pp. 277–307). Boston: Elsevier. https://doi.org/10.1016/B978-0-12-800211-
7.00011-9.
Phan, A. N., & Phan, T. M. (2008). Biodiesel production from waste cooking oils. Fuel, 87,
3490–3496. https://doi.org/10.1016/j.fuel.2008.07.008.
Rand, S. (2010). Significance of tests for petroleum products (8th ed.). United States: American
Society for Testing and Materials, ASTM.
Satyanarayana, K., & Rao, P. G. (1992). Improved equation to estimate flash points of
organic compounds. Journal of Hazardous Materials, 32, 81–85. https://doi.org/10.1016/0304-
3894(92)85106-B.
Weiss, G. (1980). Hazardous chemicals data book. Park Ridge, HJ: Noyes Data Corporation.
Yasuda, M. (1931). The determination of the iodine number of lipids. Journal of Biological Chem-
istry, 94, 401–409.
Zumdahl, S. S., & Zumdahl, S. A. (2014). Chemistry (9th ed.). Brooks Cole Cengage Learning.
Chapter 20
Fischer–Tropsch (FT) Synthesis
to Biofuels (BtL Process)

Nature is like a painting of Monet, very often when you get close
to part of nature, it’s like walking up to an impressionist
painting: you see less and less. All you see is little dots. In order
to understand the meaning and the importance and the law
contained in physics, you must step back and back to see the all
and the perfection.
Robert B. Laughlin, Nobel Prize in Physics 1998.

20.1 Background Information

The FT reaction produces long-chain hydrocarbons from synthetic gas, which is a


mixture of H2 and CO, over CO- or Fe-based catalysts (Triantafyllidis et al. 2013)

nCO + 2nH2 → (−CH2 −)n + nH2 O (20.1)

A short history of this technology is given in (Schulz 1999). In the case of BtL
(Biomass to Liquid) technology, the biomass is converted to synthetic gas by the gasi-
fication process, a thermochemical conversion technology operating at 500–1200 °C
in the presence of a gasifying agent (air, oxygen, steam, CO2 , or a mixture of these
components) (Triantafyllidis et al. 2013).
A schematic view of BtL technology is given in Fig. 20.1.
A diagram showing the gasification of biomass process is given in Fig. 20.2. In
this case, the steam gasification is shown here.
The process shown in Fig. 20.2 demonstrates the gasification of a general biomass
feedstock. The feedstock is processed in a bubbling bed reactor at temperatures of
800–900 °C, and the produced char could be sent to a combustion unit, where it will
be burned to produce energy. The main reactions taking place during this process
are:

C + CO2 → 2CO (20.2)

© Springer Nature Switzerland AG 2019 287


C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_20
288 20 Fischer–Tropsch (FT) Synthesis to Biofuels (BtL Process)

Fig. 20.1 A schematic view of BtL technology

C + H2 O → CO + H2 (20.3)

CO + H2 O ↔ CO2 + H2 (20.4)

C + 2H2 → CH4 (20.5)

On the other hand, a GtL plant includes the conversion of natural gas to synthesis
gas through steam reforming, dry reforming, partial oxidation (POX), or oxidative
steam reforming (autothermal reforming (ATR)).
The steam reforming could be applied to diverse kinds of biomass, for instance,
glycerol, with the presence of catalytic agents (Huang et al. 2018). The reaction of
steam reforming is:

CH4 + H2 O = CO + 3H2 ; (20.6)

while the dry reforming is represented by:

CO2 + CH4 → 2H2 + 2CO (20.7)

Partial oxidation (POX) of biomass is an interesting method to derive energy from


a feedstock with not necessity to destroy it completely (Özdenkçi et al. 2017):

CH4 + 0.5O2 = 2H2 + CO; (20.8)

while the autothermal reforming reactions (ATR) are demonstrated as (Triantafyllidis


et al. 2013):

CH4 + H2 O → CO + 2H2 O, H R = 206 kJ/mol (20.9)

3
CH4 + O2 → CO + 2H2 O, H R = −520 kJ/mol (20.10)
2
CO + H2 O → CO2 + H2 , H R = −41 kJ/mol (20.11)

Notice that sometimes in literature the reactions are not balanced. The enthalpies
are calculated on one mole of methane. A more general representation of the reactions
in Fisher–Tropsch synthesis is given in Eqs. (20.12–20.21)
20.1 Background Information 289

Fig. 20.2 A diagram showing the steam gasification of biomass process

nCO + 2nH2 → (−CH2 −)n + nH2 O (20.12)

2CO + H2 → (−CH2 −) + CO2 , H300K = −204 kJ/mol (20.13)

3CO + H2 → (−CH2 −) + 2CO2 , H300K = −244.5 kJ/mol (20.14)

CO2 + H2 → (−CH2 −) + H2 O, H300K = −125.2 kJ/mol (20.15)

CO + H2 O → H2 + CO2 , H300K = −39.8 kJ/mol (20.16)

Alkanes : nCO + (2n + 1)H2 → Cn H2n+2 + nH2 O (20.17)

Alkenes : nCO + (2n)H2 → Cn H2n + nH2 O (20.18)

Alcohols : nCO + 2nH2 → Cn H2n+2 O + (n − 1)H2 O (20.19)

Aldehydes − ketones : nCO + (2n − 1)H2 → Cn H2n O + (n − 1)H2 O (20.20)

Carboxylic acids : nCO + (2n − 2)H2 → Cn H2n O2 + (n − 2)H2 O (20.21)

Concerning the catalyst type and H2 /CO ratio in FT synthesis, a Fe-based catalyst
has higher Water-Gas Shift activity in general (Ma et al. 2015) and also compared to
a Co-based catalyst increasing the H2 /CO ratio. In the case of Cobalt-based cataly-
sis, the products might be shifted from long-chain hydrocarbons to mostly methane
(Riedel et al. 1999). This ratio has a significant impact on CO conversion, C5+ selec-
tivity, product distribution, and chain growth probability. A low inlet H2 /CO ratio
leads to selectivity to longer hydrocarbons (increase of C5+ selectivity) and a drop
in the hydrocarbon formation rate due to the decrease in productivity of the catalyst.
A reduction of the hydrogen in the feed leads to a lower selectivity to CH4 , but an
increase in the olefin/paraffin ratio; while CO conversion increases at high H2 /CO.
290 20 Fischer–Tropsch (FT) Synthesis to Biofuels (BtL Process)

Activities of different metal catalysts on different supports and reaction rates are
given in literature (Demitras and Muetterties 1977):

Ru > Fe > Ni > Co > Rh on Alumina (20.22)

Co > Fe > Ru > Ni > Rh on silica (20.23)

We have that, on alumina, the Ru has higher activity than Fe and so on. The kind
of support used for the catalyst metal influences the activity of the metal itself. In
this way, we have that the selectivity of the C5+ hydrocarbons is higher going from
Co/TiO2 to Co/MgO as follows (Jahangiri et al. 2014):

Co/TiO2 > Co/SiO2 > Co/Al2 O3 > Co/C > Co/MgO (20.24)

In conventional FT processes, four types of reactors are used (Triantafyllidis et al.


2013):
• fixed bed multitubular reactors;
• fluidized bed reactors;
• slurry-bed reactors;
• microchannel reactors.
However, we can find also the circulating fluidized bed reactor among the suitable
options (Steynberg et al. 2004, p. 2).
As reported by the same references, fluidized bed reactors have:
• superior quality in heat transfer and the temperature control could be performed
in a better way during highly exothermic FT reactions;
• there is the possibility of using smaller catalyst particles, in this way, we can
avoid intra-particle diffusion. This phenomenon can limit the reaction rate and the
pressure drop;
• as also commonly referred to fluidized bed reactors, we have better mixing of the
catalyst particles due to fluidization and a higher gas–solid mass transfer efficiency;
• the catalyst in this case can be replaced in a shorter time and there is the possibility
of loading fresh catalyst during the run;
• high production capacity due to the possibility of using more gas flows.
As an example of slurry-bed reactor, the SASOL slurry-bed reactor is demon-
strated in Fig. 20.3.
An important drawback of the slurry reactor is the separation of the catalyst from
the waxes. In Table 20.1, the main reactors parameters in FT synthesis are given
while in Table 20.2 also the H2 /CO ratio is reported. The original references are
given within the two tables.
Some effects of the main parameters on FT synthesis are described as: (Triantafyl-
lidis et al. 2013):
Table 20.1 Main reactor parameters in FT synthesis
Catalyst Reactor Reducing agent Temperature Temperature rise Pressure (bar) Space velocity Hours (h)
(°C) (°C/min) (l · g −1 h −1
Fe/Cu/kSiO2 Stirred, Slurry Syngas H2 /CO 270 0.17 10–25 1 13
(Hao et al. 2009) = 0.67
Fe/Cu/La/SiO2 Fixed bed 5% H2 /N2 gas 400 5 Atm 15.1 1
(Pour et al. 2008) mixture
Fe–Mn ultrafine Stirred, slurry Syngas H2 /CO 275 – 35 1 32
20.1 Background Information

catalyst (Bai =2
et al. 2002)
Fe–Mn catalyst Stirred, slurry Syngas H2 /CO 280 – 15–20 0.23 48
(Ji et al. 2001) =1
FeCrAlY foam Microchannel 5% H2 /He 350 – Atm – 12
(Chin et al. 2005)
Fe–Mn catalyst Fixed bed Syngas H2 /CO 400 – 25–30 1 32
(Liu et al. 2007) =2
Cu–Ru/Al2 O3 Stirred tank, H2 400 2 Atm 1.2 24
(Sari et al. 2009) slurry
Co–Ru/Al2 O3 Fixed bed H2 400 1 Atm – 12
(Tavasoli et al.
2010)
Co–Re/Al2 O3 Fixed bed H2 350 1 1 – 16
(Borg et al.
2007)
(continued)
291
Table 20.1 (continued)
292

Catalyst Reactor Reducing agent Temperature Temperature rise Pressure (bar) Space velocity Hours (h)
(°C) (°C/min) (l · g −1 h −1
Co–Re/Al2 O3 Fixed bed H2 250-350 – Atm – 8
(Oukaci et al.
1999)
Co–Re/Al2 O3 Fixed bed H2 /He = 0.5 Step 1: 100 Step 1: 2 Atm – Step 1: 1
(Das et al. 2003) Step 2: 350 Step 2: 1 Step 2: 10
Co–Pt/Al2 O3 Fixed bed H2 350 – Atm – 8
(Xu et al. 2005)
20 Fischer–Tropsch (FT) Synthesis to Biofuels (BtL Process)
Table 20.2 Main reactor parameters in FT synthesis with conversion and H2 /CO ratio
Catalyst Reactor Temperature Pressure (bar) H2 /CO Space velocity (l · g −1 h −1 CO conversion
(°C) (%)
Fe–Mn ultrafine catalyst (Bai Stirred tank, 260–300 15–31 0.65–2 1–2.5 91–95
et al. 2002) slurry
Fe–Mn catalyst (Ji et al. Stirred tank, 260–290 9.3–25.3 0.8–2.5 1–6.6 60–80
20.1 Background Information

2001) slurry
Fe–Mn catalyst (Liu et al. Fixed bed 280–340 22.5 1.01–2.74 0.8–5.6 64–88
2007)
Fe–Mn–K/SiO2 (Zhang et al. Slurry reactor 250 15 1.35–1.5 – 59–74
2006)
Fe–Mn–Cu–K/SiO2 Slurry reactor 250 15 1.35–1.5 – 62–76
(Zhang et al. 2006)
Co–Ru/Al2 O3 (Dalai and Stirred tank, 210–240 20–35 1–2.5 0.5–1.5
Davis 2008) slurry
Co–Ru/Al2 O3 (de Klerk and Fixed bed 210–240 25 0.5–2 0.448
Furimsky 2010)
Co–Re/Al2 O3 (Borg et al. Fixed bed 210 20 2.1 – 40–50
2007)
Co–Re/Al2 O3 (Das et al. Fixed bed 210 19.7 2 2 70–75
2003)
Co–Pt/Al2 O3 (Xu et al. Slurry 230 20 2 – 61.3
2005)
293
294 20 Fischer–Tropsch (FT) Synthesis to Biofuels (BtL Process)

Fig. 20.3 SASOL


slurry-bed reactor. Adapted
from Steynberg et al. (2004,
p. 2)

• higher temperature increases CO conversion, CH4 and C2 –C4 selectivity. The


selectivity for olefins and oxygenates is also increased. There is, however, a
decrease in the selectivity toward carbon chains with more than five carbon atoms.
• On the other hand, the increase in pressure increases the CO conversion and the
selectivity for chains with more than five carbon atoms. This means that also the
chain growth probability, α will be increased. More explanations on the meaning
for the chain growth probability will be given later at the end of the chapter.
• Increasing the H2 /CO ratio results in a higher CO conversion. It can be noticed
also an increase in the selectivity of alkenes, and a decrease in the selectivity for
carbon chains with more than five carbon atoms due to the enhancement of H2
species.
• A higher space velocity (1/residence time) results in a sharp decrease in CO con-
version, due to a decrease in the residence time of the reactants and products.
While some of the reaction kinetics proposed in literature are given in Table 20.3.
More details are given in Table 20.4 where the kind of catalyst, reactor type,
temperature, and pressure are given together with the H2 /CO ratio.
The intrinsic kinetic expressions presented by the references given in Table 20.4
are, respectively, as follows:

a PH22 PCO
−RH2 +CO =   (20.25)
1 + b PH22 PCO
−0.5
−RH2 +CO = a PH2 PCO (20.26)
20.1 Background Information 295

Table 20.3 Intrinsic kinetic expressions for the reaction between carbon monoxide and hydrogen
Intrinsic kinetic expression Catalyst used
2
apH Co/MgO/ThO2 /Diatomite Brötz (1949)
−RH2 +CO = pCO
2

0.55
apH Co/La2 O3 /Al2 O3 Pannell et al. (1981)
−RH2 +CO = 2
0.33
pCO
2 p
apH CO Co/ThO2 /Diatomite Anderson (1956a)
−RH2 +CO = 
2
2 
1+bpH
2 pCO

apH2 Co/CuO/Al2 O3 Yang et al. (1979)


−RH2 +CO = 0.5
pCO
apH2 pCO Fused Fe/K Atwood and Bennett (1979)
−RH2 +CO = pCO +bpH2 O
0.5
apH2 pCO Co/Al2 O3 Outi et al. (1981)
−RCO =  
0.5 3
1+bpCO
0.5 p 0.5
apH Co/Diatomite Sarup and Wojciechowski
−RCO = 
CO
2
2
(1989)
0.5 +cp 0.5 +dp
1+bpCO H CO
2

0.5 p
apH Co/Diatomite Wojciechowski (1988)
CO
−RCO =  2
2
1+bpCO +cpH
0.5
2

apH2 pCO Co/MgO/SiO2 Chanenchuk et al. (1991)


−RCO =
(1+bpCO )2
2 p
apH CO Precipitated Fe Deckwer et al. (1986)
−RH2 +CO = 2
pCO pH2 +bpH2 O

Table 20.4 Adapted from Yates and Satterfield (1991)


Catalyst Reactor T (°C) P (MPa) H2 /CO
type
Anderson Co/ThO2 /Diatomaceous Fixed bed 186–207 0.1 0.9–3.5
(1956b) earth
Yang et al. Co/CuO/Al2 O3 Fixed bed 235–270 0.17–5.5 1.0–3.0
(1979)
Outi et al. Co/Al2 O3 Fixed bed 250 0.1 0.2–4.0
(1981)
Sarup and Co/Diatomaceous earth Berty 190 0.2–1.5 0.5–8.3
Woj-
ciechowski
(1989)
296 20 Fischer–Tropsch (FT) Synthesis to Biofuels (BtL Process)

Fig. 20.4 Chain growth for compounds with n size on the catalyst. Adapted from Cheng et al.
(2008). The (ad) means adsorbed on the catalyst, while (g) refers to the release of compounds in
the gas phase

 
0.5 3
−RCO = a PH2 PCO
0.5
/ 1 + b PCO (20.27)
 2
−RCO = a PCO PH2 / 1 + b PCO
0.5 0.5 0.5
+ c PH0.5
2
+ d PCO (20.28)

20.2 Chain Growth Probability (Products Distribution)


for the FT Synthesis

The chain growth probability, α, related to an element with n carbon number, is


defined as:
∞
mi
αn = i=n+1
∞ ; (20.29)
m
i=n i

where mn is the molar fraction of the component with n carbon atoms. The chain
growth probability can be also written as (Cheng et al. 2008):
r g,n
αn = ; (20.30)
r g,n + rd,n

where n is the chain length, r g,n is the chain growth rate and r d,n is the chain
termination rate. These reaction rates refer to Fig. 20.4:
Taking into consideration the chain growth probability, the Anderson-
–Schulz–Flory (ASF) distribution is given by:

Mn = M1 · α n−1 ; (20.31)

where M n refers to mole fraction and M1 is a constant. Making the assumption that
α is not dependent on hydrocarbon chain length, an equation may be derived in the
following equation (Tavakoli et al. 2008),
20.2 Chain Growth Probability (Products Distribution) for the FT Synthesis 297

Fig. 20.5 The α value and product molar fraction against the chain length. The dashed lines rep-
resent the ideal ASF distribution while the solid curves are the deviated ASF distribution values.
Adapted from Cheng et al. (2008)

 
Wn
log = n log α + const (20.32)
n

where W n is the weight fraction for the component with n carbons. Notice that in
the simplified ASF distribution the chain growth probability is constant, however,
in reality this value is a function of the chain length. The linearized version of the
distribution can provide us with a straight line where the chain growth probability
can be derived. Figure 20.5 demonstrated that the values of α and Mn indeed are
different from the ideal case.
The mass fraction for different components as a function of their chain growth
probability is reported in Fig. 20.6 as a qualitative estimation. It is possible to notice
that the chain growth probability is just the distribution function for each of these
compounds.
The main process parameters such as temperature, pressure, and H2 /CO ratio have
an effect on the product distribution for the different components, the olefin/paraffin
ratio, carbon deposition, and methane selectivity. These are positively or negatively
affected by the process parameters. This is demonstrated in Table 20.5.
More information on the Sasol process and how the FT synthesis is performed at
low temperatures is given in (Espinoza et al. 1999), while the commercial process
development is described in (Geerlings et al. 1999). On the other hand, the high-
temperature FT synthesis is described in (Steynberg et al. 1999). Information on the
selection and advances in reactor design for FT synthesis is given in (Sie and Krishna
1999) while more information on the reactions’ kinetics and product distributions
are given in (Patzlaff et al. 1999; Schulz and Claeys 1999). Some patented catalysts
used in FT synthesis are compared in bubble bed column reactors by (Oukaci et al.
1999).
298 20 Fischer–Tropsch (FT) Synthesis to Biofuels (BtL Process)

0.8

Mass fracƟon [-] 0.6

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1 1.2
Chain Growth probability

Methane Ethane Propane Butane Gasoline Diesel

Fig. 20.6 Distribution functions for diverse components and their mass fraction. Qualitative esti-
mation

Table 20.5 Effect of FT process parameters on the chain growth probability, olefin/paraffin ratio,
carbon deposition, and methane selectivity
Parameter Chain growth Olefin/paraffin Carbon Methane
probability ratio deposition selectivity
Temperature Down Down Up Up
increase
Pressure increase Up Complex Complex Down
H2 /CO ratio Down Down Down Up
increase
Conversion Complex Down Up Up
increase
Space velocity Complex Up Complex Down
increase
Adapted from De Deugd (2004)

The catalyst preparation is also a very important part of the development of this
technology, and some details are given in (Ernst et al. 1999). The FT synthesis not
necessarily should be performed by using CO and H2 only, but it can be performed
also by using CO2 and H2 syngas (Riedel et al. 1999); in this case, the catalysts are
made on the basis of Iron and Cobalt.

20.3 Syngas Treatment for FT Synthesis

In the FT synthesis, the used syngas is usually treated to obtain the necessary purity
and the minimal requirements to carry on the process. The wet-cold gas cleaning
proceeds according to the scheme presented in Fig. 20.7.
20.3 Syngas Treatment for FT Synthesis 299

Fig. 20.7 Wet-cold gas cleaning process. Adapted from Triantafyllidis et al. (2013)

The so-called acid scrubbing is also utilized. In this process, the main components
used are: HCl, H2 S, HF, HCN, HBr, NH3 , and the main reactions taking place for
each of these constituents are given as follows:

HCl + NaOH → NaCl + H2 O (20.33)

HCN + NaOH → NaCN + H2 O (20.34)

2NaOH + Cl2 → NaOCl + NaCl + H2 O (20.35)

NaCN + (2NaOH + Cl2 ) → NaCNO + 2NaCl + H2 O (20.36)

H2 S + 2NaOH → Na2 S + 2H2 O (20.37)

It should be noticed that the compound NaCNO is less toxic than NaCN and can
be removed easily.
When Halogen vapors (Cl2 , F2 , Br2 ) are involved, the following reactions are
taking place:

Cl2 + 2XOH → XCl + XOCl + H2 O (at < 20 ◦ C) (20.38)

3Cl2 + 6XOH → 5XCl + XClO3 + 3H2 O (at ∼


= 75 ◦ C) (20.39)

In the previous Eq. (20.38 and 20.39), the term X can be Na or K.


The scrubbing of NH3 is usually performed by utilizing sulfuric acid according
to:

2NH3 + H2 SO4 → (NH4 )2 SO4 (20.40)

while a CO2 Scrubber involves the following reactions:

2NaOH(aq) + CO2 (g) → Na2 CO3 (aq) + H2 O(l) (20.41)


300 20 Fischer–Tropsch (FT) Synthesis to Biofuels (BtL Process)

Table 20.6 Main reactions involved in a wet flue gas desulfurization (WFGD) process
Rate-determining steps Reactions
Absorption of gaseous SO2 in liquid water SO2 + H2 O  H+ + HSO−
3

HSO− + 2−
3  H + SO3
Oxidation of HSO−
3 (liquid phase) HSO− + 2−
3 + 2 O2  H + SO4
1

Solid limestone is dissolving in acidic HSO2− 2− +


4  SO4 + H
environment (pH 5.5, Industrial process)
CaCO3  Ca2+ + CO2−
3
CO2 + H2 O  HCO−
3 +H
+

HCO− 2−
3  CO3 + H
+

H2 O  H+ + OH−
Crystallization of gypsum Ca2+ + SO2−
4 + 2H2 O  CaSO4 · 2H2 O

Adapted from De Blasio et al. (2013)

Fig. 20.8 Dry-hot gas cleaning process for removing tar components from syngas

Na2 CO3 (aq) + Ca(OH)2 (s) → 2NaOH(aq) + CaCO3 (s) H ◦ = −5.3 kJ/mol
(20.42)

Concerning the separation of SO2 from the gaseous stream, wet scrubbing is usu-
ally applied. The main reactions involved in a wet flue gas desulfurization (WFGD)
process are given in Table 20.6.
While Table 20.7 gives the gas cleaning requirements for the FT synthesis on the
basis of diverse kinds of contaminants and cleaning methods.
The dry-hot gas cleaning process is described in Fig. 20.8; this process is com-
monly applied to tar and particulate removal from syngas produced, for example,
from gasification of biomass (Zwart et al. 2010).
Two well-known tar cracking catalysts are naturally occurring minerals: dolomite
and olivine (Soomro et al. 2018). The first method is with catalyst mixed with the
feed biomass in so-called catalytic gasification or pyrolysis (in situ). In this case, tar
is removed in the gasifier itself (usually in a fluidized bed gasifier). In the second
20.3 Syngas Treatment for FT Synthesis 301

Table 20.7 FT synthesis gas cleaning requirements for diverse contaminant and cleaning methods
Contaminant Poplar (wt%) FT cleaning Cleaning Wet-cold Dry-hot
requirements efficiency cleaning cleaning
(ppb) required (%)
Particulate 1.33 0 >99.9 Cyclone Granular bed
Bag filter Metallic
Scrubber filter
Ceramic
candle filter
Cyclone
HCN + NH3 0.47 20 >99.9 Scrubber ZnO, CuO
(H2SO4) guard bed
Chemical
and physical
acid removal
(rectisol,
Selexol,
MDEA)
H2 S + COS 0.01 10 >99.9 Scrubber Safeguard
ZnO, CuO filter
guard bed ZnO, CuO
Activated guard bed
charcoal Sorption bed
COS of α−Fe
hydrolyza-
tion
Alkali 0.1 10 >99.9 Condensation Adsorption
on particles and
made by chemisorp-
cooling tion
HCl 0.1 10 >99.9 Absorption Dry fly ash
by dolomite
(in tar
cracking)
Bag filter
(reaction
with
particulates)
Scrubber
(NaOH)
Tars – 0 >99.9 Scrubber Catalytic,
(organic thermal
oil-RME) reforming
Adapted from Triantafyllidis et al. (2013)
302 20 Fischer–Tropsch (FT) Synthesis to Biofuels (BtL Process)

Fig. 20.9 Bubbling fluidized bed (BFB) (left), circulating fluidized bed (CFB) (center), fixed bed
reactor (right). Adapted from Kunii and Levenspiel (1991)

method, tar is treated downstream of the gasifier in a secondary reactor, outside of


the gasifier (fixed bed catalytic reactor).
Figure 20.9 describes some of the available tar cracking solutions. Reactors are
mainly bubbling fluidized bed (BFB), circulating fluidized bed (CFB), and fixed bed
reactors.
These are configurations used as well to perform the gasification of biomass. It is
today always more common to use integrated gasification combined cycle (IGCC)
power plant and in these configurations, the air separation unit (ASU) is one of the
most important operations which improves the efficiency of the combined cycle itself
(Jones et al. 2011).
A simple representation of a IGCC plant is reported in Fig. 20.10.
Gasification of biomass can be coupled with processes aimed at the production
of biofuels via the FT synthesis. The integration can also involve different feedstock
kinds, for instance, black liquor, denominated as BL in Fig. 20.11. To be noticed
that, in the case gasification is applied, it is required to dry the fed biomass before
gasifying it. This is a very energy-demanding process, if we consider that water has
a quite large heat capacity and a latent heat of vaporization which is reported as
2260 kJ/kg or 40 kJ/mol (Datt 2011).
In the next section, some notions will be given on the gasification process with a
focus for the supercritical water gasification process.
References 303

Fig. 20.10 Integrated gasification combined cycle (IGCC) power plant. Adapted from Shadle and
Breault (2012)

Fig. 20.11 Integration of gasification with FT synthesis and production of power

References

Anderson, R. B. (1956a). Catalysis Vol IV: Hydrocarbon synthesis, hydrogenation and Cyclization.
New York, USA: Reinhold Publishing Corporation.
Anderson, R. B. (1956b). Catalysis Vol IV: Hydrocarbon synthesis, hydrogenation and cyclization.
New York, USA: Reinhold Publishing Corporation.
Atwood, H. E., & Bennett, C. O. (1979). Kinetics of the Fischer-Tropsch reaction over iron. Indus-
trial and Engineering Chemistry Product Research and Development, 18, 163–170.
304 20 Fischer–Tropsch (FT) Synthesis to Biofuels (BtL Process)

Bai, L., Xiang, H.-W., Li, Y.-W., Han, Y.-Z., & Zhong, B. (2002). Slurry phase Fischer–Trop-
sch synthesis over manganese-promoted iron ultrafine particle catalyst. Fuel, 7th China-Japan
Symposium on C1 Chemistry, 81, 1577–1581. https://doi.org/10.1016/S0016-2361(02)00089-3.
Borg, Ø., Eri, S., Blekkan, E. A., Storsæter, S., Wigum, H., Rytter, E., et al. (2007). Fischer-
Tropsch synthesis over γ-alumina-supported cobalt catalysts: Effect of support variables. Journal
of Catalysis, 248, 89–100. https://doi.org/10.1016/j.jcat.2007.03.008.
Brötz, W. (1949). Zur Systematik der Fischer-Tropsch catalyze. Zeitscrift für Elektrochemie, 53,
301.
Chanenchuk, C. A., Yates, I. C., & Satterfield, C. N. (1991). The Fischer-Tropsch synthesis with
a mechanical mixture of a cobalt catalyst and a copper-based water gas shift catalyst. Energy &
Fuels, 5, 847–855.
Cheng, J., Hu, P., Ellis, P., French, S., Kelly, G., & Lok, C. M. (2008). A DFT study of the chain
growth probability in Fischer-Tropsch synthesis. Journal of Catalysis, 257, 221–228. https://doi.
org/10.1016/j.jcat.2008.05.006.
Chin, Y., Hu, J., Cao, C., Gao, Y., & Wang, Y. (2005). Preparation of a novel structured catalyst
based on aligned carbon nanotube arrays for a microchannel Fischer-Tropsch synthesis reactor.
Catalyst Today, Catalytic Microstructured Reactors Catalytic Microstructured Reactors, 110,
47–52. https://doi.org/10.1016/j.cattod.2005.09.007.
Dalai, A. K., & Davis, B. H. (2008). Fischer-Tropsch synthesis: A review of water effects on the
performances of unsupported and supported Co catalysts. Applied Catalysis General, 348, 1–15.
https://doi.org/10.1016/j.apcata.2008.06.021.
Das, T. K., Jacobs, G., Patterson, P. M., Conner, W. A., Li, J., & Davis, B. H. (2003). Fischer-
Tropsch synthesis: characterization and catalytic properties of rhenium promoted cobalt alumina
catalysts✩. Fuel, 82, 805–815. https://doi.org/10.1016/S0016-2361(02)00361-7.
Datt, P. (2011). Latent heat of vaporization/condensation. In V. P. Singh, P. Singh & U. K. Haritashya
(Eds.), Encyclopedia of snow, ice and glaciers (pp. 703–703). Springer Netherlands, Dordrecht.
https://doi.org/10.1007/978-90-481-2642-2_327.
De Blasio, C., Carletti, C., Westerlund, T., & Järvinen, M. (2013). On modeling the dissolution of
sedimentary rocks in acidic environments. An overview of selected mathematical methods with
presentation of a case study. Journal of Mathematical Chemistry, 51(8), 2120–2143. https://doi.
org/10.1007/s10910-013-0202-3.
De Deugd, R. M. (2004). Fischer-Tropsch synthesis revisited; efficiency and selectivity benefits from
imposing temporal and/or spatial structure in the reactor. Ponsen & Looijen B.V, Wageningen,
The Netherlands.
de Klerk, A., & Furimsky, E. (2010). Catalysis in the refining of Fischer-Tropsch Syncrude. RSC
Catalysts Series 4. RCS Publishing.
Deckwer, W., Kokuun, R., Sanders, E., & Ledakowicz, S. (1986). Kinetic studies of Fischer-Tropsch
synthesis on suspended iron/potassium catalyst—Rate inhibition by carbon dioxide and water.
Industrial and Engineering Chemistry Process Design and Development, 25, 643–649.
Demitras, G. C., & Muetterties, E. L. (1977). Metal clusters in catalysis. A new fischer-tropsch
synthesis. Journal of the American Chemical Society, 99, 2796–2797.
Ernst, B., Libs, S., Chaumette, P., & Kiennemann, A. (1999). Preparation and characterization of
Fischer-Tropsch active Co/SiO2 catalysts. Applied Catalysis General, 186, 145–168. https://doi.
org/10.1016/S0926-860X(99)00170-2.
Espinoza, R. L., Steynberg, A. P., Jager, B., & Vosloo, A. C. (1999). Low temperature Fischer-
Tropsch synthesis from a Sasol perspective. Applied Catalysis General, 186, 13–26. https://doi.
org/10.1016/S0926-860X(99)00161-1.
Geerlings, J. J. C., Wilson, J. H., Kramer, G. J., Kuipers, H. P. C. E., Hoek, A., & Huisman, H. M.
(1999). Fischer-Tropsch technology—From active site to commercial process. Applied Catalysis
General, 186, 27–40. https://doi.org/10.1016/S0926-860X(99)00162-3.
Hao, Q., Bai, L., Xiang, H., & Li, Y. (2009). Activation pressure studies with an iron-based catalyst
for slurry Fischer-Tropsch synthesis. Journal of Natural Gas Chemistry, 18, 429–435. https://doi.
org/10.1016/S1003-9953(08)60134-6.
References 305

Huang, C., Xu, C., Wang, B., Hu, X., Li, J., Liu, J., et al. (2018). High production of syngas from
catalytic steam reforming of biomass glycerol in the presence of methane. Biomass and Bioenergy,
119, 173–178. https://doi.org/10.1016/j.biombioe.2018.05.006.
Jahangiri, H., Bennett, J., Mahjoubi, P., Wilson, K., & Gu, S. (2014). A review of advanced cata-
lyst development for Fischer-Tropsch synthesis of hydrocarbons from biomass derived syn-gas.
Catalysis Science & Technology, 4, 2210–2229. https://doi.org/10.1039/C4CY00327F.
Ji, Y.-Y., Xiang, H.-W., Yang, J.-L., Xu, Y.-Y., Li, Y.-W., & Zhong, B. (2001). Effect of reaction
conditions on the product distribution during Fischer-Tropsch synthesis over an industrial Fe-Mn
catalyst. Applied Catalysis General, 214, 77–86. https://doi.org/10.1016/S0926-860X(01)00480-
X.
Jones, D., Bhattacharyya, D., Turton, R., & Zitney, S. E. (2011). Optimal design and integration of
an air separation unit (ASU) for an integrated gasification combined cycle (IGCC) power plant
with CO2 capture. Fuel Processing Technology, 92, 1685–1695. https://doi.org/10.1016/j.fuproc.
2011.04.018.
Kunii, D., & Levenspiel, O. (1991). Fluidization engineering (2nd ed.). Elsevier.
Liu, Y., Teng, B.-T., Guo, X.-H., Li, Y., Chang, J., Tian, L., et al. (2007). Effect of reaction condi-
tions on the catalytic performance of Fe-Mn catalyst for Fischer-Tropsch synthesis. Journal of
Molecular Catalysis Chemical, 272, 182–190. https://doi.org/10.1016/j.molcata.2007.03.046.
Ma, T., Imai, H., Shige, T., Sugio, T., & Li, X. (2015). Synthesis of hydrocarbons from H2 -deficient
syngas in fischer-tropsch synthesis over co-based catalyst coupled with Fe-based catalyst as
water-gas shift reaction. Journal Nanomaterial. [WWW Document].. https://doi.org/10.1155/
2015/268121.
Oukaci, R., Singleton, A. H., & Goodwin, J. G., Jr. (1999). Comparison of patented Co F-T catalysts
using fixed-bed and slurry bubble column reactors. Applied Catalysis General, 186, 129–144.
https://doi.org/10.1016/S0926-860X(99)00169-6.
Outi, A., Rautavuoma, I., & van der Baan, H. S. (1981). Kinetics and mechanism of the fischer
tropsch hydrocarbon synthesis on a cobalt on alumina catalyst. Applied Catalysis, 1, 247–272.
Özdenkçi, K., De Blasio, C., Muddassar, H. R., Melin, K., Oinas, P., Koskinen, J., et al. (2017).
A novel biorefinery integration concept for lignocellulosic biomass. Energy Conversion and
Management, 149, 974–987. https://doi.org/10.1016/j.enconman.2017.04.034.
Pannell, R. B., Kibby, C. L., Kobylinski, T. P. (1981). A steady-state study of Fischer-Tropsch
product distributions over cobalt, iron and ruthenium. Studies in Surface Science and Catalysis,
7(A), 447–459.
Patzlaff, J., Liu, Y., Graffmann, C., & Gaube, J. (1999). Studies on product distributions of iron and
cobalt catalyzed Fischer-Tropsch synthesis. Applied Catalysis General, 186, 109–119. https://
doi.org/10.1016/S0926-860X(99)00167-2.
Pour, A. N., Zamani, Y., Tavasoli, A., Kamali Shahri, S. M., & Taheri, S. A. (2008). Study on
products distribution of iron and iron–zeolite catalysts in Fischer-Tropsch synthesis. Fuel, 87,
2004–2012. https://doi.org/10.1016/j.fuel.2007.10.014.
Riedel, T., Claeys, M., Schulz, H., Schaub, G., Nam, S.-S., Jun, K.-W., et al. (1999). Comparative
study of Fischer-Tropsch synthesis with H2 /CO and H2 /CO2 syngas using Fe- and Co-based cata-
lysts. Applied Catalysis General, 186, 201–213. https://doi.org/10.1016/S0926-860X(99)00173-
8.
Sari, A., Zamani, Y., & Taheri, S. A. (2009). Intrinsic kinetics of Fischer-Tropsch reactions over
an industrial Co–Ru/γ-Al2O3 catalyst in slurry phase reactor. Fuel Processing Technology, 90,
1305–1313. https://doi.org/10.1016/j.fuproc.2009.06.024.
Sarup, B., & Wojciechowski, B. W. (1989). Studies of the fischer-tropsch synthesis on a cobalt
catalyst II. Kinetics of carbon monoxide conversion to methane and to higher hydrocarbons.
Canadian Journal of Chemical Engineering, 67, 62–74.
Schulz, H. (1999). Short history and present trends of Fischer-Tropsch synthesis. Applied Catalysis
General, 186, 3–12. https://doi.org/10.1016/S0926-860X(99)00160-X.
Schulz, H., & Claeys, M. (1999). Kinetic modelling of Fischer-Tropsch product distributions.
Applied Catalysis General, 186, 91–107. https://doi.org/10.1016/S0926-860X(99)00166-0.
306 20 Fischer–Tropsch (FT) Synthesis to Biofuels (BtL Process)

Shadle, L. J., & Breault, R. W. (2012). Integrated gasification combined cycle (IGCC). In W.-
Y. Chen, J. Seiner, T. Suzuki & M. Lackner (Eds.), Handbook of climate change mitigation
(pp. 1545–1604). Springer US, New York, NY. https://doi.org/10.1007/978-1-4419-7991-9_40.
Sie, S. T., & Krishna, R. (1999). Fundamentals and selection of advanced Fischer-Tropsch reactors.
Applied Catalysis General, 186, 55–70. https://doi.org/10.1016/S0926-860X(99)00164-7.
Soomro, A., Chen, S., Ma, S., & Xiang, W. (2018). Catalytic activities of nickel, dolomite, and
olivine for tar removal and H2 -enriched gas production in biomass gasification process, Catalytic
activities of nickel, dolomite, and olivine for tar removal and H2 -enriched gas production in
biomass gasification process. Energy Environment 0958305X18767848. https://doi.org/10.1177/
0958305X18767848.
Steynberg, A. P., Dry, M. E., Davis, B. H., & Breman, B. B. (2004). Fischer-Tropsch reactors,
Chap. 2. In A. Steynberg & M. Dry (Eds.), Studies in surface science and catalysis, Fischer-
Tropsch Technology (pp. 64–195). Elsevier. https://doi.org/10.1016/S0167-2991(04)80459-2.
Steynberg, A. P., Espinoza, R. L., Jager, B., & Vosloo, A. C. (1999). High temperature Fischer-
Tropsch synthesis in commercial practice. Applied Catalysis General, 186, 41–54. https://doi.
org/10.1016/S0926-860X(99)00163-5.
Tavakoli, A., Sohrabi, M., & Kargari, A. (2008). Application of Anderson–Schulz–Flory (ASF)
equation in the product distribution of slurry phase FT synthesis with nanosized iron catalysts.
Chemical Engineering Journal, 136, 358–363. https://doi.org/10.1016/j.cej.2007.04.017.
Tavasoli, A., Pour, A. N., & Ahangari, M. G. (2010). Kinetics and product distribution studies on
ruthenium-promoted cobalt/alumina Fischer-Tropsch synthesis catalyst. Journal of Natural Gas
Chemistry, 19, 653–659. https://doi.org/10.1016/S1003-9953(09)60133-X.
Triantafyllidis, K., Lappas, A., & Stöcker, M. (2013). The role of catalysis for the sustainable
production of bio-fuels and bio-chemicals. Elsevier.
Wojciechowski, B. W. (1988). The kinetics of Fischer-Tropsch synthesis. Catalysis Reviews–Science
and Engineering, 30, 629.
Xu, D., Li, W., Duan, H., Ge, Q., & Xu, H. (2005). Reaction performance and characterization
of Co/Al2O3 Fischer-Tropsch catalysts promoted with Pt, Pd and Ru. Catalysis Letters, 102,
229–235. https://doi.org/10.1007/s10562-005-5861-7.
Yang, C., Massoth, F. E., Oblad, A. G. (1979). In E. L. Kugler & F. W. Steffgen (Eds.), Hydrocarbon
synthesis from carbon monoxide and hydrogen. Washington, D.C.: ACS.
Yates, I. C., & Satterfield, C. N. (1991). Intrinsic kinetics of the Fischer-Tropsch synthesis on a
cobalt catalyst. Energy & Fuels, 5, 168–173. https://doi.org/10.1021/ef00025a029.
Zhang, C.-H., Yang, Y., Teng, B.-T., Li, T.-Z., Zheng, H.-Y., Xiang, H.-W., et al. (2006). Study of an
iron-manganese Fischer-Tropsch synthesis catalyst promoted with copper. Journal of Catalysis,
237, 405–415. https://doi.org/10.1016/j.jcat.2005.11.004.
Zwart, R., Van der Heijden, S., Emmen, R., Dall Bentzen, J., Ahrenfeldt, J., Stoholm, P., et al.
(2010). Tar removal from low-temperature gasifiers. Risø DTU, DFBT and Anhydro, Denmark:
ECN and Dahlman from the Netherlands and Dall Energy.
Chapter 21
Notions of Biomass Gasification

If I don’t do something really new, I get bored


Susumu Tonegawa, Nobel Prize in Medicine in 1087.

21.1 Preliminary Notions

Some information on the gasification of biomass process was given previously in


the second chapter of this part of the book, and here, only a preliminary introduction
will be provided about the general process with more focus on one particular kind
of gasification which considers water as the medium at supercritical conditions.
Gasification of biomass can be considered to follow different stages. The first step
would be the drying of the feedstock followed by pyrolysis. The pyrolysis can be then
followed by oxidation and reduction, depending on where air or oxygen is utilized
during this process. In pyrolysis, which is reported to take place in a wide range of
temperatures between 200 and 600 °C (Zhang et al. 2015), the biomass feedstock
is having reactions of decomposition and as the last result, the char is then left as
the part not containing volatiles. The formed char is then reacting within the section
where we have higher temperatures (Fig. 21.1).
Reactions taking place during gasification of biomass can be exothermic and
endothermic. The reactions written below are mainly taken from (De Blasio et al.
2015; De Blasio and Järvinen 2017; Özdenkçi et al. 2017). The exothermic reactions
are written as follows.
Combustion:

CH3 −CH2 − · · · + O2 → CO2 (21.1)

Partial oxidation:

CH3 −CH2 − · · · + O2 → CO (21.2)

© Springer Nature Switzerland AG 2019 307


C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_21
308 21 Notions of Biomass Gasification

Fig. 21.1 Stages of the


gasification process. Adapted
from Chen et al. (2018)

Methanation:

CH3 −CH2 − · · · + H2 → CH4 (21.3)

where with CH3 –CH2 –… it is referred to the general biomass and char. While the
endothermic reactions are written as:
Steam–carbon reaction:

CH3 −CH2 − · · · + H2 O → CO + H2 (21.4)

Boudouard reaction:

CH3 −CH2 − · · · + CO2 → 2CO (21.5)

Particular examples and additional reactions are given here with data on their
standard enthalpy of reaction. For the water-gas shift (WGS) and CO methanation
reactions, we have (Stangeland et al. 2017):

CO(g) + H2 O(g) ↔ CO2(g) + H2(g)  H298.15K


0
= −41.2 kJ/mol (21.6)

CO(g) + 3H2(g) ↔ CH4(g) + H2 O(g) H298.15K


0
= −206.1 kJ/mol (21.7)
21.1 Preliminary Notions 309

The CO2 methanation reaction is reported as (Stangeland et al. 2017):

CO2 + 4H2 → CH4 + 2H2 O (H = −165 kJ/mol) (21.8)

For steam reforming of acetic acid:

C2 H4 O2 (g) → 2CO(g) + 2H2 (g) H298.15K


0
= 221.2 kJ/mol (21.9)

The decarboxylation of formic acid and acetic acid is instead written as:

HCOOH(g) → CO2(g) + H2(g) H298.15K


0
= −14.9 kJ/mol (21.10)

CH3 COOH(g) → CH4(g) + CO2(g) H298.15K


0
= −36.1 kJ/mol (21.11)

In addition, the following reaction can also occur:

C(s) + 2H2(g) ↔ CH4(g) H298.15K


0
= −74.8 kJ/mol; (21.12)

and the coke gasification reactions:

CO(g) + H2(g) ↔ C(s) + H2 O(g) H298.15K


0
= −131.3 kJ/mol; (21.13)

CO2(g) + 2H2(g) ↔ C(s) + 2H2 O(g) H298.15K


0
= −90.1 kJ/mol. (21.14)

The most common types of gasifiers are generally classified as (Sikarwar et al.
2016)
– fluidized;
– entrained flow gasifiers.
And, on the other hand, they also are subject to further classification. In fact, the
fluidized kind can be divided into:
– bubbling fluidized bed;
– circulating fluidized bed;
– dual fluidized bed.
The entrained flow gasifiers can be classified as:
– top-fed gasifiers;
– side-fed gasifiers.
Regarding the nomenclature of these kinds of reactors, their names are quite
descriptive of their functioning. As a matter of fact, the fluidized kind of reactor has
the so-called fluidized bed, and, depending on the particle size distribution of the
bed and the feeding flow (gas), they behave differently. In this section, only some
information will be given regarding the basic principles of these kinds of reactors,
310 21 Notions of Biomass Gasification

and for a more detailed treatise, the reader can refer to the available literature (Kunii
and Levenspiel 1991). The main difference between the reactors was depicted in
Fig. 10.9; the circulating fluidized bed reactor differs from the others, since there is a
“recirculation” of the solid particles. This makes it different in comparison with the
bubbling kind since in the last case the solid particle are less fluidized but they are
staying in place within the same area of the reactor while this is not happening in the
case when a recirculation is taking place by means of a cyclone or suitable devices.
These reactors have a series of advantages and disadvantages. For instance, the
fixed bed reactors are commonly referred to be more simple and also more economic.
It is also referred that the tar, which could have been formed during the gasification
process, is completely degraded in this kind of reactors, and this is because we have
very high temperatures (around or more than 1000 °C), (Sikarwar et al. 2016). This
is an important feature of these types of gasifier, but it also gives more costs. On the
other hand, the circulating fluidized bed reactors and the dual fluidized bed ones have
a more complicated design, and therefore, the costs are also higher. Nevertheless, it
is also clear that the heat transfer mechanisms and the temperature gradients within
the reactors are much more favorable.
Without going into detail with the physical principles behind the utilization of
a particular bed, the inert mass is needed to influence the heat capacity inside our
system (the inner volume of the reactor). For this reason, the temperature gradients
between the reacting particles and the spaces where there is no reaction will be
significantly decreased with beneficial outcomes.
An additional factor to be taken into account is the mixing of the particles; fluidiza-
tion increases the degree of mixing of the particles and increases the reactions’ effi-
ciency. Nevertheless, the mixing and the transport phenomena are indeed increased
when the particles are smaller and, on the other hand, a recirculation is then necessary.
Some examples of commercially available and operating gasifiers are as follows,
and they will be illustrated with their basic principles of functioning (Breault 2010):
– The GE Energy Gasifier;
– ConocoPhillips E-Gas;
– Shell;
– Siemens;
– KBR Transport;
– British Gas Lurgi;
– Lurgi Multi-Purpose Gasifier;
– Lurgi Mark IV Gasifier;
– U-Gas, Synthesis Energy Systems;
– High Temperature Winkler Gasifier;
– PRENFLO™ Gasifier/Boiler.
The GE gasifier is a type of reactor produced by the company General Electric
and can be utilized for coal gasification. As a representative illustration, Fig. 21.2
gives some information about the process.
The process involves a cocurrent flow of the feedstock (a coal slurry in this case)
and the oxygen which is coming from an air separation plant unit. As gasifying
21.1 Preliminary Notions 311

Fig. 21.2 A simple schematics of the GE Energy Gasifier. Adapted from De Agarwal et al. (2018)

agent, we have steam in this case, steam has a large importance since it functions as
transport agent, it helps with the heat transfer-related phenomena, and it is involved
in the gasification reactions which were previously described (water-gas shift, steam
reforming and related). The steam participates actively within the heterogeneous
reactions between steam and carbon. The feeding system, in this case, is on the top
of the reactor, while at the bottom, the low-temperature pyrolysis takes place and the
syngas is released for its use and collection.
The ConocoPhillips process is depicted in Fig. 21.3, and in this case, we have
some small difference in comparison with the previous described process. The slurry
of feedstock and the oxygen are inserted at the bottom of the reactor where we have
the first stage of the process. Eventually, char from other processes or recovered
from the same system is inserted also within the first stage, the high-temperature
gasification process.
The Shell gasifier is illustrated in Fig. 21.4. This gasifier can be also classified as
a “Water wall” gasifier, and this is for the fact that at the wall, there is a series of
water tubes as it is represented in Fig. 21.4.
A water-walled reactor offers additional possibilities in regard to the maintenance
of the reactor and also concerning the control of the operating conditions. As a matter
of fact, by having this kind of arrangement, it is possible to control in a better way
the inner temperature of the reactor. The durability is also increased, since there is
312 21 Notions of Biomass Gasification

Fig. 21.3 Drawing of the ConocoPhillips E-gas gasifier

Fig. 21.4 Shell gasifier with a description of the constituting parts of the reactor wall. Adapted
from Lisandy et al. (2016)
21.1 Preliminary Notions 313

less temperature gradients related to the refractory material. However, also this kind
of arrangements is subject to failure.
Siemens has been also active within the development of the gasification technol-
ogy, and in Fig. 21.5 there is a schematics of the system adopted in this case.
In reference to Fig. 21.5, the feedstock and the gas are inserted from the same side
and the first stage of the process is the development of a high-temperature section; in
this case, we have indeed high temperature, in the range of 1300–1800 °C, (Görsch
et al. 2011).
It should be noted that, in gasification, there is a difference if the water is injected
in one section or another within the reactor. In this case, the water is injected at the
beginning of the second section, and we inject here the quench water.
One of the main reactions driven by the injected water is the water-gas shift
reaction, WGSR, for which a simple schematics is reported in Fig. 21.6.
This indeed has a contribution especially in the case the ultimate goal is to a have
a syngas rich in hydrogen.
The KBR Transport Gasifier (TRIGTM ) operating principles are similar to the
circulating fluidized bed combustion systems. The steam, feedstock (for instance
coal), air, and sorbent are mixed into the mixing zone, while oxygen is inserted
at the bottom of the main part of the gasifier. The products are elevated up into a

Fig. 21.5 Siemens


gasification reactor with
quench water system
314 21 Notions of Biomass Gasification

Fig. 21.6 Schematics of the quench process and its contribution to the WGSR

Fig. 21.7 Representation of the KBR Transport Gasifier (TRIGTM ). Adapted from Ariyapadi et al.
(2008)

conduct called riser, and after this, the bigger particles are firstly removed into a
disengage unit and after this, into a cyclone. The particles collected are sent back to
the gasification zone by going first into a stand pipe and then a J-leg. The produced
gases are processed further with separation and cleaning processes (Fig. 21.7).
The British Gas Lurgi Gasifier is described in a simple way in Fig. 21.8.
This case differs from the others since we have here a different direction of the
flow for the feedstock and for the injected steam/air/oxygen. It is also beneficial to
remember that the inlet of the reactor contains still the bed material and in this case
the walls are constituted by refractory material.
The company Lurgi AG has also developed a Multi-Purpose Gasifier, MPG, for
which a simple schematics is reported here (Fig. 21.9).
The simple arrangements of this gasifier allow for a certain flexibility on the kinds
of feedstock used. The claim in this case is that the flash point of the feedstock is
21.1 Preliminary Notions 315

Fig. 21.8 Description of a


British Gas Lurgi Gasifier.
Adapted from
Krishnamoorthy and Pisupati
(2015)

Fig. 21.9 Lurgi MPG.


Adapted from Koss and
Schlichting (2005)
316 21 Notions of Biomass Gasification

not limited to a certain range, the particles should be around one millimeter, and the
system allows for using diverse kinds of fuels which could be inserted at the same
time into the reactor even if they are not miscible. The Lurgi MPG gives the possibility
of operating within two modes, the boiler mode and the quench mode; in both cases,
different kinds of feedstock can be used, but there are some limitations for the case of
the boiler mode. In the boiler mode, we have high pressure for the steam and higher
costs, while for the quench configuration we have medium pressure steam and lower
costs in case we would like to focus on the production of hydrogen. The raw gas
produced needs still to have a treatment process; for instance, a desulfurization unit
is then necessary.
Mitsubishi Heavy Industries has also developed gasification systems. For instance,
a particular system suitable for municipal solid waste, MSW, is shown in (Fig. 21.10).
In this case, we have a fluidized bed of sand where the combustion air is injected.
The gasification is taking place at relatively low temperature and between 450 and
600 °C. The system allows for the separation of heavy metals which can be recycled.
The company Synthesis Energy Systems has developed a method denominated as
U-Gas® which is a gasifier with a dry feed injection. A simple schematics is shown
in (Fig. 21.11).
Regarding the positive aspects of this technology, the claim in this case is that
there is good mixing and the conversion is also high since the residence time is main-
tained at sufficient levels. The moderate temperatures allow for a longer durability.

Fig. 21.10 A system developed by Mitsubishi Heavy Industries for gasification of MSW. Adapted
from Mitsubishi Heavy Industries (2019)
21.1 Preliminary Notions 317

Fig. 21.11 U-Gas method developed by synthesis energy systems. Adapted from Lau (2009)

Table 21.1 Fuel properties Fuel parameter range Tested range


range for the U-Gas gasifier
M (wt%) 1–41
VM (wt%) 3–69
FC (wt%) 6–83
S (wt%) 0.2–4.6
Free swelling index 0–8
Ash (wt%) <42%
Ash softening, °F 1900–2500
HHV (Btu/lb) 5500–14,000
where M is the moisture, VM is the volatile matter, FC is the fixed
carbon, and S is the sulfur
Adapted from Lau (2009)

The range of properties for the feedstock and the tested range values are given in
Table 21.1.
The High Temperature Winkler Gasifier, HTW, is working following the principles
of the Winkler Generator which was developed in the beginning of last century. The
difference between the old and the new method is that the new version is pressurized.
This method was developed (the new version) by ThyssenKrupp and Rheinische
Braunkohlenwerke AG (Toporov and Abraham 2015) (Fig. 21.12).
The system allows for the processing of diverse kinds of feedstock, and the higher
pressure (10 bar) leads to better reaction rates. On the other hand, higher temperature
318 21 Notions of Biomass Gasification

Fig. 21.12 A scheme for the High Temperature Winkler Gasifier. Adapted from Toporov and
Abraham (2015)

increases the gas yields and rates; however, in this case, the temperatures are still
quite moderate with respect to other gasifiers and this gives more durability to this
system. In Finland, there is one gasifier of this kind in Oulu.
The HTW and the PRENFLO® gasifiers are both described within a report pub-
lished by ThyssenKrupp and available to the public (ThyssenKrupp 2019). A flow
diagram of the process is reported in Fig. 21.13.
This system uses a particle size range of 0.1 mm, and the gasification media are
oxygen and steam. We have here a temperature which is higher than the ash melting
temperature, and this allows for the removal of the ashes as slag. In the mentioned
reference, ThyssenKrupp refers that the gas produced is composed mainly of CO and
H2 gases. These gases are then sent to a boiler where steam is generated to produce
electricity. The gas is further processed to eliminate dust and impurities.
As can be noticed, the possible solutions for producing syngas from a variety of
feedstock are many. In this first edition of the manuscript, the author will limit the
treatise to one type of gasification: the supercritical water gasification.
21.2 Supercritical Water Gasification 319

Fig. 21.13 PRENFLO® Gasifier/Boiler. Adapted from ThyssenKrupp (2019)

21.2 Supercritical Water Gasification

Supercritical water, SCW, is used today already in a number of processes. SCW is


used present within normal and nuclear power plants; it is normally used to drive
turbines in order to generate electricity. In this case, they are nominally called super-
critical water reactors (Cervi and Cammi 2018). SCW is also used for other applica-
tions; for instance, nanoparticles and new compounds are produced by these meth-
ods. When oxygen is used together with SCW, this allows for the destruction of
hazardous waste by supercritical water oxydation, SCWO. SCW allows for a wide
range of selection of the process conditions since parameters like temperature affect
all of its properties: density, dielectric constant, specific heat capacity, diffusivity,
viscosity, and more.
SCWG is quite effective at high temperatures where a very good conversion can
be recorded; here, a visual example is given for the SCWG of weak black liquor at
700 °C. In the figure below, the sample before gasification is shown along with the
produced liquid effluent (Fig. 21.14).
Water has indeed remarkable properties; for instance, its density can be actually
reduced by decreasing its temperature, like in the case when ice is formed. Ice has a
lesser density compared to the one of liquid water, and this allows for the ice to be
formed on top of rivers or the sea. This is quite an important feature since it allows
for the organisms in water to survive even if the weather is very cold and cold for
320 21 Notions of Biomass Gasification

Fig. 21.14 Visual example


for SCWG of black liquor
and liquid products.
Photograph by Cataldo De
Blasio

a very long time. Water has been said to be the most useful among the engineering
fluids in the literature (He et al. 2015) and not only at low temperatures, but also
at high temperatures and pressures, water demonstrates interesting properties. For
instance, water is referred to have an unusual high temperature for its critical point
(Finney 2004) and this is probably because of the strong molecular bonds of water
and specifically between hydrogen and oxygen in the molecule.
At supercritical conditions, i.e., temperatures and pressures above 374 °C and
218 atm, the thermodynamic properties of water are changing drastically starting
from its density which is dropping considerably at these conditions. This is shown in
Fig. 21.15 where data are taken from the International Association for the Properties
of Water and Steam IAPWS Formulation 1995 for the Thermodynamic Properties of
Ordinary Water Substance for General and Scientific Use (Wagner and Kretzschmar
2008).
The properties of water are changing because the structure of the water itself is
changing when supercritical conditions are reached. Computational molecular mod-
eling, CMM, is in use today to explain the structure of water at these conditions. This
is done mainly because it is difficult to have precise information from experimental
methods. Nevertheless, considering experimental and theoretical studies, some addi-
tional information is given on what could be the structure of water at supercritical
conditions. It is mentioned that the percolating network of hydrogen bonds in this
system is actually broken and there is still parts of intact clusters of hydrogen atoms
bonded together; this could be surrounded by water molecules which are starting to
loose their bonds (Kalinichev 2017).
The variation of its density is quite an interesting feature when we consider the
production of electricity by expanding it in a turbine generator. However, the con-
ditions for this kind of application should be carefully planned. As a matter of fact,
it is indeed a fact that the density is decreasing and therefore it could give an extra
expansion within the turbine, but the changes in other properties have to be taken
into account as well. For instance, water is referred to be quite corrosive especially
21.2 Supercritical Water Gasification 321

Fig. 21.15 Density of water as a function of temperature at 250 bar. Data from Wagner and Kret-
zschmar (2008)

Fig. 21.16 Isothermal variation of ionic product for water. Left, 374 °C. Right: 650 °C. Data from
Wagner and Kretzschmar (2008)

in the near supercritical region and this could damage metals. The change in the heat
capacity in the near supercritical region should be also considered very well.
The change in these parameters is considered more specifically in the follow-
ing pages starting with the ionic product of water and its dissociation constant
(Fig. 21.16).
The viscosity of supercritical water is reported usually to be low (Yang et al.
2019). In Fig. 21.17, data for the viscosity against the density of water are reported
at a temperature of 386 °C.
In Fig. 21.18, the specific heat capacity against temperature is reported at two
different pressures, 24 and 36 MPa. From the figure, it can be observed that the
pressure does a certain “containment” effect on the heat capacity variation.
Instead, the variation of the density against the pressure is demonstrated in
Fig. 21.19 where two temperatures are considered, 410 and 374 °C.
The following plot demonstrates the energy required to bring water at 700 °C at
a pressure of 30 MPa (Kruse 2008). On the other hand, to bring water to the same
322 21 Notions of Biomass Gasification

Fig. 21.17 Density versus viscosity of water at 386 °C. Data from Wagner and Kretzschmar (2008)

Fig. 21.18 Specific heat of SCW against temperature at 24 MPa, Left, and 36 MPa, Right

Fig. 21.19 Density of SCW plotted against pressure at 410 °C, left, and 374 °C, right

temperature, but at atmospheric pressure, it will require that we first bring the water
at the boiling temperature and then wait for all the water to pass to the steam phase.
This is because the boiling is taking place at constant temperature. After all the water
has gone to the steam phase, then it will be possible to heat the medium further to
superheated steam. This is shown in the same reference to have a higher consumption
with respect of doing this at 300 bar. This is one reason why in our laboratory we
choose to preheat the SCWG reactor at a sufficient pressure, for instance, 250 bar
(Fig. 21.20).
21.2 Supercritical Water Gasification 323

Fig. 21.20 Enthalpy change against temperature at 30 MPa. Adapted from Kruse (2008)

21.2.1 SCWG Kinetics and Thermodynamics

In the literature, it is reported that organic reactions with water can follow mainly
two mechanisms depending if the water is at its normal conditions or at supercritical
conditions. In fact, it has also been mentioned that the reactions involved in SCWG
follow a radical mechanism of reaction.
In the case of ionic species, we can list the following phenomena, and some of
them are reported to happen also in SCW:
• autoprotolysis (Farajtabar and Gharib 2009);
• protonation (Ryan et al. 1996);
• deprotonation;
• dehydration (Antal et al. 1987);
• tautomerization (Martskainen 2009);
• acetal formation or acetalization (Hoffmann and Conradi 1998);
• aldol condensation (Sasaki et al. 2002).
When there are radicals, we can list the following series of reactions:
• ionization (Portela et al. 2001);
• C–C bond beta scission (Ratkiewicz and Truong 2012);
• isomerization (for instance cis-trans) (Randolph and Carlier 1992);
• addition of radical molecules;
• dehydration (Yamamoto et al. 2018);
• substitution;
• termination.
324 21 Notions of Biomass Gasification

And the following reactions related to the water molecule could also be taken into
consideration:

H2 O  OH + H (21.15)

H2 O + H  OH + H2 (21.16)

H2 O + O  OH + OH (21.17)

OH + HO2  H2 O + R (21.18)

H2 O + HO2  OH + H2 O2 (21.19)

When simulating reaction conditions for SCWG of biomass, it is beneficial to


have suitable relation for the related equations of state EoS. One example is the
Peng–Robinson EoS which in this case is applied to a pure component and not to a
mixture at supercritical conditions:

RT a
P= − ; (21.20)
V −b 2
V + 2bV − b2

where V is the molar volume of the medium, and it is related to the compressibility
factor, which is defined as:

PV
Z= (21.21)
RT
where R is the universal gas constant, P is the pressure, and T is the temperature of
the medium. In the Peng–Robinson formula, the coefficients a and b are derived from
the critical temperature and pressure and also from the acentric factor. Expressions
for these terms can be found in the literature (Castello and Fiori 2011), and some
examples of a, b, and acentric factor are given as follows:
   2
0.45724R 2 Tc2   T
a= 1 + 0.37464 + 1.54226ω − 0.26992ω2 1 − ;
Pc Tc
(21.22)
0.07780RTc
b= ; (21.23)
Pc

where the subscript c is related to the critical state and ω is the Pitzer acentric factor
defined as:

P vap (Tr = 0.7)


ω = −1 − log10 (21.24)
Pc
21.2 Supercritical Water Gasification 325

In the previous formula, the term Tr is referring to the reduced temperature Tr =


T /Tc .
The van der Waals’ binary mixing rules are also used when there is a mixture
of components in the supercritical medium. With these relations, it is possible to
evaluate average coefficients for the mixture.

N N
am = xi x j ai j (21.25)
i=1 J =1
N N
bm = xi x j bi j (21.26)
i=1 J =1

In the equations written previously, the term x refers to the molar fraction of the
i and j components. The terms ai j and bi j can be estimated from:

ai j = aii a j j (21.27)
 
bii + b j j
bi j = (21.28)
2
where aii , a j j , bii , and b j j , are the coefficients described in Eqs. (21.22) and (21.23)
since aii , for instance, is the coefficient a related to one element (interaction between
the element i and itself) and the same reasoning can be done for the element j.
In real conditions, the partial pressure Pi of one component is expressed in terms of
the fraction of the component, the overall pressure of the system, and a proportionality
coefficient called fugacity.

f i = P xi ϕi (21.29)

The fugacity term for one component is related to its compressibility factor, and
here, it will be shown one example of correlation from the literature (Zhou and Zhou
2001).

  A∗ B∗
ln f i = Z − 1 − ln Z − B ∗ − ∗ ln 1 + (21.30)
B Z

This is also referred as the Redlich–Kwong (SRK) equation where:

P/Pc
A∗ = 0.4278 · α ·  2 ; (21.31)
T
Tc
P/Pc
B ∗ = 0.0867 · (21.32)
T /Tc
 
α 0.5 = 1 + m 1 − Tr0.5 ; (21.33)
326 21 Notions of Biomass Gasification

Fig. 21.21 General procedure for evaluating the composition of a mixture at equilibrium

m = 0.480 + 1.574ω − 0.176ω2 ; (21.34)

As a summary, in order to obtain the composition of a mixture at equilibrium, one


procedure could be by following the steps shown in Fig. 21.21.

21.2.2 Catalysis in SCWG

In SCWG of biomass, different kinds of catalysts are utilized to promote reactions


and improve the gasification rate and yield. Also in the case of SCWG, we can have
homogeneous and heterogeneous catalysis. Among the heterogeneous catalyst, we
have:
• nickel (Wang et al. 2018)—this catalyst can be supported also on zirconia
(Ni/ZrO2 ). This kind of catalyst is referred in the literature to have beneficial
effects for the methanation of CO and CO2 (da Silva et al. 2012, p. 2);
• ruthenium (Ru/Al2 O3 ), (Ru/ZrO2 ) (Byrd et al. 2008);
• palladium (Osada et al. 2006);
• Ru/TiO2 catalyst (Osada et al. 2012);
21.2 Supercritical Water Gasification 327

• iron;
• copper (Gizir et al. 2003);
• zinc (Resende and Savage 2010);
• zirconium (Watanabe et al. 2001);
• Raney nickel (Ni–Al Alloy) (Nguyen et al. 2014);
• platinum (Pt/Al2 O3 ) (Karakuş et al. 2013).
Among the homogeneous catalysts, we have:
• alkali hydroxides (Nanda et al. 2016);
• carbonates (Na2 CO3 ) (Nanda et al. 2016);
• KOH and KHCO3 (Yakaboylu et al. 2015).
Natural minerals:
• NaHCO3 · Na2 CO3 · 2H2 O also known as Trona (Yanik et al. 2008);
• K2 CO3 (Louw et al. 2016);
• Ca(OH)2 (Wang and Takarada 2001).
It has been reported that even black liquor itself can be considered as a catalyst
(Rönnlund et al. 2011), and this is because of its high content of catalyst salts.

21.2.3 Salts and Inorganics in SCWG of Biomass

Salts and inorganic compounds play a role as catalysts, as previously mentioned, and
by SCWG, they could be even recovered, since as it will be shown in the following
pages, the dielectric constant of water is fast dropping at supercritical conditions and
this provoke that salts and inorganics are precipitating within the reactor.
Among the main salts to be considered in SCWG of organic fractions, the sodium
and potassium salts play an important role when feedstock like black liquor is pro-
cessed. With respect to the sodium salts, we can recognize:
• sodium acetate—CH3 COONa;
• sodium formate—HCOONa;
• sodium lactate—C3 H5 NaO3 ;
• sodium oxalate—Na2 C2 O4 ;
• sodium carbonate—Na2 CO3 ;
• sodium bicarbonate—NaHCO3 .
Data on the sodium formate and sodium acetate reaction in SCWG can be found
in the literature (Onwudili and William 2009, 2010).

2HCOONa + H2 O → 2H2 + Na2 CO3 + CO2 H298.15K


0
= 50.6 kJ/mol
(21.35)
2CH3 COONa + H2 O → 2CH4 + Na2 CO3 + CO2 H298.15K
0
= −14.4 kJ/mol
(21.36)
328 21 Notions of Biomass Gasification

We can also have that there is higher production of hydrogen at high temperatures,
and the following reaction can also occur
CH3 COONa + 3H2 O → 4H2 (g) + NaHCO3 (s) + CO2 (g) H298.15K
0
= 89.9 kJ/mol
(21.37)
Salts and inorganics precipitate in SCWG, and this is an evident during the exper-
iments. As a matter of fact, the dielectric constant of water drops to almost 1 at
supercritical conditions, and this means that water is no more a polar compound.
Therefore, salts and inorganics precipitate giving the possibility of recovering these
compounds (Fig. 21.22).
Nevertheless, this is not very easy in practice, even if several solutions have been
proposed in the literature. SCWG offers the possibility of using diverse reactor setups.
For instance, in the literature, scientists have proposed the usage of diverse reactor
zones and in each of these regions there could be different conditions (Susanti et al.
2010).
As an example of solubility decrease of organic salts in SCW, Fig. 21.23 shows
the solubility of sodium carbonate as a function of the temperature (Fig. 21.23).
Precipitating salts and inorganics are creating also problems to the normal oper-
ations of the reactor. Plugging can occur, and because of this, we could experience
different fluctuations in pressure and temperatures within the reactor. The plugging

Fig. 21.22 Dielectric constant of water against temperature at different pressures


21.2 Supercritical Water Gasification 329

Fig. 21.23 Na2 CO3 solubility against temperature. Adapted from Jones et al. (1976)

problem is not frequently studied in the literature with regard to the SCWG of biomass
even if there is some study cases (De Blasio et al. 2019).
Figure 21.24 describes the dynamics of what is taking place during the plugging
of one reactor entry and the sudden release of the clogging wall due to the incoming
pressure. Because of the wall release, a sudden injection of a quantity of feedstock
at much lower temperature will take place. The lower temperature will bring the
conditions to subcritical with a proportional increase of the volume density; however,
since the reactor walls will continue to pump energy into the reactor, these conditions
are rapidly changing causing stress to the reactor walls, inlet, outlet, and related valves
(Fig. 21.24).
To test SCWG of biomass, several diverse solutions have been adopted.
Figure 21.25 demonstrates one example of setup used in the literature (De Blasio
et al. 2019).
The reactors used are listed as follows, starting from the smallest to the pilot-scale
reactors configurations.
– capillary reactors, batch mode of operation (Sricharoenchaikul 2009; Potic et al.
2004);
– micro-reactors, or medium-size batch mode of operation (Castello et al. 2017);
– micro-reactors, continuous mode of operation, (Cao et al. 2011; De Blasio et al.
2015);
– fluidized bed reactors (simulation studies) (Yao and Lu 2017);
– medium-size reactors for laboratory use (Lu et al. 2012);
– Pilot-scale reactors (Yakaboylu et al. 2015).
The difference among them is the size of the reactor, but also the mode of operation.
In the literature, it is not uncommon to find studies which were done in batch mode of
330 21 Notions of Biomass Gasification

Fig. 21.24 Pressure drop due to plugging of reactor and sudden release of the blockage wall

Fig. 21.25 Schematics of a SCWG setup for laboratory-scale experiments. Adapted from De Blasio
et al. (2019)
21.2 Supercritical Water Gasification 331

operation. However, in this case, a remark should be done on the conditions adopted.
When batch conditions are adopted, the sample is normally inserted into the reactor
at a certain time, and after this, the reactor should be brought to the target reactor’s
conditions. However, during this time, the temperature and pressure of the system
will also change, and therefore, it will be difficult to really say when the experiment
is started or not. In addition, during the heating of the reactor, there could have
been reactions which could have changed the structure and composition of the feed
sample, and therefore, when the experiment has started, we cannot talk about the
same kind of sample.

References

Antal, M. J., Brittain, A., DeAlmeida, C., Ramayya, S., & Roy, J. C. (1987). Heterolysis and homol-
ysis in supercritical water. In Supercritical Fluids, ACS Symposium Series. American Chemical
Society (pp. 77–86). https://doi.org/10.1021/bk-1987-0329.ch007.
Ariyapadi, S., Shires, P., Bhargava, M., & Ebbern, D. (2008). KBR’s transport gasifier (TRIG)—An
advanced gasification technology for SNG production from low-rank coals. In Presented at the
Twenty-fifth Annual International Pittsburgh Coal Conference, KBR, Pittsburgh, USA.
Breault, R. W. (2010). Gasification processes old and new: A basic review of the major technologies.
Energies, 3, 216–240. https://doi.org/10.3390/en3020216.
Byrd, A. J., Pant, K. K., & Gupta, R. B. (2008). Hydrogen production from glycerol by reforming in
supercritical water over Ru/Al2O3 catalyst. Fuel, 87, 2956–2960. https://doi.org/10.1016/j.fuel.
2008.04.024.
Cao, C., Guo, L., Chen, Y., Guo, S., & Lu, Y. (2011). Hydrogen production from supercritical water
gasification of alkaline wheat straw pulping black liquor in continuous flow system. International
Journal of Hydrogen Energy, 2010 Asian/APEC BioH2, 36, 13528–13535. https://doi.org/10.
1016/j.ijhydene.2011.07.101.
Castello, D., & Fiori, L. (2011). Supercritical water gasification of biomass: Thermodynamic con-
straints. Bioresource Technology, 102, 7574–7582. https://doi.org/10.1016/j.biortech.2011.05.
017.
Castello, D., Rolli, B., Kruse, A., & Fiori, L. (2017). Supercritical water gasification of biomass
in a ceramic reactor: Long-time batch experiments. Energies, 10, 1734. https://doi.org/10.3390/
en10111734.
Cervi, E., & Cammi, A. (2018). Stability analysis of the supercritical water reactor by means of the
root locus criterion. Nuclear Engineering and Design, 338, 137–157. https://doi.org/10.1016/j.
nucengdes.2018.08.004.
Chen, Z., Li, Y., Lai, D., Geng, S., Zhou, Q., Gao, S., et al. (2018). Coupling coal pyrolysis with
char gasification in a multi-stage fluidized bed to co-produce high-quality tar and syngas. Applied
Energy, 215, 348–355. https://doi.org/10.1016/j.apenergy.2018.02.023.
da Silva, D. C. D., Letichevsky, S., Borges, L. E. P., & Appel, L. G. (2012). The Ni/ZrO2 catalyst
and the methanation of CO and CO2 . International Journal of Hydrogen Energy, 37, 8923–8928.
https://doi.org/10.1016/j.ijhydene.2012.03.020.
De Agarwal, S., Moholkar, V. S., & Thallada, B. (2018). Coal and biomass gasification: Recent
advances and future challenges (1st ed.). Singapore: Springer.
De Blasio, C., De Gisi, S., Molino, A., Simonetti, M., Santarelli, M., & Björklund-Sänkiaho, M.
(2019). Concerning operational aspects in supercritical water gasification of kraft black liquor.
Renewable Energy, 130, 891–901. https://doi.org/10.1016/j.renene.2018.07.004.
De Blasio, C., & Järvinen, M. (2017). Supercritical water gasification of biomass. In Encyclopedia
of sustainable technologies (pp. 171–195). Abraham M Editor.
332 21 Notions of Biomass Gasification

De Blasio, C., Lucca, G., Özdenkci, K., Mulas, M., Lundqvist, K., Koskinen, J., et al. (2015). A
study on supercritical water gasification of black liquor conducted in stainless steel and nickel-
chromium-molybdenum reactors. Journal of Chemical Technology and Biotechnology, 91(10),
2664–2678. https://doi.org/10.1002/jctb.4871.
Farajtabar, A., & Gharib, F. (2009). Autoprotolysis constants determination of water-methanol
mixtures and solvent effect. Journal of Taibah University for Science, 2, 7–13. https://doi.org/10.
1016/S1658-3655(12)60002-8.
Finney, J. L. (2004). Water? What’s so special about it? Philos. Transactions of the Royal Society
B: Biological Sciences, 359, 1145–1328. https://doi.org/10.1098/rstb.2004.1495.
Gizir, A. M., Clifford, A. A., & Bartle, K. D. (2003). The catalytic role of transition metal salts on
supercritical water oxidation of phenol and chlorophenols in a titanium reactor. Reaction Kinetics
and Catalysis Letters, 78, 175–182. https://doi.org/10.1023/A:1022586505112.
Görsch, C., Hannemann, F., Hammer, T., & Meyer, B. (2011). Impact of the water quench on the
syngas composition obtained from entrained flow gasification of carbonaceous fuels. In 28th
international pittsburgh coal conference. Presented at the Pittsburgh Coal Conference. Siemens,
Pittsburgh, USA.
He, M., Su, C., Liu, X., Qi, X., & Lv, N. (2015). Measurement of isobaric heat capacity of pure
water up to supercritical conditions. The Journal of Supercritical Fluids, 100, 1–6.
Hoffmann, M. M., & Conradi, M. S. (1998). Are there hydrogen bonds in supercritical methanol and
ethanol? The Journal of Physical Chemistry B, 102, 263–271. https://doi.org/10.1021/jp9726706.
Jones, D. G., Slater, J. E., Staehle, R. W. (1976). High temperature high pressure electrochemistry
in aqueous solutions. In International Corrosion Conference Series Volume 4 of National Asso-
ciation of Corrosion Engineers: NACE. National Association of corrosion engineers, University
of Surrey, Guildford.
Kalinichev, A. G. (2017). Universality of hydrogen bond distributions in liquid and supercritical
water. Journal of Molecular Liquids, 241, 1038–1043. https://doi.org/10.1016/j.molliq.2017.06.
114.
Karakuş, Y., Aynacı, F., Kıpçak, E., & Akgün, M. (2013). Hydrogen production from 2-propanol
over Pt/Al2O3 and Ru/Al2O3 catalysts in supercritical water. International Journal of Hydrogen
Energy, 38, 7298–7306. https://doi.org/10.1016/j.ijhydene.2013.04.033.
Koss, U., & Schlichting, H. (2005). Lurgi’s MPG Gasification plus Rectisol® gas purification–Ad-
vanced process combination for reliable syngas production. San Francisco, California, US: Pre-
sented at the Gasification Technologies.
Krishnamoorthy, V., & Pisupati, S. V. (2015). A critical review of mineral matter related issues during
gasification of coal in fixed, fluidized, and entrained flow gasifiers. Energies, 8, 10430–10463.
https://doi.org/10.3390/en80910430.
Kruse, A. (2008). Supercritical water gasification. Biofuels, Bioproducts and Biorefining, 2,
415–437. https://doi.org/10.1002/bbb.93.
Kunii, D., & Levenspiel, O. (1991). Fluidization engineering (2nd ed.). Elsevier.
Lau, F. (2009). Commercial development of the SES U-gas gasification technology. In Presented
at the gasification technologies conference, synthesis energy systems. US: Colorado Springs.
Lisandy, K. Y., Kim, R.-G., Hwang, C.-W., & Jeon, C.-H. (2016). Sensitivity test of low rank
Indonesian coal utilization using steady state and dynamic simulations of entrained-type gasifier.
Applied Thermal Engineering, 102, 1433–1450. https://doi.org/10.1016/j.applthermaleng.2016.
04.040.
Louw, J., Schwarz, C. E., & Burger, A. J. (2016). Catalytic supercritical water gasification of primary
paper sludge using a homogeneous and heterogeneous catalyst: Experimental vs thermodynamic
equilibrium results. Bioresource Technology, 201, 111–120. https://doi.org/10.1016/j.biortech.
2015.11.043.
Lu, Y., Guo, L., Zhang, X., Ji, C. (2012). Hydrogen production by supercritical water gasification
of biomass: Explore the way to maximum hydrogen yield and high carbon gasification efficiency.
International Journey of Hydrogen Energy, International Conference on Renewable Energy (ICRE
2011), 37, 3177–3185. https://doi.org/10.1016/j.ijhydene.2011.11.064
References 333

Martskainen, O. (2009). Tautomerism and fragmentation of biologically active hetero atom


(O,N)—containing acylic and cyclic compounds under electron ionization. Painosalama OY.
Mitsubishi Heavy Industries. (2019). Gasification process | mitsubishi heavy industries environ-
mental & chemical engineering [WWW Document]. URL https://www.mhiec.co.jp/en/products/
recycle/city/meltingsystem/contents/gasification.html. Accessed January 24, 19.
Nanda, S., Dalai, A. K., & Kozinski, J. A. (2016). Supercritical water gasification of timothy grass
as an energy crop in the presence of alkali carbonate and hydroxide catalysts. Biomass and
Bioenergy, 95, 378–387. https://doi.org/10.1016/j.biombioe.2016.05.023.
Nguyen, H. T., Lu, H., Kobayashi, E., Ishikawa, T., & Komiyama, M. (2014). Raney-Nickel catalyst
deactivation in supercritical water gasification of ethanol fermentation stillage and its mitigation.
Topics in Catalysis, 57, 1078–1084. https://doi.org/10.1007/s11244-014-0272-x.
Onwudili, J. A., & Williams, P. T. (2009). Role of sodium hydroxide in the production of hydrogen
gas from the hydrothermal gasification of biomass. International Journal of Hydrogen Energy, 2nd
International Conference on Hydrogen Safety 34, 5645–5656. https://doi.org/10.1016/j.ijhydene.
2009.05.082.
Onwudili, J. A., & Williams, P. T. (2010). Hydrothermal reactions of sodium formate and sodium
acetate as model intermediate products of the sodium hydroxide-promoted hydrothermal gasifi-
cation of biomass. Green Chemistry, 12, 2214–2224. https://doi.org/10.1039/C0GC00547A.
Osada, M., Sato, O., Watanabe, M., Arai, K., & Shirai, M. (2006). Water density effect on lignin
gasification over supported noble metal catalysts in supercritical water. Energy & Fuels, 20,
930–935. https://doi.org/10.1021/ef050398q.
Osada, M., Yamaguchi, A., Hiyoshi, N., Sato, O., & Shirai, M. (2012). Gasification of sugar-
cane bagasse over supported ruthenium catalysts in supercritical water. Energy & Fuels, 26,
3179–3186. https://doi.org/10.1021/ef300460c.
Özdenkçi, K., De Blasio, C., Muddassar, H. R., Melin, K., Oinas, P., Koskinen, J., et al. (2017).
A novel biorefinery integration concept for lignocellulosic biomass. Energy Conversion and
Management, 149, 974–987. https://doi.org/10.1016/j.enconman.2017.04.034.
Portela, J. R., Nebot, E., & Martı́nez de la Ossa, E. (2001). Kinetic comparison between subcritical
and supercritical water oxidation of phenol. Chemical Engineering Journal, 81, 287–299. https://
doi.org/10.1016/S1385-8947(00)00226-6.
Potic, B., Kersten, S. R. A., Prins, W., & van Swaaij, W. P. M. (2004). A high-throughput screen-
ing technique for conversion in hot compressed water. Industrial and Engineering Chemistry
Research, 43, 4580–4584. https://doi.org/10.1021/ie030732a.
Randolph, T. W., & Carlier, C. (1992). Free-radical reactions in supercritical ethane: A probe of
supercritical fluid structure. Journal of Physical Chemistry, 96, 5146–5151. https://doi.org/10.
1021/j100191a072.
Ratkiewicz, A., & Truong, T. N. (2012). Kinetics of the C–C bond beta scission reactions in alkyl
radical reaction class. Journal of Physical Chemistry A, 116, 6643–6654. https://doi.org/10.1021/
jp3018265.
Resende, F. L. P., & Savage, P. E. (2010). Effect of metals on supercritical water gasification of
cellulose and lignin. Industrial and Engineering Chemistry Research, 49, 2694–2700. https://doi.
org/10.1021/ie901928f.
Rönnlund, I., Myréen, L., Lundqvist, K., Ahlbeck, J., & Westerlund, T. (2011). Waste to energy
by industrially integrated supercritical water gasification—Effects of alkali salts in residual by-
products from the pulp and paper industry. In: Energy, 5th Dubrovnik Conference on Sustainable
Development of Energy, Water & Environment Systems, (Vol 36, pp. 2151–2163). https://doi.
org/10.1016/j.energy.2010.03.027.
Ryan, E. T., Xiang, T., Johnston, K. P., & Fox, M. A. (1996). Excited-state proton transfer reactions
in subcritical and supercritical water. Journal of Physical Chemistry, 100, 9395–9402. https://
doi.org/10.1021/jp9604204.
Sasaki, M., Furukawa, M., Minami, K., Adschiri, T., & Arai, K. (2002). Kinetics and mechanism of
cellobiose hydrolysis and retro-aldol condensation in subcritical and supercritical water. Industrial
and Engineering Chemistry Research, 41, 6642–6649. https://doi.org/10.1021/ie020326b.
334 21 Notions of Biomass Gasification

Sikarwar, V. S., Zhao, M., Clough, P., Yao, J., Zhong, X., Memon, M. Z., et al. (2016). An overview
of advances in biomass gasification. Energy & Environmental Science, 9, 2939–2977. https://doi.
org/10.1039/C6EE00935B.
Sricharoenchaikul, V. (2009). Assessment of black liquor gasification in supercritical water. Biore-
source Technology, 100, 638–643. https://doi.org/10.1016/j.biortech.2008.07.011.
Stangeland, K., Kalai, D., Li, H., & Yu, Z. (2017). CO2 Methanation: The effect of catalysts
and reaction conditions. In Energy procedia, 8th international conference on applied energy,
ICAE2016, 8–11 Oct 2016 (Vol. 105, pp. 2022–2027), Beijing, China. https://doi.org/10.1016/j.
egypro.2017.03.577.
Susanti, R. F., Veriansyah, B., Kim, J.-D., Kim, J., & Lee, Y.-W. (2010). Continuous supercritical
water gasification of isooctane: A promising reactor design. International Journal of Hydrogen
Energy, 35, 1957–1970. https://doi.org/10.1016/j.ijhydene.2009.12.157.
ThyssenKrupp. (2019). Gasification technologies.
Toporov, D., & Abraham, R. (2015). Gasification of low-rank coal in the high-temperature winkler
(HTW) process. Journal of the South African Institute of Mining and Metallurgy, 115, 589–597.
https://doi.org/10.17159/2411-9717/2015/V115N7A5.
Wagner, W., & Kretzschmar, H.-J. (2008). International steam tables—Properties of water and
steam based on the industrial formulation IAPWS-IF97: Tables, algorithms, diagrams, and CD-
ROM electronic steam tables—All of the equations of IAPWS-IF97 including a complete set
of supplementary backward equations for fast calculations of heat cycles, boilers, and steam
turbines (2nd ed.). Berlin Heidelberg: Springer.
Wang, J., & Takarada, T. (2001). Role of calcium hydroxide in supercritical water gasification of
low-rank coal. Energy & Fuels, 15, 356–362. https://doi.org/10.1021/ef000144z.
Wang, C., Zhu, W., Chen, C., Zhang, H., Lin, N., & Su, Y. (2018). Influence of reaction conditions
on the catalytic activity of a nickel during the supercritical water gasification of dewatered sewage
sludge. Journal of Supercritical Fluids, 140, 356–363. https://doi.org/10.1016/j.supflu.2018.07.
018.
Watanabe, M., Inomata, H., Smith, R. L., & Arai, K. (2001). Catalytic decarboxylation of acetic acid
with zirconia catalyst in supercritical water. Applied Catalysis General, 219, 149–156. https://
doi.org/10.1016/S0926-860X(01)00677-9.
Yakaboylu, O., Harinck, J., Smit, K. G., & de Jong, W. (2015). Supercritical water gasification of
biomass: A literature and technology overview. Energies, 8, 859–894. https://doi.org/10.3390/
en8020859.
Yamamoto, N., Murakami, K., Kimthet, C., Wahyudiono, W., Onwona-Agyeman, S., Kanda, H.,
et al. (2018). Utilization of sub and supercritical water reactions in resource recovery of biomass
wastes. Engineering Journal, 22, 31–44. https://doi.org/10.4186/ej.2018.22.5.31.
Yang, X., Cheng, K., & Jia, G. (2019). The molecular dynamics simulation of hydrogen bonding in
supercritical water. Physica A: Statistical Mechanics and its Applications, 516, 365–375. https://
doi.org/10.1016/j.physa.2018.10.022.
Yanik, J., Ebale, S., Kruse, A., Saglam, M., & Yüksel, M. (2008). Biomass gasification in super-
critical water: II. Effect of catalyst. International Journal of Hydrogen Energy, 33, 4520–4526.
https://doi.org/10.1016/j.ijhydene.2008.06.024.
Yao, L., & Lu, Y. (2017). Supercritical water gasification of glucose in fluidized bed reactor: A
numerical study. International Journal of Hydrogen Energy, 42, 7857–7865. https://doi.org/10.
1016/j.ijhydene.2017.03.009.
Zhang, J., Liu, J., & Liu, R. (2015). Effects of pyrolysis temperature and heating time on biochar
obtained from the pyrolysis of straw and lignosulfonate. Bioresource Technology, 176, 288–291.
https://doi.org/10.1016/j.biortech.2014.11.011.
Zhou, L., & Zhou, Y. (2001). Determination of compressibility factor and fugacity coefficient of
hydrogen in studies of adsorptive storage. International Journal of Hydrogen Energy, 26, 597–601.
https://doi.org/10.1016/S0360-3199(00)00123-3.
Part III
Review of Some Concepts
of Thermodynamics
and Support Material
Chapter 22
Preliminary Concepts

This, to be sure, deprives the work of the character of an original


contribution to science, and stamps it rather as an introductory
text-book on Thermodynamics for students who have taken
elementary courses in Physics and Chemistry, and are familiar
with the elements of the Differential and Integral Calculus.
A sentence written by Max Planck about his book, Treatise
on Thermodynamics, translated by A. Ogg in 1903. Longmans,
green and Co. Publishers.

22.1 Definition of Mole

Let us assume that a quantity of mass is entering a reactor at a known mass flow
(kg/s). In case the substance entering the reactor is known, then it would be possible
to calculate the number of moles of that substance. In other words, it will be possible
to calculate the number of molecules that are entering that reactor as a function of
time. This is useful, as it was stated in the manuscript, since the reactor volume can
vary (meaning that the volume of the matter in the reactor can actually vary during
the reaction time).
By definition, one mole represents the same number of particles present in 12 g
of carbon-12. For this reason, this is referred as g-mole. This number is called Avo-
gadro’s number or Avogadro’s constant (N A ) and is equal to 6.022 × 1023 particles.
Sometimes, it is referred in the literature to kg-mole instead of g-mole and since
1 kg = 1000 g, then the number of molecules in a kg-mole is 1000 g-mole.
From the definition, it comes natural that one mole of a pure substance has a mass
in grams equal to its molecular mass, also known as molecular weight (MW). This
is often referred to as the molar mass.
The volume of a substance is directly proportional to its number of moles. To
give one example, one mole of an ideal gas has a volume of: 22.4 L at standard
temperature and pressure, 0 °C (273 K), and 101.3 kPa (1 atm) (Prigogine 1968;
Levine 2008).

© Springer Nature Switzerland AG 2019 337


C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_22
338 22 Preliminary Concepts

22.2 Conservation of Mass, Closed Systems, Degree


of Advancement of Reaction

Chemical reactions are almost everywhere in engineering processes. It is therefore of


outmost importance to understand some preliminary concepts concerning chemical
reactions. It has been stated in this manuscript that, in case we consider a balance on
the number of moles entering a reactor, the volume of the matter inside the reactor will
indeed not remain the same and it is normally an assumption to consider it as constant.
In this case, we will consider here the definition of the degree of advancement of a
reaction. Let us define the change in mass for a component:

dm γ = νγ Mγ dε (22.1)

Where
νγ is the stoichiometric coefficient (dimensions of mole).
dm γ is the change in mass of a component γ .
Mγ is the molar mass.
The mass in closed systems is conserved even if there are chemical reactions
involved, with changes in mole numbers.
 

dm = νγ Mγ dε = 0 (22.2)
γ

22.3 About Internal Energy and Enthalpy

In this section, we will take into account two important concepts normally used in
energy technology and chemical engineering. First, let us consider here the enthalpy
“which is the formally incorrect combination of two different kinds of magnitudes”
(Astarita 1993). The enthalpy is defined as:

H = U + PV (22.3)

Differentiating Eq. (22.3) we have:

dH = dU + pdV + V d p (22.4)

Now, it is important to notice that the internal energy, U, is not a function of the
volume, and therefore, the terms stating a variation of volume such as dV are going
to be zero if the volume is not changing. We have:
22.3 About Internal Energy and Enthalpy 339

dU dU
dH = dT + dV + pdV + V dP (22.5)
dT v dV T

And therefore:

H = c · dT (22.6)

where c is a generic specific heat

cv dT = u (22.7)

This is because we do not have the term P · V in case we consider only U.


Therefore:

dU
cdT = dT + pdV + V dP (22.8)
dT v
and

c p dT = du + pdV + vd p = dh (22.9)

This is derived by following the definition of enthalpy.

22.4 Relations Between Specific Heat at Constant Volume


and Pressure

We know that:

P ·v = R·T (22.10)

where R is the universal gas constant and equal to 8.314 J/(mol K). We also have
from previous derivations:

cv dT = u (22.11)

In addition, we know also that:

c p dT = u + pdv + vd p = h (22.12)

Substract Eq. (22.11) from Eq. (22.12), we obtain:

c p dT − cv dT = u + pdv + vd p − u (22.13)
340 22 Preliminary Concepts

where

pdv + vd p = RdT (22.14)

To be noticed that Eq. (22.14) comes straight from Eq. (22.10) differentiating with
respect to the variables P and T. And the following relation is derived:

c p − cv = R (22.15)

In case we consider an isoentropic process the ratio between the specific heat
parameters at constant pressure and constant volume:
cp
=k (22.16)
cv

Equation (22.16) also has a straightforward derivation from Eqs. (22.11) and
(22.12). This is because if the entropy is not changing, then there is no variation of
volume or pressure on the system and also the following relation is valid:

P V k = constant (22.17)

22.5 The First Law of Thermodynamics

The change in energy contained in a system during a time interval is equal to the net
energy transfer in the system by heat transfer during the time interval minus the net
amount of energy given out of the system by work during the time interval. This is
stated as:

dKE dPE dU
+ + = Q̇ − Ẇ (22.18)
dt dt dt

where KE and PE are the kinetic energy and the potential energy, while Q̇ and Ẇ
are, respectively, the flows of heat and work. To be noticed that the work done on
a system is positive and this comes directly from the fact that when the pressure
is constant and we apply this pressure on a volume, the work will be described by
W = −P · dV and since dV would be approximated as (Vfinal − Vinitial ), if we do a
compression on the volume, this work will be positive.
A general balance should always be remembered when starting doing calculations
on mass and energy balance:

Input − Output +/− Generation (example: chemical reaction) = Accumulation


(22.19)

Notice again that the term “generation” refers, for instance, to the number of
moles in case the balance is done on them.
22.6 Some Relations Between Pressure and Temperature in Gases. Gas Laws 341

22.6 Some Relations Between Pressure and Temperature


in Gases. Gas Laws

For an ideal gas, the following notorious formula is applied:

P V = n RT (22.20)

where the term R was mentioned already, and it refers to the universal gas constant:

J
R = 8.3144621 (22.21)
mol K

and n is the number of moles. One mole is formed by: 6.02214129 × 1023
molecules/mol−1 .
If the temperature is constant, we derive Boyle’s law from Eq. (22.20)

P1 V1 = P2 V2 (22.22)

where 1 and 2 refer to the states of the system. On the other hand, in case the pressure
is constant, we have the Charles law:

V1 /V2 = T1 /T2 (22.23)

P1 V1 /T1 = P2 V2 /T2 (22.24)

If n1 = n2 and V 1 = V 2 , we derive the Gay-Lussac Law. The n’s and the V ’s


cancel in the above original expression and we have:

P1 /T1 = P2 /T2 (22.25)

22.7 A Remark on Reversible Processes

Processes, described by equations which are invariant to the sign we give to the
variable of time, are reversible.
342 22 Preliminary Concepts

For instance, the following equation:

1 ϑ 2u ϑ 2u ϑ 2u ϑ 2u
= + + (22.26)
c2 ϑt 2 ϑ x2 ϑ y2 ϑ z2

Instead, processes, described by equations where the sign given to the time makes
a difference, are irreversible. Such as:

1 ϑT ϑ2T ϑ2T ϑ2T


= + + (22.27)
α ϑt ϑ x2 ϑ y2 ϑ z2

22.8 Pressure and Internal Energy, the Kinetic Model

Consider Fig. 22.1 where a simple space is given as a box. In the box, we imagine
there is an object moving back and forth along a distance L with a velocity wi . With
the letter i, we refer to the ith particle of a system.
Remember that the velocity is a vector and it has a module (the quantity that
describes how big the vector is) and a direction. We consider now one molecule
going back and forth and a box (cube) of length L.

Fig. 22.1 Simple


representation of a space.
The box with a moving
object in it
22.8 Pressure and Internal Energy, the Kinetic Model 343

Momentum = mwi − (−mwi ) = 2mwi (22.28)

The period will be

2L
t= (22.29)
wi

The frequency will be:

number of movement wi
Frequency = = (22.30)
time 2L
The force associated with the particle is:

wi mwi2
Force = 2m · wi · = (22.31)
2L L
To calculate the pressure, remember that the pressure is defined as the force/surface
where the force is applied.

mwi 2 1
Pressure of one molecule i = · (22.32)
L 3L 2
In Eq. (22.32), the 1/3 is there since we are considering three pairs of faces, in
other words, all the box. In case:

N
n∗ = (22.33)
L3
We have:
N
m wi 2 1  2
P= ·n· i=1
= · ρ · ŵ (22.34)
3 N 3
where
n* is the number of molecules/m3 .
m · N = m tot in the space L 3 .
m · n ∗ = mass
m3
which is the density.

22.9 Specific Heat and Universal Gas Constant

From previous equations, it is possible to write:

1
PV = m Tot w 2 (22.35)
3
344 22 Preliminary Concepts

And therefore,

21
PV = m Tot w 2 (22.36)
32
We remember the ideal gas law:

P V = n RT and H = U + P V (22.37)

This expression will be written to consider M as the molecular weight:

1 3 1 3
Mw 2 = · · n RT = RT (22.38)
2 2 n 2
In case we consider a system closed, this will give:

3 1
cv dT = dU = n RdT = m tot w 2 (22.39)
2 2
This is the derived relation between the temperature and the kinetic energy.
In case the system is able to do work at P constant:

3 5
dH = c p dT = dU + pdV = n RdT + nRdT = n RdT (22.40)
2 2
Now, let us consider a particle traveling a piece of space d = space travelled.
Its velocity is:

d
vav = (22.41)
t
and the acceleration will be:

v
aav = (22.42)
t
The force on average will be:

v
Fav = m · aav = m (22.43)
t

and its work:



v
work = Fav d = m d (22.44)
t

Or
22.9 Specific Heat and Universal Gas Constant 345

va − vb
work = Fav d = m d (22.45)
t

With the average speed over the time interval:

1
d = (va + vb )t (22.46)
2
And therefore, the following expression is obtained:

1 2 
work = va − vb2 (22.47)
2

22.10 Change of State for an Ideal Gas

For this section, it is good to recall the Avogadro number N A = 6.022×1023 particles.
It is also beneficial to observe that

R
KB = ≈ 1.38 · 10−23 J/K (22.48)
NA

This is a straight derivation from the ideal gas law. In case we have an isochoric
process (V = const) (Fig. 22.2).
In this case, we necessarily have:

W1→2 = 0 (22.49)

Since we do not have a variation of the volume for our system, however, a change
in internal energy is still possible obviously:

Fig. 22.2 Constant volume


system with varying pressure
346 22 Preliminary Concepts

Fig. 22.3 Simple


representation of an isobaric
process

3
Q 1→2 = N K B (T2 − T1 ) > 0 (22.50)
2
where

N (molecules) = n(moles) · N A

This is also more simply written as:

Q 1→2 = C V T (22.51)

and since the work done on the system is zero, we have that the internal energy is
just a function of the heat variation in the system considered.

dU = Q 1→2 (22.52)

For an isobaric process, we have that the pressure is constant during the transfor-
mation and the volume is instead varying (Fig. 22.3).
In this case, the volume of the system considered is changing, and therefore, the
work done on the system should be taken into account:
2
W1→2 = − ∫ P(V, T )dV = −P(V2 − V1 ) < 0 (22.53)
1

and the change in heat content of the system will be:

5
Q 1→2 = N K B (T2 − T1 ) > 0 (22.54)
2
In this case, the quantity Q 1→2 is also clearly:
22.10 Change of State for an Ideal Gas 347

Fig. 22.4 Isothermal


process. The points 1 and 2
are on the same curve in this
case: the isothermal P-V
curve

Q 1→2 = C P T (22.55)

and therefore, the overall internal energy of the system will be:

dU = W1→2 + Q 1→2 (22.56)

For an isothermal process, we have a transformation where the temperature is


remained constant (Fig. 22.4).
In this case, there will be no variation for the internal energy for the system
considered:

dU = 0 (22.57)

2 V2 dV V2
W1→2 = − ∫ P(V, T )dV = −N K B T ∫ = −N K B T ln (22.58)
1 V1 V V1

and from previously written equations, it follows that:

Q 1→2 = −W1→2 (22.59)

In case we generalize and describe with i, the initial stage and with f the final
stage of the process:

Vf
Wi− f = −N K B T ln (22.60)
Vi

In the case, we have compression:

Vi > V f ;
348 22 Preliminary Concepts

and this will imply:

Wi− f > 0

And in case of expansion, the system does work to the environment and we have:

Vi < V f

with

Wi− f < 0

22.11 Adiabatic Transformation of an Ideal Gas

When a system is considered to be adiabatic, this means that we are talking about a
thermally isolated system.

Q 1→2 = 0 (22.61)

and

dU = W1→2 (22.62)

In order to calculate W 1–2 , we need the function of the pressure which is dependent
upon the volume and the temperature, P(V, T ), and therefore, for an adiabatic process,
it is still valid the relation:
2
W1→2 = − ∫ P(V, T )dV (22.63)
1

Let us consider the parameter f which gives the degrees of freedom for a particular
gas and, in other words, it gives the number of independent motions a molecule can
have.
f is considered to be 3 for monatomic molecules, 5 for diatomic molecules, and 6
for polyatomic. This could be simply explained by the fact that if we are considering
only one atom, of course, there will be three possibilities to move along the special
axes. On the other hand, for a diatomic molecule, there will be three possibilities to
move (for the overall molecule) along the x, y, and z axes and two possibilities for
rotation. Notice that if a diatomic molecule will rotate along the axes connecting the
two atoms, this will be not considered a rotation. For the case of nonlinear molecules,
we have that the overall molecule can, also in this case, move along the three main
22.11 Adiabatic Transformation of an Ideal Gas 349

axes, and furthermore, it can rotate along all the three axes. This will give a total of
six configurations for the system at hand.
More generally, according to Maxwell’s law of equipartition of energy the internal
energy will be expressed as (Lima and Plastino 2000):

f
U= N KBT (22.64)
2
This is because Maxwell’s law is stating that each degree of freedom has an
equal amount of energy when it is in thermodynamic equilibrium, and this amount
of energy will be proportional to the Boltzmann constant and the temperature of the
system. A derivation for the law of equipartition for kinetic energy, thermodynamic
equilibrium, and more than one dimension is given by Crawford (1986).
And this will mean:

f
dU = N K B dT = −PdV (22.65)
2
From the relation:

P V = N KBT (22.66)

we differentiate both terms with respect to the variables involved:

PdV + V dP = N K B dT (22.67)

and this will give:

2
PdV + V dP = − PdV (22.68)
f

which in turn gives:



dV 2 dP
1+ + =0 (22.69)
V f P

Now, if we consider the substitution:

2
γ =1+ (22.70)
f

Integrating Eq. (22.69) we have:

V dV P dP
γ ∫ +∫ =0 (22.71)
V1 V P1 P
350 22 Preliminary Concepts

Fig. 22.5 Adiamatic process


of compression of an ideal
gas

And finally:
 γ 
V P1
ln = ln (22.72)
V1 P

Which implies:

γ
P V γ = P1 V1 = const (22.73)

In an adiabatic process (Fig. 22.5), the work done by the gas system on the
environment is done at the expenses of its thermal energy, and this implies that its
temperature will decrease. As a summary and in relation to the parameter γ , the
following expression is obtained:
2
W1→2 = − ∫ P(V, T )dV
1
γ V2
V2 P1 V1 γ 1
=−∫ γ
dV = −P1 V1 V 1−γ
V1 V 1−γ V1
 
γ 1 1 1
= P1 V1 − γ −1 (22.74)
γ − 1 V2γ −1 V1

Considering Eq. (22.71), we have that γ = 1 + 2/3 ≈ 1.67 for monatomic


molecules, 1 + 2/5 = 1.4 for diatomic molecules, and 1 + 2/6 ≈ 1.33 for polyatomic
molecules. This is obtained not considering the vibrational degrees of freedom.
22.12 Energy Balance for a Given System: Macroscopic Balances 351

22.12 Energy Balance for a Given System: Macroscopic


Balances

In this section (Fig. 22.6), the overall energy balance in a system will be given in
accordance with what it was mentioned on a general mole/mass balance. Let us
consider the following volume and always the general relation:

In − Out +/− Gen = Accumulation

The overall energy balance for the system will be:



d(Utot + K tot + tot ) 1
= ρ1 U1 v1  + ρ1 v13 + ρ1 1 v1  S1
dt 2

1
− ρ2 U2 v2  + ρ2 v23 + ρ2 2 v2  S2
2
+ ( p1 v1 S1 − p2 v2 S2 + Wm + Q (22.75)

In Eq. (22.75), the terms Utot + K tot + tot are, respectively, the overall internal,
kinetic, and potential energy change in the system. In case the balance is done within
Sects. 22.1 and 22.2, the flow of energy is crossing the surfaces 1 and 2 at the inlet and
outlet, respectively. ρ1 is the density of the medium carrying the energy at Sect. 22.1,
U1 will be the specific internal energy of the medium at that section (notice that this
overall term has to have dimensions of Utot /time). v1  and v2  are, respectively, the
velocities at Sects. 22.1 and 22.2. S1 and S2 are the surfaces where the flow is taking
place at Sects. 22.1 and 22.2 (Biron Bird et al. 1960). Q and W are the rate at which
heat is added/subtracted to the system and the work that is done or received by the
system.

Fig. 22.6 Volume with its


boundaries
352 22 Preliminary Concepts

Equation (22.75) is written here with a descriptive purpose, and it is not the scope
here to give a detailed treatment of the balances involved when macroscopic balances
are considered. However, it is useful for the reader to notice here the analogy between
heat and mass transfer.

References

Prigogine, I. (1968). Thermodynamics of irreversible processes (3rd ed.). New York: Wiley.
Levine, I. N. (2008). Physical chemistry (6th ed.). McGraw-Hill Education.
Astarita, G. (1993). Thermodynamics: A view from outside. Fluid Phase Equilibria, 82, 1–14.
https://doi.org/10.1016/0378-3812(93)87123-I.
Lima, J. A. S., Plastino, A. R. (2000). On the classical energy equipartition theorem. Brazilian
Journal of Physics, 30. http://dx.doi.org/10.1590/S0103-97332000000100019.
Crawford, F. S. (1986). Elementary derivation of the law of equipartition of energy. American
Journal of Physics, 55, 180.
Biron Bird, R., Stewart, W. E., & Lightfoot, E. N. (1960). Transport phenomena. Singapore: Wiley.
Chapter 23
Introduction to Entropy and Second Law

Everyone knows that heat can produce motion.


That is, possesses vast motive-power no one can doubt, in these
days when the steam-engine is everywhere so well known.
Carnot (1897).

23.1 Mixing of One or Two Gases at the Same T and P

Consider the setup given by Fig. 23.1, we have two gases in two boxes and the system
formed by the two boxes is adiabatic, meaning that the system is not exchanging heat
with the environment. The stage 1 refers to the gases when they are separated: The
gas 1 is on the left, and the gas 2 is on the right of the figure. Additionally, let us also
assume that the two gases have the same identical temperature.
In case:

Q = 0

and

T = 0

A system tends to evolve from a state at lower probability to a state at higher


probability, this means that when we let the two gases to mix with each other, there
will be more probability that the two gases will be completely mixed, and therefore,
the uncertainty to locate a molecule in the final stage is doubled with respect to the
first state.
The process described previously still involves a variation of entropy as it will be
discussed later in this section.
Suppose that the heat transfer is going from a cold body to a warm body, then
the first principle is still agreed; however, we know from experience that this is

© Springer Nature Switzerland AG 2019 353


C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_23
354 23 Introduction to Entropy and Second Law

Fig. 23.1 Adiabatic mixing of two ideal gases

Fig. 23.2 Object moving under the effect of a force field, gravity

impossible. The second law of thermodynamics is founded on the principle of non-


conservation of a quantity called entropy.
The potential, mechanical, and thermal energies are transferred from a higher
value to a lower value of the intensive variables: temperature, pressure, and coordi-
nates. And therefore, it can be noticed the analogy between these quantities. For this
purpose, let us consider the system described in Fig. 23.2.
In Fig. 23.2, we have an object (a ball) standing on a hill, the hill is at a height
of H with respect to the lowest available point. The ball in the initial state A has
potential energy, but it does not have kinetic energy since it is not moving. When the
23.1 Mixing of One or Two Gases at the Same T and P 355

ball proceeds downhill, then it has kinetic and potential energy. At the bottom, the
ball is still not moving again, and in this case, it does not have kinetic and neither
potential energy.
If the system and the environment have the same value of their intensive variables,
then it is impossible to have work. In addition, we want to know what is the preferable
level at which the energy is.
Concerning the entropy concept, we can also think about a more simple derivation
of it. Let us consider Fig. 23.3.
In Fig. 23.3, a general variable function f (x, Y ) = h is described as dependent
upon two variables x and y. In order to arrive to the final value, any path could be
selected, and in particular, the following relations are valid:
y2 x2
h = ∫ S y (y, x1)dy + ∫ Sx (y2 , x)dx (23.1)
y1 x1

And
y2 x2
h = ∫ S y (y, x2)dy + ∫ Sx (y1 , x)dx (23.2)
y1 x1

h is a function of state in case its variation does not depend on the route we choose,
in addition, it has to be that the function:

dh = H (x, y)dx + F(x, y)dy (23.3)

Is an exact differential if:


 
dg
H (x, y) = (23.4)
dx y

Fig. 23.3 Initial and ending


state of a variable with two
possible routes to arrive to
the final point
356 23 Introduction to Entropy and Second Law

and
 
dg
F(x, y) = (23.5)
dy x

Notice that this is necessary since the path described by dh in this case should be
referred to the x and y coordinates. Of course, we could think about non-perpendicular
coordinates and a different system of reference, and however, it is convenient here
to keep the x-y orthogonal axes.

Example

Consider the function of two variables,

z = x 2 y + 3y 2 (23.6)

The differential will be then:


   
ϑz ϑz  
dz = dx + dy = (2x y)dx + x 2 + 6y dy (23.7)
ϑx y ϑy x

In case we consider the points P = (0, 0) and P = (2, 1). It will be possible to
calculate z along the Route 1 (referring to Fig. 23.3). We have:
1 2
1  2 y 2  x 2  6
(z)1 = ∫ x12 + 6y dy + ∫(2x y2 )dx = 6  + 2  = + 4 = 7 (23.8)
0 0 2 0 2 0 2

Along the Route 2, the calculation will be as follows:


1
1  2 y 2  6
(z)2 = ∫ x22 + 6y dy + ∫(2x y1 )dx = 0 + 4y|10 + 6  = + 4 = 7 (23.9)
0 0 2 0 2

In the case of the quantity heat and work, they are not variable of state, or state
functions. Let us consider the following expression. In case the external pressure is
not constant, we will have to consider the pressure as a variable, and it cannot be
positioned outside the integral operation:

vf
W = − ∫ Pest dV (23.10)
vi

Nevertheless, the internal energy is a state function:


2
∫ dU = U2 − U1 = U (23.11)
1
23.1 Mixing of One or Two Gases at the Same T and P 357

and by doing the partial differentiation for the definition of internal energy seen
previously:

n RT
δqrev = dU − δwrev = C V (T )dT + PdV = C V (T )dT + dV (23.12)
V
Dividing by the temperature, we have:

δqrev C V (T )dT nR
= + dV (23.13)
T T V
And this is a state function since the differential in this case is an exact differential.
A function of this kind is also called “property.”
The term:

δqrev
= S(V, T ) (23.14)
T
is called entropy and it is a function of state or state variable.
Integrating Eq. (23.14), we have:

S(V, T ) = Cv log T + n R log V + const (23.15)

23.2 Statistical Derivation of Entropy

In this section, the concept of entropy will be treated by its statistical meaning.
Imagine two boxes filled with two different gases (D and E) and they are put into
communication by a conduct which is closed in the beginning. The two gases are at
the same temperature and the system is adiabatic, meaning that there is no exchange
of heat. The system boundaries are represented by the dashed line in Fig. 23.4.
In this system, there will be still variation of entropy, even if the temperature will
be the same, and therefore, the definition given by Eq. (23.15) is not complete. The
increase in entropy is the rise from a low-probability state to a high-probability state.
When we have the two gases mixed, there is 50% chance that one particle is in
the right or left part, and for this reason, each of the particles can be considered as a
Bernoulli random variable.
Each of the particles could be imagined as tosses of a coin, and in our case, the
number of tosses is very high (consider the particles contained in one mole of gas).
It can be written that the entropy is a function of the probability.

S = f ( p) (23.16)

Entropy is an extensive function of state, so referring to Fig. 23.4 it can be written:


358 23 Introduction to Entropy and Second Law

Fig. 23.4 Mixing of two gases at the same temperature and adiabatic conditions

S1+2 = S1 + S2 (23.17)

To understand this better, when we consider the overall probability for two events
to happen, the overall probability is the product of the single probabilities. For exam-
ple, we toss a coin and we have 0.5 chance to get one side of the coin. If we want to
calculate the probability associated with two tosses, the probability to have the save
side for two time is equal to p1 · p2 = 0.5 · 0.5. Now, let us consider the probability
density function associated with the two states: The state 1 is the state when the
two gases are still separated, of course, this will be impossible practically, but still
the probability can be calculated. The state 2 is when the molecules are all mixed
in the available space. Therefore, for the state 1 there will be a probability 1 and
an associated probability density function f ( p1 ) and the same for the state 2 with
g( p2 ). The overall probability density function will be:

h( p1+2 ) = f ( p1 ) + g( p2 ) (23.18)
23.2 Statistical Derivation of Entropy 359

where p1 and p2 are probabilities (to not be confused with pressures). According to
what was said previously:

h( p1 · p2 ) = f ( p1 ) + g( p2 ) (23.19)

The only way for this to happen is:

f ( p1 ) = k ln( p1 ) + a; g( p2 ) = k ln( p2 ) + b; h( p1 · p2 ) = k ln( p1 · p2 ) + c


(23.20)

for these reasons:

S = k · ln( p) + a (23.21)

Consider two gases of equal volumes, we refer to 1 as the state of non-mixed


gases and to 2 as the state of mixed gases, the probability of state 2 is higher and it
will be considered to be the maximum possible (100%).
 
p2
S = S2 − S1 = k ln( p2 ) + a − k ln( p1 ) − a = k ln (23.22)
p1

The probability for each of the molecules to be on one side of the system will be
0.5 = 1/2, and the total probability that all the molecules for gas 1 and 2 will be on
one side of the system will be:
  Nd  2Nd
1 1
p1 = (1/2) Nd
= (23.23)
2 2

where, N d , is the number of d molecules. Considering the difference of these state


functions, we will have:
 
1
S2 − S1 = k ln  2Nd = 2Nd kln(2) (23.24)
1
2

In case we consider the Boltzmann constant, the meaning is much general

R 8.314 J/(mol K)
k= = (23.25)
NA 6.022 × 1023 mol−1

where N A is the Avogadro number. Notice that Nd /N A is equal to the number of moles
n d and we derive the expression of the state function entropy as directly proportional
to the Boltzmann constant:

S = 2n d R ln(2) (23.26)
360 23 Introduction to Entropy and Second Law

It is also sometimes written in the literature that for irreversible mixing:


   
V V
S = n a R ln + n b R ln (23.27)
Va Vb

23.3 Additional Remarks on Entropy with Variation


in Moles, i.e., Chemical Reactions

The entropy can be considered as a function of the volume, the moles associated
with each component in a reaction (the stoichiometric coefficient), and the internal
energy. In this way, we have:

S = f (U, V, n γ ) (23.28)

and for this reason, doing the total differential with respect to these variables and
considering γ = a, b, c, . . ., etc. where a, b, c, etc. are the different species present
in the mixture, we have:
     
ϑS ϑS ϑS
dS = ϑU + dV + ϑn a (23.29)
ϑU V n γ ϑ V U nγ ϑn a V,U,n b,c......

where n b,c... indicates that all the other species b, c, d, etc. are constant except for the
specie n a associated with the a component.
Let us write the more clear relation:

n γ  = n b,c......
nγ = na (23.30)

Following the definitions (Callen 1985; Wightman 1979):

S = f (U, V ) = Entropy
ϑU
T = = Temperature
ϑS
ϑU
P=− = Pressure (23.31)
ϑV
Notice that these relations were derived also previously. Equation (23.29) will be
written as:

dU p μγ
dS = + dV − dn γ (23.32)
T T γ
T
23.3 Additional Remarks on Entropy with Variation … 361

where the chemical potential is defined by


 
ϑS
μγ = −T (23.33)
ϑn γ U ·V n γ 

Because we have from previously:

dU + pdV
dS = (23.34)
T
we substitute this in the definition of chemical potential, Eq. (23.33) and we remember
that the enthalpy is:

H = U + pV, (23.35)

the Helmholtz free energy:

F = H −TS (23.36)

and the Gibbs free energy

G = H −TS (23.37)

In this way, we will have derived that:


     
ϑH ϑF ϑG
μγ = = = (23.38)
ϑn γ S Pn γ  ϑn γ T V n γ  ϑn γ T Pn γ 

23.4 Clausius Statement and Kelvin–Planck Statement


of the Second Law

The following are two commonly referred statements of the second law of the ther-
modynamics which can be found in the literature. The Clausius statement and the
one given by Kelvin–Planck:
Clausius: “It is impossible for any system to operate in such a way that the sole
result would be an energy transfer by heat from a cooler to a hotter body.”
Kelvin–Planck: “It is impossible for any system to operate in a thermodynamic
cycle and deliver a net amount of energy by work to its surroundings while receiving
energy by heat transfer from a single thermal reservoir.” (Fig. 23.5).

Wcycle ≤ 0 (for a cycle and a single reserve) (23.39)


362 23 Introduction to Entropy and Second Law

Fig. 23.5 Simple


representation of the
Kelvin–Planck statement of
the second law

Fig. 23.6 Example of


combined thermodynamic
systems

The configurations could be different, and it would be important in this case to


define precisely the boundaries for the systems considered.
The Clausius inequality is expressed as:
 
∂Q
≤0 (23.40)
T boundary

With reference to Fig. 23.6, since we have:


 
∂ Q ∂Q
= (23.41)
Tres T

The energy balance related to the combined cycle would be:

dE c = ∂ Q  − ∂ Wc (23.42)

and by consequence:
 
∂Q
∂ Wc = Tres − dE c (23.43)
T

Integrating over the cycle, to work produced will be:


23.4 Clausius Statement and Kelvin–Planck Statement of the Second Law 363
 
∂Q
Wc = Tres − dE c (23.44)
T

23.5 Entropy Balances

Entropy can be treated as any other entity, and therefore, it is possible to proceed
with the related balances:

ϑ̇ S(control volume)
Ṡ(in) + Ṡ(gen) = Ṡ(out) + (23.45)
ϑθ
and more specifically, the balance on a control volume will be:
   
 Q̇   Q̇  ϑ S(control volume)
 + ṁs + ṡ(gen) =  + ṁs + (23.46)
T  T  ϑθ
in in out out

where the first term on the left and on the right is related to the flux of thermal
entropy, then we have a flux of entropy due to the convection, and finally, we have a
generation of entropy.
For closed systems, the entropy balance would be as follows:

Q̇ ϑ S(closed system)
+ Ṡ(gen) = (23.47)
T ϑθ
In case the heat and the entropy are specific on the unit mass:
q
+ s(gen) = s(closed system) (23.48)
T
In case we consider that the observation time is infinitesimal:

ϑq
+ ϑs(gen) = ϑs(closed system) (23.49)
T
and for open and stationary systems:

Q̇ ·
+ ms + Ṡ(gen) = ṁs (23.50)
T in out

While for isolated systems, the difference in entropy is equal to the entropy which
is generated within the system:

S(isolated system) = S(gen · Isolated system) > 0


E = 0
364 23 Introduction to Entropy and Second Law

References

Callen, H. (1985). Thermodynamics and an introduction to thermostatistics (2nd ed.). New York:
Wiley.
Carnot, N. L. S. (1897). Reflections on the motive power of heat. London: Wiley.
Wightman, A. (1979). Convexity and the notion of equilibrium state in thermodynamics and statis-
tical mechanics. In Convexity in the theory of lattice gases. Princeton U. Press.
Chapter 24
Thermodynamics in Chemical Reactions
Engineering

For example, if the system contains two masses of the same


substance, not in contact, nor connected by other masses
consisting of or containing the same substance or its
components, an infinitesimal increase of the one mass with an
equal decrease of the other is not to be considered as a possible
variation in the state of the system.
On the Equilibrium of Heterogeneous Substances. J. Willard
Gibbs 1878.

24.1 Reaction Rate and Its Dependence on Temperature

Many processes are driven by equilibrium reactions and for this reason thermo-
dynamics is particularly useful in this kind of systems. Consider the reaction of
equilibrium:

k1 [A][B] = k−1 [C][D] (24.1)

where the terms in square brackets refer to the concentration of the different com-
ponents while the constants k1 and k−1 are the forward and the backward chemical
reaction constants.
The equilibrium constant for the reaction (24.1) is expressed as:

[C][D] k1
= Kc = (24.2)
[A][B] k−1

The van’t Hoff equation gives the relation between the equilibrium constant K c and
the temperature, however, before introducing that equation, we will look here at its
derivation.

© Springer Nature Switzerland AG 2019 365


C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_24
366 24 Thermodynamics in Chemical Reactions Engineering

24.2 Derivation of the van’t Hoff Equation

From the definition of the Gibbs free energy equation of state:

dG = −S dT + V dP (24.3)

at constant pressure, this will be written as:


 
ϑG
= −S (24.4)
ϑT P

remembering the definition of the Gibbs free energy:

G = H −T S (24.5)

The following expression is derived:


 
ϑG
G = H +T (24.6)
ϑT P

dividing by the temperature T:


 
G H ϑG
= + (24.7)
T T ϑT P

G
Differentiating the term T
with respect to the temperature at constant pressure, we
have:
    
ϑ GT G 1 ϑG
=− 2 + (24.8)
ϑT T T ϑT P
P

Putting in evidence the term 1/T and considering Eq. (24.7), Eq. (24.8) will be:
  
ϑ GT H
=− 2 (24.9)
ϑT T
P

Considering that:
   
ϑ(G/T ) ϑ(G/T )
= −T 2
(24.10)
ϑ(1/T ) P ϑT P

we have
24.2 Derivation of the van’t Hoff Equation 367
 
ϑ(G/T )
=H (24.11)
ϑ(1/T ) P

or taking into account, the change in Gibbs free energy


 
ϑ(G/T )
= H (24.12)
ϑ(1/T ) P

In case, we have standard conditions:


  
ϑ G 0 /T
= H 0 (24.13)
ϑ(1/T )
P

Substituting into Eq. (24.13), the expression for the standard variation of the Gibbs
free energy:

G 0 = −RT ln K eq (24.14)

and dividing by (–R):


 
ϑ ln K eq
= −H 0 /R (24.15)
ϑ(1/T ) P

Now consider the following reaction

A+B ↔C (24.16)

and the case of gaseous species reacting, we have:


nC
[C]
Kc = = V
nA nB (24.17)
[A][B] V V

and considering the ideal gas law:


nC P xC P
Kc = n RT
= RT
(24.18)
nA P
n RT
· nn RT
BP
· RT
xA P xB P
RT

pC
Kc = (RT )−(vProd. −vreag. ) (24.19)
p A · pC

Finally:

K c = K P (RT )−(vProd. −vreag. ) (24.20)


368 24 Thermodynamics in Chemical Reactions Engineering

For the reaction of Eq. (24.16),

K c = K P · RT (24.21)

and for standard conditions:

K P0 = K C0 · (RT )v (24.22)

where ν is the difference between the moles of products and the moles of reagents.
This will be equal to:

ln K P0 = ln K C0 + v(ln R) + v(ln T ) (24.23)

24.3 Derivation of the Arrhenius Equation

Differentiating with respect to the temperature and at constant pressure, we have:


   
ϑ ln K P0 ϑ ln K C0 v
= + (24.24)
ϑT P ϑT P T

and therefore:
 
ϑ ln K P0 H 0
= (24.25)
ϑT P RT 2

with:
 
ϑ ln K C0 H 0 v
= − (24.26)
ϑT P RT 2 T

this is also equal to:


 
ϑ ln K C0 H 0 v H 0 − v RT
= − = (24.27)
ϑT P RT 2 T RT 2

Remember that the standard internal energy is related to the enthalpy by:

U 0 = H 0 − v · RT (24.28)

We have finally:
 
ϑ ln K C0 U 0
= (24.29)
ϑT P RT 2
24.3 Derivation of the Arrhenius Equation 369

To simplify, we will use the total differential here, however, it should be noted that we
are just taking into account the variation of the reaction rate constants as a function
of the temperature. If we consider the equilibrium reaction as a sum of the forward
and backward reactions and the reaction rate constants as k1 and k−1 , then:

d(ln k1 ) d(ln k−1 ) U 0


− = (24.30)
dT dT RT 2
Consider that the variation of internal energy is the difference between the energies
related to the forward and the backward reaction:

E 1 − E −1 = U 0 (24.31)

this will give:

d(ln k1 ) E1
= (24.32)
dT RT 2
and

d(ln k−1 ) E −1
= (24.33)
dT RT 2
Generally, the relation between the reaction rate constant, the temperature, and the
internal energy (considered for a particular direction of the reaction) is written as
follows. It should be noticed that the pressure is considered constant:

d(ln k) E
= (24.34)
dT RT 2
which is written as:

E
ln(k) = − + const (24.35)
RT
and finally, we have derived the relation between the reaction rate constant and the
temperature at a particular pressure (Arrhenius Equation):

k = A · e− RT
E
(24.36)
370 24 Thermodynamics in Chemical Reactions Engineering

24.4 Chemical Potential

In case, we have the van’t Hoff equation with two parameters:


E0
k = A T m · e− RT (24.37)

Taking the logarithm on both sides:

E0
ln k = ln A + m ln T − (24.38)
RT
considering also the differential on both sides, we have the van’t Hoff commonly
written form:

d ln k m E0 E 0 + m RT
= − 2
= (24.39)
dT T RT RT 2
and Arrhenius form:

d ln k Ea
= (24.40)
dT RT 2
With E a = E 0 + m RT . And from Eq. (24.36), we remember:

Ea
ln(k) = ln(A) − (24.41)
RT
From Eq. (24.39):

E0
ln(k) = ln(A) − m − (24.42)
RT
Now, imposing and also remembering Eq. (24.38):

ln(A) − m = ln A + m ln T (24.43)

we have:

ln(A) = ln A + m ln(T ) + m (24.44)

and finally:

A = A T m em (24.45)

We have seen previously that for a three-dimensional space we have that the average
kinetic energy for one molecule is:
24.4 Chemical Potential 371

1 3
m(tot) · v2 = N K B T (24.46)
2 2
Recall the Boltzmann constant:

R
KB = ≈ 1.38 · 10−23 J/K (24.47)
NA

and N is the total number of molecules considered (not moles). In case we consider
only one direction (not three) and one molecule, we have.

1 1
ε= m(tot) · v2 = K B T (24.48)
2 2

24.5 Statistical Implications of the Boltzmann Distribution


Law

If we consider the Boltzmann distribution law, we have that the number of molecules
having an energy state εi is:
ε
− K iT
ni = λ · e B (24.49)

lambda is the absolute activity:


μ
λ = e KB T (24.50)

and therefore, μ is the chemical potential:

μ = K B T ln(λ) (24.51)

Consider a probability density function for the values of the velocities of one molecule
with respect to the center axis:

1 − u2
p(u) =  1/2 · e 2σ 2 (24.52)
2π σ 2

A representation of the probability density function for the values of the velocity of
a particle is given in Fig. 24.1.
From the theory of statistics, we have that the variance is:

T
σ 2 = u2 = KB · (24.53)
m
372 24 Thermodynamics in Chemical Reactions Engineering

Fig. 24.1 Probability


density function for the
velocity values of a molecule
in a box and one direction

where m is the mass of the particle/molecule. Equation (24.53) is valid since the mean
value of the velocity is zero and then the variance becomes the mean of the squared
values of the velocity, which we have in the equation written before. Therefore:

m 1/2 mu 2
− 2K
p(u) = · e BT (24.54)
(2π K B T )1/2

and:

m 1/2 − εx
p(u) = · e KB T (24.55)
(2π K B T ) 1/2

The probability that one molecule has a velocity between u and u + d(u) is:

m 1/2 − εx
p(u)d(u) = · e KB T du (24.56)
(2π K B T ) 1/2

From the expression for the kinetic energy of a molecule:

1  2 
ε= m u  (24.57)
2
we have:
 1/2

u= (24.58)
m

and:

1 1 2 1
du =  2ε 1/2 dε = dε (24.59)
2 m (2εm)1/2
m
24.5 Statistical Implications of the Boltzmann Distribution Law 373

We remember:

m 1/2 − εx
p(u)d(u) = · e KB T du (24.60)
(2π K B T ) 1/2

and Eq. (24.59), and we obtain:

m 1/2 1 − εx
p(u)d(u) = · e KB T dε (24.61)
(2π K B T ) (2ε · m)
1/2 1/2

which gives:

1 − εx
p(u)d(u) = · e KB T dε (24.62)
2(π · ε · K B T ) 1/2

and this will give:

dN x 1 − εx
= 2 p(u)d(u) = · e KB T dεx (24.63)
N (π · εx · K B T ) 1/2

which is the fraction of molecules having energies from εx and εx + dεx in the
direction x.
In two dimensions, this will give:

dN 1 − εx − y
ε
= 1/2 · e KB T e KB T dεx dε y (24.64)
N π · εx · ε y · K B T

Writing: ε = εx + ε y and integrating between 0 and ε:

dN 1 − ε
= · e KB T dε (24.65)
N (K B T )

the fraction of molecules having energy larger than ε∗ is obtained by integrating


between ε∗ and infinite giving:

− Kε
f∗ = e BT (24.66)

If we have two states of activation 1 and 2 and:


ε
− K iT
n i = λe B (24.67)

This will give:

− ( K2
ε −ε1 )
n 2 /n 1 = e BT (24.68)
374 24 Thermodynamics in Chemical Reactions Engineering

and finally:

n 2 /n 1 = e− RT
Ea
(24.69)

Considering:

ε∗ = ε2 − ε1 (24.70)

and taking into account:

E a = Na ε ∗ (24.71)

where Na is the Avogadro number.


Chapter 25
Some Parameters and Properties
of Biomass Fuels

Life is movement.
(Aristotle, 4th century BC), as cited by Brügemann and
Gerds-Ploeger (2013).

25.1 Some Useful Figures on Biomass Feedstock

To have the fundamental understanding of biofuels engineering and technology, con-


cepts like energy content, calorific values, and related properties need to be under-
stood for a necessary assessment of the biomass feedstock first and the produced
biofuel later.
The following part will give some more information on what is the composition
of common biomass with particular focus on crops and plants in general; some
properties will be given in terms of composition and energy content. Since this section
of the manuscript is mainly consisting of support material, data on the properties of
diverse kinds of biomass are given here in the form of tables, and these figures are
taken from the literature.
If the intention is to utilize energy crops as one mean for energy conversion, the
first step would be to assess how much of crops we are actually producing per unit
time and per unit of land. To be remembered that land utilization is actually one
of the biggest problems in the production of energy crops to be used in energy and
production of biofuels. We have to remember also that in the world, there are large
portions of population who actually do not have enough food. Fortunately, we can
think about third-generation biorefinery for the future.
A common parameter used to assess the production of biomass for different
species is the Net Primary Production (NPP), value and the Energy Storage Capacity
(ESC), of biomass (Gholz 1982).
It is estimated that the global NPP is around 100 Gton of carbon per year (Ksenzhek
and Volkov 1998). Considering that the specific energy storage capacity is around
41.3 kJ/g of fixed carbon, the energy stored in photosynthetic biomass is around

© Springer Nature Switzerland AG 2019 375


C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_25
376 25 Some Parameters and Properties of Biomass Fuels

4 × 1021 J/year while the total solar energy on earth is estimated around 2.75 ×
1024 J/year.
Concerning the chemical composition of wood and biomass, a suitable description
of the main composition and characteristics of biomass in terms of lignin for instance
can be found in the literature together with an approximate evaluation of the main
constituents (Demirbaş 2005). This gives:
• 75% carbohydrates, cellulose and hemicellulose;
• 25% lignin;
• lower calorific value of about 18 GJ/ton at 100% dry matter. As comparison,
conventional oils have 42–44 GJ/ton.
As an example here, the composition of common types of biomass is given in
Table 25.1 in terms of cellulose, hemicellulose, lignin, and ash.
Plants can be cultivated for the purpose of energy conversion; as a matter of fact,
this solution is being implemented in diverse countries in the world and Europe
(Jezierska-Thöle et al. 2016). The biomass produced is assessed on the basis of its
properties and also on the basis of how much of the same feedstock is produced per
year and per hectare (ha) (Jankowski et al. 2016). Another estimation of how efficient
is the biomass production in terms of incoming radiative energy can be done on the
basis of the so-called radiation use efficiency (RUE); this gives a good estimation of
how effective is a determinate kind of plant or organism in utilizing light energy. An
example of RUE estimation can be found in (Chimonyo et al. 2018). It is therefore
very important to assess the biomass in terms of radiation use efficiency or related
quantities.
Let us assume here that the produced biomass will be suitable for combustion
in power plants. Before the utilization of the feedstock, drying is strictly necessary
when moisture content is higher than 65% according to industrial recommendations.
However, less than 50% is suggested in practice. In fresh wood, the moisture content

Table 25.1 Composition of common types of biomass, percentages in weight


Cellulose Hemicellulose Lignin Ash Other
Saccharum bagasse 48 22 20 6 4
Wheat 54 26–30 16–18 7–8 6.8
Bamboo 34–36 16–16 26–28 3 9.6
Pine 42 24 27 0.2 7.6
Poplar 70 23 4 4 28
Eucalyptus 49 15 28 0.4 34
Miscanthus 44 24 17 1.5 4
Phalaris 28 22 14 8
Lucerne 25 23 12 6
Betula 41 33 21 <1
Data from Rayaprolu (2009)
25.1 Some Useful Figures on Biomass Feedstock 377

is around 30–50%, and it reduces to 18–25% after one-year drying (Serup et al.
2005).
The ash content in wood is around 2.5%, and there is low amount of sulfur.
The moisture content is a function of wood age and type. These information are
obtained by communicating with wood-driven power plant managers in Finland.
The suggestion is in this case to have the power plant quite near the origin of the
feedstock and max at around 100 km.
Energy crops can be classified as “above ground” and “below ground” or for
instance they can give a high-energy-specific product, for instance ethanol from
sugarcane or oilseed rape and can be classified accordingly.
Common values for the radiation use efficiencies (RUE) and energy conversion
efficiencies are given as follows (Pessarakli 2005):

• Sugarcane, for instance with a 18-month harvest interval, has a RUE of


2.3 g/(MJ m2 ).
• Sorghum has a RUE of approximately 3.2 g/(MJ m2 ).
• Maize has a RUE of approximately 2.5 g/(MJ m2 ).
• Miscanthus has a RUE of approximately 1.4 g/(MJ m2 ).
• Miscanthus sacchariflorus has a RUE of approximately 3.6 g/(MJ m2 ).
• Willow has a RUE of approximately 1.4 g/(MJ m2 ).
• Poplar has a RUE of approximately 1.0 g/(MJ m2 ).
• Eucalyptus has an ECE of approximately 5.8%.
• Populus has a RUE of approximately 0.5 g/(MJ m2 ).
• Pinus has a RUE of approximately 1.3 g/(MJ m2 ).
• Cannabis has a RUE of approximately 2.26 g/(MJ m2 ).
• Rice has a RUE of approximately 0.93 g/(MJ of total radiation).
• Potato has a RUE of approximately 1.6 g/(MJ of total radiation).
• Wheat has a RUE of approximately 1.3 g/(MJ of total radiation).
• Soybean has a RUE of approximately 1.20 g/(MJ of total radiation).
• Peanut has a RUE of approximately 0.98–2.24 g/(MJ of total radiation).

Table 25.2 shows some values of proximate analysis for peat, wood, paper, and coal
as comparison.

Table 25.2 Proximate values for wood, paper, peat, and coal. A comparison
Volatiles Moisture Carbon Ash Fuel HHV
(wt%) (wt%) (wt%) (wt%) ratio (MJ/kg)
Peat 65 7 20 8 0.30 22
Wood 85 6 8 1 0.10 19
Paper 75 4 11 10 0.15 13
Coal 30 5 45 20 1.5 26
Data taken from Rayaprolu (2009)
378 25 Some Parameters and Properties of Biomass Fuels

When biomass is utilized for energy conversion, for instance when wood or wood
chips are used for combustion in a power plant. The most important properties con-
sidered for energy recovery are as follows:

• moisture content;
• density;
• heating value;
• particle size distribution;
• ash content and properties;
• chemical composition;
• amount of volatiles.

And the main kind of analyses usually done on the original feedstock is the ultimate
and proximate analyses. The lower heating value (LHV) and the higher heating value
(HHV) are evaluated firstly in a calorimetric bomb as we have seen in the first part of
this book. The LHV and HHV are also encountered in the literature as net calorific
value (NCV) and gross calorific value (GCV).
Figure 25.1 is a simple representation of how the calorific value is determined.
When the HHV is determined in a calorimetric bomb, it should be noticed that
if water is contained in the original biomass feedstock, it will be condensed also
in the combustion chamber. Additionally, any water that is also formed during the
combustion process will condense in the same chamber (see part 1 of the manuscript).
The net calorific value is evaluated from the GCV taking into consideration the
moisture and the hydrogen content. The GCV and the NCV are expressed also by
empirical formulas in the literature. For instance, an expression for the gross calorific
value (GCV) is given as follows (Reed and Golden 1988):

GCV = 34.1 C + 132.2 H + 6.8 S − 1.53 A − 12.0 (O + N) kJ/g (25.1)

GCV = [146.6 C + 568.8 H + 29.4 S − 6.6 A − 51.5 (O + N)] × 102 Btu/lb


(25.2)

where C, H, S, A, O, and N are expressed in wt%/100, and they refer to carbon,


hydrogen, sulfur, ash, oxygen, and nitrogen present in the fuel. The latent heat of
vaporization (at 25 °C) of water is reported as 2.444 kJ/g, while m and H2 are

Fig. 25.1 A simple scheme


of the gross calorific value
determination
25.1 Some Useful Figures on Biomass Feedstock 379

the moisture and hydrogen in wt%/100. Equation (25.1) is also known as Milne’s
formula, and it is valid for GCV on dry basis. The HHV and the LHV can be expressed
on the basis of different parameters, for instance on dry basis (d.b.), dry and ash free
basis (d.a.f.), and as received (a.r.). The net calorific value (NCV) is calculated usually
from the GCV as follows:

NCV = GCV − (m w + 9H ) · latent heat (25.3)

Therefore, it is possible to write (Channiwala and Parikh 2002):

HHVar = HHVdry · (1 − xw ) (25.4)

HHVdry = HHVdaf · (1 − xash ) (25.5)

LHVdry = HHVdry − 2.443 · 8.936 · xH (25.6)

LHVar = LHVdry (1 − xw ) − 2.443 · xw (25.7)

LHVar = HHVar − 2.443 · {8.936 · xH (1 − xw ) + xw } (25.8)

In Eqs. (25.5)–(25.8), xw and xH are the fraction in weight of water and hydrogen,
respectively.
Sometimes in the literature, it is also possible to find the correlation in the fol-
lowing form (Van Loo and Koppejan 2008):
 
MJ
NCV = GCVdry (1 − xw ) − 2.443 · xw − 2.443 · xH · 8.936(1 − xw ) , w.b
kg
(25.9)

Table 25.3 gives examples of ultimate and proximate analyses for additional kinds
of biomass feedstock.

25.2 Overall Energy (Heat) Balances

25.2.1 Design Variables Affecting Thermochemical Processes

Biomass can contain a large number of compounds; this depends on what kind
of biomass is taken into consideration and what is the process where the biomass
feedstock is utilized. Table 25.4 is a very general example of what are some effects
of biomass compounds on thermochemical processes. These of course are just a
Table 25.3 Proximate and ultimate analysis of some biomass feedstock with HHV values
380

Material Water (wt%) FC (wt%, VM (wt%, Ash (wt%, C (wt%, H (wt%, O (wt%, N (wt%, HHV
d.b.) d.b.) d.b.) d.b.) d.b.) d.b.) d.b.) (MJ/kg, d.b.)
Coconut 8.27 22.1 77.19 0.71 50.22 5.70 43.37 0.00 20.50
shell
Akhrot shell 8.76 18.78 79.98 1.20 49.81 5.64 42.94 0.41 20.01
Groundnut 8.10 21.60 72.70 5.70 48.59 5.64 39.49 0.58 19.85
shell
Eucalyptus 16.40 21.30 75.35 3.35 46.04 5.82 44.49 0.30 18.64
wood
Bamboo 11.50 11.24 86.80 1.95 48.76 6.32 42.77 0.20 20.55
wood
Mango 14.10 11.36 85.64 2.98 46.24 6.08 44.42 0.28 19.17
wood
Jujube wood 12.50 14.14 83.63 2.32 47.63 6.12 43.78 0.15 19.77
Plywood 13.79 15.77 82.14 2.09 48.13 5.87 42.46 1.45 18.95
Block wood 12.16 14.59 83.32 2.09 46.90 6.07 43.09 0.95 18.26
Millet straw 5.06 16.45 78.28 5.27 43.71 5.85 45.16 0.01 18.05
Wheat straw 8.87 10.98 82.12 6.90 42.95 5.35 46.99 0.00 17.99
Jowar straw 7.50 15.15 75.97 8.88 42.10 5.60 43.38 0.04 17.95
Rice straw 8.10 13.91 65.70 20.38 35.68 4.62 39.14 0.28 14.85
(continued)
25 Some Parameters and Properties of Biomass Fuels
Table 25.3 (continued)
Material Water (wt%) FC (wt%, VM (wt%, Ash (wt%, C (wt%, H (wt%, O (wt%, N (wt%, HHV
d.b.) d.b.) d.b.) d.b.) d.b.) d.b.) d.b.) (MJ/kg, d.b.)
Rice husk 8.47 16.95 61.81 21.24 38.50 5.20 34.61 0.45 14.69
bran
Coconut coir 19.77 28.82 66.02 5.16 43.36 4.98 44.87 1.63 18.07
pith
Sugarcane 51.01 13.15 83.66 3.20 45.48 5.96 45.21 0.15 18.73
bagasse
Millet grain 5.20 14.50 77.10 8.40 40.56 5.24 45.50 0.40 15.21
waste
25.2 Overall Energy (Heat) Balances

Cotton gin 4.99 14.97 84.41 1.61 42.66 6.05 49.50 0.18 17.49
waste
Adapted from Channiwala and Parikh (2002)
381
382 25 Some Parameters and Properties of Biomass Fuels

Table 25.4 Properties of biomass and their effect on thermochemical processes


Property Effect on
Moisture Heating value, ignition, plant design, power plant
efficiency (Bahadori et al. 2014)
HHV, LHV Plant design, mass flows, volumes, suitability of fuel,
average burning rate, maximum solid temperature
(Yang et al. 2005)
Volatiles Plant design, mass flows, volumes, suitability of fuel.
Fluidization (Saastamoinen 2015)
Ash content Fuel suitability, disposal, safety, deposition, efficiency,
safety. Technological and environmental problems
(Vassilev et al. 2017)
Bulk density Plant design, transport, handling. Sometimes biomass
densification is necessary (Manickam et al. 2006)
Particles density Plant design, transport, handling, efficiency
Particles size and distribution Plant design, transport, handling, efficiency
Content of carbon, hydrogen, oxygen HHV, LHV, Plant design and operations. Oxygen is
decreasing the ignition time for engine fuels,
emissions are reduced, however NOx and fuel
consumption may be increased (Song et al. 2016)
Cl content Emission control processes, corrosion, ash melting
temperature. Composition of ash deposits (Liu et al.
2000)
N, content NOx emissions, process design
S, content SO2 emissions, flue gas treatment. Composition of ash
deposits (Liu et al. 2000)
F, content HF emissions and corrosion
K, content Corrosion, emission, ash processing and melting
temperature
Na, content Corrosion, ash melting temperature, ash composition
Mg, content Increases ash melting temperature, ash processing
Ca, content Increases ash melting temperature, ash processing
P, content Ash processing
Heavy metals Safety, emissions, ash processing

minimal part of the effects that the compounds could have on a process (for instance,
a biomass power plant) and therefore the effects given are not sufficient, but they
should be considered as descriptive of their importance.
Let us consider here a common power plant which could utilize biomass as fuel,
for instance, woodchips. The heat produced from the combustion of the fuel will
be utilized to heat up water to the required conditions necessary to run a turbine, in
case electricity is produced, or when district heating is in place. The main variables
considered when the design of the facilities is to be performed are:
25.2 Overall Energy (Heat) Balances 383

• moisture content;
• calorific value (here some empirical formulas can be seen for GCV and NCV);
• temperature, residence time;
• stoichiometry;
• mixing.
The moisture amount in the biomass varies largely on the basis of the biomass type and
biomass storage. Values for biomass moisture content can be found in the literature
along with the effect of moisture on biomass treatment processes (Motta et al. 2018).
Commonly, it is necessary that biomass undergoes the process of drying before it
can be used in thermochemical processes; this is because the moisture content can
reduce the adiabatic combustion temperature, and for this reason, the residence time
in the combustion chamber will be increased. One typical example is the drying of
black liquor in pulp and paper industry to utilize it in recovery boilers. Nevertheless,
some options have been proposed in the literature to utilize black liquor directly
in gasification processes; for instance, alternative methods like supercritical water
gasification (SCWG), where it would be possible to utilize the feedstock directly into
the reactor with no need of drying (De Blasio et al. 2015).
Typical moisture and ash data on wood biomass are given in the literature (Van
Loo and Koppejan 2008), and these are typical recommendations also given by
power plants’ managers. These parameters are affecting heavily the ignition front
propagation speed and the bed temperature for biomass-fired power plants (Arce
et al. 2013). As mentioned also previously and as a summary here:

• 30–50% is the moisture content in fresh wood; this reduces to 18–25% after one-
year drying.
• Drying is strictly necessary when moisture is higher than 65%; however, less than
50% is the suggested percentage.
• The ash content in wood is normally below 2.5%.
• The nitrogen content is very low.
• Wood contains very low amount of sulfur.
• Moisture is a function of wood age and type.
• Transportation of wood is not economical for distances longer than 100 km.

To give one example, Table 25.5 gives a comparison of the moisture, ash, volatile
matter, HHV, fixed carbon, and elemental analysis for diverse kind of trees.

25.2.2 Additional Data on Moisture and Ash Content of Fuels

Additional data are reported here on the moisture and ash content of different types
of biomass. As mentioned, these are two parameters which affect largely the HHV
and LHV; some values are reported in Table 25.6 where these values are given for
diverse feedstock kinds to make a comparison.
384

Table 25.5 Moisture, ash, volatile matter, HHV, fixed carbon and elemental analysis for diverse kind of trees
Specie Clone H2 O (%) Ash (%) VM (%) FC (%) HHV (J/g) C H (%) S (%) N (%) O (%)
Willow Olof 6.406 1.174 76.662 15.757 18.46 45.39 6.37 0.16 0.16 47.91
Tordis 7.11 1.19 75.33 16.37 18.32 45.49 6.36 0.16 0.19 47.80
Poplar 1214 6.45 2.18 75.99 15.39 18.19 45.11 6.26 0.16 0.15 48.33
AF2 6.02 1.70 76.90 15.37 18.42 45.58 6.26 0.15 0.15 47.86
Beaupré 6.19 1.86 76.63 15.32 18.39 45.70 6.32 0.14 0.11 47.73
Chestnut 6.07 0.92 74.72 18.30 18.00 46.04 6.18 0.19 0.25 47.35
Maritime pine 6.95 0.60 76.58 15.98 19.37 47.78 6.26 0.65 0.50 44.82
Adapted from Álvarez-Álvarez et al. (2018)
25 Some Parameters and Properties of Biomass Fuels
25.2 Overall Energy (Heat) Balances 385

25.2.3 Properties of Agrofuels and Typical Air Requirements

Some of the properties related to biofuels’ feedstock derived from agriculture are
given in Table 25.7
Typical air requirements for agrofuel-based HHV are given in Table 25.8.
Peat represents the part of biomass which is not decomposed completely, and by
some definitions, it should be composed by around 30% of dead organic material
(Clarke and Rieley 2010). There has been some discussion on this particular kind
of biomass and if we should consider Peat as a renewable source of energy. Peat is
considered to be a form of renewable biomass by some countries; however, there is
still a discussion about if considering peat as renewable source or not. In Sweden and
in Finland, wood and peat are the only indigenous fuels according to some literature
(Rayaprolu 2009) and in Nordic countries a large part of the ground could be utilized
for peat production (Virtanen et al. 2003). The content of carbon sequestrated within

Table 25.6 Heating values, moisture and ash content of some examples of biomass
HHV (MJ/kg) LHV (MJ/kg) Moisture (%) Ash (fraction in
mass)
Charcoal 30.44 28.33 6.361 0.0231
Firewood 19.04 15.6 16.041 0.008525
Croton husk 17.48 15.39 10.482 0.101
briquette
Carbonized croton 27.78 25.3 8.200 0.0745
husk
Data from Turpeinen (2016)

Table 25.7 Different Parameter Range


property values for agrofuels
Theoretical air requirement (kg/kcal) 0.00122
HHV (kcal/kg) 4500–5500
Fouling-factor 1.005–1.015
Theoretical CO2 (%) 20.2–20.75
Moisture (%) 10
H2 (dry and ash free) (%) 5.6–5.8
VM (d.a.f.) (%) 72–90
where VM is the volatile matter and d.a.f. means dry and ash free
Adapted from Rayaprolu (2009)

Table 25.8 Typical air Coal Agrofuel


requirements for agrofuel
expressed as kg/kcal 0.00131–0.00136 0.00122–0.00124
Data from Rayaprolu (2009)
386 25 Some Parameters and Properties of Biomass Fuels

peat is important considering regions like Finland. In fact, it was estimated that the
total carbon content in Finnish peat was 5.5 Pg in 1950, and the total carbon seques-
tration into Finnish peat has increased from 2.2 to 3.6 Tg ha−1 in 2010 (Minkkinen
et al. 2002). Peat can be of course processed; one of the technologies used for this
purpose is the wet carbonization process (Lau et al. 1987).
The main characteristics of peat are described as follows in Table 25.9; peat has
high amount of moisture, 80–95%, and can be reduced to 35–55% when drying is
applied.
The ultimate analysis of this fuel is described in Table 25.10.
Concerning the agrofuels, in this case there is a great variety of them and they can
vary very much in composition. Additionally, they are also seasonal and their growth
is very much affected by the climate in which they are produced. Furthermore, the
degradation of this kind of feedstock is quite fast. It is also quite obvious that if a
power plant should be run on agrofuels, then the power output cannot be very large
but in a range that could be 5–30 MW.

Table 25.9 Main properties Peat property Value (approximately)


of peat
Moisture 90% of water (Lau et al. 1987). 10%
(briquette), 50% (mulled), 25% (air dried)
(Rayaprolu 2009)
Density (kg/m3 ) 240–400 (air dried), 480–960 (Briquette)
Heating value 19.8 MJ/kg (dry basis) (Lau et al. 1987)
Ash (%) 2–10
Volatiles (%) 55–70
Carbon (%) 50–60
Hydrogen (%) 6
Oxygen (%) 30–40
Sulfur (%) 0.3–0.5
Nitrogen 1.2–1.5
Data mostly taken from Rayaprolu (2009)

Table 25.10 Ultimate Element %


analysis on dry and ash free
basis of peat N2 1.2–1.5
S 0.3–0.5
O2 30.0–40.0
H2 5.5–7.0
C 53.0–63.0
Adapted from Rayaprolu (2009)
25.2 Overall Energy (Heat) Balances 387

Agrofuels have usually a gross calorific value which is higher than 13,500 kcal/kg
and for this value, the air requirement is less than 1.21 kg/1000 kcal of GCV (Gujarat
Pollution Control Board, Paryavaran Bhavan 2014).
It is also quite reasonable that agrofuels are suited for stoker boilers and bubbling
beds but not for bubbling fluidized bed boilers, and this has to do with their density
and consistency.
They are efficient for ignition and combustion, with volatile matter (VM), of
70–90% (Rayaprolu 2009). The amount of ash is limited in this case, and their
higher amount of oxygen lowers the requirement of air for combustion. The furnace
exit gas temperature is higher than coal-fired boilers when grates are used; this is
due to higher amount of VM, volatile matter.
With respect to the combustion properties of minor agrofuels, Table 25.11 gives
a small set of data.
While, regarding the moisture content and additional parameters of wood-related
parts, Table 25.12 gives some more additional data.
Table 25.13 gives further data of proximate and ultimate analysis conducted on
additional kinds of samples.
While Table 25.14 gives the HHV and LHV for selected agricultural residues,
crops, forest and urban residues.
As a comparison to the previous data given, a list of thermodynamic properties is
reported for alkane fuels in Table 25.15.

Table 25.11 Combustion properties of minor agrofuels


Coconut Coffee Cotton Cotton Palm nut Peanut
shell husks seed husks pods shells shells
HHV 4220 3720 4055 3610 3720 4335
(kcal/kg)
LHV 3900 3395 3720 3305 3840 4000
(kcal/kg)
H2 O (%) 10.0 12.0 11.0 12.0 10.0 10.0
Ash (%) 5.0 3.0 3.0 7.0 3.0 2.0
Theoretical 5.2 5.0 5.0 4.45 5.15 5.35
air (kg/kg)
Theoretical 20.4 20.5 20.5 20.5 20.5 20.4
CO2 by
volume %
Adapted from Rayaprolu (2009)
388 25 Some Parameters and Properties of Biomass Fuels

Table 25.12 Moisture content, HHV, LHV, density, and calorific value for wood-related biomass
Moisture HHV LHV Bulk density Calorific
content (MJ/kg, d.b.) (MJ/kg, d.b.) (kg, w.b./m3 ) value
(wt%) (MJ/m3 )
Olive 53.0 22.6 8.5 650 5530
residues
(from
three-phase
production)
Olive 63.0 21.5 6.1 1130 6890
residues
(from
two-phase
production)
Straw (winter 15.0 18.7 14.5 120 1740
wheat)
High- 50.0 19.8 8.0 240 1920
pressure bales
sawdust
Cereals 15.0 18.7 14.5 175 2540
Bark 50.0 20.2 8.2 320 2620
Grass (high- 18.0 18.4 13.7 200 2740
pressure
bales)
Woodchips 50.0 19.8 8.0 350 2800
(softwood)
Woodchips 30.0 19.8 12.2 250 3050
(softwood
predried)
Woodchips 50.0 19.8 8.0 450 3600
(hardwood)
Woodchips 30.0 19.8 12.2 320 3900
(hardwood
predried)
Wood pellets 10.0 19.8 16.4 600 9840
Adapted from Van Loo and Koppejan (2008)

An explanation of the parameters given in Table 25.15 can be found in Table 25.16.
For additional information on thermodynamic data for fuels, the reader can refer
also to additional literature sources (Rumble 2018; Bartok and Sarofim 1991).
25.2 Overall Energy (Heat) Balances 389

Table 25.13 Proximate and ultimate analyses for corn, straw, husks, and wood chips
Corn Corn Corn Straw Straw Rice Wood Bark
stems cobs grain pellets husk chips
Particle 360 385 420 650 1400 740 920 880
density
(kg/m3 )
Ash (%) 2.60 2.40 1.88 6.00 6.10 22.70 0.75 1.81
Bulk 158 174 235 100 790 120 590 430
density
(kg/m3 )
LHV 15.80 16.30 15.20 16.60 16.50 12.90 9.55 6.52
(MJ/kg)
Moisture 9.20 8.00 11.90 8.50 8.50 4.10 41.50 58.00
(%)
VM (%) 67.30 69.10 75.00 61.60 61.40 59.70 47.94 31.79
FC (%) 20.90 20.50 11.20 24.20 24.00 13.50 9.82 8.45
C (wt%) 42.16 45.57 43.72 41.61 37.81 47.62 48.30
H2 6.28 5.93 7.18 6.70 4.68 6.11 6.34
(wt%)
O2 45.70 45.89 45.73 44.40 33.52 44.34 39.76
(wt%)
N2 0.71 0.74 1.24 0.66 0.28 0.65 1.29
(wt%)
S (wt%) 0.01 0.02 0.01 0.01 0.03 0.00 0.00
Ash
compo-
sition
(wt%)
SiO2 32.74 43.00 59.52 97.62 6.45 5.88
Al2 O3 5.05 8.40 2.18 0.01 0.66 0.20
TiO2 0.02 0.40 0.02 0.01 0.02 0.02
Fe2 O3 0.70 5.05 8.76 0.19 2.05 0.33
CaO 2.63 5.10 6.52 0.01 73.05 80.72
MgO 6.83 8.10 2.26 0.59 0.54 0.23
K2 O 29.24 15.30 13.16 0.87 4.50 2.57
Na2 O 1.38 0.90 1.60 0.11 0.17 0.10
P2 O5 4.68 13.00 5.07 0.44 1.90 1.30
SO3 0.12 0.10 0.36 0.25 1.93 0.95
Adapted from Rayaprolu (2009)
390 25 Some Parameters and Properties of Biomass Fuels

Table 25.14 HHV and LHV for selected agricultural residues, crops, forest, and urban residues
HHV (MJ/kg) LHV
Corn stalksa, b, f 17.6–18.5 16.8–18.1
Sugarcane bagassea, b, f 17.3–19.4 17.7–17.9
Wheat strawa, b, f 16.1–18.9 15.1–17.7
Hulls, shells, pruningsb, c 15.8–20.5
Miscanthusf 18.1–19.6 17.8–18.1
Switchgrassa, c, f 18.0–19.1 16.8–18.6
Bamboof 19.0–19.8
Black locusta, f 19.5–19.9 18.5
Eucalyptusa, b, f 19.0–19.6 18.0
Hybrid poplara, c, f 19.0–19.7 17.7
Willowb, c, f 18.6–19.7 16.7–18.4
Hardwoodb, f 18.6–20.7
Softwooda, b, c, d, e, f 18.6–21.1 17.5–20.8
Municipal solid wasteb, f 13.1–19.9 12.0–18.6
Refuse-derived fuelsb, f 15.5–19.9 14.3–18.6
Newspaperb, f 19.7–22.2 18.4–20.7
Corrugated paperb, f 17.3–18.5 17.01
Waxed cartonsb 27.3 25.3
where the apices a, b, c, d, e, and f refer to the following references:
a Energy Efficiency & Renewable Energy Department, USA (2018)
b Jenkins (1993)
c Jenkins et al. (1998)
d Tillman (1978)
e Bushnell et al. (1990)
f ECN, The Netherlands (2018)

25.3 Stoichiometry

When there is complete combustion of hydrocarbons (for instance), CO2 is normally


produced from the carbon present in the biomass fuel. Water, instead, is produced
from the oxidation of the hydrogen molecules of the feedstock. However, when the
combustion is not perfect, there can be formation of CO in the flue gases. In case we
consider a hydrocarbon with general composition Cn Hm , the stoichiometric reaction
of oxidation is:
 m m
Cn Hm + n + O2 → nCO2 + H2 O (25.10)
4 2
However, it is necessary to take into consideration also the nitrogen amount when air
is utilized during combustion; in this case, the reaction is described as Eq. (25.11)
Table 25.15 Thermodynamic properties for selected alkane compounds
Name Sg Tb p0 CPg Cpl Tig HHV LHV h fg a/f Res. Mot. h f0 pc Tc
Methane 0.466 −161 2.21 537 55.54 50.05 510 17.2 120 120 −74.4 45.4 190
Ethane 0.572 −89 1.75 472 51.90 47.51 489 16.1 115 99 −83.8 48.2 305
Propane 0.585 −42 12.8 1.62 2.48 470 50.32 46.33 432 15.7 112 97 −104.7 41.9 370
25.3 Stoichiometry

n-Butane 0.579 0 3.51 1.64 2.42 365 49.51 45.73 386 15.5 94 90 −146.6 37.5 425
Isobutane 0.557 −12 4.94 1.62 2.39 460 49.36 45.58 366 15.5 102 98 −153.5 36 408
n-Pentane 0.626 36 1.06 1.62 2.32 284 49.00 45.34 357 15.3 62 63 −173.5 33.3 470
Isopentane 0.620 28 1.39 1.60 2.28 420 48.91 45.25 342 15.3 93 90 −178.5 33.5 460
n-Hexane 0.659 69 0.34 1.62 2.27 233 48.67 45.1 335 15.2 25 26 −198.7 29.7 507
Isohexane 0.662 50 0.503 1.58 2.20 421 48.45 44.89 305 15.2 104 94 −207.4 30.9 500
n-Heptane 0.684 99 0.11 1.61 2.24 215 48.44 44.93 317 15.2 0 0 −224.2 27.0 540
Triptane 0.690 81 0.23 1.60 2.13 412 48.27 44.76 289 15.2 112 101 29.2 531
n-Octane 0.703 126 0.04 1.61 2.23 206 48.26 44.79 301 15.1 20 17 −250.1 24.6 569
Isooctane 0.692 114 0.12 1.59 2.09 418 48.12 44.65 283 15.1 100 100 −259.2 25.4 544
n-Nonane 0.718 151 0.01 1.61 2.21 48.12 44.69 288 15.1 274.7 22.6 595
n-Decane 0.730 174 0.01 1.61 2.21 48.00 44.6 272 15.1 −300.9 20.9 618
Isodecane 0.768 161 1.61 2.21 15.1 113 92 24.8 623
n-Undecane 0.740 196 1.60 2.21 47.90 44.52 265 15.0 −327.2 19.4 639
(continued)
391
Table 25.15 (continued)
392

Name Sg Tb p0 CPg Cpl Tig HHV LHV h fg a/f Res. Mot. h f0 pc Tc


n-Dodecane 0.749 216 1.60 2.21 47.84 44.57 256 15.0 −350.9 18.0 658
n-Tridecane 0.756 236 1.60 2.21 246 15.0 17.0 676
n-Tetradecane 0.763 253 1.60 2.21 239 15.0 14.2 693
n-Pentadecane 0.768 271 1.60 2.21 232 15.0 15.0 707
n-Hexadecane 0.773 287 1.60 2.21 47.61 44.31 225 15.0 −456.1 13.9 722
n-Heptadecane 0.778 302 1.60 2.21 221 15.0 12.8 733
n-Octadecane 0.782 317 1.60 2.21 47.54 44.26 214 15.0 11.8 748
Adapted from Borman and Ragland (1998)
25 Some Parameters and Properties of Biomass Fuels
25.3 Stoichiometry 393

Table 25.16 Parameters used in Table 25.15


Parameter Dimensions
Sg Specific gravity, density of the component at 20 °C specific to the density of water
at 4 °C. For gases is determined at the boiling point of the liquefied gas
Tb Boiling point temperature at 1 atm in °C
p0 Vapor pressure at 38 °C and 1 atm
CPg Specific heat of gas at 25 °C [kJ/(kg K)]
Cpl Specific heat of liquid at 25 °C [kJ/(kg K)]
Tig Auto-ignition temperature (°C)
HHV kJ/kg
LHV kJ/kg
h fg Heat of vaporization at 1 atm and boiling point temperature (kJ/kg)
a/f Stoichiometric air-to-fuel mass ratio
Res. Research octane number
Mot. Motor octane number
h f0 Enthalpy of formation at 25 °C (kJ/g mol)
pc Critical point pressure (atm)
Tc Critical point temperature (K)

 m m  m
Cn Hm + n + (O2 + 3.78 N2 ) → nCO2 + H2 O + 3.78 n + N2
4 2 4
(25.11)

The mole fractions of the products in case of complete combustion of the general
hydrocarbon are:

n m/2
yCO2 =   yH2 O =  
4.78 n + 4 + m/4
m
4.78 n + m4 + m/4
 
3.78 n + m4
yN2 =  
4.78 n + m4 + m/4

References

Álvarez-Álvarez, P., Pizarro, C., Barrio-Anta, M., Cámara-Obregón, A., Bueno, J., Álvarez, A., et al.
(2018). Evaluation of tree species for biomass energy production in northwest Spain. Forests, 9,
160. https://doi.org/10.3390/f9040160.
Arce, M., Saavedra, Á., Míguez, J., Granada, E., Cacabelos, A., Arce, M. E., et al. (2013). Biomass
fuel and combustion conditions selection in a fixed bed combustor. Energies, 6, 5973–5989.
https://doi.org/10.3390/en6115973.
394 25 Some Parameters and Properties of Biomass Fuels

Bahadori, A., Zahedi, G., Zendehboudi, S., & Jamili, A. (2014). Estimation of the effect of biomass
moisture content on the direct combustion of sugarcane bagasse in boilers. International Journal
of Sustainable Energy, 33, 349–356. https://doi.org/10.1080/14786451.2012.748766.
Bartok, W., & Sarofim, A. F. (1991). Fossil fuel combustion: A source book (1st ed.). Wiley-
Interscience.
Borman, G. L., & Ragland, K. W. (1998). Combustion engineering. McGraw-Hill.
Brügemann, J., & Gerds-Ploeger, H. (2013). Life is movement (Aristotle, 4th century BC). Nether-
lands Heart Journal, 21(10), 427–428. https://doi.org/10.1007/s12471-013-0468-x.
Bushnell, D. J., Haluzok, C., & Dadkhah-Nikoo, A. (1990). Biomass fuel characterization : Testing
and evaluating the combustion characteristics of selected biomass fuels. Final Report May 1,
1988–July, 1989 (No. DOE/BP-1363). Portland, OR (USA): USDOE Bonneville Power Admin-
istration; Corvallis, OR (USA): Oregon State Univ. Dept. of Mechanical Engineering. https://doi.
org/10.2172/6910422.
Channiwala, S. A., & Parikh, P. P. (2002). A unified correlation for estimating HHV of solid, liquid
and gaseous fuels. Fuel, 81, 1051–1063. https://doi.org/10.1016/S0016-2361(01)00131-4.
Chimonyo, V. G. P., Modi, A. T., & Mabhaudhi, T. (2018). Sorghum radiation use efficiency and
biomass partitioning in intercrop systems. South African Journal of Botany, 118, 76–84. https://
doi.org/10.1016/j.sajb.2018.06.009.
Clarke, D., & Rieley, J. (2010). Strategy for responsible peatland management. Saarijärvi: Interna-
tional Peat Society.
De Blasio, C., Lucca, G., Özdenkci, K., Mulas, M., Lundqvist, K., Koskinen, J., et al. (2015). A
study on supercritical water gasification of black liquor conducted in stainless steel and nickel-
chromium-molybdenum reactors. Journal of Chemical Technology and Biotechnology, 91(10),
2664–2678. https://doi.org/10.1002/jctb.4871.
Demirbaş, A. (2005). Estimating of structural composition of wood and non-wood biomass samples.
Energy Sources, 27, 761–767. https://doi.org/10.1080/00908310490450971.
ECN, The Netherlands. (2018). Phyllis2—Database for biomass and waste (WWW Document).
https://www.ecn.nl/phyllis2/. Accessed August 22, 2018.
Energy Efficiency & Renewable Energy Department, USA. (2018). Bioenergy Technolo-
gies Office|Department of Energy (WWW Document). https://www.energy.gov/eere/bioenergy.
Accessed August 22, 2018.
Gholz, H. L. (1982). Environmental limits on aboveground net primary production, leaf area, and
biomass in vegetation zones of the Pacific northwest. Ecology, 63, 469–481. https://doi.org/10.
2307/1938964.
Gujarat Pollution Control Board, Paryavaran Bhavan. (2014). Pollution control guidelines for con-
version of boilers/utilities from natural gas to solid fuels (coal, lignite, agro fuels etc).
Jankowski, K. J., Dubis, B., Budzyński, W. S., Bórawski, P., & Bułkowska, K. (2016). Energy
efficiency of crops grown for biogas production in a large-scale farm in Poland. Energy, 109,
277–286. https://doi.org/10.1016/j.energy.2016.04.087.
Jenkins, B. (1993). Properties of biomass. In Biomass energy fundamentals (EPRI Report).
Jenkins, B. M., Baxter, L. L., Miles, T. R., Jr., & Miles, T. (1998). Combustion properties of biomass.
Fuel Processing Technology, 54, 17–46.
Jezierska-Thöle, A., Rudnicki, R., & Kluba, M. (2016). Development of energy crops cultivation
for biomass production in Poland. Renewable and Sustainable Energy Reviews, 62, 534–545.
https://doi.org/10.1016/j.rser.2016.05.024.
Ksenzhek, O. S., & Volkov, A. G. (1998). Plant energetics. San Diego, California: Academic Press.
Lau, F. S., Roberts, M. J., Rue, D. M., Punwani, D. V., Wen, W. -W., & Johnson, P. B. (1987). Peat
beneficiation by wet carbonization. International Journal of Coal Geology, 8, 111–121. https://
doi.org/10.1016/0166-5162(87)90026-7.
Liu, K., Xie, W., Li, D., Pan, W. -P., Riley, J. T., & Riga, A. (2000). The effect of chlorine and
sulfur on the composition of ash deposits in a fluidized bed combustion system. Energy Fuels,
14, 963–972. https://doi.org/10.1021/ef990197k.
References 395

Manickam, I. N., Ravindran, D. D., & Subramanian, D. P. (2006). Biomass densification methods
and mechanism. Cogeneration and Distributed Generation Journal, 21, 33–45. https://doi.org/
10.1080/15453660609509098.
Minkkinen, K., Korhonen, R., Savolainen, I., & Laine, J. (2002). Carbon balance and radiative forc-
ing of Finnish peatlands 1900–2100—The impact of forestry drainage. Global Change Biology,
8, 785–799. https://doi.org/10.1046/j.1365-2486.2002.00504.x.
Motta, I. L., Miranda, N. T., Maciel Filho, R., & Wolf Maciel, M. R. (2018). Biomass gasification in
fluidized beds: A review of biomass moisture content and operating pressure effects. Renewable
and Sustainable Energy Reviews, 94, 998–1023. https://doi.org/10.1016/j.rser.2018.06.042.
Pessarakli, M. (2005). Handbook of photosynthesis. Boca Raton: CRC Press, Taylor & Francis
Group.
Rayaprolu, K. (2009). Boilers for power and process. Boca Raton: CRC Press, Taylor & Francis
Group.
Reed, T. B., & Golden, A. D. (1988). Handbook of biomass downdraft gasifier engine systems.
Rumble, J. (2018). CRC handbook of chemistry and physics. CRC Press, Taylor & Francis Group.
Saastamoinen, J. (2015). Release profile of volatiles in fluidised bed combustion of biomass. Journal
of Fundamentals of Renewable Energy and Applications, 5, 1–12. https://doi.org/10.4172/2090-
4541.1000148.
Serup, H., Kofman, P. D., Falster, H., Gamborg, C., Gundersen, P., Hansen, L., et al. (2005). Wood
for energy production. Dublin: Coford.
Song, H., Quinton, K., Peng, Z., Zhao, H., Ladommatos, N., Song, H., et al. (2016). Effects of
oxygen content of fuels on combustion and emissions of diesel engines. Energies, 9, 28. https://
doi.org/10.3390/en9010028.
Tillman, D. (1978). Wood as an energy resource. New York, USA: Academic Press. https://doi.org/
10.1016/B978-0-12-691260-9.X5001-0.
Turpeinen, T. (2016). Developing biofuels production from food industry wastes in the rural area
of Kenya. Savonia: University of Applied Sciences Technology and Transport.
Van Loo, S., & Koppejan, J. (2008). The handbook of biomass gasification and co-firing. Earthscan.
Vassilev, S. V., Vassileva, C. G., Song, Y. C., Li, W. Y., & Feng, J. (2017). Ash contents and
ash-forming elements of biomass and their significance for solid biofuel combustion. Fuel, 208,
377–409. https://doi.org/10.1016/j.fuel.2017.07.036.
Virtanen, K., Hänninen, P., Kallinen, R. L., Virtiainen, S., Herranen, T., & Jokisaari, R. (2003). The
peat reserves of Finland in 2000. Geological Survey of Finland. Vammalan Kirjapaino Oy.
Yang, Y. B., Ryu, C., Khor, A., & Yates, N. (2005). Effect of fuel properties on biomass combustion.
Part II. Modelling approach—Identification of the controlling factors. Fuel, 84, 2116–2130.
https://doi.org/10.1016/j.fuel.2005.04.023.
Chapter 26
A Simple Estimation of the Efficiency
for a Biomass Power Plant

Nobody really knows about dark matter. Dark matter is a myth.


Kary Mullis, Nobel Prize in Chemistry 1993.

26.1 Mass Balances

We start here with the determination of simple mass balances and some considerations
on the stoichiometric coefficients when combustion reactions are involved. After this,
the energetic efficiencies will be determined for the main steps in the process.
Considering the combustion process and the flue gases generated, their composi-
tion can be calculated in a easier way when there is equilibrium. However, in reality,
the equilibrium is not a state commonly reached and the processes have to be studied
considering gas-phase kinetics. The main calculations are related to:
• flows and composition;
• overall fuel conversion reaction;
• thermal performance and efficiency.
The mass flow of the flue gas is evaluated in case the elemental composition of the
fuel and the quantity of air used for the combustion are known. In case we assume
complete combustion, the oxygen required can be calculated as follows (Van Loo
and Koppejan 2008):
   
kg (O2 ) MO2 X H MO2 MO2  
m̄ O2 ,air = Xc + + Xs − X O 1 − X H2 O · λ
kg (fuel, Waf) Mc 4 MH Ms
(26.1)

kg
where in this case Waf is referred to be wet and ash-free fuel, Mi kmol is the
molecular mass of the element i, X i is the fraction of the i element in dry and ash-

© Springer Nature Switzerland AG 2019 397


C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_26
398 26 A Simple Estimation of the Efficiency for a Biomass …

free fuel (daf), while X H2 O is the water mass fraction in the fuel, and λ is the excess
air ratio.
The N2 from air is calculated in the same way as:
 
kg (N2 ) YN2 ,Air
m̄ N2 ,air = m̄ O2 ,air (26.2)
kg (fuel, waf) YO2 ,Air · MO2

where (YO2 ,Air ) is the volume fraction of O2 in air (0.21), while waf means in this
case water and ash-free

YN2 ,Air = 1 − YO2 ,Air (26.3)


 
kg (Air)
m̄ Air = m̄ O2 ,air + m̄ N2 ,air (26.4)
kg (fuel, waf)

26.2 Thermal Performance and Efficiency

The energy which is stored in the fuel is released during the combustion process. In
this sense, a simple energy balance is set up as follows:

Energy in fuel + Preheat energy = Energy in flue gas + Energy losses (26.5)

The fraction of energy which is lost is due to the heat given to the environment, and
in addition, there is a fraction of energy lost because of incomplete combustion. The
energy stored in the fuel is evaluated from the HHV. On the other hand, to evaluate
the energy contained in the flue gas, it is necessary to evaluate the enthalpy for each
of the species: for instance CO2 , water which could be present in the flue gases in
the form of steam, SO2 , N2 , O2 .
The enthalpy of the substance leaving our system (in this case the power plant) is
a function of the temperature at which the substance is leaving (Levine 2008):
 
J a2 a3 a4 a5 a6 R u
h i (T ) = T a1 + T + T 2 + T 3 + T 4 + (26.6)
kgi 2 3 4 5 T Mi

where R = 8.3145 J/(mol K) is the universal gas constant, T = gas temperature, K


and the coefficients ai are depending on the temperature.
The total efficiency is a function of the calorific value for the fuel, the enthalpies
of air inlet and fuel inlet, and the sum of heat losses by radiation, convection, and
conduction.
Regarding Eq. (26.6), the coefficients for enthalpy calculation of the flue gas are
given in Tables 26.1 and 26.2.
The standard heat of reaction is the sum of the heats of formation of the reaction
products minus the heats of formation of the reactants. For instance, as an example
Table 26.1 Enthalpy calculation and related coefficients for Eq. (26.6) for selected components. Data from Van Loo and Koppejan (2008)
CO2 H2 O
≤1000 K >1000 K ≤1000 K >1000 K
26.2 Thermal Performance and Efficiency

a1 2.35677352 3.85746029 4.19864056 3.03399249


a2 0.00898459677 0.00441437026 −0.0020364341 0.00217691804
a3 −0.00000712356269 −0.00000221481404 0.00000652040211 −0.000000164072518
a4 2.45919022E−9 5.23490188E−10 −5.48797062E−9 −9.7041987E−11
a5 −1.43699548E−13 −4.72084164E−14 1.77197817E−12 1.68200992E−14
a6 −48371.9697 −48759.166 −30293.7266999999 −30004.2971
399
400

Table 26.2 Enthalpy calculation and related coefficients for Eq. (26.6) for selected components
SO2 N2 O2
≤1000 K >1000 K ≤1000 K >1000 K ≤1000 K >1000 K
a1 2.911438 5.254498 3.298677 2.92664 3.78245636 3.28253784
a2 0.008103022 0.001978545 0.0014082404 0.0014879768 −0.00299673416 0.00148308754
a3 −0.00000690671 −0.0000008204226 −0.000003963222 −0.000000568476 0.00000984730201 −0.000000757966669
a4 0.000000003329015 1.576383E−10 −0.000000005641515 1.0097038E−10 −9.68129509E−09 2.09470555E−10
a5 −8.777121E−13 −1.1204512E−14 −2.444854E−12 −6.753351E−15 3.24372837E−12 −2.16717795E−14
a6 −36878.81 −37568.85 −1020.8999 −922.7977 −1063.94356 −1088.45772
Data from Van Loo and Koppejan (2008)
26 A Simple Estimation of the Efficiency for a Biomass …
26.2 Thermal Performance and Efficiency 401

Table 26.3 Coefficients for


a      
specific heat calculations b ×102 c ×105 d ×109
related to methanol, CO, and
CH3 OH 4.55 2.186 −0.291 −1.92
hydrogen
CO 6.726 0.04001 0.1283 −0.5307
H2 6.952 −0.04576 0.09563 −0.2079
Adapted from Hicks and Chopey (2012)

we could consider here the following reaction taking place at 600 K and 10.13 MPa
(Hicks and Chopey 2012).

CO(g) + 2H2 (g) → CH3 OH(g) (26.7)

For this reaction:

H298
0
= −48.08 − [−26.416 + 2(0)] = −21.664 kcal/(g mol) (26.8)

The values for Eq. (26.8) are given in the literature since these are standard heat
of formation for the components taking part in the reaction. The variation for the
specific heat related to the overall reaction is given by:

C p = a + b · T + c · T 2 + d · T 3 (26.9)

where in reference to Table 26.3 the variation for the coefficients used in Eq. (26.9)
are given by:

a = 4.55 − 6.726 − 2(6.952) = −16.08

b = [2.186 − 0.04001 + 2(0.04576)]10−2

c = −0.61056 × 10−5

d = −0.9735 × 10−9

The calculation for the enthalpy of the gas at a determinate temperature follows
the Kirchhoff’s law:

T

H (T ) = H 0 (T = 298) + a + b · T + c · T 2 + d · T 3 dT
298
(26.10)

And therefore, for this example, the total variation for the enthalpy of reaction is
given as follows:
402 26 A Simple Estimation of the Efficiency for a Biomass …

H (T = 600) = −2.39 × 104 cal/(g mol)

The integral in Eq. (26.10) gives results in cal/(g mol K). Additional examples
can be found in textbooks of thermodynamics or physical chemistry (Levine 2008).

26.3 Incomplete Combustion

The energy which is lost because of incomplete combustion is evaluated by consid-


ering the calorific value of the species taken into consideration and lost at the exit of
the power plant. For the carbon monoxide, we have:
     
kJ kJ kg(CO)
Ē CO = CVCO m̄ CO (26.11)
kg(fuel, waf) kg(CO) kg(fuel, waf)

where CVCO is the calorific value of CO which is 10,102 kJ/(kg CO) at 25 °C. The
energy lost due to uncombusted carbon is:
     
kJ kJ kgC
Ē C = CVC m̄ c (26.12)
kg fuel(waf) KgC kg fuel (waf)

where the carbon, C, has a calorific value of 34,910 kJ/kg


The energy balance between the inlet fuel and the flue gases is estimated as:

NCV · ṁ F + [h F (TF ) − h F (TAmb )]ṁ F + [h air (Tair ) − h air (Tamb )]ṁ air
 
= [h FG (TFG ) − h FG (TAmb )]ṁ FG + Qi + Ē j · ṁ F (26.13)
i j

where the energy necessary to heat up the air and the fuel is:

[h F (TF ) − h F (TAmb )]ṁ F + [h air (Tair ) − h air (Tamb )]ṁ air (26.14)

where NCV is the net calorific value of the fuel, hF , hAir , and hFG are the
 enthalpies
of fuel, air and flue gas, respectively, and T F is the fuel temperature.
 i Q i is the
sum of heat loss by radiation, convection, and conduction, and j Ē j · ṁ F is the
sum of heat losses in unburned components.
26.3 Incomplete Combustion 403

The thermal efficiency is evaluated from:



Q ch + i Q i
ηth = 1 −
NCV · ṁ F + [h F (TF ) − h F (TAmb )]ṁ F + [h air (Tair ) − h air (Tamb )]ṁ air
(26.15)

where Q ch is defined as follows with T ch indicating the chimney inlet temperature:

Q ch = [h FG (Tch ) − h FG (TAmb )]ṁ FG (26.16)

The combustion efficiency is calculated from:



j Ē j · ṁ F
ηcomb = 1 − (26.17)
NCV · ṁ F

Considering the thermal and the combustion efficiency, the total efficiency is
finally evaluated from:
 
Q ch + i Qi + j Ē j · ṁ F
ηtot = 1 −
NCV · ṁ F + [h F (TF ) − h F (TAmb )]ṁ F + [h air (Tair ) − h air (Tamb )]ṁ air
(26.18)

References

Hicks, T. G., Chopey, N. P. (2012). Handbook of chemical engineering calculations. McGraw-Hill.


Levine, I. N. (2008). Physical chemistry (6th ed.). McGraw-Hill Education.
Van Loo, S., Koppejan, J. (2008). The handbook of biomass gasification and co-firing. Earthscan.
Chapter 27
Some Data on Oxidation and Reduction
States and Half-Cell Reactions

Why the chicken crossed the road? To get on the other side.
Peter Diamond, Nobel Prize in Economics 2010.

27.1 Reduction and Oxidation States, Electron Affinities,


and Ionization Potentials

Taking into account the topics treated during this first part of the manuscript, it will
be beneficial to demonstrate some basic concepts of oxidation and reduction. It is
known that when one atom is going to be oxidized, this means that it will lost some
electrons. On the other hand, when an atom or molecule receives electrons, it will
be reduced. Therefore oxygen, which has an oxidation number equal to zero in its
molecular state, is considered as oxidizing other types of atoms like carbon, which
is therefore oxidized. Carbon will displace some of its electrons and will assume
a positive charge, while oxygen will have a negative charge. In organic chemistry,
this is normally described by the concept of oxidized state. Table 27.1 shows some
oxidation states for different kind of molecules.

Table 27.1 Reduced and oxidized states of some elements with examples (Drapcho et al. 2007)
Reduced state Oxidized state
Element Oxidation number Example Oxidation number Example
S −2 H2 S +6 SO4 2−
N −3 NH3 +5 NO3 −
O −2 H2 O, C6 H12 O6 0 O2
H 0 H2 +1 H2 O
C −4 CH4 +4 CO2

© Springer Nature Switzerland AG 2019 405


C. De Blasio, Fundamentals of Biofuels Engineering and Technology,
Green Energy and Technology, https://doi.org/10.1007/978-3-030-11599-9_27
406 27 Some Data on Oxidation and Reduction States …

Table 27.2 Ionization potential (first value) and electron affinities (second value) in eV (Jones and
Fleming 2014)
H He
13.60 24.59
0.75 0
Li Be B C N O F Ne
5.39 9.32 8.30 11.26 14.53 13.62 17.42 21.56
0.62 0 0.24 1.27 0 1.47 3.34 0
Na Mg Al Si P S Cl Ar
5.14 7.65 5.99 8.15 10.49 10.36 12.97 15.75
0.55 0 0.46 1.24 0.77 2.08 3.61 0

After acknowledging that atoms in molecules have an oxidation state, it is natural


the question of how much energy is actually required to pull one electron from a
molecule. The ionization potential is the energy required to move away one electron
from the atom; this quantity is measured in eV/atom or kcal/mole, joules/mole …,
etc. When the first electron is removed from an atom, this will be referenced as first
ionization potential and Table 27.2 gives some values for it.
While Table 27.3 gives a more detailed representation of the ionization potentials
expressed in kJ/mol.
Concerning the representation of an atom and its electrons, an atom is represented
by its energy levels (shell number). The shell number describes how far we are from
the center of the atom. For each of these energy levels, there are particular orbitals
which have a particular number and shapes. The number of orbitals present in each
of the energy levels is determined by the energy level itself. Referring to Table 27.4,
if we are in the first energy level, then only one type of orbital is there and there
is only one single orbital, the spherical one. In case we consider the second upper
energetic level, we will have two kinds of orbitals: the type s and the type p. The
type p can have 3 orientations in the space. This is given by the m number. The spin
number concerns the way an electron can spin, and there are only two values for this
parameter; this means that within one orbital we can have only two electrons and not
more.
Table 27.3 Periodic table of the ionization potentials (Jones and Fleming 2014)
H He
1312 2372
Li Be B C N O F Ne
520 900 801 1086 1402 1314 1681 2081
Na Mg Al Si P S Cl Ar
496 738 578 786 1012 1000 1251 1521
K Ca Sc Ti V Cr Mn Fe Co Ni Cu Zn Ga Ge As Se Br Kr
419 590 631 658 650 653 717 759 758 737 745 906 579 762 947 941 1140 1351
Rb Sr Y Zr Nb Mo Tc Ru Rh Pd Ag Cd In Sn Sb Te I Xe
403 549 616 660 664 685 702 711 720 805 731 868 558 709 834 869 1008 1170
Cs Ba La Hf Ta W Re Os Ir Pt Au Hg Tl Pb Bi Po At Rn
356 503 538 680 761 770 760 840 880 870 890 1007 589 716 703 812 890 1037
27.1 Reduction and Oxidation States, Electron Affinities …

Fr Ra Ac Rf Db Sg Bh Hs Mt Ds Rg Uub Uut Uuq Uup Uuh Uus Uuo


384 509 499
Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu
527 523 530 536 543 547 593 565 572 581 589 597 603 524
Th Pa U Np Pu Am Cm Bk Cf Es Fm Md No Lr
587 568 584 597 585 578 581 601 608 619 627 635 642 473
407
408 27 Some Data on Oxidation and Reduction States …

Table 27.4 Quantum representation of atoms, quantum numbers (Jones and Fleming 2014)
n (Shell number) l type of orbital, can ml [orientation in s (spin) Orbital type
have values (0 … n space, (−l … +l)]
− 1)
1 0 0 ∓ 21 1s

2 0 0 ∓ 21 2s

2 1 −1, 0, +1 ∓ 21 2p

3 0 0 ∓ 21 3s

3 1 −1, 0, +1 ∓ 21 3p

3 2 −2, −1, 0, +1, +2 ∓ 21 3d

27.2 Electronegativity

The electronegativity coefficient is a measure of the capacity to attract an electron.


For example, with reference to Table 27.5 the carbon is more electronegative than
hydrogen; this is why in C–H bonding we have that the oxidation state of carbon is
negative and H is positive.
We need to know the capability of a particular specie to attract or to loose electrons
in case we would like to use that particular element in a fuel cell or galvanic cell, or
any other kind of setup where there are redox reactions. We have seen previously in
the first part of this manuscript that the generation of biomass has a lot in common
with galvanic cells and the coupling of specific half-cell reactions. Here some more
data are given in Table 27.6. For a more complete list of half-cell reactions for
non-biological systems, the readers can refer to some of the available handbooks on
chemistry or physics (Haynes 2010).
Table 27.5 Electronegativity values for the components of the periodic table (Jones and Fleming 2014)
H He
27.2 Electronegativity

2.20
Li Be B C N O F Ne
0.98 1.57 2.04 2.55 3.04 3.44 3.98
Na Mg Al Si P S Cl Ar
0.93 1.31 1.61 1.90 2.19 2.58 3.16
K Ca Sc Ti V Cr Mn Fe Co Ni Cu Zn Ga Ge As Se Br Kr
0.82 1.0 1.36 1.54 1.63 1.66 1.55 1.83 1.88 1.91 1.90 1.65 1.81 2.01 2.18 2.55 2.96 2.9
Rb Sr Y Zr Nb Mo Tc Ru Rh Pd Ag Cd In Sn Sb Te I Xe
0.82 0.95 1.22 1.33 1.6 2.16 1.9 2.2 2.28 2.20 1.93 1.69 1.78 1.80 2.05 2.1 2.66 2.6
Cs Ba La Hf Ta W Re Os Ir Pt Au Hg Tl Pb Bi Po At Rn
0.79 0.89 1.10 1.3 1.5 2.36 1.9 2.2 2.20 2.28 2.54 2.00 1.62 1.87 2.02 2.0 2.2
Fr Ra Ac Rf Db Sg Bh Hs Mt Ds Rg Uub Uut Uuq Uup Uuh Uus Uuo
0.7 0.9
Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu
527 523 530 536 543 547 593 565 572 581 589 597 603 524
Th Pa U Np Pu Am Cm Bk Cf Es Fm Md No Lr
587 568 584 597 585 578 581 601 608 619 627 635 642 473
409
410 27 Some Data on Oxidation and Reduction States …

Table 27.6 Standard


Half cell E 0 (V )
electrochemical potentials for
biological half-cell reactions Succunate + CO2 + 2H+ + 2e− ↔ −0.670
α − ketoglutarate + H2 O
Acetate + 2H+ + 2e− ↔ Acetaldehyde −0.580
2H+ + 2e− ↔ H2 −0.421
α − ketoglutarate + CO2 + 2H+ + 2e− ↔ isocitrate −0.380
Cystine + 2H+ + 2e− ↔ 2cysteine −0.340
NAD+ + 2H+ + 2e− ↔ NADH + H+ −0.320
NADP+ + 2H+ + 2e− ↔ NADPH + H+ −0.324
Acetaldehyde + 2H+ + 2e− ↔ ethanol −0.197
Pyruvate + 2H+ + 2e− ↔ lactate −0.185
Oxaloacetate + 2H+ + 2e− ↔ malate −0.166
FAD + 2H+ + 2e− ↔ FADH2 +0.031
Fumarate + 2H+ + 2e− ↔ succinate +0.031
Ubiquinone + 2H+ + 2e− ↔ ubiquinol +0.045
2cytochrome box + 2e− ↔ 2cytochrome bred +0.070
2cytochrome cox + 2e− ↔ 2cytochrome cred +0.254
2cytochrome a3(ox) + 2e− ↔ 2cytochrome a3(red) +0.385

1
2 O2 + 2H+ + 2e− ↔ H2 O +0.816

Adapted from (Karp, 2009)

References

Drapcho, C. M., Nhuan, N. P., Walker, T. H. (2007). Biofuels engineering process technology.
McGraw-Hill Professional.
Haynes, W. M. (2010). CRC handbook of chemistry and physics (91st ed.) CRC Press, Taylor &
Francis Group.
Jones, M., Fleming, S. A. (2014). Organic chemistry (5th edn). W. W. Norton & Company.
Karp, G. (2009). Cell and molecular biology: Concepts and experiments (6th ed.). New York: Wiley.

You might also like