Bideterioration of Concrete
Bideterioration of Concrete
of ConCrete
Biodeterioration
of ConCrete
Thomas Dyer
University of Dundee
Division of Civil Engineering
Dundee, Scotland, UK
p,
A SCIENCE PUBLISHERS BOOK
Cover Acknowledgement
CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742
For permission
For permission toto photocopy
photocopy oror use
use material
material electronically
electronically from
from this
this work,
work, please
please access
access www.copyright.com
www.copyright.com
(http://www.copyright.com/) or
(http://www.copyright.com/) or contact
contact the
the Copyright
Copyright Clearance
Clearance Center,
Center, Inc.
Inc. (CCC),
(CCC), 222
222 Rosewood
RosewoodDrive,
Drive,Danvers,
Danvers,
MA 01923, 978-750-8400. CCC is a not-for-profit organization that provides licenses and registration for
MA 01923, 978-750-8400. CCC is a not-for-profit organization that provides licenses and registration for aa variety
variety
of users. For organizations that have been granted a photocopy license by the CCC, a separate system of
of users. For organizations that have been granted a photocopy license by the CCC, a separate system of payment payment
has been arranged.
has been arranged.
Trademark Notice:
Trademark Notice: Product
Product or
or corporate
corporate names
names may
may be
be trademarks
trademarks or
or registered
registered trademarks,
trademarks, and
and are
are used
usedonly
onlyfor
for
identification and
identification and explanation
explanation without
without intent
intent to
to infringe.
infringe.
Library
Library of
of Congress
Congress Cataloging‑in‑Publication
Cataloging‑in‑Publication Data
Data
Names:
Names:Alava,
Names: Dyer, Juan
Alava, Thomas
Juan Joseì,
Joseì,D.,editor.
author.
editor.
Title:
Title:Tropical
Title: Tropical pinnipeds
Biodeterioration
pinnipedsof :: bio-ecology, threats,
threats, and
concrete / Thomas
bio-ecology, conservation
Dyer,
and University /of
conservation / editor,
Dundee
editor,
Juan
Juan Joseì Alava,
Joseì of
Division CivilFaculty
Alava, Faculty of
of Science,
Engineering, Science, Institute
Institute
Dundee, for
for the
Scotland, Oceans
Oceans and
theUK. and
Fisheries,
Fisheries,
Description:The University
The Boca Raton,of
University ofFLBritish
: CRCColumbia,
British Columbia, Vancouver,
Vancouver,
Press, [2016] Canada.
Canada.
| "A science publishers book."
Description: Boca
Description: Boca Raton,
Raton, FL :: CRC
CRC Press,
FLreferences Press, 2017.
2017. || “A
“A science
science publishers
publishers book.”
book.”
| Includes bibliographical and index.
||Identifiers:
Includes
Includes bibliographical
bibliographical references
references and
and index.
index.
LCCN 2017005278| ISBN 9781498709224 (hardback : alk. paper) |
Identifiers: LCCN
Identifiers: LCCN 2016054170|
2016054170| ISBN
ISBN 9781498741392
9781498741392 (hardback
(hardback :: alk.
alk. paper)
paper) ||
ISBN9781498741408
ISBN 9781498709231 (e-book)
(e-book
ISBN 9781498741408
Subjects: LCSH: (e-book :: alk.
alk. paper)
paper)
Concrete--Biodegradation.
Subjects: LCSH:
Subjects: LCSH: Seals
Seals (Animals)--Tropics.
(Animals)--Tropics. || Pinnipedia.
Pinnipedia.
Classification:LCC
Classification: LCC TA440 .D94 2016 || DDC 620.1/3623--dc23
Classification: LCC QL737.P6
QL737.P6 T76 T76 2017
2017 | DDC
DDC 599.79--dc23
599.79--dc23
LC record
LC record available at https://lccn.loc.gov/2017005278
LC record available
available at
at https://lccn.loc.gov/2016054170
https://lccn.loc.gov/2016054170
Visit the
Visit the Taylor
Taylor &
& Francis
Francis Web
Web site
site at
at
http://www.taylorandfrancis.com
http://www.taylorandfrancis.com
and the
and the CRC
CRC Press
Press Web
Web site
site at
at
http://www.crcpress.com
http://www.crcpress.com
To
Judith, Angus and Oscar
Preface
Dedication v
Preface vii
1. Introduction 1
2. The Chemistry of Concrete Biodeterioration 7
3. Bacterial Biodeterioration 77
4. Fungal Biodeterioration 153
5. Plants and Biodeterioration 212
6. Damage to Concrete from Animal Activity 265
Index 279
Chapter 1
Introduction
Biodeterioration of Concrete
Around 1899 the outfall sewer serving Los Angeles—which had been built
only four years previously—started showing signs of severe deterioration
[1]. The concrete lining of the sewers and the lime mortar used in the
brickwork was undergoing significant expansion, before crumbling away.
The bricks themselves appeared unaffected, save for those which spalled
away as a result of the considerable pressure exerted by the expansion.
This might have been attributed to aggressive substances dissolved in the
sewage itself, were it not for the inconvenient fact that the corrosion was
occurring above the waterline.
Early in the 20th century, the construction of brick-lined sewers was
superseded by the use of concrete pipes. However, the problem of corrosion
persisted. Corrosion above the waterline suggested that a gas was causing
the corrosion and the most likely candidate was hydrogen sulphide (H2S),
a malodorous and potentially lethal gas produced by sulphate-reducing
bacteria in sewage. This theory appeared to be supported by the fact that
rates of deterioration were significantly reduced by measures taken to
prevent the release of H2S. In the Los Angeles case, modification of the
sewers to keep them fully flooded—effectively turning them into a septic
tank—solved the problem. Moreover, in 1929, reduced rates of deterioration
were observed in concrete sewers in Orange County and El Centro County,
California, after chlorination of the sewage was initiated [2].
However, H2S—whilst being extremely hazardous—is actually not
corrosive to hardened cement. The most likely explanation for the damage
was that the gas was being converted into sulphuric acid. Chemical reactions
capable of causing this conversion were proposed, but were not (and, in fact,
could not) be demonstrated experimentally. The unsatisfactory nature of
the proposed mechanism of acid formation led Guy Parker (a bacteriologist
who was part of a team facing a similar problem in concrete sewers in
2 Biodeterioration of Concrete
Deterioration rates vary considerably and will depend on the type and
number of organisms involved, as well as the environmental conditions
and the composition and properties of the concrete. However, in cases
where long service lives are required, even slow rates of deterioration
may be unacceptable. This is certainly true in the case of much of our
infrastructure—such as bridges and sewers—but becomes fundamental
in the case of containment applications such as nuclear waste storage and
oil well decommissioning, where service lives of hundreds or thousands
of years may be required.
This Book
This book aims to examine, as widely as possible, the different ways in
which living organisms can compromise the durability of concrete, and to
also explore means by which deterioration can be prevented or controlled.
Following on from the discussion above, we might be led to assume
that the presence of life in close proximity to concrete should always lead to
problems. This is not the case—in most instances living organisms will have
little impact on the durability of a structure. Indeed, a number of instances
are discussed in this book which their presence has a protective effect.
This raises the more general issue of whether it is appropriate to attempt
to remove a particular species from around a structure simply to improve
durability. This partly depends on what sort of organism is involved, but
this is not the whole issue: the organism in question will make up part of
a much larger ecosystem, and its removal may have significant impacts
on the local environment as a consequence. Furthermore, whilst the use
of biocides as a means of protection from biodeterioration and biofouling
has been widely practiced, the secondary toxic effects to other species
potentially significant distances away from the point of use have led to this
approach being questioned. Legislation, plus a more general desire to act
in an environmentally responsible manner, is leading to a move away from
biocides. This book has been written with this trend in mind.
One aspect of the colonization of concrete structures by living organisms
which this book attempts to avoid is the matter of aesthetics. This is largely
because of the difficulties in making judgements about what constitutes an
aesthetically disagreeable state. It is notable that opinions in this respect
vary considerably. A plant climbing the side of a building is considered
attractive to many, but is disagreeable to others. The discolouration of
a facade by lichen, fungi or algae is considered by many to be ugly and
requiring cleaning, whilst to others it is the building’s ‘patina’. Thus, this
author has chosen to steer clear of adopting any position on this issue.
The book is divided into six chapters (including this introduction).
Because chemical processes play such an important role in many forms
Introduction 5
References
[1] Olmstead FH and Hamlin H (1900) Converting portions of the Los Angeles outfall sewer
into a septic tank. Engineering News and American Railway Journal 44: pp. 317–318.
[2] Goudey RF (1928) Odor control by chlorination. California Sewage Works Journal 1:
87–101.
[3] Parker CD (1945) The corrosion of concrete. 1. The isolation of a species of bacterium
associated with the corrosion of concrete exposed to atmospheres containing hydrogen
sulphide. Australian Journal of Experimental Biology and Medical Science 23: 81–90.
[4] Hueck HJ (1965) The biodeterioration of materials as part of hylobiology. Material und
Organismen 1: 5–34.
[5] Woese C, Kandler O and Wheelis M (1990) Towards a natural system of organisms:
proposal for the domains Archaea, Bacteria, and Eucarya. Proceedings of the National
Academy of Science 87: 4576–4579.
Chapter 2
2.1 Introduction
Many of the biodeterioration processes described in subsequent chapters
involve the production of chemical substances by organisms. These
substances are, in some cases, products of excretion processes, whilst in
other cases are produced with the purpose of modifying an organism’s
environment to its benefit. The influences of these substances on concrete
are diverse, since reactions between cement hydration products in the
concrete yield products whose characteristics vary widely.
Many of these substances are produced by more than one kingdom or
domain, and so it makes sense to discuss the chemistry of these substances
in the context of concrete first, and to subsequently use this chapter as a
reference for subsequent discussion. Thus, this chapter provides information
on the aqueous chemistry of various biogenic products in the presence of
the most relevant elements present in concrete. It then examines, in generic
terms, the various mechanisms that can lead to deterioration of concrete
through chemical reaction.
Before introducing specific substances, a brief introduction is provided
to the reader regarding some of the concepts used to characterize their
behaviour. These are acidic strength, the formation of metal complexes in
solution and the solubility product. The chapter employs a form of solubility
diagram to illustrate the influence of the substances on cementitious systems.
For this reason, an introduction to these diagrams is included, along with a
brief description of the method used to construct such diagrams for this book.
HA ⇌ A– + H+
to completion when it is dissolved in water.
For many acids, this reaction will reach equilibrium with only a
proportion of molecules deprotonated. The extent to which deprotonation
occurs for a given acid is expressed in terms of the acid dissociation constant
(Ka). This constant is defined by the equation:
[H+] [A–]
Ka =
[HA]
With square brackets denoting concentration. For convenience, Ka is
often expressed as its negative logarithm (pKa). A strong acid is one with a
pKa value of less than –1.74.
The pH of the resulting solution is given be the equation:
pH = –log[H+]
pH and pKa are intrinsically related, and the pH of a solution containing
acid HA is described by the equation:
[A–]
pH = pKa + log
[AH]
1.0
0.8
0
i=
<{
0:: 0.6
L1J
u
z
<{
0
z 0.4
::::l
aJ
<{
0.2
0.0
2 4 6 8 10 12 14
pH
s':(
citrate3-
H(citrate}'-
0.8 H2(citrater
; ""' H3 (citrate)
0 I \
~
'-'- 0.6
I I
I \
\ > o"'}-o
L1J I. / I
I '
0
u
z
<{
)' I
'I \I
~ 0.4 I.
::::l 1 I I
aJ I . I
<{ / I 1
I I I
0.2 I I
I \ I
I \
\ \
0.0 +--""----,-"~--".=::o-~----,-'-'-"~----,-----,-----1
2 4 6 10 12 14
pH
interactions. This effective concentration is called the activity (ɑ) of the ion.
The activity can be defined as:
[i]
αi = γi [iθ]
I = 12 ∑[i]zi2
where zi is the charge on ionic species i. The ionic strength equation must
include all ions present in a solution.
At low ionic strengths (I < 0.001 M) the Debye-Hückel equation can be
used to obtain a value for the activity coefficient:
log γi = –Az 2i √I
where A is a constant which is independent of the ion involved, but
which changes with temperature and pressure. At 25°C and 1 atmosphere
pressure, its value is 0.5085 [1]. At higher ionic strengths, different forms
of the Debye-Hückel equation must be used. Where the ionic strength is
in the range 0.001 ≤ I < 0.1, the extended Debye-Hückel equation applies:
–Az 2i √I
logγi =
1 + Ba0 √I
where B = a constant dependent on temperature and pressure; and
a0 = the theoretical hydrated radius of the ion (Å).
The hydrated radius of an ion is the radius within which water
molecules are closely bound to the ion.
Within the range 0.1 ≤ I < 0.7, the Davies equation [2] should be used:
–Az 2i √I
logγi = + bI
1 + √I
12 Biodeterioration of Concrete
The original form of the Davies equation used a fixed value of b, and
this value has been revised several times, but is commonly 0.3. Using a
fixed value will introduce some inaccuracy at ionic strengths lower than
0.1 and also for ions with valencies of more than 1. This has subsequently
been addressed through the development of ion-specific values of b. Where
such an approach is adopted, the Davies equation is referred to as the
Truesdell-Jones equation [3].
A number of approaches for obtaining activity coefficients for solutions
of ionic strength, including the Pitzer ion interaction and the B-dot equation.
These are of lesser relevance to the issues under discussion, but of great
value when dealing with environments containing solutions containing
brine-like concentrations of highly soluble salts.
Log Ksp is influenced by crystal size, temperature and pressure. It is
often the case that the influence of crystal size, rightly or wrongly, is often
ignored. Pressure usually has a relatively minor influence over the solubility
product. Temperature, however, has potentially a much greater influence
and, in systems where temperature may vary, this effect will usually need
to be accommodated in solubility calculations.
The influence of temperature on solubility is the result of Le Chatelier’s
principle, which states that a change to a system of concentration,
volume, temperature or pressure will cause that system to oppose the
change. Dissolution reactions can either be exothermic (giving out heat)
or endothermic (taking in heat) and this will change the temperature of
the fluid in which the reaction is occurring. Thus, in accordance with Le
Chatelier’s principle, as the temperature of a system increases, the solubility
product of a compound which undergoes an exothermic reaction will
decline, whilst the solubility of a compound with an endothermic reaction
will increase.
The influence of temperature (expressed in Kelvin) is described by the
van’t Hoff equation:
∆HR0 1 1
logKsp,T2 – logKsp,T1 = ( –
2.303R T1 T2 )
where Ksp,T2 = the solubility product at a temperature of T2;
Ksp,T1 = the solubility product at a reference temperature T1;
∆HR0 = the standard enthalpy of reaction (J/mol); and
R = the gas constant (J/K mol).
In the context of aqueous chemistry, the reference temperature is
frequently 25ºC. ∆HR0 is dependent on temperature, but where temperatures
are close to the reference temperature (± 20 oK), the value for the reference
temperature may be assumed with little compromise in the accuracy of the
predicted change in log Ksp.
The Chemistry of Concrete Biodeterioration 13
When a metal ion dissolves in water in the manner described in the previous
section it may simply remain in solution as a cation. However, where other
molecules or ions are present in solution, there exists the possibility that the
cation will form a metal complex with one or more of these entities, which
are referred to as ligands when interacting in this way. A complexation
reaction can be written in the form
M + L ⇌ ML
The complex ML may not be the only complex which can form. For
instance, a second ligand may be involved:
ML + L ⇌ ML2
In a similar manner to the acid dissociation constant, the equilibrium
of the first reaction can be described using the equation:
[ML]
K1 =
[M][L]
where K1 is said to be a stability constant, and is specifically the association
constant of the complex. Similarly, the equilibrium of the second reaction
is described thus:
14 Biodeterioration of Concrete
[ML2]
K2 =
[ML][L]
These association constants are described as stepwise constants: they
describe the individual steps required to move from M to ML2. Of course,
the reaction could be described in terms of both stepwise reactions occurring
together:
M + 2L ⇌ ML
In which case the equilibrium of the reaction can be characterized by
the cumulative constant, β12:
[ML] [ML2] [ML2]
β12 = K1K2 = ∙ =
[M][L] [ML][L] [M][L]2
It is, again, often convenient to express stability constants in terms of
log K.
The ionic form that a molecule adopts in solution will define to a large
extent what complexes it can form with metals. For instance, calcium will
form a complex (of the ML type) with the acetate ion (C2H3O2–), but not
with its protonated form. This has significant implications, since it means
that the concentration of metal complexes in solution will vary depending
on pH. This is shown in Figure 2.3 which shows the distribution of the
Ca(acetate) complex only appears in significant quantities once the pH is
sufficiently high to yield large concentrations of deprotonated acetic acid
(as previously seen in Figure 2.1). Another point to note about this plot is
10~ .---------------------------------------------------,
8x1o-•
""'=
0
E
z
0 6x10 9
i=
<(
0::
1- Ca(acetate(
z / ------------------,
UJ
0
z
4x10-9
/ / ' \
0
0
! \
I
2x1o-• I
I
I
I
/
--"'
0
2 4 6 8 10 12 14
pH
Figure 2.3 Concentrations of complexes formed by calcium in an acetic acid solution
as a function of pH. Concentrations: acetic acid: 100 mmol/l; Ca: 0.0001 mmol/l.
The Chemistry of Concrete Biodeterioration 15
that calcium also forms a complex with hydroxide ions (CaOH+), whose
stability constant is higher than that of Ca(acetate), but whose appearance
–
in significant quantities is observed only once OH concentrations are high.
As for solubility, stability constants are also influenced by temperature
and the effect can again be described by the van’t Hoff equation. The salt
effect also influences stability constants and, as a result, it is usually the
convention to state the ionic strength at which a solubility product was
measured.
Particularly stable complexes are formed where the ligand is able to
form multiple bonds through different co-ordinating atoms with the same
metal ion. This type of interaction is known as chelation, and the stability
of the resulting complex is dependent on the size and number of ‘chelate
rings’ formed. A chelate ring is the closed structure formed by the metal
ion, a co-ordinating atom in the ligand molecule, through the sequence of
bonded atoms that link the co-ordinating atoms, and back to the metal ion.
A higher number of chelate rings is likely to increase the stability of the
surface complex formed. Moreover, the number of ring members plays an
important role in complex stability, with relative stability usually following
the sequence:
5-membered > 6-membered > 7-membered/4-membered.
\
0
~
.2l
'0
c
C1l
~
0
-2 0..
ro
~ -4
0>
0
....J
Ca2 +
1\
+I
0
C1l
-6 0
2 3 4 5 6 7 8 9 10 11 12 13 14
pH
Figure 2.4 Solubility diagram for calcium.
The solubility diagrams used in this chapter have been drawn using
the computer program MEDUSA which uses the stability constants and
solubility products defined for a system. These constants have been
obtained from various sources in the literature. Unless otherwise stated
they correspond to a temperature of 25ºC. In the case of stability constants,
wherever possible the values for an ionic strength of 0 have been selected.
Where the effect of acidic species is investigated, all solubility diagrams
have been constructed for an acid concentration of 0.1 mol/l.
The clinker that is used to make Portland cement consists of four main
phases: tricalcium silicate (3CaO.SiO2), dicalcium silicate (2CaO.SiO2),
tricalcium aluminate (3CaO.Al2O3) and tetracalcium aluminoferrite (4CaO.
Al2O3.Fe2O3). This clinker is ground with a quantity of calcium sulfate—
usually either as gypsum (CaSO4.2H2O) or anhydrite (CaSO4).
The Chemistry of Concrete Biodeterioration 17
(higher CaO) and siliceous (low levels of CaO) forms. Both forms are mainly
composed of SiO2 and Al2O3. Whilst some of the minerals found in FA
(notably quartz, mullite, hematite and magnetite) are of very low solubility,
and react to a limited extent, a large proportion of the material consists of a
glassy substance which reacts with portlandite formed by Portland cement
to form CSH and aluminoferrites. This reaction is known as a pozzolanic
reaction, and the high SiO2 and Al2O3 content of the material mean that,
again, the Ca/Si ratio of the CSH gel is reduced and more aluminoferrite
phases are formed.
Silica fume is a by-product of the manufacture of silicon for the
electronics sector. It is usually almost 100% SiO2 and undergoes a pozzolanic
reaction similar to FA, leading to a modification in the composition of CSH.
The use of all of these materials in combination with Portland cement
will also lead to a reduction in the quantity of portlandite formed, since the
Portland cement is diluted, and may be reduced further if the material in
question undergoes a pozzolanic reaction. As a result, the pH of the pore
solutions of hardened cement pastes containing GGBS, FA and SF will be
lower than that for Portland cement alone.
Another type of cement which is gaining popularity in many parts
of the world are the calcium aluminate cements. These cements contain
mainly Al2O3 and CaO, although the proportions vary from product to
product, unlike Portland cement whose composition has gradually evolved
to a relatively uniform composition globally. After hydration, hardened
calcium aluminate cements consist largely of CaO.Al2O3.10H2O and 2CaO.
Al2O3.8H2O, with some Ca3Al2O9.6H2O and amorphous Al(OH)3. The cement
will gradually undergo a process of conversion whereby CaO.Al2O3.10H2O
and 2CaO.Al2O3.8H2O decompose to the last two products. Conversion has
been a cause of concern in the past, since the conversion leads to a loss in
volume of the cement, leading to the formation of porosity and a loss in
strength. However, better understanding of the material means that design
processes can take into account conversion. Moreover, as will be seen in
later chapters calcium aluminate cements can potentially play a role in
achieving resistance to certain forms of biodeterioration.
A variant of calcium aluminate cements are the calcium sulfoaluminate
cements. These cements also contain a quantity of sulfate which means
that the hydration products formed are usually ettringite, monosulfate and
amorphous Al(OH)3 [5].
oxygen (0, 2–), nitrogen (3–, 0, 3+, 5+) and sulphur (6+, 4+, 0, 1–, 2–). A
number of other trace metals also display redox behavior.
The transition from a high oxidation state (e.g., Fe3+) to a lower one
(Fe2+)—reduction—requires the species to accept electrons (e–). This process
can be summarized by the reaction:
Fe3+ + e– ⇌ Fe2+
The Fe3+ here acts as an oxidizing agent—it accepts electrons. This
reaction is referred to as a redox couple, but is not a complete chemical
reaction, since there is no explanation of the origin of the electron. The
electron must come from a reducing agent. In aqueous solution the reducing
agent may be water:
H+ + ¼O2 + e– ⇌ ½H2O
This is also a redox couple, and combining the two gives:
Fe3+ + ½H2O ⇌ Fe2+ + H+ + ¼O2
This is now a complete reaction, with Fe3+ and O2 acting as oxidizing
agents and H2O and Fe2+ acting as reducing agents.
The equilibrium state of a redox couple gives a useful measure of
whether a system is oxidizing or reducing. Since redox reactions involve
transfer of electrons, it is common for this equilibrium to be expressed in
terms of the theoretical voltage associated with the reaction. This voltage is
known as the redox potential (Eh). If we consider a solution containing two
oxidized species (A and B) undergoing reduction to produce two reduced
species (C and D):
aA + bB + ne– ⇌ cC + dD
Eh can be determined using the equation:
a
Eh = E° + RT In [A]c [B] d
b
nF [C] [D]
where E° = the standard potential of the reaction (V);
R = the gas constant (8.3145 J/molK);
T = temperature (K); and
F = the Faraday constant (96,485 J/V g eq).
It is also common for the redox potential to be expressed in terms of
the negative logarithm of the electron concentration (pE):
Eh
pE = – log10[e –] =
0.05916
The redox potential of a solution is dependent on what redox couples are
present, and their concentration. Each redox couple has a characteristic Eh
and the redox potential of the solution will reflect the predominant couple.
20 Biodeterioration of Concrete
If the solution experiences conditions which will alter the Eh—for instance,
the introduction of a reducing agent—the concentration of the predominant
couple will determine the solution’s ability to resist this change, with a
high concentration providing greater resistance. Where there is a large
concentration of a particular redox pair, the solution is said to have a high
redox capacity, or to be ‘well-poised’. Once the capacity of the predominant
pair is exceeded, the redox potential will shift with relative ease to that of
the redox couple with the nearest Eh.
The redox potential can play an important role in determining the
oxidation state of elements present in smaller quantities. For instance, if
chromium is present in small quantities alongside iron in larger quantities,
the redox reaction
Fe2+ + Cr3+ ⇌ Fe3+ + Cr6+
will be established. Thus, the relative proportions of Fe2+ and Fe3+ will
dictate the relative proportions of Cr3+ and Cr6+. This, in turn, has significant
implications for solution chemistry, since—as is the case for chromium—the
oxidation state of an element will often influence its solubility.
Redox potential is also influenced by pH, with the redox potential
rising with increasing pH.
In Portland cement, there are usually only a small number of dissolved
redox pairs present, in relatively small concentrations. These couples include
O2/H2O, Fe3+/Fe2+, SO42–/SO32–.
This means that the pore solution within Portland cement is poorly-
poised. The redox potential of Portland cement is typically around Eh = +100
– +200 mV. However, where quantities of sulfides are present—for instance,
if ground granulated blastfurnace slag has been used in conjunction with
Portland cement—the redox potential may be much lower: –400 mV [6]. In
constructing solubility plots in this chapter, an Eh of +150 mV was used.
Al
Fe(II)
Fe(III)
Fe(OH) 4
–
Fe + 4H2O ⇌ Fe(OH)4 + 4H
3+ – +
–21.588 [7]
Table 2.2 Solubility and molar volume data for compounds relevant to the interaction of
hydrated Portland cement with water.
Aluminium and iron are in many ways very similar elements, and as a result
are able to substitute for each other in the structures of many compounds.
This is certainly true of the AFm and AFt phases of cement. One of the key
distinguishing features is that, whilst both elements can exist in a number
of oxidation states, iron is commonly encountered in two—Fe2+ and Fe3+—
whereas aluminium usually exists in only the Al3+ state.
Figure 2.5 shows the solubility diagram for aluminium. A variety of
complexes are formed between aluminium and hydroxide ions—as shown
in Table 2.1—and most of these are encountered in the diagram. A significant
proportion of this diagram is occupied by the mineral gibbsite (Al(OH)3),
which only becomes soluble under very high and very low pH conditions.
Crystalline gibbsite is, in fact, seldom encountered in hydrated Portland
cement, meaning that it cannot be identified using techniques such as X-ray
diffraction. Instead, an amorphous precipitate is formed.
The solubility diagram for iron is shown in Figure 2.6. As for aluminium,
much of the diagram’s area is occupied by a solid phase, in this case
ferrihydrite. Ferrihydrite is an oxyhydroxide compound with a variable
composition. The official formula ascribed to it by the International Mineral
Association is 5Fe2O3.9H2O, but the water content varies considerably.
Because of this, the formula Fe(OH)3 is often used for simplicity. It has been
proposed that ferrihydrite may, in fact, be a mixture of multiple phases,
although a single phase structure has also been proposed. The reason for
this uncertainty is that ferrihydrite is precipitated as nano-scale particles
and so appears amorphous when studied using X-ray diffraction.
The persistence of ferrihydrite even at relatively low pH means that
concrete and cement undergoing acid attack will, for many acids, develop a
The Chemistry of Concrete Biodeterioration 23
0.--.----------------------------------------------,
Gibbsite (s)
-2
~
Cl -4
0
...J
AI(OH)4-
-6
-8+---,---,---~--~~,_--,---,---,---,---,---,-~
2 3 4 5 6 7 8 9 10 11 12 13 14
pH
0.-----------.-------------------------------,
-2
Qi'
~
Cl
-4 Ferrihydrite (s)
0
...J
Fe'•
-6
Fe(OH);
-8
2 3 4 5 6 7 8 9 10 11 12 13 14
pH
in Table 2.5, along with other solid compounds potentially formed between
calcium, aluminium, iron and sulphate ions.
The solubility diagram of the system comprising calcium, aluminium
and sulphate is shown in Figure 2.7. At high calcium concentrations
the dominant component of the system is portlandite. As the pH falls,
portlandite is replaced by solid gypsum (CaSO4.2H 2O). At calcium
concentrations below the solubility limit of gypsum, CaSO4 in aqueous
solution is predominant, except at high pH, where ettringite is stable. In
the absence of aluminium, the solubility diagram essentially remains the
Table 2.5 Solubility and molar volume data for compounds relevant to the interaction of
hydrated Portland cement with sulphuric acid.
0
ortlandite (s
f::ttri
CaS04 .2H 20 (s) >?.g,i:
('!(.s;
-2
ro
~ -4 CaS04
Cl
0
_J Ca 2+
-6
-8 +--------,~-----.-------,--------,-------.-------~
2 4 6 8 10 12 14
pH
Figure 2.7 Solubility diagram for Ca and Al in the presence of sulphuric acid, with
respect to calcium-bearing species and phases. The concentrations of Al and SO42–
are both fixed at 100 mmol/l.
same, with the exception that ettringite and monosulphate are, of course,
absent. Exchanging iron for aluminium, the solubility diagram retains a
similar form (Figure 2.8).
Figure 2.9 examines the Ca-Al-SO4-H2O system from the perspective
of the predominant forms of Al. Ettringite is, again, evident under high
pH conditions. As for the case when no sulphate is present, gibbsite is
dominant at intermediate pH values, whilst the solid phase Jurbanite
(Al(OH)SO4.5H2O) is, in theory, precipitated under more acidic conditions.
However, in reality, the formation of this phase has not been reported. It has
been proposed that in natural soils, a mixture of amorphous Al(OH)3 and
basaluminite (Al4(OH)10SO4.5H2O) is actually formed [17]. Whether this is
the case for hydrated Portland cement is not known, and it is feasible that
Al3+ and SO42– ions are, in fact, present within the silica gel remaining from
the decalcification of CSH gel.
In the Ca-Fe-SO 4-H 2O system (Figure 2.10), the dominance of
ferrihydrite is somewhat reduced relative to the situation where sulphate
is absent. This is due to the formation of FeSO4 in solution. At very low pH,
and in the absence of oxygen, at least in theory, solid iron (II) sulfide forms.
The Chemistry of Concrete Biodeterioration 27
0
I ~ortland1te (s
rn Ca2 CaS04
~ -4
Cl
0
....J
-6
-8+--l----,------,-------,------,-------,-----~
2 4 6 8 10 12 14
pH
Figure 2.8 Solubility diagram for Ca and Fe in the presence of sulphuric acid,
with respect to calcium-bearing species and phases. The concentrations of Fe and
SO42– are both fixed at 100 mmol/l.
-2
'i-0
.. ,/~
~
Gibbsite (s)
~
Cl
-4
0
_J
Ettringite (s)
Also;
-6
AI(OH)4 -
-8
2 4 6 8 10 12 14
pH
Figure 2.9 Solubility diagram for Al and Ca in the presence of sulphuric acid,
with respect to aluminium-bearing species and phases. The concentrations of
Ca and SO42– are both fixed at 100 mmol/l.
28 Biodeterioration of Concrete
,CD
N(f)
-2 :§:
-6
Fe(OH)4 -
-8
2 4 6 8 10 12 14
pH
Figure 2.10 Solubility diagram for Fe and Ca in the presence of sulphuric acid,
with respect to iron-bearing species and phases. The concentrations of Ca and
SO42– are both fixed at 100 mmol/l.
nitrate
Table 2.7 Stability constants of complexes formed by calcium and iron (III) ions
in water containing nitric acid.
Table 2.8 Solubility and molar volume data for compounds relevant to the interaction of
hydrated Portland cement with nitric acid.
\
\
\ \ Nitrate I Nitrite AFm (s)
-2
\
\
\
ro Ca2 • \
~ -4
Cl \
0
....J \
\
\
\
2 3 4 5 6 7 8 9 10 11 12 13 14
pH
Figure 2.11 Solubility diagram for Ca and Al in the presence of nitric acid
and nitrous acid, with respect to calcium-bearing species and phases. The
concentrations of Ca and NO3–/NO2– are all fixed at 100 mmol/l. Solid line =
nitric acid; dashed line = nitrous acid.
of the nitrate AFm phase has been found in the literature, but its existence
is possible.
Nitrous acid is considerably weaker than nitric acid (Table 2.9), dissociating
to form nitrite ions. It forms even weaker complexes with calcium
than nitric acid, and does not form complexes with aluminium or iron
(Table 2.10). In the presence of calcium, nitrite ions can form the salt calcium
nitrite monohydrate (Table 2.11), although this compound is relatively
soluble. Thus, as for nitric acid, Figures 2.4–2.6 act as solubility diagrams
for nitrous acid with single metal ions. A nitrite AFm phase also exists, with
similar characteristics to the nitrate AFm phase. A calcium/aluminium/
nitrite solubility diagram is also plotted in Figure 2.11, alongside that for
nitrate. As for nitrate AFm, it is possible that an iron-bearing form of the
nitrite AFm phase exists, but this has yet to be documented.
nitrite
Ca
Table 2.11 Solubility and molar volume data for compounds relevant to the interaction of
hydrated Portland cement with nitrous acid.
and the temperature. This dissolved CO2 is often referred to as carbonic acid
(H2CO3), although in reality it will be dissociated to bicarbonate or carbonate
ions to a degree determined by pH and temperature. The dissociation
constants for carbonic acid at 20°C are given in Table 2.12.
The relationship between the concentration of dissolved carbonic acid
—[H2CO3]—and the partial pressure of CO2–PCO2—is given by the equation:
0
"{ortlandite (s)
-2
Calcite (s)
ro
~ -4
~"'---
Cl
0
_J Ca'·
+
00
-6
I
0 "' CaC03
-8
2 4 6 8 10 12 14
pH
Figure 2.12 Solubility diagram for Ca in the presence of carbonic acid.
14
12 1:---- __
\
10 I \
\
\
--- -- ------
I 8 I
c.
\
ll
Water + 0.001 moles portlandite
Water
4
I
2 ,_-------,--------,--------,--------,-------~
0.0 0.2 0.4 0.6 0.8 1.0
Figure 2.13 Change in pH of water versus the partial pressure of CO2 in contact
with it for two scenarios: (i) pure water; (ii) water in contact with portlandite.
Formic acid is a carboxylic acid with the formula HCOOH, making it the
simplest of carboxylic acid compounds. It is monoprotic and is relatively
weak (Table 2.15). It is capable of forming complexes with Ca, Al and
Fe(III) ions (Table 2.16). Although the acid mainly forms weak complexes,
Al3(OH)2(CHO2)6+ is the exception to this. Formic acid forms salts with Ca, Al
and Fe(II), all of which are relatively soluble in water (Table 2.17). Aluminium
formate and iron (III) formate have previously been thought to have the
formulae Al(CHO2)3∙3H2O and Fe(CHO2)3∙H2O respectively. However, a
recent crystallographic study has concluded that the compounds have the
formulas Al(CHO2)3(CH2O2)0.25(CO2)0.75∙0.25H2O and Fe(CHO2)3(CH2O2)0.25
(CO2)0.75∙0.25H2O, where CH2O2 is a fully protonated formic acid molecule
The Chemistry of Concrete Biodeterioration 35
Table 2.14 Solubility and molar volume data for compounds relevant to the interaction of
hydrated Portland cement with carbonic acid.
4OH– + 24H2O
-2 Portlandite (s)
Calcite (s)
Monocarbonate (s)
ro
~ -4
Ol
0
...J Ca2-
OM
-6 u
I
u"' CaC03
-8
2 4 6 8 10 12 14
pH
Figure 2.14 Solubility diagram for Ca and Al in the presence of carbonic acid,
with respect to calcium-bearing species and phases. The concentrations of Al
and CO32– are fixed at 100 mmol/l.
Monocarbonate (s)
Gibbsite (s)
-2
<(
Ol -4
0
...J Al3 +
AI(OHJ;
-6
2 3 4 5 6 7 8 9 10 11 12 13 14
pH
Figure 2.15 Solubility diagram for Ca and Al in the presence of carbonic acid, with
respect to aluminium-bearing species and phases. The concentrations of Ca and
CO32– are fixed at 100 mmol/l.
The Chemistry of Concrete Biodeterioration 37
Calcite (s)
-2 Portlandite (s)
-6
0
(.)
:r:
"'
(.) CaC0 3
-8
2 4 6 8 10 12 14
pH
Figure 2.17 Solubility diagram for Ca and Fe in the presence of carbonic acid,
with respect to calcium-bearing species and phases. The concentrations of Fe
and CO32– are fixed at 100 mmol/l.
38 Biodeterioration of Concrete
Siderite (s)
-2
Monocarbonate (Fe) (s)
Q) Ferrihydrite (s)
~ -4
0
_J
-6
2 3 4 5 6 7 8 9 10 11 12 13 14
pH
Figure 2.18 Solubility diagram for Ca and Fe in the presence of carbonic acid,
with respect to iron-bearing species and phases. The concentrations of Ca and
CO32– are fixed at 100 mmol/l.
Table 2.16 Stability constants of complexes formed by calcium, aluminium and iron
(III) ions in water containing formic acid.
Table 2.17 Solubility and molar volume data for compounds relevant to the interaction of
hydrated Portland cement with formic acid.
+ 0.75CO2(g) + 0.25CH2O2 +
0.25H2O
+ 0.75CO2(g) + 0.25CH2O2 +
0.25H2O
Like formic acid, acetic acid is a carboxylic acid. The acetic acid molecule
contains two carbon and is slightly weaker than formic acid (Table 2.18).
The acetate ion forms weak complexes with Ca, Al and Fe(II) ions,
but forms stronger complexes with Fe(III) (Table 2.19). Many of the salts
formed by acetic acid with Ca, Al and Fe are soluble (Table 2.20). Aluminium
40 Biodeterioration of Concrete
\ ~
2
'0
c
~"'
o._
-2
ro
~ -4
Cl
0
....J
Ca2 + Ca(Formatet
\
'I
0
-6 "'
(.)
2 3 4 5 6 7 8 9 10 11 12 13 14
pH
Figure 2.19 Solubility diagram for Ca in the presence of formic acid.
0.--.----------------------------------------------~
Gibbsite (s)
-2
:l
Cl -4
AI3 (0H)2(Formate)6 '
0
....J
A I(OH)4 -
-6
-8 +---~-------,---.--L,----,---.---,---,----,---.--~
2 3 4 5 6 7 8 9 10 11 12 13 14
pH
H
acetate
H O
Table 2.20 Solubility and molar volume data for compounds relevant to the interaction of
hydrated Portland cement with acetic acid.
stable complexes with Fe (III), but, again, this has a negligible influence over
the manner in which solid phases are precipitated (Figure 2.23).
~
~
Jll
uc
'2 "'
0
-2 ()_
rn
Q
Ol
0
_J
-4
Ca2 + Ca(Acetate(
1\
+I
0
-6 "'
0
2 3 4 5 6 7 8 9 10 11 12 13 14
pH
0,--,----------------------------------------~
Gibbsite (s)
-2
~
Ol -4
0
_J
AI( Acetate)'•
AI(OH)4-
-6
2 3 4 5 6 7 8 9 10 11 12 13 14
pH
this extends to their behaviour in solution with metal ions. Thus, wherever
possible, constants relating to lactic acid have been obtained for the L-form.
Lactic acid contains two –OH groups, but only the carboxylate group
normally undergoes deprotonation (Table 2.21). However, interaction with
44 Biodeterioration of Concrete
0,-----------.---------------------------------,
-2
-6 Fe(Acetatet
2 3 4 5 6 7 8 9 10 11 12 13 14
pH
Table 2.22 Stability constants of complexes formed by calcium and aluminium ions in water
containing lactic acid.
Al
Al3+ + CH3CH(OH)CO2– ⇌ Al(CH3CH(OH)CO2)2+ 1.21
Al3+ + 2CH3CH(OH)CO2– ⇌ Al(CH3CH(OH)CO2)2+ 2.72 [47]
Al + 3CH3CH(OH)CO2 ⇌ Al(CH3CH(OH)CO2)3
3+ –
4.92 (DL-Lactate)
Table 2.23 Solubility and molar volume data for compounds relevant to the interaction of
hydrated Portland cement with lactic acid.
0.--.----------------------------------------------~
Gibbsite (s)
-2
~
Cl -4
.3 AI(CH 3CHOC02)(Lactate) AI(OH);
-6
-8+---.---.---.---.---.---.---+---.---.---.---.-~
2 3 4 5 6 7 8 9 10 11 12 13 14
pH
Figure 2.24 Solubility diagram for Al in the presence of lactic acid.
Oxalic acid—C2O4H2—is a relatively strong acid that can shed two protons
from two carboxylate groups. The acid dissociation constants of oxalic acid
are provided in Table 2.27. The stability constants of complexes formed
between the acid and Ca and Fe ions are provided in Table 2.28.
Oxalic acid is capable of forming a number of salts with calcium of
relatively low solubility, the only difference in composition being the
amount of associated water of crystallization. Three of these compounds are
The Chemistry of Concrete Biodeterioration 47
Table 2.25 Stability constants of complexes formed by calcium and iron(II) and (III)
ions in water containing glycolic acid.
Table 2.26 Solubility and molar volume data for compounds relevant to the interaction of
hydrated Portland cement with glycolic acid.
0
Ca2 •
\ ~
l!l
'6
c:
"'
t:
0
-2 D..
rn
~ -4
Cl
0
__J
Ca(Giycolatet
1\
+I
0
-6 u "'
-8+---~-,---,---.--,---,---,--,---,------~~
2 3 4 5 6 7 8 9 10 11 12 13 14
pH
Ferrihydrite (s)
-2
/
Qj'
~
-4
Ol
0
__J Fe2-+
Fe(CH,OCO, ), (Giycolate( ~
'i:
0
Q)
lJ._
-6
-8+---.-~.---.---.---.---.---.---.---.---.---.-L-4
2 3 4 5 6 7 8 9 10 11 12 13 14
pH
Table 2.28 Stability constants of complexes formed between the oxalate ion relevant to cement
chemistry.
AlOH(Oxalate)2 2–
Al + 2C2O
3+
4
2–
+ H2O ⇌ AlOH(C2O ) 4 2
2–
+H
+
6.8 [7]
The low solubility of calcium oxalate means that this solid phase
persists even at low pH (Figure 2.27). Whilst the formation of complexes
with oxalate ions destabilizes gibbsite and ferrihydrite, the low solubility
of the Al and Fe oxalate salts means that solid phases, again, are present in
significant quantities—at least at higher metal concentrations—across the
full pH range (Figures 2.28 and 2.29).
Oxalic acid is one of the few organic compounds known to form
complexes with silicon. However, the complex is weak, having a stability
constant for the reaction H4SiO4 + C2O42– ⇌ Si(C6H5O7)(OH)42– of 0.04 [61].
0,-----------------------------------~~~~~
'\{ortlandite (s)
Ca(Oxalate).H 2 0 (s)
Ca(Oxalate)t
ro
2.
Cl
-4
.3
-6
-8+-------.------.-------,------.-------,-----~
2 4 6 8 10 12 14
pH
Figure 2.27 Solubility diagram for Ca in the presence of oxalic acid.
Gibbsite (s)
Al 2 (0xalate)3 .4H,O (s)
-2
<(
Cl -4
0
_J
AI(Oxalate)/ -
AI(OH)4-
-6
AI(OH) 2 (0xalater
-8
2 4 6 8 10 12 14
pH
Figure 2.28 Solubility diagram for Al in the presence of oxalic acid.
52 Biodeterioration of Concrete
0
Fe 3 (0H) 6 (s)
Fe 2 (0xalate) 3 .5H,O (s)
-2
Fe(OH) 4-
-8
2 4 6 8 10 12 14
pH
Only the salt calcium pyruvate is known to form (Table 2.32), and
quantitative data for this compound is seemingly lacking. However, on the
assumption that calcium pyruvate is slightly soluble, a solubility of 5 g/l
has been assumed, leading to the solubility diagram shown in Figure 2.30,
indicating predominance of this compound at higher Ca concentrations
over a relatively wide pH range.
Hydrogen Succinate
succinate
54 Biodeterioration of Concrete
ro
~ -4
Cl
[\
0 Ca2•
__J Ca(Pyruvate)
()
"'0
I+
-6
-8 +---~-------.---.---.---.---.---.---.---..-~--~
2 3 4 5 6 7 8 9 10 11 12 13 14
pH
Malic acid possesses two carboxylate groups and a hydroxyl group, with
only the carboxylate groups normally undergoing deprotonation (Table 2.36).
It forms an extremely varied series of complexes with aluminium and iron
(III) (Table 2.37) where further deprotonation occurs. It forms iron (II) and
calcium salts. Whilst there is little data on several of these, the dihydrate
and trihydrate calcium salts are known to be of relatively low solubility
(Table 2.38) leading to its predominance at higher calcium concentrations
on the calcium solubility diagram (Figure 2.32).
Tartaric acid is a chiral compound, with the levotartaric (L-) form being
the naturally occurring form. In addition, the mirror-image molecule,
dextrotartaric acid (D-) and mesotartaric acid can be synthesized artificially.
The molecule possesses two carboxylate groups, and two hydroxyl groups,
although it is only the carboxylate groups which normally undergo
deprotonation (Table 2.39).
Calcium forms complexes with both the partly and fully deprotonated
molecule, whilst aluminium and iron only form complexes once the
molecule is fully deprotonated. Calcium hydrogen tartrate and calcium
The Chemistry of Concrete Biodeterioration 55
Table 2.34 Stability constants of complexes formed between the succinate ion relevant to
cement chemistry.
Ca
Al
Fe(II)
Fe(III)
\ l
0
"~·'fe
f'.s;
Ca(Succinate).3H 2 0 (s)
-2
---1
Cii'
~ -4
0)
0 Ca2 +
\
...J
Ca(Succinate) ·:r:
0
0"'
-6
-8 +----,---,---,L--,,---,---,----,---,---,---,----r-~
2 3 4 5 6 7 8 9 10 11 12 13 14
pH
tartrate, but on the assumption that it is soluble in water, tartaric acid has
the effect of solubilizing gibbsite under lower pH conditions (Figure 2.34).
In the case of iron, both the Fe(II) and Fe(III) salts are of low solubility
(Table 2.41). However, it is the Fe(II) salt which has a significant effect on
the solubility diagram (Figure 2.35)—it leads to a persistence of a solid iron
phase to a low pH for higher iron concentrations.
Fumaric acid is structurally similar to succinic acid, with the exception that
the central carbon-carbon bond in its molecule is a double bond (Table 2.45).
The lack of complexes identified as being formed by fumaric acid (Table
2.46) implies that this structural difference compromises the compound’s
ability to do so. It must be stressed that the absence of data partly indicates
either a lack of experimental investigation into this compound or practical
difficulties in conducting measurements. However, it is likely that the
complexes formed are weak: using the same stability constant prediction
method used for pyruvic acid, the estimated stability constant values are
2.3, 2.8 and 3.2 for 1:1 complexes of Al, Fe(II) and Fe(III).
Whilst the salts formed by fumaric acid are only sparingly to slightly
soluble (Table 2.47), they do not appear as predominant species using the
parameters used to form solubility plots in this chapter.
Table 2.37 Stability constants of complexes formed between the malate ion relevant to cement
chemistry.
Ca
Al
Al2(Malate H ) +
–2
2Al + 2[C4H4O5] ⇌ Al2[C4H2O5][C4H3O5]
3+ 2– –
1.78
(Malate H+–1) –
+ 3H+ [69]
Al2(Malate H+–2)22– 2Al3+ + 2[C4H4O5]2– ⇌ Al2[C4H2O5]22– + 4H+ –4.46
Fe(II)
Fe(III)
Fe2(Malate H ) +
–1 2
2Fe + 3[C4H4O5] ⇌
3+ 2–
17.85 [70]
(Malate)2– Fe2[C4H3O5]2[C4H4O5]2– + 2H+
~
Ca(Malate) 3H2 0 (s)
"IJo:·
''&(&;
-2
ro
~ -4
OJ
0 Ca2+
_J
Ca(Malate)
-6 .___ .6
"'
()
-8
2 3 4 5 6 7 8 9 10 11 12 13 14
pH
O - -
O O
O OH O O
O OH O OH
Table 2.40 Stability constants of complexes formed by calcium, aluminium and iron(II) and
(III) ions in water containing tartaric acid.
(Table 2.48), but the hydroxyl groups may also deprotonate in the presence
of iron(III) (Table 2.49). The salts of gluconic acid are soluble (Table 2.50).
However, the calcium salt is close to the edge of this category and, whilst
it does not appear on a solubility diagram using the parameters adopted
in this chapter, could be precipitated at higher acid concentrations than
the 0.1 mol/l used.
0,-----------------------------------~--------,
Ca(Tartrate).4Hp (s)
~
-2 \
ca>•/'-.l----
----------------------------------------1
rn
~ -4
Cl
0
...J Ca(Tartrate)
-6
-8 +--LT---,---,---,--,---,---,---,---,---,---.--~
2 3 4 5 6 7 8 9 10 11 12 13 14
pH
0,--,------------------------------------------,
Gibbsite (s)
-2
<r:
Cl -4
0
...J
AI(Tartrate) 2-
AI(OH)4 -
-6
-8 +---.----,---,---.---.~L-.---,---,---,----,---,--~
2 3 4 5 6 7 8 9 10 11 12 13 14
pH
Table 2.41 Solubility and molar volume data for compounds relevant to the interaction of
hydrated Portland cement with tartaric acid.
CH(OH)CO22–
Iron (II) Fe(OOCCH(OH)CH(OH) –8.23 [78] unknown –
tartrate CO2)∙2.5H2O ⇌ Fe2+ +
OOCC(OH)CH(OH)CO22– +
2.5H2O
Iron (III) Fe2(OOCCH(OH)CH(OH) –14.525* unknown –
tartrate CO2)3∙H2O ⇌ 2Fe3+ +
3OOCC(OH)CH(OH)CO22– +
H 2O
*based on solubility data provided by a number of chemical suppliers, but original source
unclear.
0.-----------.-------------------------------,
-2
Fe(II)(Tartrate)2.5H 20 (s)
Q) Ferrihydrite (s)
~ -4
0
-'
-B Fe 2
Fe(II)(Tartrate)
2 4 5 6 7 8 9 10 11 12 13 14
pH
Figure 2.35 Solubility diagram for Fe in the presence of tartaric acid.
The Chemistry of Concrete Biodeterioration 63
Table 2.44 Solubility and molar volume data for compounds relevant to the interaction of
hydrated Portland cement with butyric acid.
pKa2 pKa1
Table 2.47 Solubility and molar volume data for compounds relevant to the interaction of
hydrated Portland cement with fumaric acid.
Table 2.50 Solubility and molar volume data for compounds relevant to the interaction of
hydrated Portland cement with gluconic acid.
Table 2.53 Solubility and molar volume data for compounds relevant to the interaction of
hydrated Portland cement with propionic acid.
Ca + C6H5O7 + H ֖ Ca(C6H6O7)
2+ 3– +
9.26
Ca + C6H5O73– + 2H ֖ CaH(C6H7O7)
2+ +
12.26
[33]
Al Al3+ + C6H5O73– ֖ Al(C6H5O7) 9.97
Fe + C6H5O7 + H ֖ Fe(C6H6O7)
3+ 3– + +
14.4
Table 2.56 Solubility and molar volume data for compounds relevant to the interaction of
hydrated Portland cement with citric acid.
Ca2• \
Ca3 (Citrate) 2 .4H,O (s)
~C/ite (sj
~
(
-2
ro
~ -4
Ol + Ca(Citrater
0
_l 2 2
~
Q_
"'
.l::
i:3
I N I'
-6 "'
0 "'
0
-8
2 4 6 8 10 12 14
pH
Figure 2.36 Solubility diagram for Ca in the presence of citric acid.
The Chemistry of Concrete Biodeterioration 69
-8+---L---,---L---,-------.-~----.-------.-----~
2 4 6 8 10 12 14
pH
Ferrihydrite (s)
-2
~ :§'
C) -4
0
...J
~
0 Fe(Citrater
:C
())
lL
-6
Fe(OHJ;
-8
2 4 6 8 10 12 14
pH
Various factors will play a role, including the solubility of the salt and the
pH range over which it remains stable. As a result, salt formation requires
individual investigation for specific acids. Where data are available, this
is attempted in Chapter 4, which looks at fungal deterioration of concrete.
This is because fungi are largely responsible for the production of acids
involved for this type of deterioration.
2.6 References
[1] Drever JI (1997) The Geochemistry of Natural Waters. 3rd Ed., Prentice Hall, Upper
Saddle River.
[2] Davies CW (1962) Ion Association. Butterworths, Washington.
[3] Truesdell AH and Jones BF (1973) WATEQ, a computer program for calculating chemical
equilibria of natural waters. US Geological Survey Journal of Research 2: 233–248.
[4] Chen JJ, Thomas J, Taylor HFW and Jennings HM (2004) Solubility and structure of
calcium silicate hydrate. Cement and Concrete Research 34(9): 1499–1519.
[5] Winnefeld F and Lothenbach B (2010) Hydration of calcium sulfoaluminate cements—
experimental findings and thermodynamic modelling. Cement and Concrete Research
40(8): 1239–1247.
[6] Glasser FP (1992) Chemistry of cement-solidified waste forms. pp. 1–39. In: Spence RD
(ed.). Chemistry and Microstructure of Solidified Waste Forms. Lewis Publishers, Boca
Raton, FL, USA.
[7] U.S. Environmental Protection Agency (1999) MINTEQA2/PRODEFA2, A Geochemical
Assessment Model for Environmental Systems—User Manual Supplement for Version
4.0. National Exposure Research Laboratory, Athens, GA, USA.
[8] Haynes WM (2014) CRC Handbook of Chemistry and Physics. 95th Ed., CRC Press,
Boca Raton, Florida, USA.
[9] Marshall WL and Jones EV (1966) Second dissociation constant of sulfuric acid from 25
to 350° evaluated from solubilities of calcium sulfate in sulfuric acid solutions. Journal
of Physical Chemistry 70: 4028–4040.
[10] Damidot D and Glasser FP (1993) Thermodynamic investigation of the CaO-Al2O3-
CaSO4-H2O system at 25ºC and the influence of Na2O. Cement and Concrete Research
23: 221–238.
[11] Balonis M and Glasser FP (2009) The density of cement phases. Cement and Concrete
Research 39: 733–739.
[12] Roberts WL (1990) Encyclopedia of Minerals. 2nd Ed., Van Nostrand Reinhold.
[13] Möschner G, Lothenbach B, Rose J, Ulrich A, Figi R and Kretzschmar R (2008) Solubility
of Fe–ettringite (Ca6[Fe(OH)6]2(SO4)3.26H2O). Geochimica et Cosmochimica Acta 72:
1–18.
[14] Dilnesa BZ, Lothenbach B, Renaudin G, Wichser A and Kulik D (2014) Synthesis and
characterization of hydrogarnet Ca3(AlxFe1 − x)2(SiO4)y(OH)4(3 − y). Cement and Concrete
Research 59: 96–111.
[15] Singh SS (1980) Thermodynamic properties of synthetic basic aluminite
[Al4(OH)10SO4.5H2O] from solubility data. Canadian Journal of Soil Science 60: 381–384.
[16] van Breemen N (1973) Dissolved aluminum in acid sulfate soils and in acid mine waters.
Soil Science Society of America Proceedings 37: 694–697.
[17] Jones AM, Collins RN and Waite TD (2011) Mineral species control of aluminum
solubility in sulfate-rich acidic waters. Geochimica et Cosmochimica Acta 75: 965–977.
[18] Kolthoff I (1959) Treatise on Analytical Chemistry. Interscience Encyclopedia, New
York.
[19] Bailar JC and Trotman-Dickenson AF (1973) Comprehensive Inorganic Chemistry.
Pergamon Press, Oxford.
The Chemistry of Concrete Biodeterioration 73
[20] Balonis M (2010) The influence of inorganic chemical accelerators and corrosion
inhibitors on the mineralogy of hydrated portland cement systems. Ph.D. thesis,
University of Aberdeen.
[21] Renaudin G and Francois M (1999) The lamellar double-hydroxide (LDH) compound
with composition 3CaO.Al2O3.Ca(NO3)2.10H2O. Acta Crystallographica C 55: 835–838.
[22] Speight J (2005) Lange’s Handbook of Chemistry. 16th Ed., McGraw-Hill, London.
[23] Ribar B, Divjakovic V, Herak R and Prelesnik B (1973) A new crystal structure study of
Ca(NO3)2.4H2O. Acta Crystallographica B 29: 1546–1548.
[24] Herpin P and Sudarsanen K (1965) Crystal structure of aluminum nitrate hydrate.
Bulletin de la Societe Francaise de Mineralogie et de Cristallographie 88: 595–601.
[25] Schmidt H, Asztalos A, Bok F and Voigt W (2012) New iron(III) nitrate hydrates:
Fe(NO3)3.xH2O with x = 4, 5 and 6. Acta Crystallographica C 68: i29–i33.
[26] Park J-Y and Lee Y-N (1988) Solubility and decomposition kinetics of nitrous acid in
aqueous solution. Journal of Physical Chemistry 92: 6294–6302.
[27] Brooker MH and DeYoung BS (1980) Measurement of the stability constants for aqueous
calcium and magnesium nitrite by Raman spectroscopic methods. Journal of Solution
Chemistry 9: 279–288.
[28] Plummer LN and Busenberg E (1982) The solubilities of calcite, aragonite and vaterite
in CO2-H2O solutions between 0 and 90°C, and an evaluation of the aqueous model of
the system CaCO3-CO2-H2O. Geochimica and Cosmochimica Acta 46: 1011–1040.
[29] Rodriguez-Blanco JD, Shaw S and Benning LG (2011) The kinetics and mechanisms of
amorphous calcium carbonate (ACC) crystallization to calcite, via vaterite. Nanoscale
3: 265–271.
[30] Graf DL (1961) Crystallographic tables for the rhombohedral carbonates. American
Mineralogist 46: 1283–1316.
[31] Reardon EJ (1990) An ion interaction-model for the determination of chemical-equilibria
in cement water-systems. Cement and Concrete Research 20: 175–192.
[32] Lothenbach B, Matschei T, Möschner G and Glasser FP (2008) Thermodynamic modelling
of the effect of temperature on the hydration and porosity of Portland cement. Cement
and Concrete Research 38: 1–18.
[33] Martell AE and Smith RM (2001) Critical Selected Stability Constants of Metal Complexes
Database, Version 6.0 for Windows; National Institute of Standards and Technology,
April.
[34] International Union of Pure and Applied Chemistry, IUPAC-NIST Solubility Database,
Version 1.1, National Institute of Science and Technology, 2015. http://srdata.nist.gov/
solubility/index.aspx.
[35] Nitta I and Osaki K (1948) Crystal structure of calcium formate. X-sen Osaka University
5: 37–42.
[36] Tian Y-Q, Zhao Y-M, Xu H-J and Chi C-Y (2007) CO2 template synthesis of metal formates
with a ReO3 net. Inorganic Chemistry 46: 1612−1616.
[37] Weber G (1980) Iron(II) formate dihydrate. Acta Crystallographica B 36: 3107–3109.
[38] Pöllmann H, Kaden R and Stöber S (2014) Crystal structures and XRD data of new
calcium aluminate cement hydrates. pp. 75–85. In: Fentiman CH, Mangabhai RJ and
Scrivener KL (eds.). Calcium Aluminates: Proceedings of the International Conference
2014. IHS BRE Press.
[39] Saury C, Boistelle R, Dalemat F and Bruggeman J (1993) Solubilities of calcium acetates
in the temperature range 0–100°C. Journal of Chemical Engineering Data 38: 56–59.
[40] Klop EA, Schouten A, van der Sluis P and Spek AL (1984) Structure of calcium acetate
monohydrate, Ca(C2H3O2)2.H2O. Acta Crystallographica C 40: 51–53.
[41] Balarew Chr, Stoilova D and Demirev L (1974) Untersuchung einiger dreistoffsysteme
vom typ Me(OCOCH3)2–CH3COOH–H2O bei 25°C (Me = Ni, Co, Mg, Mn, Ca). Zeitschrift
für Anorganische und Allgemeine Chemie 410: 75–87.
[42] Klop EA and Spek AL (1984) Structure of calcium hydrogen triacetate monohydrate,
CaH(C2H3O2)3.H2O. Acta Crystallographica C 40: 1817–1819.
74 Biodeterioration of Concrete
[43] Lewis RJ (2007) Hawley’s Condensed Chemical Dictionary. 15th Ed., Wiley-Interscience,
Hoboken NJ, USA.
[44] Alcala R and Garcia JF (1973) Revista de la Academia de Ciencias Exactas. Fisico-
Quimicas y Naturales de Zaragoza 28: 303.
[45] Hamada YZ, Carlson B and Dangberg J (2005) Interaction of malate and lactate with
chromium(III) and iron(III) in aqueous solutions. Synthesis and Reactivity in Inorganic,
Metal-Organic and Nano-Metal Chemistry 35: 515–522.
[46] Dawson RMC (1959) Data for Biochemical Research. Clarendon Press, Oxford.
[47] Marklund E, Sjöberg S, Öhman L-O, Salvatore F, Niinistö L and Volden HV (1986)
Equilibrium and structural studies of silicon(IV) and aluminium(III) in aqueous solution.
14. Speciation and equilibria in the aluminium(III)lactic acid-OH-system. Acta Chemica
Scandinavica 40: 367–373.
[48] Apelblat A, Manzurola E, van Krieken J and Nanninga GL (2005) Solubilities and
vapour pressures of water over saturated solutions of magnesium-l-lactate, calcium-l-
lactate, zinc-l-lactate, ferrous-l-lactate and aluminum-l-lactate. Fluid Phase Equilibria
236: 162–168.
[49] Bombi GG, Corain B, Sheikh-Osman AA and Giovanni C Valle (1990) The speciation
of aluminum in aqueous solutions of aluminum carboxylates. Part I. X-ray molecular
structure of Al[OC(O)CH(OH)CH3]3. Inorganica Chimica Acta 171: 79–83.
[50] Adu-Wusu K (2012) Literature review on impact of glycolate on the 2H evaporator and
the effluent treatment facility (ETF). Savannah River National Laboratory, Aiken SC,
USA.
[51] Sillen LG and Martell AE (1971) Stability Constants of Metal-Ion Complexes, Supplement
No. 1. The Chemical Society, London.
[52] Martell AE and Smith RM (1989) Critical Stability Constants, Vol. 6 Second Supplement.
Plenum, London.
[53] Martell AE and Smith RM (1977) Critical Stability Constants, Vol. 3 Other Organic
Ligands. Plenum, London.
[54] Streit J, Tran-Ho LC and Koenigsberger E (1998) Solubility of calcium oxalate hydrates
in sodium chloride solutions and urine-like liquors. Monatshefte für Chemie 129:
1225–1236.
[55] Weast RC, Astel MJ and Beyer WH (1986) CRC Handbook of Chemistry and Physics.
67th Ed., CRC Press, Boca Raton, Florida, USA.
[56] Christodoulou E, Panias D and Paspaliaris I (2001) Calculated solubility of trivalent
iron and aluminum in oxalic acid solutions at 25°C. Canadian Metallurgical Quarterly
40: 421–432.
[57] Hochrein O, Thomas A and Kniep R (2008) Revealing the structure of anhydrous calcium
oxalate, Ca[C2O4], by a combination of atomistic simulation and Rietveld refienement.
Zeitschrift Anorganische Allgemeine Chemie 634: 1826–1829.
[58] Tazzoli V and Domeneghetti C (1980) The crystal structures of whewellite and
wheddellite: re-examination and comparison. American Mineralogist 65: 327–334.
[59] Deganello S, Kampf AR and Moore PB (1981) The crystal structure of calcium oxalate
trihydrate: Ca(H2O)3(C2O4). American Mineralogist 66: 859–865.
[60] Echigo T and Kimata M (2008) Single crystal X-ray diffraction and spectroscopic studies
on humboldtine and lindbergite: weak Jahn-Teller effect of Fe2+ ion. Physical Chemistry
of Minerals 35: 467–475.
[61] Öhman LO, Nordin A, Sedeh IF and Sjöberg S (1991) Equilibrium and structural studies
of silicon(IV) and aluminium(III) in aqueous solution. 28. Formation of soluble silicic
acid–ligand complexes as studied by potentiometric and solubility measurements. Acta
Chemica Scandinavica 45: 335–341.
[62] Hancock RD and Martell AE (1989) Ligand design for selective complexation of metal
ions in aqueous solution. Chemical Reviews 89(8): 1875–1914.
[63] Panel on Food Additives and Nutrient Sources, Calcium acetate, calcium pyruvate,
calcium succinate, magnesium pyruvate magnesium succinate and potassium malate
The Chemistry of Concrete Biodeterioration 75
added for nutritional purposes to food supplements. The EFSA Journal 1088: 2009, pp.
1–25.
[64] Alliey N, Venturini-Soriano M and Berthon G (1996) Aluminum-succinate complex
equilibria and their potential implications for aluminum metabolism. Annals of Clinical
and Laboratory Science 26: 122–132.
[65] Karipides A and Reed AT (1980) The structure of diaquasuccinatocalcium(II)
monohydrate. Acta Crystallographica B 36: 1377–1381.
[66] Mathew M, Takagi S, Fowler BO and Markovic M (1994) The crystal structure of calcium
succinate monohydrate. Journal of Chemical Crystallography 24: 437–440.
[67] Xu T-G, Xu D-J, Wu J-Y and Chiang MY (2002) catena-Poly[[tetraaquairon(II)]-[mu]-
succinato-[kappa]2O:O’]. Acta Crystallographica C 58: m615–m616.
[68] Kim Y-J and Jung D-Y (1999) Fe5(OH)2(C4H4O4)4 : hydrothermal synthesis of microporous
iron(II) dicarboxylate with an inorganic framework. Bulletin Korean Chemical Society
20: 830.
[69] Berthon G (2002) Aluminium speciation in relation to aluminium bioavailability,
metabolism and toxicity. Coordination Chemistry Reviews 228(2): 319–341.
[70] Timberlake CF (1964) Iron–malate and iron–citrate complexes. Journal of the Chemical
Society 5078–5085.
[71] Bränden C-I and Söderberg B-O (1966) The crystal structure of Ca-malate-dihydrate,
CaC4H4O5.2H2O. Acta Chemica Scandinavica 20: 730–738.
[72] Lenstra TH and Van Havere W (1980) Calcium di(hydrogen 1-malate) hexahydrate.
Acta Crystallographica B36: 156–158.
[73] Schmittler H (1968) Cell dimensions of some salts of malic acid. Acta Crystallographica
B24: 983–984.
[74] Bertron A and Duchesne J (2013) Attack of cementitious materials by organic acids
in agricultural and agrofood effluents. pp. 131–173. In: Alexander M, De Belie N and
Bertron A (eds.). Performance of Cement—based Materials in Aggressive Aqueous
Environments. RILEM State-of-the-Art Report TC 211-PAE. Springer, Dordrecht,
Netherlands.
[75] Hawthorn FC, Borys I and Ferguson RB (1982) Structure of calcium tartrate tetrahydrate.
Acta Crystallographica B 38: 2461–2463.
[76] Seidell A (1919) Solubilities of Inorganic and Organic Compounds. 2nd Ed., Van
Nostrand, New York.
[77] O’Neil MJ (2001) The Merck Index: An Encyclopedia of Chemicals, Drugs, and
Biologicals. 13th Ed., Merck, Whitehouse Station, NJ, USA.
[78] Cayot P, Guzun-Cojocaru T and Cayot N (2013) Iron fortification of milk and dairy
products. pp. 75–89. In: Preedy VR, Srirajaskanthan R and Patel VB (eds.). Handbook
of Food Fortification and Health: From Concepts to Public Health Applications. Volume
1 Springer, New York.
[79] Valora A, Reguera E and Sanchez-Sinencio F (2002) Synthesis and X-ray diffraction
study of calcium salts of some carboxylic acids. Powder Diffraction 17: 13–18.
[80] Weast R (1966) CRC Handbook of Chemistry and Physics. 47th Ed., Chemical Rubber
Co, Cleveland, OH, USA.
[81] Gupta MP, Prasad SM, Sahu RG and Sahu BN (1972) The crystal structure of calcium
fumarate trihydrate CaC4H2O4.3H2O. Acta Crystallographica B28: 135–139.
[82] Pallagi A, Tasi ÁG, Peintler G, Forgo P, Pálinkó I and Sipos P (2013). Complexation of
Al(III) with gluconate in alkaline to hyperalkaline solutions: formation, stability and
structure. Dalton Transactions 42(37): 13470–13476.
[83] Pecsok RL and Sandera J (1955) The gluconate complexes. II. The ferric-gluconate
system. Journal of the American Chemical Society 77(6): 1489–1494.
[84] Vavrusova M, Munk MB and Skibsted LH (2013) Aqueous solubility of calcium L-lactate,
calcium D-gluconate, and calcium D-lactobionate: importance of complex formation
for solubility increase by hydroxycarboxylate mixtures. Journal of Agricultural and
Food Chemistry 61(34): 8207−8214.
76 Biodeterioration of Concrete
[85] Cicek V (2012) Characterization studies of mild steel alloy substrate surfaces treated by
oxyanion esters of α-hydroxy acids and their salts. International Journal of Chemical
Science and Technology 2(3): 244–260.
[86] Marklund E, Öhman L-O and Sjöberg S (1989) Equilibrium and structural studies of
silicon(IV) and aluminium(III) in aqueous solution. 20. Composition and stability of
aluminium complexes with propionic acid and acetic acid. Acta Chemica Scandanavica
43: 641–646.
[87] Wise DL, Trantolo DJ, Lewandrowski K-U, Gresser JD and Cattaneo MV (2000)
Biomaterials Engineering and Devices: Human Applications: Volume 2. Orthopedic,
Dental, and Bone Graft Applications. Springer, Dordrecht, Netherlands.
[88] Young Hee K (1967) Basic iron (III) alkanoate complexes. Masters Thesis—Paper 7135.
Missouri University of Science and Technology, Rolla, MO, USA.
[89] Herdtweck E, Kornprobst T, Sieber R, Straver L and Plank J (2011) Crystal
structure, synthesis, and properties of tri-calcium di-citrate tetra-hydrate
(Ca3(C6H5O7)2(H2O)2).2H2O. Zeitschrift für Anorganische und Allgemeine Chemie 627:
655–659.
[90] Sheldrick B (1974) Calcium hydrogen citrate trihydrate. Acta Crystallographica B 30:
2056–2057.
[91] Froment DH, Buddington B, Miller NL and Alfrey AC (1989) Effect of solubility on the
gastrointestinal absorption of aluminum from various aluminum compounds in the
rat. Journal of Laboratory and Clinical Medicine 114: 237–242.
[92] Dyer T (2016) Influence of cement type on resistance to organic acids. Magazine of
Concrete Research, in press.
[93] Carde C and François R (1997) Effect of the leaching of calcium hydroxide from cement
paste on mechanical and physical properties. Cement and Concrete Research 27(4):
539–550.
[94] Parkhurst DL and Appelo CAJ (2013) Description of input and examples for PHREEQC
version 3—A computer program for speciation, batch-reaction, one-dimensional
transport, and inverse geochemical calculations. In U.S. Geological Survey Techniques
and Methods, Book 6. US Geological Survey, Denver, CO, USA, Chapter A43. See http://
pubs.usgs.gov/tm/06/a43 (accessed 08/12/2016).
Chapter 3
Bacterial Biodeterioration
3.1 Introduction
Bacteria are single-celled organisms falling within the Prokaryota grouping.
Prokaryote cells differ from the cells of other living organisms in their
lack of a membrane-bound nucleus, membrane-bound organelles and
mitochondria. As discussed in Chapter 1, the Prokaryote grouping has been
subdivided into Archaea and Bacteria, but for the purpose of this book the
two can be discussed together.
Bacteria comprise the largest proportion of biomass on the planet, and
their presence on the Earth’s surface (and to some distance beneath it) is
practically ubiquitous. Indeed, bacteria have been found to be capable of
existing in extremely hostile conditions.
The deterioration of concrete in contact with bacterial communities
is principally through chemical reactions between cement and aggregate
and substances produced by these organisms. However, it is of benefit to
discuss the metabolism and reproduction of these organisms first, since
these aspects have important implications with regards to interactions
between bacteria and concrete.
(dX
dt
)
µ=
X
and the maximum specific growth rate is the specific growth rate at a
concentration of substrate at or above saturation. The saturation constant
is the substrate concentration required for a specific growth rate which is
half that of the maximal rate.
Thus, growth is dependent on the population of bacteria and the
concentration of the growth limiting substrate. The concentration of
substrate will decline in a manner described by the equation
dS µ SX
= max
dt Y(Ks – S)
Where Y is a parameter known as the growth yield. The mass of bacteria
will change in accordance with the plot shown in Figure 3.1. The initial
Bacterial Biodeterioration 81
z
0
"" \ BIOMASS
SUBSTRATE
~
1-
z
(f) \ w
(f) 0
\ z
~ \ 0
Q 0
(}) w
\ I-
\ ~
\ 1-
(f)
\ aJ
:::>
\ (f)
\
\
TIME
interaction between proteins in the cytoplasm of the cell and those in the
outer membrane, leading to a movement of the outer membrane not unlike
that of caterpillar tracks on vehicles.
The motion of bacteria is directed by local stimuli including
concentrations of dissolved chemical species (chemotaxis), light (phototaxis)
and oxygen (aerotaxis). Motion may be directed towards or away from such
stimuli, depending on the nature of the bacteria.
3.4.1 pH
High and low pH conditions are toxic to most bacteria. Whilst some species
can survive under very extreme pH conditions, growth is fundamental to
the survival of these organisms. Most bacteria are neutrophilic—they grow
at optimal rates around a pH of 7. A small number of bacteria are able to
survive at pHs as low as 1, and are referred to as being acidophilic. Alkaliphilic
bacteria are capable of growth in the range of between 8.5 and just over 11,
but their optimal rate of growth is less than this upper limit [21]. The pore
solution of relatively young concrete will typically be in the range of pH
11–13, meaning that a new concrete surface is unlikely to support bacterial
growth of any kind.
A number of different processes can occur in concrete which will act
to reduce pH at the concrete surface. Probably the most common of these
reactions is carbonation. This reaction, between carbon dioxide gas and
cement hydration products in the presence of water, leads to the conversion
of portlandite (Ca(OH)2) to calcium carbonate:
CO2 + H2O H2CO3
18
16 •
0
w/c = 0.5; PC content = 18%
w/c = 0.6; PC content = 14%
E
12 "
E 10
I -
I-
c.. 8
LJ.J
Cl
6
"
4
0
0 4 6 8 10 12 14 16
TIME , years
communities [26, 27, 28, 29]. This is also reflected in the results of studies
examining the relationship between the development of populations
of cyanobacteria (and other organisms) on concrete surfaces, where no
clear relationship was identified [30, 31]. However, the surface porosity
of concrete does appear to influence the rate at which it is populated by
bacteria, with a higher volume of porosity (via a higher water/cement ratio)
leading to higher rates of colonization (Figure 3.3) [30, 31, 32]. The reason
for this is unclear. Whilst it is possible that the pores provide features on
the surface for bacteria to become lodged, it is most likely that the porosity
simply provides a reservoir for moisture, which the bacteria need.
The previous discussion identified that many of the molecules produced
by bacteria in the formation of biofilms possess functional groups capable
of forming bonds with metal ions. This has two important implications
from the perspective of bacterial attachment to concrete surfaces. Firstly, it
is likely that this ability may well provide a mechanism for the formation
of strong chemical bonds between biofilms and cement hydration products
at the surface. One study has observed the formation of covalent bonds
between biofilms produced by Pseudomonas aeruginosa and TiO2 surfaces
[33]. The compound responsible was identified as pyoverdine, which is
a peptide siderophore. Siderophores are molecules which are capable of
chelating iron, but also capable of forming strong bonds with other metal
ions. In the case of pyoverdine, this appears to be via catecholate functional
groups. Whilst this species has not been identified as one which is involved
in deterioration of concrete, molecules with catecholate groups and similar
100
,.
~
0
w-
u 80
<I: /
u... /
0::
::J /
/
.-/
(/)
u... 60 /
0
w /
<!J /
<I: /
0:: 40
w /
>
0 /
/
u /
z 20
/ .
<I: /
w /
:2; /
/
W' /
0
0.3 0.4 0.5 0.6 0.7 0.8
are produced by many bacteria, making the formation of strong bonds with
metal oxides in concrete a very likely possibility. It is also believed that
divalent cations, such as calcium, can act as bridges between polysaccharide
molecules, rendering biofilms stronger [34].
So far discussion of colonization of concrete by bacteria has been
limited to surfaces. However, the fact that concrete is a porous material
means that the possibility of bacteria penetrating further into the material
exists. Concrete pores cover a wide range of diameters from nanometers to
hundreds of micrometers. The pores in concrete take two forms: capillary
pores and gel pores.
Gel pores are present in the CSH gel produced by the hydration of
cement and have dimensions of between 0.5 and 10 nm. Capillary pores are
the remnants of water-filled space between cement grains from when the
concrete was in its fresh state. Much of the original space will be occupied
by cement hydration products in hardened concrete, but what remains
forms a network of pores, and has dimensions in the 10 nm to 10 µm range.
Bacteria range widely in size between species, generally within the
range 0.5–1.0 µm [35]. This means that for many bacteria, much of the
porosity—and certainly all of the gel porosity—is inaccessible. It is often
tempting to imagine porosity in concrete as a series of tunnels of circular
cross-section and uniform diameter running from one end of the material
to another. The reality is much more complex—porosity consists of a
highly interconnected network with high constrictivity (i.e., rapid change
in diameter over a length of porosity). Taking this into account, cement
porosity should be even less accessible to bacteria.
However, bacteria are able to penetrate beneath the surface of concrete,
albeit gradually. One reason for this is that, as the concrete deteriorates, pore
sizes will increase as their surfaces are dissolved. Additionally, cracks may
form, leading to the interior becoming even more accessible. One study
has examined bacterial communities in the deteriorated layers of asbestos
cement pipes used for the transport of drinking water for over 50 years [36].
The researchers identified four layers—an outer layer where significant
deterioration had occurred, two intermediate layers, and a deterioration
front beyond which the cement was essentially unaffected. The bacteria
were characterised in terms of their grouping (slime-forming, iron related,
heterotrophic, acid-producing and sulphate-reducing) and populations of
bacteria in each layer were estimated (Table 3.1). The most populous layer
was the innermost layer, which contained mainly slime-forming bacteria.
3.4.3 Nutrients
Table 3.1 Estimated populations of different bacteria in the deteriorated layers within
an asbestos cement pipe used for carrying drinking water for a period of 52 years [11].
Total thickness of deteriorated zone was between 3.1 and 4.6 mm, but thickness of
individual layers is not specified.
1.0
/--
Hco; I
1
H,co, /co,'"
0.8 I
I
I
I
z I
0 0.6 I
i= ··I
,,
0:: I
0 I
£l..
0 I .
0:: 0.4 I
£l.. I
I
I
I
0.2 I
I
I
I
_,/
0.0
4 6 10 12 14
pH
fly ash for use in concrete limits total P2O5 levels to less than 5.0% by mass,
and soluble phosphate to 100 mg/kg [43].
The alkali metals potassium and sodium are present in all concrete
constituents. Regardless of the issue of biodeterioration, the alkali content of
concrete is of concern from a durability perspective, since certain susceptible
aggregates may undergo alkali-aggregate reactions at high alkali contents.
Alkalis in Portland cement are usually of greatest concern, since they are
typically present in highly soluble sulphate salt form on the surface of
cement grains, usually converting to alkali hydroxides during hydration.
In cement, alkali content is expressed as the equivalent of sodium oxide
(Na2O), which is calculated using the equation:
Na2Oeq = Na2O + 0.658K2O
A study of Portland cement manufactured in the USA and Canada in
2004 found the average level of alkalis to be 0.70% Na2Oeq, for Type I cement,
with a maximum of 1.20. Other types of cement displayed slightly lower
averages. Fly ash and silica fume typically possess alkali levels somewhat
higher than this [44, 45, 46], whilst slag generally contains lower quantities
[47].
Alkalis in aggregate are often in a much less soluble form, frequently
present in silicate minerals such as feldspars, mica and clay-forming
minerals. However, these can gradually be released under the high pH
conditions of concrete [48] and, as discussed previously, bacteria can
potentially accelerate their release through the production of organic acids
and/or complex-forming compounds. Alkali may also be introduced into
concrete dissolved in mix water.
Iron is present in Portland cement in relatively large quantities (typically
around 2–4% by mass as Fe2O3). In the hydrated state it is present largely
as the AFm and AFt phases and as a substituted ion in CSH gel. As the
pH of concrete is reduced, the AFm and AFt phases will decompose (see
Chapter 2). However, the iron will remain in a relatively insoluble form as
ferrihydrite, until much lower pH conditions are reached. The presence of
organic acids capable of forming complexes with iron may increase the pH
at which iron becomes soluble. Many aggregates will contain quantities
of iron, although commonly this is present as hematite (Fe2O3) or possibly
magnetite (Fe3O4) which are essentially insoluble. Again, complexation by
organic acids and other compounds may render iron more soluble.
Magnesium is also present in Portland cement, although its levels are
again limited by industrial standards. EN 197-1 limits MgO in Portland
cement clinker to 5.0% by mass, whilst ASTM C150 limits it to 6.0% in all
cements covered by the standard [49]. The European standard for fly ash
in concrete limits MgO to less than 4.0% by mass [43], whilst the equivalent
standard for slag limits it to 18% [50].
90 Biodeterioration of Concrete
(SO42–) dissolved in sewage. The resulting sulphide ion (S2–) rapidly reacts
with hydrogen dissolved in the sewage to give hydrogen sulphide (H2S).
The solubility of sulphide in water is dependent on pH. This is because
H2S undergoes dissociation at higher pHs to give bisulphide (HS–) ions and
a variety of sulphide ions at even higher pHs, increasing the capacity of
the water to contain dissolved sulphide. The dissociation constants of H2S
at different temperatures are given in Table 3.2.
Where the pH of sewage is above 7, its capacity for dissolved sulphide
is relatively high. However, where the pH is lower—which is often the
case for sewage—the capacity is limited and H2S escapes as gas into the air
space above the sewage in the pipe. The reason for the low pH is also tied
to microbial activity—organic matter in the sewage is being broken down
into organic acid intermediates, which will also act as a source of carbon to
the sulphate reducing bacteria [57]. It can also be seen from Table 3.2 that
higher temperatures will also encourage the release of H2S.
The H 2S above the sewage now experiences a different chemical
environment: at the surface of the concrete pipe is usually a thin layer of
condensed water which, being in contact with concrete has a relatively high pH.
This means that the layer of moisture can dissolve relatively large quantities
of H2S. Oxygen in the atmosphere above the sewage can react with the H2S
to form a range of different solid compounds including elemental sulphur,
thiosulfates and polysulphates (tetrathionates, pentathionates, etc.) [58].
It is at this point that the second group of bacteria play a role. These
are the species Thiobacillus which are also present at the concrete surface.
Thiobacillus is one of a group of species known as colourless sulphur
bacteria, or sulphur-oxidising bacteria. These organisms are chemotrophs
and obtain their energy by oxidising inorganic compounds. Many reactions
can be used for this purpose, as outlined in Table 3.3. The products of most
of these reactions is either sulphuric acid (H2SO4) or compounds which can
subsequently by used in reactions which yield sulphuric acid. Another
common group of sulphur-bearing compounds in sewer pipes are the
mercaptans—organic compounds with a thiol (–SH) group. These, however,
do not appear to be usable by thiobacilli [60].
Bacteria identified in or on samples of concrete taken from corroded
sewer pipes from three studies are shown in Table 3.4. Common to all these
studies is Thiobacillus thiooxidans. Also included in the table are the pH
ranges within which different Thiobacillus species can grow. It should be
noted that Thiobacillus thiooxidans is only able to grow in conditions of low
pH, and that none of the other species grow under pH conditions greater
than 9 or 10. For this reason, it is necessary for the surface of the concrete,
which would otherwise have a pH of 11 or more, to become less alkaline.
The dissolution of acidic H2S in the water layer will play a small role in
the neutralization of the cement hydration products which are responsible
for the high pH. However, the largest contribution is likely to come from
carbonation—CO2 concentrations in the airspace in the pipe are in most cases
Table 3.3 Reactions used by colourless sulphur-oxidising bacteria to obtain energy [61].
Compound Reaction
Table 3.4 Bacteria identified from samples of corroded concrete from sewer pipes.
decalcify. The overall effect of this is that considerable mass is lost from the
concrete surface and its strength is substantially reduced.
The effect of the above process is shown in Figure 3.6, which shows
the results of a geochemical model in which hydrated cement paste is
brought into contact with sulphuric acid. It can be seen that the outer part
of the cement paste is completely dissolved, followed by a layer in which
decalcified CSH is present. Beyond this layer are deposits of small quantities
of ettringite and substantial quantities of gypsum, followed by unaffected
hydrated cement. These modelled results match well with the results of
powder X-ray diffraction carried out on the acid-affected layers of concrete
specimens [65]. Under conventional sulphate attack, it would be not unusual
10°
""0 10-1
E 1Q-2
6 10-3
w
>
_J 10-4
0 10-5
(/)
(/)
10-6 SO/
0
Ca
;::
_J 10-7
------ Si
10-8 - - -- - AI
0
f-- 1Q-9 --- Fe
0.020
1 ----
pH
12
10
8 I
c_
6
4
"'Cll
0
0 .015 ·-·-·-" 2
E
>--
f--
i= 0.010
z
<(
::J
a
0 .005
0 2 3 4 5 6 7 8 9 10
Figure 3.6 Concentrations of dissolved elements (top) and quantities of solid phases
obtained from geochemical modelling of sulphuric acid attack of hydrated Portland
cement. Model conditions: acid concentration = 0.1 mol/l; volume of acid solution = 4 l;
mass of cement 80 g; diffusion coefficient = 5 × 10–13 m2/s.
Bacterial Biodeterioration 95
0.25
Before exposure
Sulfur bacteria
"'E
.!2 Sulfuric acid
0.20
"'E
(.)
>--
1-
U5 0.15
0
0::
0 ' ' '
[}_
'
LlJ
>
0.10 ' "\
i= \
::5 \
::::> \
~ 0.05 \
::::>
()
"' ..... ~
0.00
1x10° 10x10° 100x10° 1x103 10x103 100x103 1x106
PORE DIAMETER, nm
Figure 3.7 Cumulative pore size distributions obtained using mercury intrusion
porosimetry for concrete specimens placed in a reactor containing sulphur oxidizing
bacteria or submerged in a 0.15 mmol/l solution of sulphuric acid for 14 days [67].
96 Biodeterioration of Concrete
40.----------------------------------------------.
• Cement content = 570 kg/m2
- -o- Cement content= 310 kg/m2
30
::oR
0
(/)
g
(/)
20
(/)
(/)
<(
:2:
10
0
0.0 0.5 1.0 1.5 2.0 2.5 3. 0 3.5
pH
Figure 3.8 8-week mass loss from concrete exposed to sulphuric acid solutions with
different concentrations (expressed in terms of the resulting pH). Specimens were
not abraded [68].
20 , -----------------------------------------------,
•
--o-
Finite concentration
Replenished
15
::oR
0
(/)-
(/)
0 10
...J
(/)
(/)
<(
:2:
0~------------.------.------.------.-----,,-----~
2.0
?F. //0
(/) / /0
/
0 /
_J
9/
(/)
_...-- --•
/
(/) /
<( 0.5 /
~ 0 /
0/ 0
_...-·- •
o--·-·-
/
//
0.0
-0.5 - f - - - , - - - , - - - , - - - , - - - - , - - - - , - - - - , - - - - - - 1
80 100 120 140 160 180 200 220 240
TIME, days
Figure 3.10 Mass loss from concrete specimens stored in a reactor containing
simulated wastewater and sulphur oxidizing bacteria [70].
to mimic the effect of the flowing water in sewers. The extent to which
this mimics reality is debatable, given that corrosion occurs above the
water-line. However, this level will in fact vary with time and so corroded
zones are likely to experience flowing water periodically. The rate of
mass loss resulting from brushing is greater than that without brushing
(Figure 3.11). Clearly the majority of this increase is simply the result of
removal of material. However, it should also be remembered that the
reaction products may protect the underlying concrete from further attack,
and brushing will diminish this effect.
Concrete characteristics which influence the magnitude to which
sulphuric acid attack causes deterioration include cement content, water/
cement ratio, cement type and aggregate type.
Since sulphuric acid attack is largely a cement deterioration process, the
cement content plays a major role in determining the response of concrete.
As the cement content increases, it increases the availability of calcium,
which is an essential component in the formation of gypsum and ettringite.
Moreover, it is largely the cement which will be dissolved as a result of the
process of acidolysis. Thus, as cement content increases, so too does the
rate of deterioration (Figure 3.12).
The relationship between water/cement ratio and concrete durability
under sulphuric acid attack runs counter to expectation. A reduced water/
cement ratio yields higher strengths and reduced porosity, both of which
provide greater resistance against many forms of deterioration. This appears
Bacterial Biodeterioration 99
10000
• Unbrushed
- -o - Brushed
8000
'E
0, 6000
CJi
(f)
0
--'
(f) 4000
(f)
~
::2
2000
TIME, days
Figure 3.11 Mass-loss versus time plots for concrete specimens exposed to a
pH 1.5 sulphuric acid solution with and without periodic abrasion using a
brush [71].
35
• -- ""
30 - --o -
--y--
pH 1
pH 0.6
pH 0.3
/ .....-
__ .,...- -- >r --
25 /
-
/ _-0
:oR ~
0
/ _o - - o- -
u) /
(f) 20 y"' _-er
__o-
0 o---
--'
(f)
(f) 15
~
::2
10
0
250 300 350 400 450 500 550 600
Figure 3.12 Mass loss from concrete specimens with different Portland cement
contents after exposure to sulphuric acid solutions for a period of 60 days [68].
100 Biodeterioration of Concrete
to not be the case for sulphuric acid, regardless of whether the concrete
undergoes abrasion [72, 73, 74, 75]. This can probably be attributed to two
factors. Firstly, the conventional means of achieving low water/cement
ratios during concrete mix design is to increase the cement content, yielding
the increase in vulnerability discussed above. Secondly, the increased
porosity at high water/cement ratios may offer a larger volume of space
within which the reaction products are able to precipitate leading to a delay
in the onset of expansive forces. Moreover, the accumulation of reaction
products in this way is likely to offset mass loss from fragmentation and
acidolysis.
However, results obtained from sulphuric acid exposure experiments
do not present the whole picture, and Figure 3.10 indicates that, where
attack derives from sulphur bacteria, a low water/cement does reduce the
rate of attack. This issue is discussed further in the chapter when measures
to limit rates of deterioration are discussed.
Cementitious materials used in combination with Portland cement tend
to provide enhanced resistance to sulphuric acid in comparison to concrete
containing only Portland cement. This is likely to be for two reasons. Firstly,
the cement fraction will possess a reduced calcium content, thus limiting the
quantities of gypsum and ettringite formed. Secondly, many such materials
have the effect of refining the pore structure of concrete either through fine
particles occupying the space between coarser Portland cement particles,
or through the formation of reaction products as a result of pozzolanic or
latent hydraulic reactions. Table 3.5 summarises the findings of studies
examining the performance of concrete containing ground granulated
blastfurnace slag (GGBS), fly ash (FA), silica fume (SF), natural pozzolanas
(NP) and metakaolin (MK) with regards to sulphuric acid resistance. GGBS
appears to have only a minor influence over resistance, presumably due
to its relatively high calcium content. Fly ash, whose presence will reduce
calcium content considerably more, improves performance to a much
greater extent (Figure 3.13). Whilst silica fume and metakaolin improve
performance in most of the studies included in Table 3.5, this cannot be
attributed to reduction in calcium content, since they are used at relatively
low levels, and so refinement of porosity is likely to be the mechanism of
enhanced resistance.
Calcium aluminate cements have also been found, in some instances,
to display enhanced resistance to sulphuric acid. However, results vary
considerably from study to study, possibly due to variation in composition
of these materials. A field study in which calcium aluminate cement (CAC)
was used in concrete mixtures in sewer pipes alongside Portland cement
concrete pipes indicated considerably less deterioration from the CAC
material when evaluated through inspection [69]. However, laboratory
experiments conducted by the same researchers did not replicate these
results. The same study also placed specimens made from sewer pipe
Bacterial Biodeterioration 101
80
60
I
•
I
--
-·-
Mass loss, 100% PC
-o- Mass loss, 70% PC /30% FA
Expansion, 100% PC
--0- Expansion, 70% PC /30% FA
0.8
0.6
0.4
•
I
~ I ~
(/)~ 40 I z
r...rt
0.2
(/)
_.o----0 0
0_j Ui
--o-- o - z
(/) -o- -o- o- -o- -o- -o- -o-- -o- <{
(/) 0.0 [L
20
<{
::;! ><
L.U
-0.2
0
-0.4
~20 -0.6
0 10 20 30 40
TIME, months
Figure 3.13 Mass loss and expansion of Portland cement and Portland cement/
fly ash concrete specimens exposed to a 2% sulphuric acid solution with periodic
replenishment to maintain pH [78].
lining mixes made using both CAC and Portland cement in sewers and
measured mass loss. The CAC lining materials performed better than their
Portland cement equivalents. A similar study in which concrete specimens
were suspended in sewers found similar results when CAC concrete was
compared against Portland cement and Portland cement/GGBS blend
concrete [86].
One reason for enhanced resistance of CAC concrete, where this is
achieved, is most probably the lower calcium content of the material. The
results of geochemical modelling of sulphuric attack of a typical hydrated
calcium sulphoaluminate cement is shown in Figure 3.14. Comparing
these results to those obtained for Portland cement (Figure 3.6) it is evident
that the peak quantity of gypsum produced is lower for the calcium
aluminate cement. However, a substantial quantity of this mineral is
still present, and so only minor improvement in performance might be
expected. It would also appear that the high aluminate content of the
cement has the effect of suppressing bacterial growth: it has been shown
that aluminium concentrations in excess of 350–600 mg/l stop growth of
Thiobacillus thiooxidans [87]. This would certainly explain the relatively poor
performance of CAC concrete when exposed to sulphuric acid solutions,
compared to performance in the field.
The possibility of using limestone aggregates as a means of enhancing
sulfuric acid resistance has been explored, on the basis that the calcium
carbonate in these materials is capable of neutralising acids [76, 88].
The presence of limestone does tend to enhance resistance, although it
Table 3.5 Findings of research conducted to evaluate resistance of concrete containing pozzolanic and latent hydraulic materials to sulphuric acid attack. FA =
fly ash; SF = silica fume; NP = natural pozzolana; MK = metakaolin.
Cement Study
[72] [71] [73] [74] [75] [76] [78] [80] [81] [82] [83] [84] [84]
H2SO4
Outcome Slight Negligible
improvement change
FA Level 7.5–22.5% 33% Optimum 10%–70% 13.5–50%
siliceous siliceous fly at 50% siliceous siliceous
fly ash ash, 7% SF siliceous fly fly ash fly ash
ash, 10% SF
Acid Conc. 1% 1% 1% 2% with 8.7%
replenished replenished periodic H2SO4
weekly weekly additions
Acid Conc. 1% 2%
Outcome Enhanced
up to 10%
of cement
Bacterial Biodeterioration 103
104 Biodeterioration of Concrete
100
"'= 10-1
0
E 10·2
c:i 10·3
~=::.-=~-=-~"=-~~~~-~~;_-
w \ .·
~ 10·4 .. /
0
(f)
(f)
10-5
""'' - - SO/
0 10·6 ------
I ./" ---- Ca
_J 10•7
/
~
0
10·8 1/ - · -- -
Si
AI
I- 10·9 --- Fe
. --·-- 12
0.020 pH,.. . /
10
8 I
0.
6
4
rJ)
Q)
0 .015
2
0
E
>--
I-
i= 0 010
z
<(
;:)
0
0 .005
0 2 3 4 5 6 7 8 9 10
Figure 3.14 Concentrations of dissolved elements (top) and quantities of solid phases
obtained from geochemical modelling of sulphuric acid attack of a typical hydrated calcium
sulphoaluminate cement. Model conditions: acid concentration = 0.1 mol/l; volume of acid
solution = 4 l; mass of cement 80 g; diffusion coefficient = 5 × 10–13 m2/s.
12
10
•
--o-
Siliceous aggregate
Limestone aggregate
8
;,R
0
(f)-
(f) 6
0_J
jJ
/
(f)
(f) 4
<{
:::2: /
;:f
2
_o---
--0--- ------
0 -- o---
-2
0 20 40 60 80 100 120 140 160 180
TIME, days
Figure 3.15 Mass loss from concrete specimens containing different aggregate
exposed to a 1% sulphuric acid solution [76].
Bacterial activity can also produce nitric acid. As with sulphuric acid
formation, the process is relatively complex, involving a chain of interaction
between various bacterial groups. The starting point of biogenic nitric acid
attack is the compound urea (CO(NH2)2). Urea can be used by a group of
bacteria known as urobacteria. Urobacteria hydrolyse urea for energy,
producing ammonia (NH3):
106 Biodeterioration of Concrete
Table 3.6 Ammonia-oxidising bacteria found in water distribution and treatment environments.
feature heavily in Tables 3.6 and 3.7. However, a growing problem in recent
history has been the growth in concentrations of atmospheric ammonia.
The rise in atmospheric ammonia concentrations is the result of human
activities, with such sources including coal combustion, fertilizer use and
livestock production. Ammonia released into the atmosphere will tend to
react with pollution-derived acids in the atmosphere (sulphuric acid, nitric
acid, etc.) to form ammonium salts. These salts will be precipitated either
as particulates or dissolved in precipitation [97]. As a result, high levels of
ammonia are frequently found in porous building materials with surfaces
exposed to the exterior environment [98].
Most nitrifying bacteria are obligate lithoautotrophs (with the exception
of Nitrobacter, which can grow heterotrophically [90]), and so no source of
organic matter is required. They are also obligate aerobes, needing oxygen
for the nitrifying reactions. They are, however, killed by visible blue
light and long-wavelength UV light [100]. As a result, they form biofilms
containing significant quantities of EPS to act as a barrier to light.
In sandstone (which is typically more porous), Nitrosovibrio have been
found at depths of 30 cm below the surface [99]. Nitrifying bacteria are also
capable of penetrating concrete: whilst the ‘slime forming’ bacteria found at
some depth inside asbestos cement water pipes discussed in Section 3.4.2
are not specifically identified, it is likely that they are nitrifying bacteria.
When nitric acid is brought into contact with hardened Portland cement,
the effect is more straightforward than that for sulphuric acid. The cement
hydration products will undergo acidolysis, leaving a decalcified zone at
the surface. The results of geochemical modelling of the process of nitric
acid attack are shown in Figure 3.16.
The results from the model indicate why the water pipe study discussed
in Section 3.4.2 found the most nitrifying bacteria at the interface between
the unaffected cement and the deteriorated material. The production of
nitric acid, if unchecked, will lead to a drop in pH which will eventually
108 Biodeterioration of Concrete
101
<:::: 10°
0
E 10-1
cl 1Q·2
UJ
>
_J
10-3
0 10-4
(/)
(/) 10-5
0 10-6
_J
~
10·'
Ca
Si \_ _____ _
0 10-8 - ·- - • AI
I- 10-9 --- Fe
1
12
r·-- ·--·--·--·--·--·--·
. pH
0.020 10
I
8
I I
Q_
6
I
4
If)
Q)
0 .015 I
2
0 I
E · -·- · - · -·-·-·- · - · -·- · ----~ 0
>-"
I-
i= 0.010
z hydrite
<t:
~
0
0 2 3 4 5 6 7 8 9 10
60
50
E
E
I 40
I-
a_
w
0
z 30
Q
(f)
0
0:: 20
0::
0
(.)
10
0
00 0.1 0.2 0.3 0.4 0.5 0.6
CONCENTRATION , mol/1
Figure 3.17 Depth of corrosion of Portland cement paste specimens exposed
for 200 days in nitric acid solutions maintained at various concentrations [101].
110 Biodeterioration of Concrete
0.16
Before exposure
0.14 Nitrogen bacteria
"'E
0 Nitric acid
"'"0E 0.12 ' -... ,.
~
1- 0.10
" "\
\
Ui \
0 \
0::: "\
0 0.08 .\
0... \
L.U \
> 0.06 \
i=
<(
--'
=> 0.04
,.
\.
I
:2
=>
u
"""
0.02 "
0.00
1x10° 10x10° 100x1 oo 1x103 10x1 03 100x103 1x106
PORE DIAMETER, nm
Figure 3.18 Cumulative pore size distributions obtained using mercury intrusion
porosimetry for concrete specimens placed in a reactor containing nitrifying
bacteria or submerged in a 0.15 mmol/l solution of nitric acid for 14 days [67].
16
-~ -------- v 60 days
14
E 12
E
:i"
1- 10
c..
L.U
2
... ---- - - - - · - - - - - - - - - - - - - - - - - - - · 7 days
0
0.3 0.4 0.5 0.6
--·
14
E
12
·--- - ------ - ~- -
E
z- 10
0
Ui
0 8
0::
0::
0
u 6
LL
0
I
f- 4
()._
w
0
2
0
0.5 1.0 1.5 2.0 2.5 3.0 3.5
Cement Study
[73] [103] [104] [83] [84]
Slag Level
Acid
Conc.
Outcome
FA Level 7.5–22.5%
siliceous
fly ash
Acid 0.01%
Conc.
Outcome Reduction
SF Level 7.5–30% 5–30% 0.2%
Acid All 1% 0.013% Concentrated
Conc. and 20%
Outcome Negligible Enhancement Enhancement
change
NP Level 40% ‘true 28% trass
pozzolana’ + 6%
calcareous
fly ash
Acid 2:1 0.3% to
Conc. sulphuric/ 0.8%
nitric, pH =
3.5, periodic
additions to
maintain pH
Outcome Reduction Reduction
MK Level 7.5–30%
Acid 0.01%
Conc.
Outcome Negligible
change
101
"'
0
10°
E 10"1
cl
LU
>
...J
10·2
---
10-3 -!=~~~ ·~·--~~·--------- - - - - --
0 10·4
(/)
(/) 1Q·5
No; IL ----------
0 10·6
Ca 1
...J
10·7 ------ Si I____ _
~ 10·8 --- · AI
0
1- 10·9 --- Fe v
1
12
0.020 10
8 I
c.
6
4
en 0.015
Ql
2
0
-·-·-·-·-·
)
E -·-·- ·- - ·- - 0
~
1-
i= 0.010
z
<(
::J
a
0005
0 2 3 4 5 6 7 8 9 10
Figure 3.21 Concentrations of dissolved elements (top) and quantities of solid phases
obtained from geochemical modelling of nitric acid attack of a typical hydrated calcium
sulphoaluminate cement. Model conditions: acid concentration = 0.1 mol/l; volume of
acid solution = 4 l; mass of cement 80 g; diffusion coefficient = 5 × 10–13 m2/s.
Acetic acid
100 10
• Mass loss
Deteriorated depth
I
I 0
80 ---o 8 E
----
E
-- 0
::C
--
;,R 1-
0 9- ll.
(/)- 60 6 w
(/) /
/ 0
0 / 0
....1 -()
/
w
(/) / 1-
~
/
___________ _.
(/)
<{ 40 4
a' / o 0
~
I - · --- -
a:w
cj _.----- 1-
w
/r
20 I
I
_.
... 2 0
0 t
0 100 200 300
0
TIME, days
Figure 3.22 Mass loss and depth of deterioration of Portland cement paste specimens
exposed to acetic acid [116]. Acid concentration = 0.28 mol/l, with pH adjusted to
4.0 using sodium hydroxide.
attack, deterioration will typically occur at a slower rate with acetic acid,
since it is a weaker acid.
Since acidolysis is the main mode of attack, deterioration takes the
form of decalcification at the surface, leading to a thick layer of what is
essentially silica gel being the only remaining material at the outside of the
cement paste. Figure 3.23 shows micro-CT scans through various cement
paste specimens exposed to acetic acid solutions. The dark outer layer is
the silica gel layer. In the case of the Portland cement specimen, there is a
second layer between the wholly-decalcified material and the unaffected
cement, where partial decalcification has occurred, which is not present
in the case of the other cements in this figure, due to their lower calcium
content. It should be noted that the silica gel layer has cracked substantially
in all cases. This is not the direct result of acid attack, but of shrinkage
during drying. However, the magnitude of the cracking shown in these
images clearly illustrates the notable susceptibility of the layer to shrinkage.
Taking the diameter of the wholly unaffected cement as a measure of
resistance, the figure indicates that there is some enhancement in resistance
to acetic acid attack through the use of fly ash. This has also been observed
with concrete [117], with enhanced performance with silica fume reported
frequently, also (Table 3.9). Calcium aluminate cements a sulfoaluminate
cement—CSA—in the case of Figure 3.23 typically provide enhanced
resistance, for the same reasons as for nitric acid.
Bacterial Biodeterioration 117
Figure 3.23 micro-CT scans from cement paste specimens exposed to a 0.1 M
solution of acetic acid for 90 days [118].
Cement Study
[117] [103] [73] [80] [81]
FA Level 5%
Acid
conc. 1.7%
Outcome Enhancement
SF Level 10% 5–30% 7.5–30% 78.5–30% 15%
Acid
conc. 1.7% 14% 5% 5% 5%
Outcome Enhancement Enhancement Reduction Enhancement Enhancement
MK Level 7.5–30%
Acid
conc. 5%
Outcome Negligible
change/
Reduction
Lactic acid
Lactic acid does not form insoluble salts with calcium, aluminium or iron,
and forms weak complexes with these ions (although the aluminium
complexes are somewhat stronger than those with other ions). For this
reason, the deterioration of concrete in contact with lactic acid is almost
purely a process of acidolysis.
An example of mass loss characteristics from lactic acid attack are
shown in Figure 3.24, whilst micro-CT scans from cement paste specimens
118 Biodeterioration of Concrete
made from PC, PC/FA and CSA cements are shown in Figure 3.25. It should
be noted that the specimens in this figure show less deterioration than
those exposed to acetic acid. From a theoretical perspective, this seems
improbable, since lactic acid has a lower acid dissociation constant (see
Chapter 2) and is, thus, the stronger of the two. A study examining the
composition of organic acid solutions brought into contact with hardened
Portland cement paste [120] has found that, despite the lower pH of the
lactic acid solution, the concentration of calcium is lower than that of acetic
acid (as well as butyric, isobutyric and propionic acids).
100,-----------------------------------------------~
80
I ~ PC concrete
85% PC / 15% Silica fume concrete
I
0~
(/)- 60
(/)
0
-'
(/)
(/)
<( 40
~
20
0
0 50 100 150 200
TIME , days
Figure 3.24 Mass loss from concrete specimens made with Portland cement
and a Portland cement/silica fume blend exposed to lactic acid [81]. Specimens
were scrubbed periodically with a wire brush.
The reason for this is almost certainly related to the fact that the lactate
ion forms complexes with Al, leading to an increase in the concentration of
this element in solution. Higher Si concentrations were measured compared
to the other acid, which is at least partly the result of the lower pH obtained
with lactic acid. However, neither of these effects directly explain the lower
Ca concentrations. It is possible that, since CSH gel will normally contain a
quantity of aluminium, the removal of this element from the gel as a result
of complexation means that calcium takes its place.
Both figures show that the use of pozzolanic material and CSA enhances
resistance to lactic acid attack. Slag also gives improved resistance, as will
be discussed later in this section.
Propionic acid
Very little research has been conducted into the interactions of propionic
acid with cement and concrete. However, experiments using of a number
of the common organic acids formed by bacteria have been conducted to
examined the pH and dissolved species detected after exposure of crushed
PC paste specimens to acid solutions of the same molar concentration [120].
It was found that there was very little difference between the behaviour of
propionic acid compared with either acetic, butyric or isobutyric acid. This
is not surprising, since all of these acids are structurally similar and have
very similar acid dissociation constants (see Chapter 2).
Butyric acid
Figure 3.26 micro-CT scans showing cross-sections through cement paste specimens
exposed to a 0.1 M solution of butyric acid for 90 days [118].
600
N
E
0
- Gravel
c=J Limestone
1
0, 500 1
E
<(
w
0::: 400
<(
!:::
z ,---
::J 300
0:::
w
a..
C/) 200
C/)
0
.....J
C/)
C/) 100
<(
~
Figure 3.27 Mass loss from concrete specimens made using either PC, PC/GGBS
or calcium aluminate cement (CAC) as the cement fraction exposed to a solution
containing both lactic acid (32 mg/l) and acetic acid (4 mg/l). All water/cement ratios
are 0.45, except the CAC mixes (w/c = 0.40). PC/GGBS proportions unstated, but by
definition must be 36–95% by mass [121].
Bacterial Biodeterioration 121
to have little effect in the case of the PC-only mixes, but which produces
lower levels of deterioration for lower cement content where GGBS is
present.
Figure 3.28 compares performance in volume reduction terms of PC,
PC/FA and PC/SF mixes of the same water/cement ratio. Silica fume
is more effective than fly ash, although both give higher resistance to
deterioration compared to PC [122]. Generally, performance increases in
the sequence PC–PC/FA–PC/SF–PC/GGBS [123].
Calcium aluminate cement also seemingly provides greater resistance
to attack (Figure 3.27), although it should be noted that the CAC mixes in
this figure had lower water/cement ratios, which is also likely to play a
part. Conversion of the cement (the result of carbonation, leading usually
to a loss in strength) reduces resistance slightly, but the converted material
still performs better than the PC-only mixes [121].
The influence of aggregate type is also shown in Figure 3.27. The use of
limestone aggregate provides greater resistance when compared to siliceous
gravel for the PC and PC/GGBS mixes. This is the result of the calcium
carbonate present in limestone increasing the neutralising capacity of the
concrete. This effect is not evident in the case of calcium aluminate cement.
In a similar study, a solution mimicking the composition of pig
manure was used [124]. This comprised 0.21 mol/l of acetic acid, 0.04
mol/l propionic acid, 0.02 mol/l butyric acid, 0.01 mol/l isobutyric acid
400 ,------------------------------------------------,
-PC
c:::::J FA
- SF
w· 300
0
<(
u..
0:::
=>
(j)
!:::
z 200
=>
0:::
w
o.._
(j)
(j)
100
0_J
(j)
(j)
<(
::::
0+-----~~------~------~~----J,L------.~~--~
LEVEL, % by mass
Figure 3.28 Mass loss from concrete specimens made using either PC, PC/FA or
PC/SF as the cement fraction exposed to a solution containing both lactic acid (30
mg/l) and acetic acid (30 mg/l). All water/cement ratios are around 0.40 [122].
122 Biodeterioration of Concrete
and 0.003 mol/l of valeric acid. The solution was regularly refreshed to
maintain these concentrations. Figure 3.29 shows how the development of
deteriorated depth settles into what approximates to a linear relationship
with respect to time under these conditions. The pH of these solutions was
adjusted to 4 and 6 using sodium hydroxide, and it is evident that pH plays
an important role, with a high pH solution yielding less damage. GGBS was
found to improve performance. 10% Silica fume also enhanced resistance,
albeit to a slightly lesser extent [125].
6,------------------------------------------------,
e PC. pH4
· 0· 68% GGBS / 32% PC. pH4
E 5 __ .....,.._ __ PC. pH6
E - -? - 68% GGBS I 32% PC, pHS
I"
I-
o._
4
L.U
0
0
L.U 3
I-
<(
0::
0
0:
__.--- - - - ____
2
L.U
1-- 0
.,
L.U _ ..xv
0
0
------.. .- -:.
---:-JJ~- · · - · · - · Y
--v----
~- - ·
0~~~-v,_·--.----.----.---.----.----.---.----.--~
0 2 4 6 8 10 12 14 16 18 20
TIME, weeks
Figure 3.29 Deteriorated depth versus time for PC and PC/GGBS pastes in
a mixed acetic/propionic/butyric/isobutyric/valeric acid solution [124].
Bacterial Biodeterioration 123
consisted of two different species, one of which could not be identified and
either a species of Flavobacterium or Xanthomonas. The sulphur oxidizing
bacteria were Thiobacillus intermedius. It was concluded that the concrete
in the bioreactors was deteriorating as a result of attack from organic acids
deriving from aerobic bacteria (and, thus, presumably Flavobacterium or
Xanthomonas). It should be stressed that this does not mean that the sewer
pipes themselves were undergoing this form of deterioration, simply
that the samples taken and the conditions of the bioreactor favoured the
development of aerobic heterotrophic organisms.
Studies into the effect of heterotrophic bacteria on reinforced concrete
have observed considerable deterioration of properties and accelerated
de-passivation of the steel [127, 128]. Concrete specimens submerged in a
growth medium solution in which heterotrophic bacteria were present were
found to display lower strengths than specimens submerged in sterilised
water from the same source [127]. The disparity between the strength was
as high as 36%, with the greater difference in strength corresponding to the
weaker concrete mixes, although the researchers do not state the period
of time over which this occurred, or whether the differences correspond
to losses in strength or arrested strength development. Electrochemical
measurements made on the steel reinforcement bars embedded in other
specimens found that the corrosion potential in the presence of bacteria
increased at early ages and, in most cases, remained high, whereas the
sterile specimen potentials remained low. However, potentiodynamic
polarisation measurements indicated passivation of the steel affected by
bacteria, and it was proposed that this was the result of biofilm formation
at the steel surface.
A similar study did not use a nutrient solution, but instead introduced
sodium citrate into the concrete as an admixture at the mixing stage [128].
The citrate ion can be used as a nutrient by heterotrophic bacteria. Sodium
citrate is used as a corrosion inhibitor, but not normally in concrete,
since it will lead to the formation of calcium citrate leading to expansion
and cracking. Consequently, the 28 day strengths of the concrete mixes
were reduced in direct proportion to the admixture dosage. The concrete
specimens were submerged in volumes of water removed from a pond
—to ensure the presence of heterotrophic bacteria—for a period of 90
days. The strength of all specimens increased over this period, although
strength gain was proportionally less in the specimens containing higher
quantities of citrate. Corrosion potential measurements indicated a rise in
corrosion potential with citrate dosage, and gravimetric measurements on
the reinforcement indicated a higher rate of corrosion above sodium citrate
dosages of 1.0%.
Whilst both these last two studies suggest that the presence of
heterotrophic bacteria have the potential to cause significant damage to
concrete and its reinforcement, features of the experimental design and
124 Biodeterioration of Concrete
interpretation of results in both cases means that this aspect still needs
further investigation.
3.5.4 Biofilm formation as a form of deterioration
on the façades of buildings which are facing the dominant wind direction
(which are consequently most likely to be the dampest side of a building)
[130]. The sides of building receiving less sunlight during the day will also
tend to retain moisture longer, promoting growth of cyanobacteria.
The material characteristic which has most influence over cyanobacterial
colonisation is the porosity of concrete, with higher levels of porosity
(corresponding to higher water/cement ratios) encouraging growth [30,
31, 32]. Surface roughness, as discussed earlier in this chapter, appears to
have less of an influence [30, 32]. Portland cement composition and the
use of pozzolanic and latent hydraulic materials have little influence over
colonization rates [30].
It should be stressed that biofilms are frequently not exclusively
composed of bacterial communities, and can contain a range of different
organisms. For this reason, biofilms will be revisited in subsequent chapters.
1.2.-------------------------------------------------,
• 100% PC unbrushed
o 50% PC /50% FA unbrushed
1.0 " 70% GGBS I 30% PC unbrushed
-v 100% PC brushed
• 50% PC I 50% FA brushed
::R 0.8 D 70% GGBS I 30% PC brushed
0
(/)-
(/)
0 0.6
...J
(/)
(/)
<t:
:2 0.4
0.2
0.0
0 20 40 60 80 100
TIME, days
Figure 3.30 Mass loss from concrete specimens exposed to a carbonic acid
solution with a pH of 4.1 [133].
from three different concrete mixes are shown—PC, and 50% PC/50% FA
and 70% GGBS/30% PC. The GGBS and FA mixes show greater resistance
to attack, with the GGBS mixes performing particularly well when abrasive
conditions are effective. It should be noted, however that rates of mass loss
are considerably smaller than for other acids discussed previously.
material of choice for the plug (or ‘barrier’) is currently cement, although
other materials can potentially be used [146]. Whilst the process of cement
plug deterioration by sulfuric acid attack is one that is a potentially real one,
it should be put into perspective—multiple barriers with combined lengths
of 50 to 100 m are typical. Additionally, a source of oxygen is necessary to
produce sulphuric acid. This will normally derive from nitrate, but since
concentrations of nitrate in such environments are typically limited, it is
likely that sulphuric acid formation would only occur for relatively short
periods of time. Nonetheless, in the UK, evaluation of barrier materials
requires testing under conditions which mimic downhole conditions to
establish rates of deterioration, to allow prediction of performance [146].
The organic acids found in waters associated with oil formations are
characteristic of those produced by heterotrophic bacteria. However, their
origins are most probably related to the exposure of oil hydrocarbons to
elevated temperatures, rather than any biological degradation [147].
Two different approaches are available with regards to limiting the damage
from sulphur bacteria. Firstly, the environment in which sulphur bacteria
are likely to establish themselves can potentially be engineered such that the
concentrations of substances which play a part in the deterioration process
are limited. Secondly, concrete itself can be designed and possibly treated
in such a manner that it possesses enhanced resistance.
Environmental control
The manner in which sewers are designed and operated can be used to limit
damage from sulphur bacteria. The removal of H2S before it comes into
contact with the sewer walls is clearly a desirable condition, and this can be
achieved through ventilation of the air space in the sewage pipe. Ventilation
130 Biodeterioration of Concrete
Thus, from the rate equation it is evident that increasing the velocity
of the stream and/or increasing the slope of the pipe reduces the rate of
development. This is believed to be for two reasons. Firstly, an increased
rate of flow will increase the rate at which re-aeration occurs, thus providing
adequate oxygen to satisfy the BOD and prevent aerobic conditions
establishing. Secondly, faster flow will tend to act to scour biofilms and
solids at the bottom of the pipe, limiting their development as a habitat for
sulphate reducing bacteria. However, some doubts have been expressed
with regards to how important this process is [148].
Where turbulence is created, re-aeration rates are further enhanced,
although it must be noted that where turbulence occurs and sulphide is
already present, this tends promote the release of H2S as gas [149].
Dissolved oxygen can be increased further by injecting air or pure
oxygen into the water, usually in features within a sewer system such as
pressurized mains pipes (‘force’ or ‘rising’ mains) or the sumps of gravity
flow sewers (wet wells) [152]. The biofilm (or slime layer) on the sewer wall
in which sulphate reducing bacteria reside will act as a barrier to oxygen.
However, increasing the dissolved oxygen concentration to above 0.5 mg/l
usually has the effect of increasing the depth of penetration sufficiently to
limit bacterial growth and allow sulphide oxidation [149]. The effect of air
injection is shown in Figure 3.31.
Given that this is the result of a process of oxygen diffusion through the
biofilm, increasing the concentration of dissolved oxygen in the water will
4 .-------------------------------------------------,
_.
I ~ Without air injection
With a1r injection
I
/
/
/
/
/
3 /
'"E /
/.
0, /
/
u.i /
0
/
LL
_J 2
::::J
(/) /
/
;::
_J
/
/
0 /
t- /
/
/
/
~ .........
..........
0~~~~-~-~-----~-~
----_-_-_-__-_-_-_&__-,-_-__-_-_-_-__-_-_~------------~
0 2 3 4
(II), manganese, zinc, copper and nickel. One possible means of reducing
hydrogen sulphide formation is to add additional of soluble metal salts
to wastewater which will precipitate sulfide salts. The most appropriate
candidates are iron (II) salts, since they are relatively cheap and are likely
to have lesser environmental implications, due to their lower toxicity in
comparison to the other suitable metals.
The addition to wastewater leads to a reaction of the form:
Fe2+ + HS– FeS + H+
However, addition of Fe (III) salts is as effective, if not more so
(Figure 3.32), since in wastewater Fe (III) will be reduced to Fe (II):
2Fe3+ + HS– 2Fe2+ + S0 + H+
Suitable soluble iron (II) compounds are iron (II) chloride (FeCl2),
iron (II) nitrate (Fe(NO3)) and iron (II) sulphate (FeSO4). Suitable iron (III)
compounds are the ferric equivalents: FeCl3, Fe(NO3)3 and Fe2(SO4)3.
It has already been shown that the formation of H2S requires acidic
conditions, and so raising the pH will limit its formation. A pH of greater
than 9 is required, since this will ensure that the vast majority of dissolved
sulphide is present as the dissociated HS– rather than H2S (Table 3.2).
This can be achieved through the addition of sodium hydroxide (NaOH).
Continuous dosing is prohibitively expensive [159], but an alternative
approach is to add large doses to achieve a pH of 12.5–13.0 for periods of
2.5
• Fe (II) addition
• •
0 Fe (Ill) addition
2.0
'§,
E
UJ-
0
.,
\
\
LL 1.5
•\•
<
...J
::J
.,
"'• ,.• •
(f)
..
0 \
LlJ
1.0 cP \
~ qCb 0
'
0
(f)
o<f o ••• • •
"••"" ,.•__
(f)
0 ~cg
0.5
or!} •
()<f3""8&
0
0 • • ~---- •
0.0
8 •
0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
Fe:S RATIO
Figure 3.32 Effect of the addition of Fe (II) and Fe (III) salts on dissolved sulphide
concentrations in wastewater samples [158].
134 Biodeterioration of Concrete
less than an hour [160]. These conditions will reduce H2S formation, but also
have the effect of severely reducing the population of sulphate reducing
bacteria in a sewer system for a number of days, after which the process
must be repeated (see Figure 3.33). One problem with this approach is that
the high pH wastewater resulting from this ‘shock’ dosing cannot undergo
the microbial digestion treatments that sewage will normally undergo
downstream, and so must be diverted and dealt with by alternative means.
Magnesium hydroxide (Mg(OH)2) has been used in a similar manner
to sodium hydroxide [161]. It has the benefit of being ‘self-buffering’—it
maintains pH at around 9.5 to 10.0 despite perturbations in the form of
additions of acidic or alkaline substances. This is because magnesium
hydroxide precipitates as a solid at a pH above around 10.0. Thus, if the
pH begins to rise, magnesium hydroxide will precipitate, removing OH–
ions from solution and reducing the pH. As the pH falls below 10.0, solid
magnesium hydroxide dissolves again, maintaining the pH above 9.5. This
characteristic is beneficial in that it prolongs the high pH conditions, thus
extending the period over which dosing is effective.
Cleaning of sewer pipes has the potential to remove both the biofilms
and deposits of solids in which sulphate reducing bacteria reside, along with
sulphur oxidizing bacteria on pipe surfaces above the waterline. A range
of different techniques can be used, which are selected on the basis of the
size of the sewer, the extent to which deposits have established themselves,
and the types of deposits present [163].
~ 120
0
TIME , days
Figure 3.33 Reduction in sulphide levels following shock dosing of a pressure
main sewage system with sodium hydroxide for two different durations [162].
Bacterial Biodeterioration 135
""'0
20 •0
Control
0.05 mmolll nitrite
E ~ 0.01 mmoll l molybdate
E \1 0.05 mmolll nitrite+ 0.01 mmolll molybdate
z-
0 15
i= ~--·---~-·-~--~~*-~~
,/..,
<(
0::: /e • // o ~
I- I I
z • I
w
.
u 10
I
I I
)f/
z d;
0
u
+I I
I J/
w I I ~
/
0 I I /
5 I
I
p /
l.L. ~__,.
....1 I
=>
(f) _.~/.:
I y ---~ / ~
~="Po~- :::r)V'k~ - I;Z. -v- -'V
0
0 100 200 300 400 500
TIME, hours
Figure 3.34 Effect of using nitrite and molybdate compounds on sulphide
formation by a pure culture of Desulfovibrio sp. sulphate reducing bacteria [165].
Material factors
Coatings
50
40
Ol
Cl) 30
(/)
0_J
(/)
(/)
<( 20
~
10
0 l l
-iy ~
Figure 3.35 Mass loss from samples taken from concrete sewer pipes, treated with
various surface coatings and exposed to sulphur oxidizing bacteria [170].
Polymer modification
Study
[179] [180] [181] [67] [182] [74]
Styrene acrylic ester SB: Lactic/ SB:
improved acetic improved
acid:
improved
Acrylic polymer SB: SB: SB:
worsened worsened worsened
NB:
improved
Styrene-butadiene SB: SB:
improved improved
Vinyl co-polymer SB: SB:
improved improved
Polysiloxane SB:
improved
NB:
improved
Polycarboxylate SB:
improved
NB:
improved
Polymethylmethacrylate H2SO4:
improved
Polystyrene H2SO4:
improved
Polyacrylonitrile H2SO4:
improved
Polyvinyl acetate H2SO4:
improved
SB = sulphur bacteria; NB = nitrifying bacteria
Bacterial Biodeterioration 143
3.7 References
[1] Bennett PC, Rogers JR, Choi WJ and Hiebert FK (2001) Silicates, silicate weathering,
and microbial ecology. Geomicrobiology Journal 18(1): 3–19.
[2] Liu W, Xu X, Wu X, Yang Q, Luo Y and Christie P (2006) Decomposition of silicate
minerals by Bacillus mucilaginosus in liquid culture. Environmental Geochemistry and
Health 28(1-2): 133–140.
[3] Hiebert FK and Bennett PC (1992) Microbial control of silicate weathering in organic-
rich ground water. Science 258(5080): 278–281.
[4] Friedrich S, Platonova NP, Karavaiko GI, Stichel E and Glombitza F (1991) Chemical
and microbiological solubilization of silicates. Acta Biotechnologica 11(3): 187–196.
[5] Duff RB, Webley DM and Scott RO (1963) Solubilization of minerals and related materials
by 2-ketogluconic acid-producing bacteria. Soil Science 95(2): 105–114.
[6] Lian B, Chen Y, Zhu L and Yang R (2008) Effect of microbial weathering on carbonate
rocks. Earth Science Frontiers 15(6): 90–99.
[7] Subrahmanyam G, Vaghela R, Bhatt NP and Archana G (2012) Carbonate-dissolving
bacteria from ‘miliolite’, a bioclastic limestone, from Gopnath, Gujarat, Western India.
Microbes and Environments 27(3): 334–337.
[8] Subrahmanyam G (2014) Bacterial diversity and activity of semiarid soils of Mahi river
basin Western India. Ph.D. Thesis, Maharaja Sayajirao University of Baroda, India, 346
p.
[9] Li W, Yu L-J, Wu Y, Jia L-P and Yuan DX (2007) Enhancement of Ca2+ release from
limestone by microbial extracellular carbonic anhydrase. Bioresource Technology 98(4):
950–953.
[10] Howell JA (1983) Mathematical models in microbiology: mathematical toolkit. In: Bazin
M (ed.). Mathematics in Microbiology. Academic Press, New York, NY, USA.
[11] Viessman W, Hammer MJ, Perez EM and Chadik PA (2009) Water Supply and Pollution
Control. 8th Ed., Pearson, Upper Saddle River, NJ, USA.
[12] Lewandowski Z (2000) Structure and function of biofilms. pp. 1–17. In: Evans LV (ed.).
Biofilms: Recent Advances in their Study and Control. Harwood Acdemic Publishers,
Amsgterdam, The Netherlands.
144 Biodeterioration of Concrete
[13] DeBeer D, Stoodley P, Roe F and Lewandowski Z (1994) Effects of biofilm structures
on oxygen distribution and mass transport. Biotechnology and Bioengineering 43(11):
1131–1138.
[14] Elasri MO and Miller RV (1999) Study of the response of a biofilm bacterial community
to UV radiation. Applied and Environmental Microbiology 65(5): 2025–2031.
[15] Freeman C and Lock MA (1995) The biofilm polysaccharide matrix: A buffer against
changing organic substrate supply? Limnology and Oceanography 40(2): 273–278.
[16] Geesey GG, Jang L, Jolley JG, Hankins MR, Iwaoka T and Griffiths PR (1988) Binding
of metal ions by extracellular polymers of biofilm bacteria. Water and Wastewater
Microbiology 20(11-12): 161–165.
[17] Fang HHP, Xu L-C and Chan K-Y (2002) Effects of toxic metals and chemicals on biofilm
and biocorrosion. Water Research 36(19): 4709–4716.
[18] Benzerara K, Menguy N, López-García P, Yoon T-H, Kazmierczak J, Tyliszczak T,
Guyot F and Brown GE (2006) Nanoscale detection of organic signatures in carbonate
microbialites. Proceeding of the National Academy of Science 103(25): 9440–9445.
[19] Decho AW (2010) Overview of biopolymer-induced mineralization: What goes on in
biofilms? Ecological Engineering 36(2): 137–144.
[20] Bardy SL, Ng SYM and Jarrell KF (2003) Prokaryotic motility structures. Microbiology
149(2): 295–304.
[21] Souza KA, Deal PH, Mack HM and Turnbill CE (1974) Growth and reproduction of
microorganisms under extremely alkaline conditions. Applied Microbiology 28(6):
1066–1068.
[22] Currie RJ (1986) Carbonation depths in structural quality concrete. Building Research
Establishment, Watford, UK.
[23] Buenfield NR and Newman JB (1984) The permeability of concrete in a marine
environment. Magazine of Concrete Research 36(127): 67–80.
[24] Hoła J, Sadowski Ł, Reiner J, Stach S (2015) Usefulness of 3D surface roughness
parameters for nondestructive evaluation of pull-off adhesion of concrete layers.
Construction and Building Materials 84: 111–120.
[25] Garbacz A and Courard L (2006) Characterization of concrete surface roughness and
its relation to adhesion in repair systems. Materials Characterization 56(4-5): 281–289.
[26] Kinnari TJ, Esteban J, Zamora N, Fernandez R, López-Santos C, Yubero F, Mariscal D,
Puertolas JA and Gomez-Barrena E (2010) Effect of surface roughness and sterilization on
bacterial adherence to ultra-high molecular weight polyethylene. Clinical Microbiology
and Infection 16(7): 1036–1041.
[27] Truong VK, Lapovok R, Estrin YS, Rundell S, Wang JY, Fluke CJ, Crawford RJ and
Ivanova EP (2010) The influence of nano-scale surface roughness on bacterial adhesion
to ultrafine-grained titanium. Biomaterials 31(13): 3674–3683.
[28] Riedewald F (2006) Bacterial adhesion to surfaces: the influence of surface roughness.
PDA Journal of Pharmaceutical Science and Technology 60(3): 164–171.
[29] Mitik-Dineva N, Wang J, Mocanasu RC, Stoddart PR, Crawford RJ and Ivanova EP
(2008) Impact of nano-topography on bacterial attachment. Biotechnology Journal 3(4):
536–544.
[30] Giannantonio DJ, Kurth JC, Kurtis K and Sobecky PA (2009) Molecular characterizations
of microbial communities fouling painted and unpainted concrete structures.
International Biodeterioration and Biodegradation 63(1): 30–40.
[31] Barberousse H, Ruot B, Yéprémian C and Boulon G (2007) An assessment of façade
coatings against colonisation by aerial algae and cyanobacteria. Building and
Environment 42(7): 2555–2561.
[32] Dubosc A, Escadeillas G and Blanc PJ (2001) Characterization of biological stains on
external concrete walls and influence of concrete as underlying material. Cement and
Concrete Research 31(11): 1613–1617.
Bacterial Biodeterioration 145
[33] McWhirter MJ, Bremer PJ, Lamont IL and McQuillan AJ (2003) Siderophore-mediated
covalent bonding to metal (oxide) surfaces during biofilm initiation by Pseudomonas
aeruginosa bacteria. Langmuir 19(9): 3575–3577.
[34] Vu B, Chen M, Crawford RJ and Ivanova EP (2009) Review bacterial extracellular
polysaccharides involved in biofilm formation. Molecules 14(7): 2535–2554.
[35] Alcamo IE, Fundamentals of microbiology. Jones and Bartlett, London, 832 p.
[36] Wang D, Cullimore R, Hu Y and Chowdhury R (2011) Biodeterioration of asbestos
cement (AC) pipe in drinking water distribution systems. International Biodeterioration
and Biodegradation 65(6): 810–817.
[37] Badger MR and Price GD (2003) CO2 concentrating mechanisms in cyanobacteria:
molecular components, their diversity and evolution. Journal of Experimental Botany
54(383): 609–622.
[38] Singh SK, Sundaram S and Kishor K (2014) Photosynthetic Microorganisms. Mechanism
for Carbon Concentration. Springer, London, UK.
[39] British Standards Institution, BS EN 197-1:2011 Cement. Composition, specifications
and conformity for common cements. British Standards Institution, London, UK.
[40] Nurse RW (1952) The effect of phosphate on the constitution and hardening of Portland
cement. Journal of Applied Chemistry 2(12): 708–716.
[41] Ifka T, Palou MT and Bazelová Z (2012) The influence of CaO and P2O5 of bone ash
upon the reactivity and the burnability of cement raw mixtures. Ceramics-Silikáty 56(1):
76–84.
[42] Eštoková A and Palaščákováa L (2012) Content of chromium and phosphorus in cements
in relation to the slovak cement eco-labelling. Chemical Engineering Transactions 26:
75–80.
[43] British Standards Institution (2012) BS EN 450-1:2012—Fly ash for concrete Part 1:
Definition, specifications and conformity criteria. British Standards Institution, London,
UK.
[44] Hubbard FH, Dhir RK and Ellis MS (1985) Pulverised-fuel ash for concrete:
Compositional characterisation of United Kingdom PFA. Cement and Concrete Research
15(1): 185–198.
[45] Shehata MH and Thomas MDA (2006) Alkali release characteristics of blended cements.
Cement and Concrete Research 36(6): 1166–1175.
[46] de Larrard F, Gorse J-F and Puch C (1992) Comparative study of various silica fumes
as additives in high-performance cementitious materials. Materials and Structures
25(149): 265–272.
[47] Hobbs DW (1986) Deleterious expansion of concrete due to alkali-silica reaction:
Influence of PFA and slag. Magazine of Concrete Research 38(1376): 191–205.
[48] Dyer TD (2014) Concrete Durability. CRC Press, Boca Raton, FL, USA.
[49] American Society for Testing and Materials, ASTM C150/C150M—Standard Specification
for Portland Cement. American Society for Testing and Materials, West Conshohocken,
Pennsylvania, USA.
[50] British Standards Institution (2006) BS EN 15167-1—Ground granulated blast furnace
slag for use in concrete, mortar and grout—Part 1: Definitions, specifications and
conformity criteria. British Standards Institution, London, UK.
[51] Ortega-Calvo JJ, Sanchez-Castillo PM, Hernandez-Marine M and Saiz-Jimenez C (1993)
Isolation and characterization of epilithic chlorophytes and cyanobacteria from two
Spanish Cathedrals (Salamanca and Toledo). Nova Hedwiga 57(1): 239–253.
[52] Tomaselli L, Lamenti G, Bosco M and Tiano P (2000) Biodiversity of photosynthetic
micro-organisms dwelling on stone monuments. International Biodeterioration and
Biodegradation 46(3): 251–258.
[53] Crispim CA, Gaylarde PM and Gaylarde CC (2003) Algal and cyanobacterial biofilms
on calcareous historic buildings. Current Microbiology 46(2): 79–82.
146 Biodeterioration of Concrete
[54] Crowther RF and Harkness N (1975) Anaerobic bacteria. pp. 65–91. In: Curds CR and
Hawkes HA (eds.). Ecological Aspects of Used Water Treatment—The Organisms and
their Ecology. Vol. 1. Academic Press, London, UK.
[55] Lin ES (2004) A Modelling Study of H2S Absorption of Pure Water and in Rainwater,
Ph.D. thesis, National University of Singapore, Singapore.
[56] Martell AE and Smith RM (2004) Critical Selected Stability Constants of Metal
Complexes Database, Version 8.0 for Windows. National Institute of Standards and
Technology, Gaithersburg, MD, USA. See http://www.nist.gov/srd/nist46.cfm
(accessed 07/06/2016).
[57] Postgate JR (1984) The Sulphate-Reducing Bacteria. 2nd Ed., Cambridge University
Press, Cambridge, UK.
[58] Roberts DJ, Nica D, Zu G and Davis JL (2002) Quantifying microbially induced
deterioration of concrete: initial studies. International Biodeterioration & Biodegradation
49(4): 227–234.
[59] Robertson LA and Kuenen JG (2005) The colorless sulfur bacteria. pp. 985–1011.
In: Balows A, Truper HG, Dworking M, Harder W and Schleifer K-H (eds.). The
Prokaryotes. 2nd Ed., Springer-Verlag, Heidelberg, Germany.
[60] Sand W (1987) Importance of hydrogen sulfide, thiosulfate, and methylmercaptan
for growth of thiobacilli during simulation of concrete corrosion. Applied and
Environmental Microbiology 53(7): 1645–1648.
[61] Nica D, Davis JL, Kirby L, Zuo G and Roberts DJ (2000) Isolation and characterization
of microorganisms involved in the biodeterioration of concrete in sewers. International
Biodeterioration and Biodegradation 46(1): 61–68.
[62] Milde K, Sand W, Wolff W and Bock E (1983) Thiobacilli of the corroded concrete walls
of the Hamburg sewer system. Journal of General Microbiology 129(5): 1327–1333.
[63] Hernandez M, Marchand EA, Roberts D and Peccia J (2002) In situ assessment of active
Thiobacillus species in corroding concrete sewers using fluorescent RNA probes.
International Biodeterioration & Biodegradation 49(4): 271–276.
[64] Hughes DC (1985) Sulphate resistance of OPC, OPC/fly ash, and SRPC pastes: pore
structure and permeability. Cement and Concrete Research 15(6): 1003–1012.
[65] Hill J, Byars EA, Sharp JH, Lynsdale CJ, Cripps JC and Zhou Q (2003) An experimental
study of combined acid and sulfate attack of concrete. Cement and Concrete Composites
25(8): 997–1003.
[66] Jiang G, Wightman E, Donose BC, Yuan Z, Bond PL and Keller J (2014) The role of iron
in sulfide induced corrosion of sewer concrete. Water Research 49: 166–174.
[67] Stanaszek-Tomal E and Fiertak M (2015) Biological and chemical corrosion of cement
materials modified with polymer. Bulletin of the Polish Academy of Sciences Technical
Sciences 63(3): 591–596.
[68] Hewayde E, Nehdi M, Allouche E and Nakhla G (2007) Effect of mixture design
parameters and wetting-drying cycles on resistance of concrete to sulfuric acid attack.
Journal of Materials in Civil Engineering 19(2): 155–163.
[69] Alexander MG and Fourie C (2011) Performance of sewer pipe concrete mixtures with
Portland and calcium aluminate cements subject to mineral and biogenic acid attack.
Materials and Structures 44(1): 313–330.
[70] Guadalupe M, Gutiérrez-Padilla D, Bielefeldt A, Ovtchinnikov S, Hernandez M and
Silverstein J (2010) Biogenic sulfuric acid attack on different types of commercially
produced concrete sewer pipes. Cement and Concrete Research 40(2): 293–301.
[71] O’Connell M, McNally C and Richardson MG (2012) Performance of concrete
incorporating GGBS in aggressive wastewater environments. Construction and Building
Materials 27(1): 368–374.
[72] Girardi F, Vaona W and Di Maggio R (2010) Resistance of different types of concretes
to cyclic sulfuric acid and sodium sulfate attack. Cement and Concrete Composites
32(8): 595–602.
Bacterial Biodeterioration 147
[73] Roy DM, Arjunan P and Silsbee MR (2001) Effect of silica fume, metakaolin, and low-
calcium fly ash on chemical resistance of concrete. Cement and Concrete Research
31(12): 1809–1813.
[74] Fattuhi NI and Hughes BP (1988) Ordinary portland cement mixes with selected
admixtures subjected to sulfuric acid attack. ACI Materials Journal 85(6): 512–518.
[75] De Belie N, Monteny J, Beeldens A, Vincke E, Van Gemert D and Verstraete W (2004)
Experimental research and prediction of the effect of chemical and biogenic sulfuric
acid on different types of commercially produced concrete sewer pipes. Cement and
Concrete Research 34(12): 2223–2236.
[76] Chang Z, Song X, Munn R and Marosszeky M (2005) Using limestone aggregates and
different cements for enhancing resistance of concrete to sulphuric acid attack. Cement
and Concrete Research 35(8): 1486–1494.
[77] Tamimi AK (1997) High performance concrete mix for an optimum protection in acidic
conditions. Materials and Structures 30(3): 188–191.
[78] Torii K and Kawamura M (1994) Effects of fly ash and silica fume on the resistance of
mortar to sulfuric acid and sulfate attack. Cement and Concrete Research 24(2): 361–370.
[79] Siad H, Mesbah HA, Khelafi H, Kamali-Bernard S and Mouli M (2010) Effect of mineral
admixture on resistance to sulphuric and hydrochloric acid attacks in self-compacting
concrete. Canadian Journal of Civil Engineering 37(3): 441–449.
[80] Durning TA and Hicks C (1991) Using microsilica to increase concrete’s resistance to
aggressive chemicals. Concrete International 13(3): 42–48.
[81] Mehta PK (1985) Studies on chemical resistance of low water/cement ratio concretes.
Cement and Concrete Research 15(6): 969–978.
[82] Kim H-S, Lee S-H and Moon H-Y (2007) Strength properties and durability aspects of
high strength concrete using Korean metakaolin. Construction and Building Materials
21(6): 1229–1237.
[83] Sersale R, Frigione G and Bonavita L (1998) Acid depositions and concrete attack: main
influences. Cement and Concrete Research 28(1): 19–24.
[84] Türkel S, Felekoğlu B and Dulluҫ S (2007) Influence of various acids on the physico–
mechanical properties of pozzolanic cement mortars. Sadhana 32(6): 683–691.
[85] Adesanya DA and Raheem AA (2010) A study of the permeability and acid attack of
corn cob ash blended cements. Construction and Building Materials 24(3): 403–409.
[86] Scrivener KL, Cabiron J-L and Letourneux R (1999) High-performance concretes from
calcium aluminate cements. Cement and Concrete Research 29(8): 1215–1223.
[87] Geoffroy VA, Bachelet M and Crovisier JL (2008) Evaluation of aluminium sensitivity of
bacteria on a biodegrading Acidithiobacillus Thiooxidans: Definition of a specific growth
medium. In: Calcium Aluminate Cements: Proceedings of the Centenary Conference,
Avignon, 30 June–2 July 2008. IHS BRE Press, Garston, UK, pp. 309–331.
[88] Jariyathitipong P, Hosotani K, Fujii T and Ayano T (2014) Sulfuric acid resistance
of concrete with blast furnace slag fine aggregate. Journal of Civil Engineering and
Architecture 8(11): 1403–1413.
[89] Thornton HT (1978) Acid attack of concrete caused by sulfur bacteria action. ACI Journal
75(11): 577–584.
[90] Bock E, Koops H-P and Harms H (1986) Cell biology of nitrifying bacteria. In: Proser
JI (ed.). Nitrification. IRL Press, Oxford, UK.
[91] Lipponen MTT, Martikainen PJ, Vasara RE, Servomaa K, Zacheus O and Kontro
MH (2004) Occurrence of nitrifiers and diversity of ammonia-oxidizing bacteria in
developing drinking water biofilms. Water Research 38(20): 4424–4434.
[92] Gieseke A, Bjerrum L, Wagner M and Amann R (2003) Structure and activity of multiple
nitrifying bacterial populations co-existing in a biofilm. Environmental Microbiology
5(5): 355–369.
[93] Lydmark P, Lind M, Sörensson F and Hermansson M (2006) Vertical distribution of
nitrifying populations in bacterial biofilms from a full-scale nitrifying trickling filter.
Environmental Microbiology 8(11): 2036–2049.
148 Biodeterioration of Concrete
[113] Bertron A and Duchesne J (2013) Attack of cementitious materials by organic acids
in agricultural and agrofood effluents. pp. 131–173. In: Alexander M, De Belie N and
Bertron A (eds.). Performance of Cement-based Materials in Aggressive Aqueous
Environments, RILEM State-of-the-Art Report TC 211-PAE. Springer, Dordrecht,
Netherlands,
[114] Yang K, Yu Y and Hwang S (2003) Selective optimization in thermophilic acidogenesis
of cheese-whey wastewater to acetic and butyric acids: partial acidification and
methanation. Water Research 37(10): 2467–2477.
[115] Johansen AG, Vegarud GE and Skeie S (2002) Seasonal and regional variation in
the composition of whey from Norwegian Cheddar-type and Dutch-type cheeses.
International Dairy Journal 12(7): 621–629.
[116] Larreur-Cayol S, Bertron A and Escadeillas G (2011) Degradation of cement-based
materials by various organic acids in agro-industrial waste-waters. Cement and
Concrete Research 41(8): 882–892.
[117] Bertron A, Duchesne J and Escadeillas G (2005) Attack of cement pastes exposed to
organic acids in manure. Cement and Concrete Composites 27(9-10): 898–909.
[118] Dyer TD (2017) Influence of cement type on resistance to attack from two carboxylic
acids. Cement and Concrete Composites (in press).
[119] Dyer TD (2017) Influence of cement type on resistance to organic acids. Magazine of
Concrete Research 69(4): 175–200.
[120] Bertron A, Duchesne J and Escadeillas G (2005) Attack of cement pastes exposed to
organic acids in manure. Cement and Concrete Composites 27(9-10): 898–909.
[121] De Belie N, Debruyckere M, Van Nieuwenburg D and De Blaere B (1997) Concrete
attack by feed acids: accelerated tests to compare different concrete compositions and
technologies. ACI Materials Journal 94(6): 546–554.
[122] De Belie N, De Coster V and Van Nieuwenburg D (1997) Use of fly ash or silica fume
to increase the resistance of concrete to feed acids. Magazine of Concrete Research
49(181): 337–344.
[123] De Belie N, Verselder HJ, De Blaere B, Van Nieuwenburg D and Verschoore R (1996)
Influence of the cement type on the resistance of concrete to feed acids. Cement and
Concrete Research 26(11): 1717–1725.
[124] Bertron A, Duchesne J and Escadeillas G (2005) Accelerated tests of hardened cement
pastes alteration by organic acids: Analysis of the pH. Cement and Concrete Research
35(1): 155–166.
[125] Bertron A, Duchesne J and Escadeillas G (2007) Degradation of cement pastes by organic
acids. Materials and Structures 40(3): 341–354.
[126] Kawai K, Teranishi S, Morinaga T and Tazawa EI (1993) Concrete deterioration caused
by anaerobic bacteria. Translation form Journal of Materials, Concrete Structures and
Pavements 21(478) in Concrete Library of JSCE, 1994, pp. 127–139.
[127] Maruthamuthu S, Saraswathi V, Mani A, Kalyanasundaram RM and Rengaswamy NS
(1997) Influence of freshwater heterotrophic bacteria on reinforced concrete. Biofouling
11(4): 313–323.
[128] Parande AK, Muralidharan S, Saraswathy V and Palaniswamy N (2005) Influence of
microbiologically induced corrosion of steel embedded in ordinary Portland cement
and Portland pozzolona cement. Anti-Corrosion Methods and Materials 52(3): 148–153.
[129] Gaylarde CC and Gaylarde PM (2005) A comparative study of the major microbial
biomass of biofilms on exteriors of buildings in Europe and Latin America. International
Biodeterioration and Biodegradation 55(2): 131–139.
[130] Barberousse H, Lombardo RJ, Tell G and Couté A (2006) Factors involved in the
colonisation of building façades by algae and cyanobacteria in France. Biofouling 22(2):
69–77.
[131] Ruttner F (1963) Fundamentals of Limnology (translated by Frey DG and Fry FEJ).
University of Toronto Press, Toronto, Canada.
[132] Terzaghi RD (1948) Concrete deterioration in a shipway. American Concrete Institute
Proceedings 44(6): 977–1005.
150 Biodeterioration of Concrete
[133] Ballim Y and Alexander MG (1990) Carbonic acid water attack of Portland cement
based matrices. In: Dhir R and Green J (eds.). Protection of Concrete: Proceedings of
the International Conference, University of Dundee, September 1990. Spon, London,
UK.
[134] Colasanti R, Coutts D, Pugh SYR and Rosevear A (1991) The microbiology programme
for UK Nirex. Experientia 47(6): 560–572.
[135] Bayer EA, Shoham Y and Lamed R (2006) Cellulose-decomposing bacteria and their
enzyme systems. pp. 578–617. In: Dworkin M, Falkow S, Rosenberg E, Schleifer K-H
and Stackebrandt E (eds.). The Prokaryotes: Volume 2: Ecophysiology and Biochemistry.
Springer, New York, NY, USA.
[136] Walczak I, Libert M-F, Camaro S and Blanchard J-M (2001) Quantitative and qualitative
analysis of hydrosoluble organic matter in bitumen leachates. Agronomie 21(3): 247–257.
[137] Libert M-F and Walczak I (2000) Effect of radio-oxidative ageing and pH on the release
of soluble organic matter from bitumen. In: Proceedings of ATALANTE 2000 Scientific
Research on the Back-End of the Fuel Cycle for the 21st Century. Avignon, France, pp.
1–4.
[138] Nakayama S, Iida Y, Nagano T and Akimoto T (2003) Leaching behavior of a simulated
bituminized radioactive waste form under deep geological conditions. Journal of
Nuclear Science and Technology 40(4): 227–237.
[139] Bertron A, Jacquemet N, Erable B, Sablayrolles C, Escadeillas G and Albrecht A (2014)
Reactivity of nitrate and organic acids at the concrete–bitumen interface of a nuclear
waste repository cell. Nuclear Engineering and Design 268: 51–57.
[140] Arter HE, Hanselmann KW and Bachofen R (1991) Modelling of microbial degradation
process: the behaviour of microorganisms in a waste repository. Experientia 47(6):
578–584.
[141] Pedersen K (1999) Subterranean microorganisms and radioactive waste disposal in
Sweden. Engineering Geology 52(3-4): 163–176.
[142] MacDonald DD, Roberts B and Hyne JB (1978) The corrosion of carbon steel by wet
elemental sulfur. Corrosion Science 18(5): 411–425.
[143] Boden PJ and Maldonado-Zagal SB (1982) Hydrolysis of elemental sulfur in water and
its effects on the corrosion of mild steel. British Corrosion Journal 17(3): 116–120.
[144] Nemati M, Jenneman GE and Voordouw G (2001) Impact of nitrate-mediated microbial
control of souring in oil reservoirs on the extent of corrosion. Biotechnology Progress
17(5): 852−859.
[145] Kodama Y and Watanabe K (2004) Sulfuricurvum kujiense gen. nov., sp. nov., a
facultatively anaerobic, chemolithoautotrophic, sulfur-oxidizing bacterium isolated
from an underground crude-oil storage cavity. International Journal of Systematic and
Evolutionary Microbiology 54(6): 2297–2300.
[146] Oil and Gas UK (2015) Guidelines on Qualification of Materials for the Abandonment
of Wells, Issue 2, Oil and Gas UK, Aberdeen, UK.
[147] Lundegard PD and Kharaka YK (1994) Distribution and occurrence of organic acids in
subsurface waters. In: Pittman ED and Lewan MD (eds.). Organic Acids in Geological
Processes. Springer-Verlag, London, UK.
[148] US Environmental Protection Agency (1985) Design Manual—Odor and Corrosion
Control in Sanitary Sewerage Systems and Treatment Plants. US Environmental
Protection Agency, Office of Research and Development, Cincinnati, OH, USA.
[149] US Environmental Protection Agency (1991) Technical Report 430/09-91-010—
Hydrogen Sulfide Corrosion in Wastewater Collection and Treatment Systems. US
Environmental Protection Agency, Washington, DC, USA.
[150] Pomeroy RD and Parkhurst RD (1977) The forecasting of sulfide build-up rates in
sewers. Progress in Water Technology 9(3): 621–628.
[151] Stewart PS, Rayner J, Roe F and Rees WM (2001) Biofilm penetration and disinfection
efficacy of alkaline hypochlorite and chlorosulfamates. Journal of Applied Microbiology
91(3): 525–532.
Bacterial Biodeterioration 151
[170] Fattuhi NI and Hughes BP (1988) Ordinary Portland cement mixes with selected
admixtures subjected to sulfuric acid attack. ACI Materials Journal 85(6): 512–518.
[171] De Muynck W, De Belie N and Verstraete W (2009) Effectiveness of admixtures, surface
treatments and antimicrobial compounds against biogenic sulfuric acid corrosion of
concrete. Cement and Concrete Composites 31(3): 163–170.
[172] Concrete Society (1997) Concrete Society Technical Report No. 50—Guide to Surface
Treatments for Protection and Enhancement of Concrete. Concrete Society, Slough, UK.
[173] Valix M, Zamri D, Mineyama H, Cheung WH, Shi J and Bustamante H (2012)
Microbiologically induced corrosion of concrete and protective coatings in gravity
sewers. Chinese Journal of Chemical Engineering 20(3): 433–438.
[174] Aguiar JB, Camōes A and Moreira PM (2008) Coatings for concrete protection against
aggressive environments. Journal of Advanced Concrete Technology 6(1): 243–250.
[175] Roy SK (2002) Coated Concrete—Why, What and How. In: Bassi R and Roy SK (eds.).
Handbook of Coatings for Concrete. Whittles, Caithness, UK.
[176] Vipulanandan C and Liu J (2002) Glass-fiber mat-reinforced epoxy coating for concrete
in sulfuric acid environment. Cement and Concrete Research 32(2): 205–210.
[177] Seneviratne A, Sergi G and Page C (2000) Performance characteristics of surface coatings
applied to concrete for control of reinforcement corrosion. Construction and Building
Materials 14(1): 55–59.
[178] Liu J and Vipulanandan C (2005) Tensile bond strength of epoxy coatings to concrete
substrate. Cement and Concrete Research 35(7): 1412–1419.
[179] Vincke E, Van Wanseele E, Monteny J, Beeldens A, De Belie N, Taerwe L, Van Gemert D
and Verstraete W (2002) Influence of polymer addition on biogenic sulfuric acid attack
of concrete. International Biodeterioration and Biodegradation 49(4): 283–292.
[180] De Belie N and Monteny J (1998) Resistance of concrete containing styrol acrylic acid
ester latex to acids occurring on floors for livestock housing. Cement and Concrete
Research 28(11): 1621–1628.
[181] Monteny J, De Belie N, Vincke E, Verstraete W and Taerwe L (2001) Chemical and
microbiological tests to simulate sulfuric acid corrosion of polymer-modified concrete.
Cement and Concrete Research 31(9): 1359–1365.
[182] Bhattacharya VK, Kirtania KR, Maiti MM and Maiti S (1983) Durability tests on polymer-
cement mortar. Cement and Concrete Research 13(2): 287–290.
[183] Beeldens A, Monteny J, Vincke E, De Belie N, Van Gemert D, Taerwe L and Verstaete
W (2001) Resistance to biogenic sulphuric acid corrosion of polymer-modified mortars.
Cement and Concrete Composites 23(1): 47–56.
[184] Gorninski JP, Dal Molin DC and Kazmierczak CS (2007) Strength degradation of polymer
concrete in acidic environments. Cement and Concrete Composites 29(8): 637–645.
[185] Li G, Xiong G, lü Y and Yin Y (2009) The physical and chemical effects of long-term
sulphuric acid exposure on hybrid modified cement mortar. Cement and Concrete
Composites 319(5): 325–330.
Chapter 4
Fungal Biodeterioration
4.1 Introduction
Fungi are eukaryotes: their cells contain nuclei and organelles both of which
are enclosed by membranes. This configuration of the cell is the same as
that for plants and animals.
Species of fungi can live in terrestrial or aquatic environments,
including seawater. Superficially fungi may seem very like plants—both
include unicellular and multicellular organisms, lack motility, and many
of the multicellular forms share some common morphological features.
However, in taxonomic terms, fungi occupy a separate kingdom to these
other organisms. There are a number of reasons for this. Firstly, unlike plants,
fungi obtain energy and carbon from organic compounds in the surrounding
environment. Secondly, the cell walls of fungi contain the protein chitin, as
opposed to plants, whose cell walls are cellulose.
Fungi can take the form of both single and multicellular organisms.
It is useful to discuss these forms separately, as the differences have some
significance to some of the topics covered later in this chapter.
4.1.1 Yeasts
Yeasts differ from other forms of fungi both structurally and in their means
of reproducing. Yeasts are unicellular, like bacteria, meaning that they are
composed of single discrete cells. Unlike bacteria, yeasts are non-motile, and
so only move when fluids in which they are present are mobile. Some yeast
are capable of attaching to surfaces, which occurs through the formation of
adhesive compounds, including sugars, at the cell surface [1].
Like bacteria, some yeasts reproduce asexually through fission.
However, it is more common for asexual reproduction to occur via budding.
Budding involves the formation of a small bud on the side of the yeast cell.
Once formed, the nucleus of the cell also splits to form two identical nuclei
(mitosis), with one half migrating into the bud. The bud then separates
154 Biodeterioration of Concrete
from the parent cell. Yeasts can also reproduce sexually, usually when the
organism faces a shortage of nutrients or other stress.
Multicellular fungi differ from yeast in that they grow in the form of a system
of tubes which are capable of growth and branch formation (see Figure 4.1).
These tubes are referred to as hyphae, with a grouping of hyphae referred
to as the mycelium. The hyphae are effectively cells, although fungi vary in
how these hyphae are structured—in some cases individual hyphae will
be structured as a single cell, whilst others are divided up into multiple
cells by walls known as septa. However, the septa are normally perforated,
allowing fluid and organelles to move between the partitions.
Beyond this definition of multicellular fungi, there exists a great deal of
structural variation, from relatively simple moulds and rusts, to complex
structures including mushrooms and puffballs. Thus, further discussion of
the physical form that such fungi take will be made as required in discussing
those which affect concrete.
4.1.3 Lichen
Lichen are composite organisms containing more than one species, where
what may be a symbiotic relationship exists. However, the mechanisms by
septum - - - --- - - - - -
septum - - - --- - - - - -
which lichen may affect a concrete substrate on which they are growing are
largely the result of the fungal component of this co-existence.
The organisms that constitute a lichen are fungi (the mycobiont) and
photosynthesising organisms (the photobiont)—either algae (Chapter 5) or
cyanobacteria (Chapter 3). Usually, there are two species involved, but in
some cases three may be present [2].
Lichen adopt many different growth types, including crustose (forming
relatively flat structures whose underside is fully attached to the substrate
on which they grow), foliose (leaf-like) and fruticose (hair-like, strap-shaped
or shrub-like) [3]. Of specific interest with regards to concrete durability
is the endolithic growth form where the lichen exists entirely beneath a
mineral surface [4].
The photobionts are embedded within the structure of the mycobiont,
either evenly distributed through the body of the lichen (the thallus), or
present as discrete layers.
The photobionts in lichens undergo photosynthesis and produce
organic compounds for the mycobiont to use as a source of energy and
carbon. It has been queried whether the relationship is truly symbiotic, on
the grounds that there does not appear to be much benefit to the photobiont.
Indeed, it has been suggested that the relationship is potentially parasitic in
nature [2]. However, it is thought that residing within the fungus provides
the photobiont with a degree of protection from extreme conditions that it
would otherwise not have [6].
polyols such that the solute concentration in the hyphae is higher than that
in seawater [6]. Moulds which grow on salty or sugary foodstuffs, such as
Aspergillus eurotium and various species of Penicillium produce glycerol, or
absorb solutes from the surrounding environment to achieve the same effect.
Under desiccating conditions, lichens also produce polyols, which
replace water molecules associated with macromolecules in their cells. This
allows them to survive long periods of desiccation in a seemingly lifeless
state, which they recover from on re-hydration [6].
Fungi, being heterotrophic do not undergo photosynthesis. Whilst they
utilise light as a means of triggering processes such as spore production
and dispersal, exposure to UV radiation is usually detrimental. For this
reason, many fungi exist with much or all of their mycelium in the dark.
For fungi growing on exposed mineral surfaces, one solution is to produce
pigments such as melanins, carotenoids and mycosporines, which provide
protection from sunlight [18]. As we will see later, another option on at least
some mineral surfaces is to send hyphae beneath the surface.
4.3.2 Reproduction
During the idiophase, once the hyphae begin to encounter the edge of the
substrate on which they are growing the process of reproduction may begin.
This usually takes the form of the formation of spore forming organs, which
can take many different forms.
Fungi are capable of both asexual and sexual reproduction, with both
types of reproduction usually playing a role in their life-cycle. Asexual
reproduction can occur through mycelial fragmentation where a mycelium
splits into two separate mycelia which grow separately from each other—
vegetative growth. More commonly, asexual reproduction occurs through
spore formation. This involves the development of spore forming organs by
the fungus. Hyphae in these organs undergo mitosis, and one of the nuclei is
encapsulated within a spore which grows from the hyphae before detaching.
The spore is released into the wider environment and may be dispersed by
the wind, by flowing water, or on the surface or in the digestive system of
animals. Some spores are motile. Once the spore is in a location conducive
to growth it will begin to form hyphae and grow into a mycelium.
Sexual reproduction occurs when hyphae from two separate mycelia
undergo meiosis to form haploid cells (cells containing a set of unpaired
chromosomes) which become fused together. The process of fusion can take
many forms, but the product is a zygotic diploid cell (containing paired
chromosomes). This cell then undergoes meiosis to form spores, which
are released. The important aspect of the spore formation process from
the perspective of concrete durability is that spores can be travel widely,
meaning that any concrete surface in contact with air or water is potentially
capable of acquiring fungal occupants.
Fungal Biodeterioration 159
4.4.1 pH
Most fungi grow at optimal rates under slightly acidic conditions. This
is illustrated in Figure 4.2 which shows the magnitude of fungal growth
in soil samples from a strip of land over which a pH gradient exists. The
optimum conditions for growth in this case are at a pH of around 5. The
optimum pH for growth, however, is dependent on the substrate in which
the fungus is growing.
Some fungus display optimal growth rates under alkaline conditions.
Some coprinus species (Coprinus radiatus, Coprinus micaceus and Coprinus
ephemerus) and carbonicolous (‘coal-inhabiting’) fungi have been found
to display optimal growth rates under pH conditions close to 8 [22, 23].
160 Biodeterioration of Concrete
_j
0 30
.,..,,
0::
w
~ 25
0
~
!
I
~
I
w I I
I I
..
20
~
1~ ·~I
z- "Ol
0~
I
w
1--
:.c 15
I
I I e
\
~ ~ I \
oo.
..
I \
a. fl ' -t
a:: 10
0
()
z : ' •-----
w : 1 ...
~
()
<(
4 6 10
pH
Figure 4.2 Growth of fungi (measured as acetate incorporation into
ergosterol) in soil samples taken from a strip of soil in which a pH
gradient exists [21].
Ammonia fungi,
early stage species Ambfyosporium botrytis 202
§ 400 --o- Ambfyosporium botrytis 220
'g --.--
......
Peziza urinophifa 165
E -----v--- Pseudombrophifa petrakii 236
.s= Tephrocybe tesquorum 190
~ - ·- D - · - Tephrocybe tesquorum 212
0 300
0, - -+-- Coprinus echinosporus 179
-----<)---- Coprinus echinosporus 215
0 -~- Coprinus phlyctidosporus NA0554
E
0
--10.-- Coprinus phlyctidosporus NA0559
::!: 200
Ol
I--
I
<..9
sUi 100
~
0
o~~~~~~==~~~============~
Ammonia fungi,
late stage species Laccaria bico/or 252
§ 200 --o- Hebefoma vinosophylfum 110
...
:0
<lJ
--.-- Hebe/oma vinosophylfum 135
E -----v--- Hebeloma radicosoides 9
.s= Hebefoma radicosoides SL610603
~
e
0)
15o
0
E
0
::!: 100
Ol
r..:
I
~
w
s 50
>-
c::
0
Col/ybia dryophila 86
§ 200 - -o - Marasmius pulcherripes 77
:0 - -.-- Lyophylfum semita/e AK9301
<lJ
E -- ---v- - - Amanita rubescens SL401801
.s= • Sui/Ius luteus SL710802
~ 150
Ol
0
E
0
::!: 100
Ol
r..:
I
(9
Ui
s 50
~
0
2 4 5 6 7 8 9 10 11 12 13 14
pH
0 weeks
...
1 . 0
:.:" • It :
,. ~
Carbonated
and leached
Non-
weathered
control
lcrn
Figure 4.4 Effect of weathering (carbonation and leaching) on the development of Conosporium
unicinatum on white Portland cement paste prisms [25].
4.4.2 Nutrients
Table 4.1 Stability constants of complexes formed between citrate ions and fungal micronutrient
elements.
Zn + C6H5O7 + H ⇌ Zn(C6H6O7)
2+ 3– +
10.20
Mn Mn + C6H5O7 ⇌ Mn(C6H5O7)
2+ 3– –
4.28
4MoO4 + 2C6H5O
2–
7
3–
+ 10H ⇌ Mo4O12(OH)4(C6H8O7)2
+ 4–
64.69
Table 4.2 Stability constants of complexes formed between oxalate ions and fungal micronutrient
elements.
Zn + C2O4 + H ⇌ Zn(C2HO4)
2+ 2– + +
5.54
Table 4.3 A selection of lichenic acids and some species which produce them [34].
Lichen Species
Parmelia stenophylla
Umbillicaria arctica
Parmelia furfuracea
Parmelia conspersa
Cladonia sylvatica
Baeomyces roseus
Caloplaca elegans
Cladonia furcata
Evernia vulpina
Acid
Baeomycic acid ͻ
O OH
O CH3
HO O
O
CH3 OH
H3C OH
Atranorin ͻ
O OH
O CH3
HO O CH3
O
CH3 O
H3C OH
Lecanoric acid ͻ
CH3
OH OH
O
HO O
O OH
CH3
Olivetoric acid ͻ
HO
OH
O
O
O
OH
H3C OH
H3C O
Gyrophoric acid ͻ
OH CH3 OH
HO O O O
O O OH
CH3 OH CH3
CH3
HO O
O
HO
Fumarprotocetraric acid ͻ ͻ
O
O O CH3
O O
HO
HO
O OH
O CH3 O
OH
Lobaric acid
O O
O OH
H3C OH
O
O
O
CH3
CH3
Physodic acid ͻ
O
O
O OH
OH
O
O
HO
H3C
CH3
Vulpinic acid ͻ
OH
O
O
O O
CH3
Ursolic acid ͻ
H3C
H3C
H3C O
OH
CH3
CH3
HO
H3C CH3
CH3
HO O
O
HO
Usnic acid ͻ ͻ ͻ
H3C
O
O
OH
H3C
H3C O
HO O
O CH3
100.----------------------------------------------.
-•- LaGrange
- -o- Savannah
--T- Atlanta
80
-v- Gainsville
(3
z
::>
_,
LL.
60
>-
co
L.LJ
<.?
<{
0:: 40
L.LJ
>
0
u
20
0 ~----.------------.------------,-----------.-----~
Figure 4.5 Fungal coverage of mortar tile specimens inoculated with fungal
isolates derived from concrete surfaces from various locations in Georgia,
USA. After inoculation, tiles were incubated in the dark with a source of
nutrition (potato dextrose broth) for 7 days [37].
the first of these. Given that nutrients are also in relatively short supply
in concrete itself, it is most probably the greater capacity of high water/
cement ratio to retain moisture that is most important.
Once attached, hyphae begin to grow over the concrete surface. The
hyphae are typically coated in a layer of EPS and exudates of secondary
metabolites, thus forming a biofilm (see Figure 4.6). At locations beneath
Fungal Biodeterioration 169
Calcium
oxalate
crystals
Table 4.4 Fungi identified growing on concrete surfaces, or successfully grown in the laboratory.
Table 4.5 Fungi identified growing on the concrete shield around the damaged reactor at
Chernobyl [47].
all taken from concrete surfaces, and also because the researchers found
different species in different proportions where levels of radiation were
higher, presenting the possibility that the findings are not representative
of a more conventional environment. Indeed, it would appear that some
melanin-forming fungi are capable of radiotrophism—obtaining energy
from ionizing radiation [60].
Nonetheless, many of the fungi identified from the Chernobyl site are
also seen in Table 4.4.
In addition to the genera and species listed in Tables 4.4 and 4.5, a
number of sterile mycelia growing on concrete surfaces have been identified
in characterization studies [45, 47, 51]. Sterile mycelia ‘mycelia sterilia’ are
fungi that do not produce spores, meaning that characterising them in
taxonomic terms is problematic.
Another fungus worthy of mention is Serpula lacrymans, or ‘dry rot’.
This is a fungus which lives and feeds on wood and can potentially cause
severe damage to timber and timber-frame structures. However, it appears
to benefit considerably from growth in close proximity to sources of calcium
and iron. These sources can potentially include mortar, calcium silicate
bricks, mineral insulation wools and concrete [61]. Where such a source
exists, the fungus will grow over the surface of the calcium source, with
possible growth of hyphae beneath the surface of the material [62]. The
calcium is taken up by the hyphae, translocated through the mycelium and
subsequently used in the formation of calcium oxalate through reaction
with oxalic acid. As discussed previously, the acid is formed with the main
objective of assisting in the breakdown of cellulose in the timber. However,
it would appear that calcium oxalate formation is also used as a means of
buffering the pH within a range which is optimal for the organism [63].
The white rot fungus Resinicium bicolor has also been found to use calcium
translocation in a similar manner [64].
Whilst no evidence can be found in the literature of damage to concrete
by Serpula lacrymans, the fact that calcium is removed from its source means
that it is likely that damage of some magnitude occurs.
Table 4.6 lists species of lichen identified growing on concrete and
mortar by researchers.
Genera Species Growth Form Reference Genera Species Growth Form Reference
Acarospora cervina Crustose [67] erysibe Crustose [65, 66]
murorum Crustose [66] turicensis Crustose [68]
subcastanea Crustose [66] Lecanora albescens Crustose [65, 66, 69]
Aspicilia Contorta ssp. Crustose [68] dispersa Crustose [66, 67, 70]
hoffmanniana
Caloplaca sp. – [70]
176 Biodeterioration of Concrete
Genera Species Growth Form Reference Genera Species Growth Form Reference
variabilis Foliose [66] viridula Crustose [68]
Xanthoria sp. – [70]
viridissima Foliose [66] candelaria Foliose [66]
Lecania sp. Crustose [66] fallax Foliose [66, 69]
cuprea Crustose [68] parietina Foliose [65, 66]
178 Biodeterioration of Concrete
Table 4.7 Instances of organic acid production produced by various strains of fungi of the
genera Aspergillus [5].
Acid
Isobutyric
Propionic
Gluconic
Ascorbic
Succinic
Fumaric
Itaconic
Tartaric
Butyric
Formic
Oxalic
Acetic
Lactic
Malic
Citric
Species
Aspergillus flavus 1 1 2
(2 strains)
Aspergillus flavipes 1 1 1 1 1 1
(1 strain)
Aspergillus niger 1 4 7 11 1 8 3 1 10 14 1 9 9
(14 strains)
Aspergillus terreus 1 1 1 1
(1 strain)
studied. It should be noted that more strains of Aspergillus niger were ex-
amined in the study than any of the other species. After oxalic and citric
acids, malic, succinic, tartaric and gluconic and butyric acids are the most
commonly encountered. The action of butyric acid on concrete has already
been examined in Chapter 3. However, the remaining compounds are dis-
cussed below.
4.6.1 Oxalic acid
The action of oxalic acid on concrete is somewhat unusual, in that it can
often have a protective effect. The reason for this can be seen in the solubility
diagrams for oxalic acid in Chapter 2: in the case of Ca, Al and Fe, a solid
phase is precipitated which persists to very low pH values. These solid
phases are calcium oxalate monohydrate (Ca(C2O4).H2O—whewellite),
aluminium trioxalate tetrahydrate (Al2(C2O4)3.4H2O) and iron (III) trioxalate
pentahydrate (Fe2(C2O4)3.5H2O).
We have seen in Chapter 3 that in the case of sulphuric acid attack the
precipitation of salts resulting from reactions between acids and cement
can be problematic if the molar volume of the precipitate is significantly
greater than that of the hydration products it replaces. Whilst the molar
volume of the iron and aluminium salts is not known, the molar volume of
whewellite is 63.8 cm3/mol. Whilst this is larger than the molar volume of
the portlandite that it will replace (32.9 cm3/mol) it is less than the molar
volume of gypsum (74.5 cm3/mol). Thus, rather than cause expansion
cracking in the manner of gypsum, a protective outer layer forms which
prevent further ingress of acid. Thus, mass loss from cement paste and
concrete specimens stored in solutions of oxalic acid is virtually zero [72].
180 Biodeterioration of Concrete
The nature of the interaction of citric acid with concrete is a process with
the potential to cause significant damage. The reason for this has little to do
Fungal Biodeterioration 181
with its strength as an acid, which is relatively low (see Chapter 2). Instead,
the main process of deterioration is the precipitation of calcium citrate
tetrahydrate (Ca3(C6H5O7)2.4H2O). There are, in fact, two different forms
of this salt—the mineral earlandite [76] and a form which is obtained by
precipitation through combining solutions of calcium and citric acid [77].
It is the second form which appears to form when citric acid comes into
contact with cement [78].
Earlandite has a molar volume of 285 cm3/mol, and it is likely that
the molar volume of the second form is similar. This is more than 8 times
greater than portlandite (see Chapter 2). The effect—at least at higher
concentrations—is an extremely rapid and progressive fragmentation at
the concrete surface [73]. Figure 4.9 shows the development of deteriorated
depth in Portland cement paste specimens with time for citric acid and
oxalic acid. Results for acetic acid are also shown to provide comparison
with an organic acid that largely causes deterioration by acidolyisis. The
rate of deterioration in the case of citric acid is substantially greater than
acetic acid.
Figure 4.10 shows mass loss from cement pastes of various types
at two different citric acid concentrations, whilst Figure 4.11 shows
micro-CT cross-sections through the specimens. In the case of the higher
concentration of citric acid, the greatest resistance is displayed by the
calcium sulphoaluminate cement, and the least by the PC/fly ash blend.
This is also reflected in the CT scans, where the dramatic disintegration
14
• Oxalic acid
0
12 I 0 Citric acid
I 'I' Acetic acid
E I
E I
I - 10 6
1- I
a.. I
w I
I
0 8
--------.,--
"'--
--
0 0
w I
1- I - ~- -
<( 6 I
0:: I /
/
0 I /
/~
ii 9/
w 4 /
1- I ,,/ 'I'
w I
0 I y/
2
9/
I/
0
0 100
- -
200 300
TIME, days
100
90
C/)
C/)
Cll
E 80
Cll
>-
""0
4-- 70
0
~
0
(/)
(/) 60 e PC, 0.10M
<(
0 65% PC /35% FA, 0.1 0 M
2:
"' GSA, 0.10 M
\1 PC, 0.01 M
50
• 65% PC /35% FA, 0.01 M
o GSA, 0.01 M
40
0 20 40 60 80 100
TIME, days
Figure 4.10 Loss of mass of cement paste specimens (water/cement ratio = 0.5) made from
Portland cement (PC), a PC/FA blend, and a calcium sulphoaluminate cement exposed to
0.10 and 0.01 M solutions of citric acid [78] plus other data.
Figure 4.11 micro-CT scans showing cross-sections through cement paste specimens
exposed to a 0.1 M solution of citric acid for 14 days [78].
of the cement is evident, with the fly ash specimen having undergone the
greatest loss of material. The reason for this is most probably related to the
relative abilities of the materials to resist fragmentation. All of the cement
pastes had the same water/cement ratios, and so it is likely that the PC/
FA blend would have the lowest strength, thus making it most susceptible
to fragmentation. Since fragmentation will uncover the unaffected cement
within, allowing it to be attacked, deterioration will progress at a faster rate.
Fungal Biodeterioration 183
14 .-----------------------------------------------------,
12
---
- - - - "'7: .
..--~-
10 /
/ .. ··
I
I
8 I
I
I I
0..
------- ---
~I
6
/ . ..-:. ~ ···
,,,.•""" .
4
ll /
PC
2 65% PC I 35% FA
GSA
0+------,,------,------,-------,------,------~
what is occurring for the CSA cement paste exposed to the dilute solution
in Figure 4.12.
The citrate ion is capable of forming relatively strong complexes with
calcium, aluminium and iron. It is also capable of forming complexes with
silicon (see Chapter 2). Within the context of concrete durability, this is of less
concern, since fragmentation will typically remove material indiscriminately
at a greater rate.
Tartaric acid is a relatively weak acid, but one which is capable of forming
insoluble calcium tartrate on contact with cement. The influence of calcium
tartrate precipitation on the integrity of cement is a mixture of positives
and negatives.
Calcium tartrate is an expansive product, with a molar volume of 144
cm3/mol. This effect is seen in Figure 4.13 which shows micro-CT scans
obtained from a series of cement paste specimens exposed to a 0.1 M
solution. The expansive nature of the product is evident from the images
obtained from PC and PC/FA pastes, with a thick product having formed
within the silica-gel layer of the specimens, which has been considerably
disrupted, leading to partial delamination from the surface.
Fungal Biodeterioration 185
Figure 4.13 micro-CT scans showing cross-sections through cement paste specimens
exposed to a 0.1 M solution of tartaric acid for 90 days [78].
14
• Tartaric acid
12 0 Malic acid
E "' Acetic acid
E v Succinic acid
::C 10
I-
D...
UJ
0 8
0
UJ
I-
<( 6
0::
0
0::
UJ 4
I-
UJ
0
2
0
0 50 100 150 200 250 300 350
TIME, days
Figure 4.14 Development of degraded depth in cylindrical Portland cement
paste specimens exposed to 0.1 M solutions of tartaric, malic, acetic and succinic
acids [72, 73].
60
----
0 Malic acid
50
"'
'V
Acetic acid
Succinic acid
................ / \1
/
~ 40 /.g 'V
0
(/)- ~ _ _ :!!
(/) V'9
--.-
v/
0_.J 30
_________
"
/ _()
(/) ,q ......
(/)
<( /~
~ 20 r.:~·~_g--0
. .
'S/ .v
~ly r:P
v/j ____ ---- --------.
..
10
!o
,_
0
~- _,. .;-.~· --
0 100 200 300 400
TIME, days
Figure 4.15 Mass loss from cylindrical Portland cement paste specimens exposed
to 0.1 M solutions of tartaric, malic, acetic and succinic acids [72, 73].
Fungal Biodeterioration 187
10°.-------------------------------------------------,
!-------------------=======l
§1l
<:::
0 10-1
E
6
w
~
:~:: c:=c:-~~~~~=~i I".
0 10-4 11 e
(f) ~·.
~ 10-' I:/
0
I- 10-a I
10-9 +----.----.----,----,----.----.----.----.----,,---~
~·-·-·-·-·-·- · - · -·-·-·-·- · - 12
I
I 10
0.03
I pH
8 I
c.
0
"'
Q)
__ ___ . / 6
E
~
0.02 - ·- ·- 4
I- Calcium malate trihydrate
2
i=
z
<(
~
a (Ca/Si=O.B)
0.01
Portland ite
0 2 3 4 5 6 7 8 9 10
Figure 4.16 Concentrations of dissolved elements (top) and quantities of solid phases obtained
from geochemical modelling of malic acid attack of hydrated Portland cement. Model
conditions: acid concentration = 0.1 mol/l; volume of acid solution = 4 l; mass of cement 80
g; diffusion coefficient = 5 × 10–13 m2/s.
10°
-=:: 10-1
0
E 10-2
0 - -·- ·
-----
-:-- ·· - ··- ·
L.U 10-3
> -'\..
· · §a·
0 "--
.
(/)
(/) 10-5
0 10-6
....J
<( ····· ····· ··· ···· · Si
I- 10-7
0 ------ AI
I- 10-8 - -- - Fe
10-9
/
.~·-·- · - · - · - · - · -·-- · -·-·-·-·-·-· 12
I
10
0.03 I
I pH 8 I
c_
C/l I
Q) 6
0 I
E ·-·-·-'
4
>-- 0.02
I-
CSH (Ca/Si=O.B) 2
i= Gibbsite
z
<(
~ Ettringite (AI)
a
0.01
CSH (Ca/Si=1.1)
Portlandite
0 2 3 4 5 6 7 8 9 10
10°
"'
0 1Q-1 Ca
E 10-2
Si
6 AI
LlJ 10-3 Fe
~ 1Q-4
0
(/)
(/) 10-5
15 1Q-6
...J
~ 1Q-7
0
1- 10"'
1o-•
, ·- ·-·-·-·-·-·-·-·-·-·-·-·- 12
(
0.03 I 10
I pH
8 ~
I
({)
Q) ) 6
0
E
0.02
- -- ·- 4
~ Calcium succinate trihydrate
2
i=
z
<(
::>
0 C3AH 6
0.01
0 2 3 4 5 6 7 8 9 10
14
E
E ,.ecitric
(/)-
12 /
>-
<{
/
/
0 /
0
0
10 /
/
1- /
<{ /
8 /
I
1- /
a.. /
L.LJ /
0 6 /
0 Succinice /
L.LJ /
1- /
<{ 4 /
0::: /
0 /
0::: 2 /
L.LJ / eMalic
1- /
L.LJ
0 Oxalic _/ ;eTartaric
0
3 4 5 6 7
Figure 4.19 Deteriorated depth at 100 days of Portland cement pastes exposed to
0.1 M solutions of various organic acids produced by fungi versus the logarithm
of each compound’s first acid dissociation constant pKa1. Deteriorated depth data
from [72, 73].
14 .-------------------------------------------------~
-Control
12 r::::::l Mycelia sterila
- Aspergillus niger
0 +-------v
• _______ I I-Y------------cr--------y-.---------------J
>0.9 0 9-0 06 0 06-0 009 <0 009
500
___._ Aspergillus niger
0 · · Mycelia sterila
400
z
6 300
<(
0
--'
l1J
0:::
::J 200
--'
<(
L.L.
100
0
0 2 4 6 8 10 12
TIME , months
Figure 4.21 Loss in flexural strength (expressed as the failure load under
three-point loading) resulting from growth of Aspergillus niger and a sterile
fungus (Mycelia sterila) [45].
194 Biodeterioration of Concrete
The most deteriorated area investigated was the basement of the bakery
where there was a high quantity of airborne flour dust, which presumably
promoted growth as a source of carbon and energy. Aspergillus niger was
prevalent in this environment.
Deterioration of concrete by fungi was also identified in a UK study. In
this case the fungi were growing below the waterline in a partially flooded
cellar [49]. The fungi involved were sampled and isolated and identified as
being Aspergillus glaucus and a Penicillium species. The isolated fungi were
then placed in flasks with growth medium and small concrete specimens. A
drop in pH of around 1.5 after 40 days was observed in the flask containing
Aspergillus glaucus, which corresponded to a loss in mass from the concrete
specimen of around 6%. Calcium was found to be the main element leached
from the specimen during mass loss. The Penicillium fungus did not produce
a drop in pH and mass loss was of a lesser magnitude.
The influence of two lichen species (Acarospora cervina and Candelariella
vitelline) growing on asbestos cement roofs has been evaluated [65]. In the
case of Acarospora cervina, the roof material on which it was growing was
found to contain traces of whewellite, indicating oxalic acid production.
This phase was not identified in the roof on which Candelariella vitelline
was growing, which is known to produce pulvinic acid derivatives instead.
Damage to the roofs was not quantified, but it was proposed that the lichen
were altering the nature of the asbestos fibres, rendering them amorphous
and less toxic. The study was conducted from the perspective of the health
hazards associated with asbestos, and so this effect, plus the physical
coverage of the roof by lichen which was presumed to prevent dispersal of
asbestos fibres into the wider environment, was considered a favourable one.
One notable feature of the results of experiments investigating fungal
deterioration of concrete is that, despite the presence of fungi known for
the production of both oxalic and citric acid (such as Aspergillus niger), only
calcium oxalate has been observed when the concrete is analysed using
powder X-ray diffraction, and only in some cases. From the group of studies
where this technique was used to analyse concrete or cement surfaces
affected by fungi or lichen, only three positively identify calcium oxalate
[38, 67, 54]. In all other cases it is absent [45, 68], although it is incorrectly
identified in one instance.
The reason for this is not clear. It is certainly true that calcium oxalate is
the less soluble of the two compounds, meaning that, where the two acids
are present, calcium oxalate will precipitate in preference to calcium citrate.
Thus, the absence of the citrate salt is conceivable. However, the absence
of calcium oxalate is harder to explain. In some instances, the absence is
simply because oxalic acid is not produced by the fungi. It is also possible
that calcium oxalate is being dissolved by other acids. However, the salt
remains insoluble even to very low pH, meaning that, if this were the case,
196 Biodeterioration of Concrete
the acid involved would need to be one which forms very strong complexes
with calcium.
Oxalic acid is also decomposed to hydrogen peroxide and carbon
dioxide when illuminated by sunlight in the presence of oxygen. Whilst
direct decomposition of calcium oxalate does not appear to occur, dissolution
of oxalate ions into water at a concrete surface would allow the gradual
removal of oxalate. The reaction is significantly catalyzed by the presence
of dissolved iron (III). It has been determined that under UV intensities of
0.65 m Einstein/m2s, 1 μM dissolved iron (III) was able to decompose oxalic
acid at a rate of 10 nM/s [80].
In at least one case, calcium oxalate was found to be absent at the fungi-
inhabited surface of a cement paste specimen, despite analytical evidence
that oxalic acid was being produced. It had previously been thought that
once calcium oxalate was precipitated, fungi were subsequently unable
to utilise the oxalate in any manner [81]. Recently, however, experiments
conducted into calcium oxalate formation by a range of fungi found
that, for some species, there was a subsequent disappearance of oxalate
crystals [82]. It was proposed that the fungi were releasing enzymes such
as decarboxylase or oxalate oxidase which were breaking down the oxalate
to produce hydrogen peroxide (H2O2) which could be used in the oxidative
decomposition of cellulose.
The fungi species found to be capable of this process were Pleurotus
tuberregium, Polyporus ciliates, Agaricus blazei, Pleurotus eryngii, Pleurotus
citrinopileatus, Pycnoporus cinnabarinus and Trametes suaveolens. It should
be stressed that none of these species have been identified growing on
concrete surfaces.
protect surfaces [67, 85]. With this in mind, the decision of whether lichen
needs to be removed on aesthetic grounds needs to be made on the basis of
the nature of the building, its surroundings and, potentially, the opinions
of its users.
Nonetheless, there are functional reasons why both fungi and lichen
may need to be removed. One of these is that, in wet conditions, colonised
surfaces may present a slip hazard. Moreover, where concrete surfaces are
adjacent to timber components in a structure, removal may be prudent to
avoid colonisation and deterioration of the wood by fungi.
3
.$
·c: I
:::J
I
.
i::' I
jg I
I
:0 I
Cii
I' 2 ., I
I
~
0 I
I
0::: I
(j I
I
(5 I
z I
::J
LL •• I
I• •
LL I
0 I
..
LlJ I
LlJ I
0::: I
(j I
LlJ
0 0 /
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
nutrient (see Section 4.4.2) the limited sulphate levels in the cement may
be the reason for the absence of fungi.
Regardless of its influence of on fungal growth rates, cement type will
also influence rates of deterioration. There is incomplete understanding
of this area with respect to the acids typically produced by fungi, and
performance is dependent on the type of acids. However, it would appear
that, in most cases PC or calcium aluminate cements offer a slight advantage.
In the case of tartaric acid, PC/fly ash blends, and possibly PC slag blends
may be more beneficial. A feature of many of the acids formed by fungi is
that they form an insoluble and often expansive salt with calcium. Moreover,
biophysical processes operate by generating stresses within concrete which
lead to fracture. There is, therefore, a very strong argument, where fungal
attack is a possibility, for employing concrete with a low water—cement
ratio, and a hence high strength to resist the stresses associated with
expansion. This has the secondary benefit of limiting rates of colonisation
and growth through reduced water absorption (see Section 4.4.3).
The mechanical aspects of attack from fungi means that the choice of
aggregate used will have a lesser influence on performance in comparison
to modes of deterioration such as bacterial attack, where acidolysis can
be checked by the neutralising capacity of calcareous aggregate. Indeed,
calcareous aggregate will presumably also form expansive salts in contact
with fungi-derived acids.
Another approach to limiting fungal growth employs titanium dioxide
(TiO2) powder. The use of TiO2 in concrete exposed to the atmosphere as a
means of maintaining a clean surface and reducing atmospheric pollution
has been a developing technology over the past 20 years. TiO2 acts as a
photocatalyst: when illuminated with sunlight the surface of the compound
reacts with oxygen and water in the atmosphere to produce oxidising and
reducing species which are subsequently able to react with and break down
various airborne pollutants including oxides of nitrogen (NOx), sulphur
dioxide (SO2) and various volatile organic compounds (VOCs) [89]. The
anatase form of TiO2 is most effective. These reactions also act on non-
volatile organic substances which have become attached to the concrete
surface, ultimately leading to their degradation to an extent which allows
them to be easily washed away from the surface by rainwater.
The effectiveness of this so-called ‘self-cleaning’ characteristic of
concrete containing TiO2 with regards to preventing the colonisation and
growth of fungi on concrete found the approach to be successful [37]. A series
of mortar tiles were prepared containing a small quantity (unspecified)
of anatase nanoparticles in the cement. The tiles were inoculated, along
with mortar tiles made with the same cement without TiO2, with various
fungal strains and incubated for a week in an environment where a nutrient
solution was sprinkled periodically onto their surface. The tiles were lit
for 6 hours with near-UV radiation followed by 6 hours of darkness. After
Fungal Biodeterioration 201
incubation, the tiles with TiO2 were found to be considerably more resistant
to fungal colonisation than the controls.
Whilst TiO2 particles in concrete appear to be effective against fungal
colonization, long-term performance is currently not well-understood.
Certainly it appears that where levels of surface contamination by dirt are
high, the material does not perform as well [90]. Moreover, where surfaces
are horizontal, and hence the self-cleaning process is not assisted by gravity,
accumulation of dirt—and presumably fungus—is still possible.
It is possible to add fungicidal admixtures at the mixing stage of concrete
production. There are a wide variety of substances with fungicidal effects.
These include compounds containing copper, zinc or boron, disinfectants
(such as sodium hypochlorite), quaternary ammonium compounds and
various organic compounds which are toxic to fungi. However, when
used as an admixture it is also necessary that a fungicide does not have a
detrimental effect on concrete properties.
Much of the guidance regarding anti-fungal admixtures identifies
polyhalogenated phenols and copper compounds as suitable candidates,
but this is somewhat outdated [91].
Research into various potential formulations for copper-based antifungal
admixtures examined a wide range of soluble candidate compounds [92].
The study measured growth of Aspergilus niger and Penicillium viridis on
the surfaces of concrete specimens made from concrete which had various
concentrations of the compounds added. Of the compounds evaluated,
copper arsenite and copper acetoarsenite were found to be most effective,
although additions of at least 5% by mass of cement were required. The
use of copper based admixtures, however, has side-effects. The presence
of copper will normally retard the hydration reactions of Portland cement
[93], and it was found that the compressive strength of the concrete mix
containing 10% copper acetoarsenite was around 80% less than the control.
Setting time was, in fact, accelerated. Additionally, copper-based admixtures
will normally impart an—often dramatic—colour change to concrete.
The main concern with such admixtures however, is the presence of
arsenic which is hazardous to human health and a possible carcinogen.
As a result, these compounds are now restricted in many countries. Other
copper-based compounds may potentially suitable, but the retardation of
cement hydration is far from ideal.
Polyhalogenated phenols have also been demonstrated as being
effective against mould growth when used as a concrete admixture [94].
Again, there are concerns regarding the human health and environmental
hazards presented by many of these compounds, and many substances are
now strictly regulated. Indeed, the main reason for exercising caution before
deciding upon the use of biocidal admixtures in concrete is the potential
for releasing ecotoxic substances into the wider environment.
202 Biodeterioration of Concrete
0
0
ADJUSTED MEAN LICHEN COVERAGE %
I
co
0
I
.I
I
I
I
Q
I ;;-
l iil
l'lE
l (ij
lcQ
IJ
1w
10
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
I
~
(J)
I
0
I
I
..,.
0
I
I
I
N
0
_I
0
Tap water
Savlon
Barium metaborate
Quatramine 80 1%
Quatramine 80 4%
Haipen 50 EC
Mitrol PT
Disodium ethylene-diaminetetraacetate
Cupric ethanolamine
Cupric sulfate
Zinc hexafluorosilicate
Susan 30
Bromosulfona
Tertiary amine
Potassium permanganate
50% Benomyl : 50% Vitavax
Sodium borate
PermaprufT
@- nwo
-ooro
5}3-<'
_ _:::J
3 3 Ol
:::l c Ol
~
3 c
c0 ::::J
3o
6_c.
(/)N
nco
c ::::J
::::J
0 Ol
uc
0 -
0
iii'
5'
it
u;
!!!.
§?;>
en Q.
::::J
!2.3
0.(1)
<D~
'8. 3
no
3:::J"
o~
"0~
c.
::::J
(/)
c
0
Figure 4.24 Lichen coverage of asbestos cement roof panels after treatment with potentially
fungicidal additives. Panels initially had c 70% lichen coverage. After treatment, panels were
left outside for a period of 83 months prior to evaluation of coverage [112].
4.9 References
[1] Buck JW and Andrews JH (1999) Attachment of the yeast Rhodosporidium toruloides
is mediated by adhesives localized at sites of bud cell development. Applied and
Environmental Microbiology 65(2): 465–471.
[2] Nash TH (2010) Introduction. In: Nash TH (ed.). Lichen Biology. Cambridge University
Press, Cambridge, UK.
[3] Büdel B and Scheidegger C (2010) Thallus morphology and anatomy. In: Nash TH (ed.).
Lichen Biology. Cambridge University Press, Cambridge, UK.
[4] Lisci M, Monte M and Pacini E (2003) Lichens and higher plants on stone: a review.
International Biodeterioration and Biodegradation 51(1): 1–17.
[5] Liaud N, Giniés C, Navarro D, Fabre N, Crapart S, Herpoël-Gimbert I, Levasseur
A, Raouche S and Sigoillot J-C (2014) Exploring fungal biodiversity: organic acid
production by 66 strains of filamentous fungi. Fungal Biology and Biotechnology 1(1):
1–10.
[6] Jennings DH and Lysek G (1996) Fungal Biology: Understanding the Fungal Lifestyle.
Bios, Oxford, UK.
[7] Gadd GM (1999) Fungal production of citric and oxalic acid: importance in metal
speciation, physiology and biogeochemical processes. Advances in Microbial Physiology
41: 47–92.
[8] Dutton MV and Evans CS (1996) Oxalate production by fungi: its role in pathogenicity
and ecology in the soil environment. Canadian Journal of Microbiology 42(9): 881–895.
[9] Gharieb MM and Gadd GM (1999) Influence of nitrogen source on the solubilization of
natural gypsum (CaSO4.2H2O) and the formation of calcium oxalate by different oxalic
and citric acid-producing fungi. Mycological Research 103(4): 473–481.
[10] Takao S (1965) Organic acid production by Basidiomycetes. I. Screening of acid-
producing strains. Applied Microbiology 13(5): 732–737.
[11] Horner HT, Tiffany LH and Cody AM (1983) Formation of calcium oxalate crystals
associated with apothecia of the discomycete Dasyscypha capitate. Mycologia 75(3):
423–435.
[12] Fink S (1991) Unusual patterns in the distribution of calcium oxalate in spruce needles
and their possible relationships to the impact of pollutants. The New Phytologist 119(1):
41–51.
[13] Hang YD, Splittstoesser DF, Woodams EE and Sherman RM (1977) Citric acid
fermentation of brewery waste. Journal of Food Science 42(2): 383–384.
[14] Pintado J, Torrado A, González MP and Murado MA (1998) Optimization of nutrient
concentration for citric acid production by solid-state culture of Aspergillus niger on
polyurethane foams. Enzyme and Microbial Technology 23(1-2): 149–156.
[15] Clark DS, Ito K and Horitsu H (1966) Effect of manganese and other heavy metals on
submerged citric acid fermentation of molasses. Biotechnology and Bioengineering
8(4): 465–471.
[16] Vandenberghe LPS, Soccol CR, Pandey A and Lebeault J-M (1999) Microbial production
of citric acid. Brazilian Archives of Biology and Technology 42(3): 263–276.
[17] Kubicke CP, Zehentgruber O, El-Kalak H and Roehr M (1980) Regulation of citric
acid production by oxygen; effects of dissolved oxygen tension on adenylate levels
Fungal Biodeterioration 207
[38] Fomina M, Podgorsky VS, Olishevska SV, Kadoshnikov VM, Pisanska IR, Hillier S
and Gadd GM (2007) Fungal deterioration of barrier concrete used in nuclear waste
disposal. Geomicrobiology Journal 24(7-8): 643–653.
[39] Bowen AD, Davidson FA, Keatch R and Gadd GM (2007) Induction of contour sensing
in Aspergillus niger by stress and its relevance to fungal growth mechanics and hyphal
tip structure. Fungal Genetics and Biology 44(6): 484–491.
[40] Maheshwari R (2016) Fungi: Experimental Methods In Biology, Second Edition.
Mycology. CRC Press, Boca Rato, FL, USA.
[41] Burford EP, Kierans M and Gadd GM (2003) Geomycology: fungi in mineral substrata.
Mycologist 17(3): 98–107.
[42] Money NP (2004) The fungal dining habit: a biomechanical perspective. Mycologist
18(2): 71–76.
[43] Bechinger C, Giebel K-F, Schnell M, Leiderer P, Deising HB and Bastmeyer M (1999)
Optical measurements of invasive forces exerted by appressoria of a plant pathogenic
fungus. Science 285(5435): 1896–1899.
[44] Gorbushina AA, Krumbein WE, Hamman CH, Panina L, Soukharjevski S and
Wollenzien U (1993) Role of black fungi in color change and biodeterioration of antique
marbles. Geomicrobiology Journal 11(3): 205–211.
[45] Perfettini JV, Revertegat E and Langomazini N (1991) Evaluation of cement degradation
induced by the metabolic products of two fungal strains. Experientia 47(6): 527–533.
[46] Gaylarde CC and Gaylarde PM (2005) A comparative study of the major microbial
biomass of biofilms on exteriors of buildings in Europe and Latin America. International
Biodeterioration and Biodegradation 55(2): 131–139.
[47] Zhdanova NN, Zakharchenko VA, Vember VV and Nakonechnaya LT (2000) Fungi
from Chernobyl: mycobiota of the inner regions of the containment structures of the
damaged nuclear reactor. Mycological Research 104(12): 1421–1426.
[48] Giannantonio DJ, Kurth JC, Kurtis KE and Sobecky PA (2009) Molecular characterization
of microbial communities fouling painted and unpainted concrete structures.
International Biodeterioration and Biodegradation 63(1): 30–30.
[49] McCormack K, Morton LHG, Benson J, Osborne BN and McCabe RW (1996) A
preliminary assessment of concrete biodeterioration by microorganisms. pp. 68–70. In:
Gaylarde CC, de Sá ELS and Gaylarde PM (eds.). Biodegradation and Biodeterioration
in Latin America. Mircen/UNEP/UNESCO/ICRO-FEPAGRO/ UFRGS, Porto Alegre,
Brazil.
[50] Lajili H, Devillers P, Grambin-Lapeyre C and Bournazel JP (2008) Alteration of a cement
matrix subjected to biolixiviation test. Materials and Structures 41(10): 1633–1645.
[51] Nielsen KF, Holm G, Uttrup LP and Nielsen PA (2004) Mould growth on building
materials under low water activities. Influence of humidity and temperature on fungal
growth and secondary metabolism. International Biodeterioration and Biodegradation
54(4): 325–336.
[52] Gu J-D, Ford TE, Berke NS and Mitchell R (1998) Biodeterioration of concrete by the
fungus Fusarium. International Biodeterioration and Biodegradation 41(2): 101–109.
[53] Koval EZ, Serebrenik VA, Roginskaya EL and Ivanov FM (1991) Mycodestructors of
building structures of inner premises of food industry enterprises. Microbiologichny
Zhurnal 53: 96–103 (in Russian).
[54] Fomina MO, Olishevska SV, Kadoshnikov VM, Zlobenko BP and Pidgorsky VS (2005)
Concrete colonization and destruction of mitosporic fungi in model experiment.
Mikrobiolohichnyĭ Zhurnal 67(2): 96–104.
[55] Park J-M, Park S-J, Kim W-J and Ghim S-Y (2012) Application of antifungal CFB to
increase the durability of cement mortar. Journal of Microbiology and Biotechnology
22(7): 1015–1020.
[56] Park J-M, Park S-J and Ghim SY (2013) Characterization of three antifungal calcite-
forming bacteria, Arthrobacter nicotianae KNUC2100, Bacillus thuringiensis KNUC2103,
and Stenotrophomonas maltophilia KNUC2106, derived from the Korean islands, Dokdo
Fungal Biodeterioration 209
[74] Giordani P and Modenesi P (2003) Determinant factors for the formation of the calcium
oxalate minerals, weddellite and whewellite, on the surface of foliose lichens. The
Lichenologist 35(3): 255–270.
[75] Salvadori O and Tretiach M (2002) Thallus-substratum relationships of silicicolous
lichens occurring on carbonatic rocks of the Mediterranean region. Bibliotheca
Lichenologica 82: 57–64.
[76] Herdtweck E, Kornprobst T, Sieber R, Straver L and Plank J (2011) Crystal
structure, synthesis, aund properties of tri-calcium di-citrate tetra-hydrate
(Ca3(C6H5O7)2(H2O)2).2H2O. Zeitschrift für Anorganische und Allgemeine Chemie 627:
655–659.
[77] Kaduk JA (2013) Crystal structures of group 2 citrate salts, abstract from ICDD Annual
Spring Meetings. Powder Diffraction 28(2): 146.
[78] Dyer T (2016) Influence of cement type on resistance to organic acids. Magazine of
Concrete Research (in press).
[79] Peynaud E (1984) Knowing and Making Wine. John Wiley, Chichester, UK.
[80] Zuo Y and Hoigné J (1994) Photochemical decomposition of oxalic, glyoxalic and
pyruvic acid catalysed by iron in atmospheric waters. Atmospheric Environment 28(7):
1231–1239.
[81] Foster JW (1949) Chemical Activities of Fungi. Academic Press, New York, NY.
[82] Guggiari M, Bloque R, Aragno M, Verrecchia E, Job D and Junier P (2011) Experimental
calcium-oxalate crystal production and dissolution by selected wood-rot fungi.
International Biodeterioration and Biodegradation 65(6): 803–809.
[83] Warscheid T and Braams J (2000) Biodeterioration of stone: a review. International
Biodeterioration and Biodegradation 46(4): 343–368.
[84] Shannon PM, Unnithan V, Bouriak S and Chachkine P (1998) Role for lichen melanins
in uranium remediation. Nature 391: 649–650.
[85] Mottershead D and Lucas G (2000) The role of lichens in inhibiting erosion of a soluble
rock. Lichenologist 32(6): 601–609.
[86] McIlroy de la Rosa JP, Warke PA and Smith BJ (2012) Lichen-induced biomodification
of calcareous surfaces: Bioprotection versus biodeterioration. Progress in Physical
Geography 37(3): 325–351.
[87] Little B, Staehle R and Davis R (2001) Fungal influenced corrosion of post-tensioned
cables. International Biodeterioration and Biodegradation 47(2): 71–77.
[88] Strigáč J and Martauz P (2016) Fungistatic properties of granulated blastfurnace slag
and related slag-containing cements. Ceramics-Silikáty 60(2): 19–26.
[89] Macphee DE and Folli A (2016) Photocatalytic concretes—The interface between
photocatalysis and cement chemistry. Cement and Concrete Research 85: 48–54.
[90] Yu JC-M (2006) Deactivation and regeneration of environmentally exposed Titanium
Dioxide (TiO2) based products. Testing report prepared for the environmental protection
department. Chinese University in Hong Kong, Shatin, Hong Kong.
[91] American Concrete Institute (2005) Specifications for Structural Concrete, ACI 301-05.
American Concrete Institute, Farmington Hills, MI, USA.
[92] Robinson RF and Austin CR (1951) Effect of copper-bearing concrete on mould.
Industrial Engineering Chemistry 43(9): 2077–2082.
[93] Dyer TD, Jones MR and Garvin S (2009) Exposure of Portland cement to multiple trace
metal loadings. Magazine of Concrete Research 61(1): 57–65.
[94] Levowitz LD (1952) Anti-bacterial cement gives longer lasting floors. Food Engineering
24(6): 57–60 and 134–135.
[95] Atwan DS (2014) Influence of types anti-fungal admixtures on the concrete mix. Nahrain
University, College of Engineering Journal 16(2): 180–191.
[96] Do J, Song H, So H and Soh Y (2005) Antifungal effects of cement mortars with two
types of organic antifungal agents. Cement and Concrete Research 35(2): 371–376.
[97] BASF (2016) MasterLife AMA 100 Integral Antimicrobial Admixture. BASF, Cleveland,
OH, USA.
Fungal Biodeterioration 211
[98] Bolkan SA, Stairiker C, Czechowski MH and Ventura M (2015) Stable aqueous solutions
of silane quat ammonium compounds. US 8956665 B2.
[99] Uysal A (2015) Self-cleaning and air cleaning cement. EP 2957546 A1.
[100] Fukushima Y, Shuto Y, Ohinata T and Okouchi S (2008) Antifungal effect of calcined
colemanite. Journal of Antibacterial and Antifungal Agents 36(5): 293–297.
[101] Park S-K, Kim J-HJ, Nam J-W, Phan HD and Kim J-K (2009) Development of anti-fungal
mortar and concrete using zeolite and zeocarbon microcapsules. Cement and Concrete
Composites 31(7): 447–453.
[102] Freed WW (1995) Reinforced concrete containing antimicrobial-enhanced fibers.
WO/1995/006086.
[103] Ramirez TH and De Leon GD (2004) Concrete-based floors and wall coverings with an
antimicrobial effect. US20040121140 A1.
[104] Navas J and Borralleras P (2005) Concrete with antibiotic and antifungal properties.
pp. 65–75. In: Sánchez Espiniza E and Garcimartin MA (eds.). Proceedings of the Vth
International Symposium on Concrete for a Sustainable Agriculture. San Lorenzo de
El Escorial, Spain.
[105] Warscheid T (2004) Prevention and remediation against biodeterioration of building
materials. pp. 29–36. In: Silva MR (ed.). Second International RILEM Workshop on
Microbial Impact on Building Materials. RILEM, Paris, France.
[106] Eyssautier-Chuine S, Gommeaux M, Moreau C, Thomachot-Schneider C, Fronteau G,
Pleck J and Kartheuser B (2014) Assessment of new protective treatments for porous
limestone combining water-repellency and anti-colonization properties. Quarterly
Journal of Engineering Geology and Hydrogeology 47(2): 177–187.
[107] Highways Agency (2007) Inspection Manual for Highway Structures: Vol. 1: Reference
Manual. Highways Agency, The Stationery Office, London, UK.
[108] May E, Lewis FJ, Pereira S, Tayler S, Seaward MRD and Allsopp D (1993) Microbial
deterioration of building stone—a review. Biodeterioration Abstracts 7(2): 109–123.
[109] Dyer TD (2014) Concrete Durability. CRC Press, Boca Raton, FL, USA.
[110] Historic Scotland, Biological Growth on Masonry: Identification and Understanding.
Historic Scotland, Edinburgh, UK.
[111] Brown SK (1982) Fibre release from corrugated asbestos cement cladding due to
weathering and cleaning. CSIRO Division of Building Research, Highett, Australia, 35 p.
[112] Martin AK, Johnson GC, McCarthy DF and Filson RB (1992) Attempts to control lichens
on asbestos cement roofing. International Biodeterioration and Biodegradation 30(4):
261–271.
Chapter 5
5.1 Introduction
The kingdom of plants incorporates a significant variety of organisms in
terms of both size and form. Their two common features are the use of
photosynthesis as a means of obtaining energy and the material that makes
up their cell walls: cellulose.
The interaction of plants with concrete structures in manners which can
be problematic are both chemical and physical. However, most physical
damage primarily derives from larger plants. For this reason, these divisions
of the plant kingdom are discussed below. Whilst usually not harmful to
concrete, the development of moss and similar plants can sometimes play
a role in the establishment of other plants on concrete surfaces, and the
division of plants in which mosses are placed (Bryophyta) is also covered.
5.1.1 Algae
The term algae covers a wide range of organisms, whose similarity and
distinction from other organisms is not easy to define. One definition is that
they obtain energy by photosynthesis using chlorophyll, and that they have
reproductive cells that are not protected by layers of sterile cells [1]. Algae
can be both unicellular and multicellular, with some of the multicellular
species, such as seaweeds, being able to grow to a considerable size. Both
unicellular and multicellular algal forms are most probably present in the
photograph in Figure 5.1. Many of the unicellular species are motile.
When algae are discussed in this chapter, it will only be in the context
of eukaryotic algae—organisms whose cells contain a nucleus and other
organelles contained within membranes. Cyanobacteria are prokaryotic
organisms—lacking a nucleus bound within a membrane—which are often
also included within the grouping of algae. Cyanobacteria are covered in
Chapter 3.
Plants and Biodeterioration 213
Figure 5.1 Algae in the form of an algal biofilm and seaweed at the
waterline of a concrete bridge.
Air bladder
5.1.3 Bryophyta
Bryophytes are a division of plants which differ from the higher plants
in that they lack a vascular system—they lack a xylem and phloem for
Air bladder
Air bladder
Air bladder
The means by which algae reproduce are diverse and complex. However,
they can be grouped into three general forms: vegetative, asexual and
sexual. Vegetative reproduction involves either cell division or division of
multicellular algae through various mechanisms. Asexual reproduction
employs spores, or similar entities, which grow into a plant genetically
identical to the parent plant. Sexual reproduction involves either isogamy
(where algal cells from different individuals form similar motile gametes
which fuse together) or heterogamy (where gametes with very different
size and/or characteristics are produced).
Algae
For algae to be brought into contact with a concrete surface, there must
be some movement of algae-bearing water past the surface. The rate of
flow plays an important role in the rate at which algae become attached to
this surface. A study which examined rates of attachment of a number of
species of algae delivered to different surfaces by flowing water travelling
at different rates found that slower rates favoured colonisation [2].
The same study also found that attachment rates were dependent on the
nature of the surface. However, the findings are somewhat counterintuitive,
with algae often attaching with greater efficiency to hydrophobic surfaces
with low free surface energies.
It has been proposed that the attachment of algae is the result of
extracellular polymeric substances (EPS) produced by these organisms. This
is supported by research which examined the nature of the polysaccharides
produced by a number of species of algae isolated from building façades [3].
The polymers were categorised in terms of whether they were released into
the environment surrounding the cell (released polysaccharides, RPS), or
were retained as the outer envelope of the cells (capsular polysaccharides,
CPS). The polymers were found to be anionic in nature and, in some
cases, reduced the surface tension of water. The kinetics of CPS adsorption
onto a glass surface and silicon wafers coated with a hydrophobic agent
were characterised. It was found that, whilst the rates of adsorption were
similar initially, the hydrophobic surface ultimately adsorbed more of the
polysaccharides.
In the case of the larger multicellular algae, colonisation is likely to be
initiated by a spore becoming attached to the surface.
Vascular plants
It has been proposed that higher plants may become established on concrete
surfaces through two mechanisms [4].
The first mechanism occurs on both horizontal and vertical surfaces and
involves seeds of plants becoming lodged in cracks or holes in the surface.
220 Biodeterioration of Concrete
Order Family Genus Species Common Dispersal Thousand Calcicole/ Optimum References
Name Seed Calcifuge pH
Weight, g
Apiales Araliaceae Hedera helix Ivy animal 69.30 Calcicole 6.0–7.5 [8, 12]
Brassicales Capparidaceae Capparis spinosa Caper animal 9.00 Calcicole 7.5–8.0 [8, 10, 16]
222 Biodeterioration of Concrete
Brassicaceae Erysimum cheiri Wallflower wind 1.40 Calcicole 5.5–7.5, but [8, 11, 12]
can tolerate
higher pH
Caryophyllales Polygonaceae Fallopia japonica Japanese wind/animal 2.065 – 3.0–8.5 [8, 15, 18]
Knotweed
Celastrales Celastraceae Euonymus europaeus Spindle- animal 40.60 Calcicole 6.0–8.0 [8]
tree
Dipsacales Caprifoliaceae Sambucus nigra Elderberry animal 26.00 Calcicole 5.5–6.5 [8, 12]
Fabales Leguminosae Robinia pseudoacacia False Gravity/root 18.00–20.00 Calcicole 4.6–8.2 [8, 13, 14]
acacia/ suckers
Black
locust
Lamiales Oleaceae Syringa vulgaris Lilac 5.60 Calcicole 6.0–7.5 [8, 12]
Ranunculales Ranunculaceae Clematis vitalba Old man’s wind 2.10 Calcicole 5.5–8.0 [8, 15]
beard
Rosales Ulmaceae Celtis australis Nettle-tree animal 203.70 Calcicole 6.0–7.8 [8, 13]
Sapindales Simaroubaceae Ailanthus altissima Tree of wind 29.40 Calcicole 5.5–8.5 [8, 13]
Heaven
Order Family Genus Species Common Name Dispersal Thousand Seed Weight, g
5.3.2 pH
0.3
0.2
• •
:!2
ui
I- 0.1
~
I
I- 0.0
s
0
~
0 -0 .1
LLJ
>
~ -0 .2
5
LLJ
~
•
-0 .3
•
-0.4 2;---------:4;----~6~----:--8-----.----
10 12
pH
Figure 5.6 Population growth rate of a species of Cryptomonas versus pH [19].
226 Biodeterioration of Concrete
100
~
0
I
~
::R
0 80 1 I
.•
I
u.i I I
u ell
<(
•
LL rp
''
0::: 60 I
:J I
(/) I
LL I
0 (])
I
UJ I
(!)
<(
0:::
40
(]!
t (
UJ I I
> I
I
(/)
0 20 I
u rp
I
I •
0
Uncarbonated
Carbonated
0
0 20 40 60 -- 80 100 120
TIME, days
Figure 5.7 Coverage by Klebsormidium flaccidum versus time for carbonated and
uncarbonated mortar surfaces [21].
140
E
::t
<{ 120
~
w /----"
z
0 100 ' I ~
I- I \
0 I \
a: I
I \
\
11. 80
>-
a: I
I \
\
<:<: I \
~ 60 I ~\
0::
11. I '
•
.--
LL.
40 II ' '-
0 II '
J:
I- __../ .
(.9
20
z
w
....1
0
2 4 6 8
pH
22
20
18 I
•
0
Shoots
Roots
I
/
/
........ -·--- ' '
~"
16
/ /
/
'
0)
~---
14 /
/ '•
J: /
(.9 12 /
Lij /
s 10 /
/
>-
•
/
a: 8
0
6
_ _ _ _ _ _ Q_ _ _ _ _ _
2 0' __..-
----- -o-- -o
0+-------.------.-------,------,-------,-----~
4.0 4.5 5.0 5.5 6.0 6.5 7.0
pH
Figure 5.9 Dry mass of the shoots and roots versus pH for Buddleia davidii plants
60 days after being transplanted to a pine bark medium. pH differences were
achieved through amendments of dolomitic lime [24].
228 Biodeterioration of Concrete
5.3.3 Nutrients
had the effect of solubilising phosphate, whilst citric acid solubilised iron
and manganese from the soils used.
Heterotrophic algae have also been found to produce organic acids.
These include formic, pyruvic, lactic and succinic acids. One study
examining acid formation by three types of green algae (Chlorella vulgaris,
Scenedesmus basilensis and a species of Chlamydomonas) [32]. Under aerobic
conditions with glucose as the source of energy, pyruvic acid was produced
in the largest quantities by Chlorella and Scenedesmus. Chlamydomonas
produced formic acid in larger concentrations under aerobic conditions,
albeit in much smaller concentrations overall. Under anaerobic conditions,
all three algal cultures produced lactic acid. These acids are most likely
formed as products of metabolising glucose, rather than as a means of
adjusting the surrounding conditions, as in the case of plant roots. However,
it should be noted that even under autotrophic conditions (i.e., in the absence
of glucose) the acids were all produced by the mixotrophic Chlorella vulgaris,
although in much smaller concentrations.
One possible example of exudation of organic acids by algae being
used functionally is in the case of cryptoendolithic algae which appear to
live in layers just beneath rock surfaces as a means of limiting the harmful
effects of extreme environmental conditions such as high or low ambient
temperatures and low humidity [33]. It has been proposed that these
recesses—and possible subsequent biodegradation of the stone—have been
formed through the exudation of acids [34]. This strategy demands that the
rock is sufficiently optically clear to permit sunlight to pass through the
surface to allow the algae to photosynthesize, with sandstones proving ideal
[33]. Whether typical concrete formulations contain suitable aggregates to
permit this is uncertain.
One nutrient required by a specific group of algae is silicon. The
group in question is the diatoms, which are unicellular algae found in
both freshwater and marine environments. The silicon is used by diatoms
to form a porous, symmetrical cell wall made of hydrated silica known as
a frustule. There is obviously a quantity of silicon present in cement and
aggregate. Generally the Si in the cement matrix will be present as CSH
gel, and will be more soluble than that present in the aggregate. However,
the solubility of Si in CSH is still low. Whilst, as will be seen in Section 5.4,
diatoms may be found in close proximity to concrete structures, diatoms do
not attach to surfaces, but instead remain suspended in water. Moreover,
there is no evidence that diatoms produce substances to render Si more
available. Thus, concrete does not provide a habitat for diatoms, and is
unlikely to be affected by their proximity.
It is generally believed that the rhizoids of moss solely perform the
function of anchors. Rhizoids appear to be capable of penetrating rock,
mortar and concrete, but seemingly only if the porosity is sufficient to allow
Plants and Biodeterioration 231
Algae
Vascular plants
5.4.1 Algae
Table 5.3 lists the species identified by studies examining algal growth
on concrete and mortar surfaces either in the field or under laboratory
conditions.
5.4.2 Bryophytes
Table 5.3 Algae identified growing on concrete surfaces, or successfully grown in the laboratory.
pressures is relatively small (between 0.7 and 2.5 MPa) [52]. However,
given that a root growing into a concrete crack will be exerting pressure
in a manner which will be resolved as a tensile stress in the concrete itself,
there still exists the potential for damage.
The tensile strength of concrete is notably low. The following equation
has been proposed for the estimation of tensile strength of concrete from
compressive cylinder strength [52]:
ft = 0.3fc 32
where ft and fc are the tensile strength and compressive cylinder strength
respectively. Thus, concrete with a cylinder strength of 30 MPa will have a
tensile strength of around 2.9 MPa. Furthermore, given that the tips of cracks
will also act as sources of stress concentration, the widening of fissures in
concrete is less of a challenge for plants than it might immediately seem.
The production of root exudates by plants is likely to exacerbate the
magnitude of damage that can be done by a plant, particularly where
the substances produced include compounds such as citric acid, whose
236 Biodeterioration of Concrete
aluminium ions in Chapter 2 indicates that their effect will largely be that
of acidolysis.
A special case with regards to interaction of vascular plants with
concrete structures is that of climbing plants (Figure 5.10). There has been a
great deal of debate with regards to whether such plants are harmful to the
walls on which they grow, with particular attention to the ivy, Hedera helix.
One study of ivy growing on a specially constructed stone test wall found
that the plant was not particularly predisposed to enter holes and cracks
[54]. However, observation of ivy on historic structures has identified many
instances where the roots enter existing voids in walls leading to damage
[54, 55]. It has been proposed that this process is promoted by shortages of
water, and that events such as the cutting off of the stem of the plant at the
base may cause smaller rootlets to permeate such voids more aggressively
to obtain water [54].
The types of voids which are likely to be amenable to plants such as
ivy are likely to be far scarcer on a concrete structure, compared to a stone
wall, but might include joints, or cracks resulting from other forms of
deterioration. Where the incursion of roots into the fabric of a structure is
unlikely, there is some evidence that ivy provides protection to buildings,
principally in its ability to capture air-borne particulates before they become
attached to the building surface. It is argued that this has the effect of
keeping the surfaces themselves clean, which may also have the additional
benefit of limiting the delivery of nutrients to the surface which might
otherwise be used by micro-organisms in biodeterioration processes [56].
pipe will leak, with environmental and efficiency implications. Secondly, the
pipe will gradually become blocked by the invading root (see Figure 5.11).
A number of studies have been conducted into this phenomenon, with
many taking the form of surveys of water networks in specific geographical
locations. Such studies have been made possible through the use of mobile
cameras that can be run through pipes to inspect their interior. One such
study, conducted in Sweden, examined both the frequency of intrusion
and related each instance of intrusion to tree and shrub species growing in
close proximity to the pipes [62]. The species found to be most frequently
responsible for intrusions were Malus floribunda (Japanese flowering
crabapple) and Populus candensis (Canadian poplar). The two main types of
pipe encountered in the survey were concrete and PVC, and it is interesting
to note that the concrete pipes were considerably more resistant to intrusion
than the polymer pipes. The survey detected 0.661 root intrusions per pipe
joint in the case of PVC, whilst the frequency was only 0.080 in the case
of concrete. This may reflect the greater mass of the concrete pipes which
would be likely to present greater resistance to the growth pressures of roots.
A similar study conducted in Poland made a similar observation,
although in this case the two types of pipe encountered were concrete and
vitrified clay [61]. The frequency of intrusion was expressed in a slightly
different form (intrusions per 100 m). Concrete pipes proved more resistant,
with an intrusion rate of 3.24 intrusions per 100 m, compared to a value of
3.99 for vitrified clay pipes. Smaller diameter pipes are also more vulnerable
to intrusion [63].
day experimental period. Evaluation of cell numbers showed that the algae
were completely eradicated by 7 days, presumably as a result of the high pH
of the culture medium (around 10) originating from the Portland cement.
Where the geopolymer concrete was exposed to the culture medium in the
absence of algae, the corrosion potential rapidly dropped to a more negative
value, indicating a greater potential for corrosion. This was presumably
either due to the lower pH of the cement limiting passivation of the steel
surface or the higher porosity of the cement matrix, or a combination of
both. However, when algae were present the corrosion potential remained
at a level comparable to the Portland cement concrete until the end of the
14-day experiment, at which point there was an abrupt drop to a more
negative value. This drop approximately coincided with the extinction
of the algae population, leading the researchers to propose that the algae
might, in fact, be protecting the steel by maintaining a higher pH through
oxygen production.
Regardless of the precise nature of the processes occurring in these
experiments, there would appear to be no reported examples of such a
process occurring in the field. The main argument against such a process
is that, for it to be effective, there would need to be considerable diffusion
of oxygen into the concrete towards the steel surface. This is because algal
activity will be limited to a zone very close to the concrete surface, due to
the need for light. Because of the uncertainties surrounding this proposed
form of biodeterioration, it is not explored further in this chapter.
Nonetheless the presence of algae on concrete surfaces is often
prominently visible and the resulting stain can be detrimental in aesthetic
terms, particularly on lighter surfaces. Moreover, the presence of algae on
horizontal surfaces on which people walk can be slippery and consequently
hazardous.
Environmental factors
Aside from the nutrient needs of algae, the organisms require water and
sunlight. Thus, where either of these are absent, algae will be able to live.
Thus, surfaces which are permanently in darkness cannot be occupied.
Figure 5.12 shows the generation time (the time required for the population
to double) of a species of algae. At low levels of light intensity, very long
generation times are observed. However, as light intensity increases—for
this particular case—the generation time increases again beyond the
optimum intensity of around 60 W/m2. The intensity of sunlight incident
on the Earth’s surface will depend on the latitude and time of year (with
greater seasonal variation for latitudes further away from the equator).
However, intensities will typically be considerably higher than 60 W/
m2 at noon on a clear day (for instance around 500 W/m2 in spring and
autumn in the South of England). However, this is unlikely to compromise
algal growth too much, and many surfaces will seldom experience the full
intensity of sunlight where surfaces are horizontal, or where there is partial
shading and cloud cover.
Parts of a concrete structure which experience low levels of moisture
are unlikely to support algal growth. The environment surrounding a
structure will also define the availability of moisture. A survey involving
statistical analysis of algal colonisation of building façades across France
found that relative humidity was the most important factor in influencing
algal growth, with higher quantities of precipitation and closer proximity
of vegetation also encouraging growth [74].
60
50
•
I
VI
>-
"'
"0
40
ui
:2;
i=
z 30
0
i=
<(
.
a::
w 20
zw
___
--·-
--------
C)
I
10
~
··-------------
0+-----~----~----~----r-----r-----r---~
0 50 100 150 200 250 300 350
Thus, interiors and exterior surfaces which receive shelter from rainfall
from other parts of a structure (such as cantilevered projections, awnings and
roof eaves) will potentially remain sufficiently dry to resist the establishment
of algae. Indeed, in designing a structure, there may exist possibilities for
incorporating features which will limit the wetting of concrete surfaces,
including the use of coping and capping, or projecting sills.
Algae were found to favour the North- and West-facing facades of
buildings in France [74]. This was attributed to the North side of buildings
receiving less sun, and consequently drying out less frequently. Favourable
colonisation of West-facing surfaces was considered the result of dominant
westerly winds bringing rain from the Atlantic Ocean. Such patterns
will clearly vary according to the geographic location of a building, and
knowledge of local conditions will at least allow identification of which
parts of a structure require measures to control growth.
Temperature ranges over which algae will grow vary between species.
However, for most species optimum growth rates occur somewhere
between 20ºC and 35ºC, although some algae continue to grow at higher
temperatures. This includes Chlorella vulgaris, which can grow on concrete
surfaces (Table 5.3) and which is capable of growth at temperatures as high
as 40ºC [75].
These ranges are clearly widely experienced as ambient temperatures,
meaning that the vast majority of concrete structures worldwide will
offer suitable environments for algae, at least at certain parts of the year.
Moreover, the control of external temperatures is not a practical option.
The amplitude of temperatures throughout the year has also been found
to play an important role, with smaller magnitudes of fluctuation limiting
evaporation rates and maintaining levels of moisture [74].
Material characteristics
100
80
GROWTH RATIO, %
60
40
20
0
l
se
T
SH
se
tro
in
a
a
on
D
at
at
ot
D
An
an
C
Bi
s
io
ed
An
op
-d
Fe
Figure 5.13 Growth rate of algae and cyanobacteria on mortar surfaces,
Figure 5.13 Growth rate of algae and cyanobacteria on mortar surfaces, expressed as the quantity
expressed as the quantity of chlorophyll a extracted as a percentage of the
of chlorophyll a extracted as a percentage of the control [42].
control [43].
Protection after construction
A study examining the effectiveness of different façade coatings measured their ability to support
growth of three different species of algae – Klebsormidium flaccidum, Chlorella mirabilis and
Stichococcus bacillaris [41]. The coatings investigated were three mortars, an acrylic water‐based
organic coating and a similarly formulated paint. Neither the paint or the organic coating contained a
biocide. The paint was considerably better at limiting growth in comparison to the other materials
(Figure 5.14). Indeed, a study characterising and quantifying algal colonisation of a large number of
building façades in France found that concrete and mortar surfaces were more amenable to the
establishment of algal populations than organic surfaces such as paint [72], since the isolation of
algae from the porosity of the concrete limits the availability of water.
Plants and Biodeterioration 251
80
//,....,...........
,'
I
"" /
/
/
.,..,.
,-0
'2ft
? I
I I
/
,-
LLi
l
/
I ;:Y
<!J I I
I /
<( I I /
a::: 60 I I I /
w I I I
> I I
-·-
I I I
0 I I I 0
I
(.)
I I I
L.U I I
(.) 40 I I Paint
';/ I I
<( / - Q - Organic coating
LL I I I
-·-
I / One-coat mortar A
a::: /
I I - ~-
..
/
I/ / 0
/
...
I/ ,~_,,
_,., ~;;::.-
1/
0
0 10 20 30 40
TIME, days
100
I
I - -· ~
I y-'V
•
I I I
I
80 I
I I
I ;f
I I
I I
I I
""u.i .o--0 I
•
0 I I
I fr ~o-/ I
60 I 'j1
I
(!)
<(
I
I
~j1 I
I
0:: I
LJ..I I /I -;:;
> I p I
-·-
0 40 I / I
u ;:r:-0' -;:;
I
f!1/"'-i,' '\. ~ I
vi ">~
I
I
I
_...,_ Water
- {)- repellent
Untreated
l ;f
,--,1/
20 Biocide
I / ; ,- \
\7
----
I Biocide then WR
~
- 'q -
II '--7
P1 I WR then biocide
TIME, weeks
refined through the use of both of these oxides coated with co-catalysts of
either platinum (Pt) or iridium (Ir) which proved to be highly effective at
limiting growth. It was demonstrated that the presence of these elements
was effective through a photocatalytic mechanism rather than one of
toxicity through limiting the light source used to wavelengths that did not
contribute significantly to the photocatalytic effect, which yielded growth at
a greater rate. Whether such an approach is practical for concrete structures
is questionable: both metals are extremely expensive.
The conclusion that must be drawn from these findings is that anatase
and the application of paint appear to offer the best means of achieving
concrete surfaces which limit algal growth. However, the use of water
repellent coatings, should not be entirely ruled out. In all cases, surface
coatings must be viewed as being non-permanent, and requiring re-
application later in a structure’s life.
Cleaning
The change in colour of the surface during this treatment was measured
using colorimetry. The treatment was found to yield a change in colour that
indicated a cleaning effect. This was not as effective as submerging the cubes
in the solution, or, indeed, submerging the cubes in a solution of sulphuric
acid. Petrographic microscopy of the surface indicated a layer of gypsum
has been formed (see Chapter 3). It is sound to assume that this layer would
eventually be weathered leaving a wholly clean surface behind. However,
whilst such a process would appear to be effective, the use of sulphuric
acid—regardless of its source—as a means of cleaning will clearly lead to
a loss of mass from the surface and might be deemed overly aggressive.
to damage the underlying concrete, cracks or joints will still probably need
to be present.
Removal of moss is most effectively achieved through physical methods
such as scraping or water jetting—removal is the important part, and
application of herbicide will not make this process any easier.
The design of concrete mixes to resist damage from plant damage
can be done with less certainty compared to the organisms examined in
previous chapters. A reduced water/cement ratio is certainly unlikely to
do any harm, could potential reduce the availability of water, and would
have the side-effect of improving concrete durability in more general terms.
Furthermore, assuming that the worst-case scenario for acid exudation from
roots is citric acid, the results reported in Chapter 4 would point to either
Portland or calcium aluminate cement.
Where vascular plants have already become established on a structure,
their removal is necessary, followed by filling of cracks with appropriate
repair materials, and joints with sealants, to prevent re-establishment. In
some cases the application of herbicides may be beneficial.
Herbicides can be categorised in terms of the manner in which they
work and can be selected accordingly. Hormone type herbicides mimic the
growth hormones of plants. They selectively target broad-leafed plants, but
do not harm narrow-leafed plants, such as grasses.
Moss and algae herbicides are particularly suited for acting on these plants.
They include acetic acid and ferrous sulphate, both of which are not
suited for use on concrete. Ferrous sulphate will stain concrete, whilst the
potentially damaging effect of acetic acid on concrete is covered in Chapter 3.
However, acetic acid products marketed as both algae/moss herbicides and
cleaning agents are sometimes marketed specifically for cleaning concrete,
on the grounds that the acid will remove the outer surface of the paving as
part of the cleaning process.
Contact herbicides kill only the part of the plant that they come into contact
with, whilst systemic herbicides will move through the plant to affect
all parts. Residual herbicides take some time either to be dispersed or
break down, thus leaving soil uninhabitable by plants for some time after
application.
Table 5.5 lists herbicides which are likely to potentially be usable on
concrete. It should be noted that most commercial products are combinations
of more than one herbicide, to tailor the product to specific applications.
Climbing plants
The extent to which roots are able to grow under pavements can be limited
by the use of root barriers. Root barriers are layers of material—running
vertically between the tree and the pavement to be protected—whose
purpose is to deflect or constrict the growth of roots [58]. Materials able to
deflect growth include plastic and glass fibre composite panels, preserved
plywood and tar paper or asphalt impregnated felt. Constricting barriers
are normally geotextiles, but can also include copper mesh. This material
possibly has an enhanced control over roots, in that copper released into
the soil is likely to have an inhibitory effect on growth. Indeed, geotextiles
impregnated with copper sulphate have also been successfully used as a
root barrier.
It has been argued that the use of root barriers compromises the stability
of trees, but a study simulating the effect of wind on trees grown with and
without barriers found that the presence of barriers actually appeared to
increase resistance to being uprooted [84]. This was attributed to the deeper
roots of these trees.
Another possible approach to limiting upward growth of roots beneath
pavements is to allow moisture in soil to evaporate, leading to lower
moisture levels at the pavement–soil interface. This can be achieved through
permeable paving. Where the paving is concrete, this can take the form
either of paviers with gaps which allow the movement of water vapour—
but still support foot traffic—or pervious concrete. Pervious concrete is
concrete containing coarse aggregate, but little or no fine aggregate [85].
This produces a material with a volume of large pores, which allow the
movement of liquid water and oxygen downwards, and water vapour
upwards. For permeable pavements to be successful, a soil which permits
percolation of water is essential, since an impermeable soil will still lead
to high moisture levels close to the surface [86].
The success of permeable pavements, however, is reported as being
somewhat inconsistent, with levels of moisture and oxygen beneath test
pavements showing little correlation between permeable and impermeable
surfaces [87].
When compacted subgrade is located below pavements, there exists
the possibility of including ‘root paths’ [86]. These are trenches in the sub-
grade lined with strip drain material leading to the other side of the paved
area. The trenches are backfilled with good quality soil before the rest of the
pavement construction is completed. The theory behind this approach is that
it offers a low-resistance path for the root to grow beneath the pavement into
an area where growth can continue without disruption of infrastructure.
Such systems are also sometimes referred to as ‘root breakout zones’ [58].
There is currently an absence of data with regards to the effectiveness of
this approach.
Plants and Biodeterioration 259
Pipes
The type of seal used at pipe joints will define the ease with which root
intrusion can occur. The previous surveys of root intrusion into pipes have,
in part, examined pipes with more traditional seals, such as textile strips
impregnated with bitumen and then sealed in place with cement. This
sealing technique has been superseded, firstly with natural rubber rings and
secondly with synthetic rubber gaskets designed to making fitting easier. A
study examining the more modern seal found that despite its composition
260 Biodeterioration of Concrete
and design, roots had little problem in breaching it [63]. Improvement was
observed, however, when the outside of the joints was further sealed with
self-vulcanising tape.
The recommendations which have arisen from this study are that care
should be taken in planting trees near pipelines, with consideration of
distances and tree species. Tree species should be those having a low ‘root
energy’—i.e., being less intrusive in nature. Common lime (Tilia europaea)
is suggested as an example by the researchers. It was also proposed that
the plant bed (the volume of soil which is excavated and replaced when a
tree is planted) should be made larger than normal. The reason why such
an approach is likely to be effective is that this volume of soil tends to be
more porous. Root growth will typically initially be limited to this zone
prior to infiltrating the denser surrounding soil, thus extending the period of
time before roots reach the pipe. Finally, the use of root barriers geotextiles
was proposed as a means of slowing the progress of roots, although it was
stressed that such barriers will typically hinder root growth rather than
stop it.
Where root intrusion has occurred, specialised cutting equipment can be
deployed in pipes to remove blockages. Root intrusions can also be removed
through the application of formulations containing the contact herbicide
metam sodium and a growth inhibitor [86]. By using a contact herbicide, the
intruding roots can be destroyed without killing the entire plant. However,
since it requires releasing quantities of herbicide into wastewater, in many
countries this approach is not permitted.
5.7 References
[1] Lee RE (2008) Phycology, 4th Ed., Cambridge University Press, Cambridge.
[2] Barberousse H, Brayner R, Botelho Do Rego AM, Castaing J-C, Beurdeley-Saudou P and
Colombet J-F (2007) Adhesion of façade coating colonisers, as mediated by physico-
chemical properties. Biofouling 23(1): 15–24.
[3] Barberousse H, Ruiz G, Gloaguen V, Lombardo RJ, Djediat C, Mascarell G and
Castaing J-C (2006) Capsular polysaccharides secreted by building façade colonisers:
characterisation and adsorption to surfaces. Biofouling 22(6): 361–370.
[4] Lisci M, Monte M and Pacini E (2003) Lichens and higher plants on stone: a review.
International Biodeterioration and Biodegradation 51(1): 1–17.
[5] English Heritage (2014) Landscape Advice Note: Vegetation on Walls. English Heritage,
London.
[6] Segal S (1969) Ecological Notes on Wall Vegetation. Junk, The Hague.
[7] Kumar R and Kumar AV (1999) Biodeterioration of Stone in Tropical Environments—
An Overview. The Getty Conservation Institute, Los Angeles, CA, USA.
[8] Kew Royal Botanic Gardens (2008) Seed Information Database version 7.1 Kew Royal
Botanic Gardens, London, UK. http://data.kew.org/sid/
[9] Çalişkan O, Mavi K and Polat A (2012) Influences of presowing treatments on the
germination and emergence of fig seeds (Ficus carica L.). Acta Scientiarum Agronomy
34(3): 293–297.
[10] Janick J and Paull RE (2008) The Encyclopedia of Fruit and Nuts. Centre for Agriculture
and Biosciences International, Wallingford, UK.
Plants and Biodeterioration 261
[11] Hintze C, Heydel F, Hoppe C, Cunze S, König A and Tackenberg O (2013) D³: The
dispersal and diaspore database—baseline data and statistics on seed dispersal.
Perspectives in Plant Ecology, Evolution and Systems 15(3): 180–192. www.seed-
dispersal.info accessed November 2016.
[12] Algoplus (2016) Plant pH Preference List, Algoplus, Jacksonville, FL, USA. www.
algoplus.net/soilpHinfo.html accessed November 2016.
[13] Ajuntament de Barcelona (2016) Trees and Palm Trees of Barcelona, Ajuntament de
Barcelona, Barcelona, Spain. w110.bcn.cat/portal/site/MediAmbient/menuitem.37
ea1e76b6660e13e9c5e9c5a2ef8a0c/index1536.html?vgnextoid=5834c50d8a35b210Vgn
VCM10000074fea8c0RCRD&vgnextchannel=5834c50d8a35b210VgnVCM10000074fea
8c0RCRD&vgnextrefresh=1&lang=en_GB accessed November 2016.
[14] Forest Ecology and Forest Management Group (2016) Tree factsheet—Robinia
pseudoacacia L. Forest Ecology and Forest Management Group, Wageningen University,
Wageningen, Netherlands.
[15] Centre for Agriculture and Biosciences International (2016) Invasive Species
Compendium. www.cabi.org/isc/datasheet/14280 accessed November 2016.
[16] Trewartha J and Trewartha S (2005) Producing Capers in Australia—Viability Study.
Australian Government Rural Industries Research and Development Corporation,
Canberra, Australia.
[17] Wacquant JP and Picard JB (1992) Nutritional differentiation among populations of the
Mediterranean shrub Dittrichia viscosa (Asteraceae) in siliceous and calcareous habitats.
Oecologia 92(1): 14–22.
[18] Beerling DJ (1991) The testing of cellular concrete revetment blocks resistant to growths
of Reynoutria japonica houtt (Japanese knotweed). Water Research 25(4): 495-498.
[19] Weisse T and Stadler P (2006) Effect of pH on growth, cell volume, and production
of freshwater ciliates, and implications for their distribution. Limnology and
Oceanography 51(4): 1708–1715.
[20] Bell TAS, Prithiviraj B, Wahlen BD, Fields MW and Peyton BM (2015) A lipid-
accumulating alga maintains growth in outdoor, alkaliphilic raceway pond with mixed
microbial communities. Frontiers in Microbiology 6: 1480–1491.
[21] Tran TH, Govi A, Guyonnet R, Grosseau P, Lors C, Garcia-Diaz E, Damidot D, Devèse O
and Ruote B (2012) Influence of the intrinsic characteristics of mortars on biofouling by
Klebsormidium flaccidum. International Biodeterioration and Biodegradation 70: 31–39.
[22] Štechová T, Hájek M, Hájková P and Navrátilová J (2008) Comparison of habitat
requirements of the mosses Hamatocaulis vernicosus, Scorpidium cossonii and Warnstorfia
exannulata in different parts of temperate Europe. Preslia 80(4): 399–410.
[23] Ogbimi AZ, Owoeye YB, Ibeyemi VO and Bofede AV (2014) Effects of pH, photoperiod
and nutrient on germination and growth of Calymperes erosum C. Muell. Gemmaling.
Journal of Botany 2014: 1–5.
[24] Gillman JH, Dirr MA and Braman SK (1998) Effects of dolomitic lime on growth
and nutrient uptake of Buddleia davidii ‘Royal Red’ grown in pine bark. Journal of
Environmental Horticulture 16(2): 111–113.
[25] Hewitt EJ and Smith TA (1975) Plant Mineral Nutrition. English Universities Press,
London, UK.
[26] Barberousse H, Tell G, Yéprémian C and Couté A (2006) Diversity of algae and
cyanobacteria growing on building façades in France. Algological Studies 120: 81–105.
[27] White PJ and Broadley MR (2003) Calcium in plants. Annals of Botany 92(4): 487–511.
[28] Rout G, Samantaray S and Das P (2001) Aluminium toxicity in plants: a review.
Agronomie 21(1): 3–21.
[29] Taylor GJ (1991) Current views of the aluminum stress response: the physiological basis
of tolerance. Current Topics in Plant Biochemistry and Physiology 10: 57–93.
[30] Ryan PR, Delhaize E and Jones DL (2001) Function and mechanism of organic anion
exudation from plant roots. Annual Review of Plant Physiology and Plant Molecular
Biology 52: 527–560.
262 Biodeterioration of Concrete
[31] Ström L (1997) Root exudation of organic acids: importance to nutrient availability and
the calcifuge and calcicole behaviour of plants. Oikos 80(3): 459–466.
[32] Kobayashi T, Inamori Y, Tanabe I and Obayashi A (1974) Organic acid production by
unicellular green algae during the heterotrophic growth. Nippon Nōgeikagaku Kaishi
48(4): 261–267.
[33] Friedmann EI (1980) Endolithic microbial life in hot and cold deserts. Origins of Life
and Evolution of Biospheres 10(3): 223–235.
[34] Viles H (1995) Ecological perspectives on rock surface weathering: Towards a conceptual
model. Geomorphology 13(1): 21–35.
[35] Garcia-Rowe J and Saiz-Jimenez C (1991) Lichens and bryophytes as agents of
deterioration of building materials in Spanish cathedrals. International Biodeterioration
28(1-4): 151–163.
[36] Klavina L, Springe G, Nikolajeva V, Martsinkevich I, Nakurte I, Dzabijeva D and
Steinberga I (2015) Chemical composition analysis, antimicrobial activity and
cytotoxicity screening of moss extracts (moss phytochemistry). Molecules 20(9):
17221–17243.
[37] Bates JW (1992) Mineral nutrient acquisition and retention by bryophytes. Journal of
Bryology 17(2): 223–240.
[38] Jayakumar S and Saravanene R (2010) Biodeterioration of coastal concrete structures
by marine green algae. International Journal of Civil Engineering 8(4): 352–361.
[39] Hughes P, Fairhurst D, Sherrington I, Renevier N, Morton LHG, Robery PC and
Cunningham L (2013) Microscopic study into biodeterioration of marine concrete.
International Biodeterioration and Biodegradation 79: 14–19.
[40] Dubosc A, Escadeillas G and Blanc PJ (2001) Characterization of biological stains on
external concrete walls and influence of concrete as underlying material. Cement and
Concrete Research 31(11): 1613–1617.
[41] De Muynck W, Ramirez AM, De Belie N and Verstaete W (2009) Evaluation of strategies
to prevent algal fouling on white architectural and cellular concrete. International
Biodeterioration and Biodegradation 63(6): 679–689.
[42] Barberousse H, Ruot B, Yéprémian C and Boulon G (2007) An assessment of façade
coatings against colonisation by aerial algae and cyanobacteria. Building and
Environment 42(7): 2555–2561.
[43] Fonseca AJ, Pina F, Macedo MF, Leal N, Romanowska-Deskins A, Laiz L, Gómez-Bolea
A and Saiz-Jimenez C (2010) Anatase as an alternative application for preventing
biodeterioration of mortars: Evaluation and comparison with other biocides.
International Biodeterioration and Biodegradation 64(5): 388–396.
[44] Crispim CA, Gaylarde PM and Gaylarde CC (2003) Algal and cyanobacterial biofilms
on calcareous historic buildings. Current Microbiology 46(2): 79–82.
[45] Rosato VG (2008) Pathologies and biological growths on concrete dams in tropical and
arid environments in Argentina. Materials and Structures 41(7): 1327–1331.
[46] Andersson MH, Berggren M, Wilhelmsson D and Öhman MC (2009) Epibenthic
colonization of concrete and steel pilings in a cold-temperate embayment: a Weld
experiment. Helgoland Marine Research 63(3): 249–260.
[47] Olivia M, Moheimani N, Javaherdashti R, Nikraz HR and Borowitzka MA (2013)
The influence of micro algae on corrosion of steel in fly ash geopolymer concrete: a
preliminary study. Advanced Materials Research 626: 861–866.
[48] Rindi F, Guiry MD, Critchley AT and Ar Gall E (2003) The distribution of some species
of Trentepohliaceae (Trentepohliales, Chlorophyta) in France. Cryptogamie Algologie
24(2): 133–144.
[49] Jukonienė I (2008) The impact of anthropogenic habitats on rare bryophyte species in
Lithuania. Folia Cryptogamica Estonica 44: 55–62.
[50] Ignatov MS, Ignatova EA, Akatova TV and Konstantinova NA (2002) Bryophytes of
the Khosta’ Taxus and Buxus forest (Western Caucasus, Russia). Arctoa 11(1): 205–214.
Plants and Biodeterioration 263
[51] Sakovich A and Rykovsky G (2011) Comparative analysis of the bryophyte floras of
Northwest Belarus concrete fortification and Carpathians. Biodiversity Research and
Conservation 24(1): 23–27.
[52] Gregory PJ (1988) Growth and functioning of plant roots. In: Wild A (ed.). Russell’s Soil
Conditions and Plant Growth. Longmans, London, UK.
[53] Raphael JM (1984) Tensile strength of concrete. Journal of the American Concrete
Institute 81(2): 158–165.
[54] Viles H, Sternberg T and Athersides A (2011) Is ivy good or bad for historic walls?
Journal of Architectural Conservation 17(2): 25–41.
[55] Elinç ZK, Korkut T and Kaya LG (2013) Hedera helix L. and damages in Tlos ancient
city. International Journal of Development and Sustainability 2(1): 333–346.
[56] Sternberg T, Viles H, Cathersides A and Edwards M (2010) Dust particulate absorption
by ivy (Hedera helix L.) on historic walls in urban environments. Science of the Total
Environment 409(1): 162–168.
[57] Lewin SZ and Charola AE (1981) Plant life on stone surfaces and its relation to stone
conservation. Scanning Electron Microscopy 1981: 563–568.
[58] Roberts J, Jackson N and Smith M (2013) Tree Roots in the Built Environment.
Arboricultural Association, Stonehouse, UK.
[59] Building Research Establishment (1999) Digest 298: Low-Rise Building foundations:
the Influence of Trees in Clay Soils. Building Research Establishment, Watford, UK.
[60] National House Building Council (2016) NHBC Standards 2017. National House
Building Council, Milton Keynes, UK.
[61] Kuliczkowska E and Parka A (2017) Management of risk of environmental failure caused
by tree and shrub root intrusion into sewers. Urban Forestry and Urban Greening 21:
1–10.
[62] Östberg J, Martinsson M, Stål Ö and Fransson A-M (2012) Risk of root intrusion by tree
and shrub species into sewer pipes in Swedish urban areas. Urban Forestry and Urban
Greening 11(1): 65–71.
[63] Ridgers D, Rolf K and Stål Ö (2006) Management and planning solutions to lack of
resistance to root penetration by modern PVC and concrete sewer pipes. Arboricultural
Journal 29(4): 269–290.
[64] Pohls O, Bailey NG and May PB (2004) Study of root invasion of sewer pipes and
potential ameliorative techniques. Acta Horticulturae 643: 113–121.
[65] Garcia-Rowe J and Saiz-Jimenez C (1991) Lichens and bryophytes as agents of
deterioration of building materials in Spanish Cathedrals. International Biodeterioration
28(1-4): 151–163.
[66] Javaherdashti R, Nikraz H, Borowitzka M, Moheimani N and Olivia M (2009) On the
impact of algae on accelerating the biodeterioration/biocorrosion of reinforced concrete:
a mechanistic review. European Journal of Scientific Research 36(3): 394–406.
[67] Jayakumar S and Saravanane R (2010) Detrimental effects on coastal concrete by Ulva
fasciata. Construction Materials 163(4): 239–246.
[68] Coombes MA, Naylor LA, Viles HA and Thompson RC (2013) Bioprotection and
Disturbance: seaweed, microclimatic stability and conditions for mechanical weathering
in the intertidal zone. Geomorphology 202: 4–14.
[69] Hughes P, Fairhurst D, Sherrington I, Renevier N, Morton LHG, Robery PC and
Cunningham L (2013) Microscopic examination of a new mechanism for accelerated
degradation of synthetic fibre reinforced marine concrete. Construction and Building
Materials 41: 498–504.
[70] Hughes P, Fairhurst D, Sherrington I, Renevier N, Morton LHG, Robery PC and
Cunningham L (2014) Microbial degradation of synthetic fibre-reinforced marine
concrete. 86(A): 2–5.
[71] Suseela M and Toppo K (2007) Algal biofilms on polythene and its possible degradation.
Current Science 92(3): 285–287.
264 Biodeterioration of Concrete
[72] Lane RA (2005) Under the microscope: understanding, detecting, and preventing
microbiologically influenced corrosion. Journal of Failure Analysis and Prevention
5(10-12): 33–38.
[73] Qin J (2005) Bio-Hydrocarbons from Algae: Impacts of temperature, light and salinity
on algae growth. Australian Government Rural Industries Research and Development
Corporation, Barton, Australia.
[74] Barberousse H, Lombardo RJ, Tell G and Couté A (2006) Factors involved in the
colonisation of building façades by algae and cyanobacteria in France. Biofouling 22(2):
69–77.
[75] Singh SP and Singh P (2015) Effect of temperature and light on the growth of algae
species: A review. Renewable and Sustainable Energy Reviews 50: 431–444.
[76] Navío JA, Macias M, Garcia-Gómez M and Pradera MA (2008) Functionalisation versus
mineralisation of some N-heterocyclic compounds upon UV-illumination in the presence
of un-doped and iron-doped TiO2 photocatalysts. Applied Catalysis B: Environmental
82(3-4): 225–232.
[77] Maury-Ramirez A, De Muyncka W, Stevens R, Demeestere K and De Belie N (2013)
Titanium dioxide based strategies to prevent algal fouling on cementitious materials.
Cement and Concrete Composites 36: 93–100.
[78] Linkous CA, Carter GJ, Locuson DB, Ouellette AJ, Slattery DK and Smith LA (2000)
Photocatalytic inhibition of algae growth using TiO2, WO3, and co-catalyst modifications.
Environmental Science and Technology 34(22): 4754–4758.
[79] Building Research Establishment (1992) BRE Digest 370: Control of Lichens, Moulds
and Similar Growths. Building Research Establishment, Watford, UK.
[80] The Concrete Society (2013) Visual Concrete—Weathering, Stains and Efflorescence.
The Concrete Society, Camberley, UK.
[81] De Graef B, Dick J, De Windt W, De Belie N and Verstraete W (2004) Cleaning of
concrete fouled by algae with the aid of thiobacilli. pp. 55–64. In: Silva MR (ed.). Second
International RILEM Workshop on Microbial Impact on Building Materials. RILEM,
Paris, France.
[82] Dyer TD (2014) Concrete Durability. CRC Press, Boca Raton, FL, USA.
[83] Building Research Establishment (1996) BRE Digest 418: Bird, Bee and Plant Damage
to Buildings. Building Research Establishment, Watford, UK.
[84] Smiley ET, Key A and Greco C (2000) Root barriers and windthrow potential. Journal
of Arboriculture 26(4): 213–217.
[85] Tennis PD, Leming ML and Akers DJ (2004) Pervious Concrete Pavements. Portland
Cement Association, Skokie, IL, and National Ready Mixed Concrete Association, Silver
Spring, MD, USA.
[86] Watson GW, Hewitt AM, Custic M and Lo M (2014) The management of tree root
systems in urban and suburban settings II: a review of strategies to mitigate human
impacts. Arboriculture and Urban Forestry 40(5): 249–271.
[87] Smiley ET, Calfee L, Fraedrich BR and Smiley EJ (2006) Comparison of structural and
noncompacted soils for trees surrounded by pavement. Arboriculture and Urban
Forestry 32(4): 164–169.
Chapter 6
6.1 Introduction
This chapter examines the impact animal activity has on the durability
of concrete. Animals are members of the kingdom Animalia. They are
eukaryotes, like fungi, but differ significantly in that they are exclusively
multicellular and are motile. This chapter differs from previous chapters in
that, given the variety of organisms within this kingdom, the complexity
of their life-cycles, plus the relatively small size of the sub-set which
cause deterioration of concrete, there is little benefit in understanding
the organisms themselves in detail. Thus, the approach taken will be to
identify types of animal documented as having caused damage to concrete
structures, with a discussion of approaches which may protect structures
vulnerable to such attack.
Additionally, the chapter focuses largely on biodeterioration resulting
from concrete being specifically targeted by an animal to satisfy some
requirement for survival. For instance, whilst it is likely that abrasion of
concrete surfaces by the hooves of cattle compromises the durability of
concrete worldwide, the cattle do not benefit from this abrasion and the
concrete is not specifically selected by the animals because it can be abraded,
or contains something that they need to live.
The majority of this chapter involves the marine animals, since these
are the most damaging to concrete, albeit under specific conditions. In most
cases the manner in which these organisms attack concrete is comparable,
and common solutions exist.
The chapter ends with a discussion of the potential for damage to
concrete from bird droppings. Whilst this falls outside of the chapter remit
defined above, this topic has been the subject of some speculation in recent
years, and this discussion aims to evaluate the facts and draw balanced
conclusions from them.
266 Biodeterioration of Concrete
100
80
;12.
u.i 60
~
<(
0::
w
>
0 40
u
20
0
0 2 4 6 10 12
TIME, months
Figure 6.1 Colonisation of a concrete pile off the West coast of Sweden [1].
Damage to Concrete from Animal Activity 267
been documented as causing damage. It has also been found that some
marine worms and sea urchins may also be capable of causing damage.
It has been observed previously that, in some cases, the colonization
of a concrete surface may at least partly be determined by the presence of
substances in the material which can be used as nutrients. In the case of
marine organisms, this is not the reason for causing damage—all of the
nutrients required by such marine animals are likely to be present in the
seawater itself either as dissolved chemical species or as other organisms or
particles of debris. Instead, in most cases, these animals are boring beneath
the surface for shelter, or the damage is a side-effect of feeding off other
organisms on the concrete surface.
Before examining each of these types of marine animal in more detail, it
is worth briefly mentioning the barnacle, whose presence on concrete would
appear to have potentially beneficial effects with regards to durability.
Barnacles are arthropods which fix themselves permanently to a hard
substrate, feeding on particles suspended in seawater. They are encased in a
carapace which is composed of calcium carbonate. Most barnacle species are
attached to the substrate by a calcium carbonate basal plate or membrane,
which effectively isolates the animal from the substrate.
The basal plate is typically a relatively dense formation, and where
barnacles are present on concrete, it has the tendency to make the surface
less permeable to ingressing chemical species. This has particular benefits
from a concrete durability perspective in limiting the movement of chloride
ions from seawater through the concrete cover towards steel reinforcement,
where their presence will eventually initiate corrosion. The effect of barnacle
coverage on the chloride diffusion coefficient of concrete is shown in
Figure 6.2.
Barnacle colonization of concrete tends to be slower than for other
surfaces, such as steel [1]. This is largely the result of the high initial pH [5].
6.2.1 Molluscs
Bivalves
Certain bivalves bore a cylindrical borehole into the substrate (usually rock)
and live in the resulting hole, enlarging it as the organism grows.
Table 6.1 lists bivalve species which have been observed to bore into
concrete. One location in this table is the Panama Canal, where Lithophaga
aristata was established as causing damage to concrete piers. However, the
same report also suggests that other species in this location may also be
268 Biodeterioration of Concrete
ro Q)
2.0
N~ 1.8
E
(.)
r-- 1.6
z
L1J
Q
LL
LL
1.4
1.2 I',
•
L1J
''
0
(.)
z
1.0
' ' •
' .....................
0 ',
0.8 ' '
(i5 ' '-,
::::> '-,
LL
LL 0.6 • ' ,,
0 ••
• ••
L1J 0.4
0
(i:
0.2
0
--'
I
(.) 00
0 20 40 60 80 100
COVERAGE,%
6.2.2 Sponges
Some sea urchins are capable of making quite substantial burrows in rock.
The burrow serves to provide shelter, and is formed by the animal securing
itself to the rock with its feet and abrading the surface with its five jaws
and rubbing with its test (shell) and spines. The motion is usually back and
forth in only one direction, with the ultimate effect of producing a burrow
which consists of a trench which is triangular in profile, with the urchin
normally located at the lowest point [16].
The purple sea urchin (Strongylocentrotus purpuratus)—which can have
a diameter of around 65 mm—has been reported to have burrowed to
significant depths into concrete breakwaters on the Californian coast [20].
Another marine borer identified as being capable of boring into concrete are
the marine worms of the class Polychaeta. Many of these worms are capable
Damage to Concrete from Animal Activity 273
of boring into calcareous rock. Initially it was believed that this was achieved
through a combination of chemical and mechanical processes achieved
through the use of calcium-chelating mucopolysaccharide secretions from
glands and the use of brush-like structures known as setae [21]. However,
experiments in which the setae were removed found that the worm could
bore without difficulty into calcareous rock, presumably through purely
chemical means [22].
A study which examined the biodeterioration of concrete blocks in
an artificial reef-type structure found that the worms appeared to favour
boring into the cement matrix rather than into the calcareous aggregate
used in the blocks [7].
It is firstly worth noting that the most aggressive forms of marine animal
attack are limited to warmer zones, presumably where the productivity of
the sea is sufficiently high to support the sort of energy requirements of
boring on the scales reported. Thus, damage from marine animals in many
parts of the world is likely to be on a scale which is of little concern.
The mechanisms employed by marine animals to bore or burrow into
concrete fall into two categories: mechanical and chemical, and different
options are available in limiting damage depending on what mechanism
is used.
The results discussed in this section point to the likelihood that using
non-calcareous aggregate in concrete is likely to limit the magnitude of
attack. However, different animal types showed different preferences. Thus,
boring sponges, which seem to exclusively bore into calcareous aggregate
would probably be severely limited in its ability to bore by the presence of
wholly siliceous aggregate. Lithophaga appears to be able to bore through
both cement and calcareous aggregate. Nonetheless, use of non-calcareous
aggregate might be expected to hamper or entirely prevent boring. This
appears to be supported by the apparent complete absence of reported
incidences of Lithophaga in concrete without calcareous aggregate.
The only detailed report of Polychaeta worms boring into concrete
appears to indicate less sensitivity to aggregate type and a possible
preference for the cement matrix, indicating that sourcing specific aggregate
types to control this form of attack may not be an option.
Where mechanical damage is observed, it is often associated with
concrete of a low quality, and simply achieving appropriate performance
through mix design and good practice in mixing, placing and curing is
likely to limit biodeterioration considerably.
Whether the mechanism used by marine animals is chemical or
mechanical, the use of physical barriers to prevent boring is likely to be an
274 Biodeterioration of Concrete
effective option. The report of attack by Lithophaga and Cliona off the coast
of Saudi Arabia examined possible barrier options [6]. These included
polymer surface wraps, underwater epoxy paints, polyester bags deployed
around the concrete structure surface into which cement grout is pumped,
and moulded fiberglass jackets. The solution that was eventually selected
was the jackets, which were secured using an epoxy grout introduced into
the annulus between the jacket and the concrete surface.
Barriers of these types are effective through the provision of an
additional layer of protection between the concrete surface and the seawater.
However, it should be noted that purple sea urchins have been reported to
have abraded through steel casings around concrete piles [20], indicating
that even a strong barrier may not entirely resolve the problem. How
much of this damage was carried out by the sea urchin and how much
was the result of abiotic corrosion—which would be likely to be occurring
in parallel—is not clear.
10
0.03
8 I
0.
en
Q) 6
0
E Original edge of cement paste
- 0.02 4
~ 2
i=
z hydrogen urate hexahydrate
<t: Ettringite (AI)
::::l Monosulfate (Fe)
a
0 2 3 4 5 6 7 8 9 10
Figure 6.5 Quantities of solid phases versus depth obtained from geochemical
modelling of uric acid attack of hydrated Portland cement. Model conditions: Excess
of solid uric acid located in acid solution around cement paste; volume of acid solution
= 4 l; mass of cement 80 g; diffusion coefficient = 5 × 10–13 m2/s.
exposed to distilled water [30]. However, the differences were of the order
of tenths of a percent, and the influence of such a difference on concrete
performance is likely to be small. It should also be pointed out that by using
leached solutions, the concentration of uric acid would have been low, due
to its low solubility. Thus, there exists the possibility that the damaging effect
of uric acid would have been greater if solid uric acid had been retained,
since it would continue to dissolve during the experiments to replace any
which had reacted with the concrete, as would happen in reality.
The chemical effect of concrete to bird droppings and uric acid
specifically is still an area that needs further investigation, but there is little
evidence of there being a serious problem.
However, one manner in which bird droppings may cause deterioration
is as a source of nutrients for other species—such as fungi, bacteria and
plants—which can damage concrete, in part through their production of
other organic acids [31]. When fungi metabolize uric acid, the end products
are urea and glyoxylic acid [32]. Urea does not affect the durability of
concrete. Whilst experimental data relating to the interaction of glyoxylic
acid with concrete is not available, it is likely that the compound behaves
in a similar manner to other monocarboxylic acids, such as acetic acid (see
Chapter 3).
It should be stressed that there are many other reasons for limiting the
deposition of bird droppings on the built environment, including those
276 Biodeterioration of Concrete
6.4 References
[1] Andersson MH, Berggren M, Wilhelmsson D and Öhman MC (2009) Epibenthic
colonization of concrete and steel pilings in a cold-temperate embayment: a field
experiment. Helgoland Marine Research 63(3): 249–260.
[2] Oil and Gas UK (2013) The Management of Marine Growth During Commissioning.
Oil and Gas UK, Aberdeen UK.
[3] Agatsuma Y (2013) Chapter 15: Stock enhancement. pp. 213–224. In: Lawrence JM (ed.).
Sea Urchins: Biology and Ecology, 3rd Ed. Academic Press, London.
[4] Yokota H, Hamada H and Iwanami M (2013) Evaluation and prediction on performance
degradation of marine concrete structures. In: Third International Conference on
Sustainable Construction Materials and Technologies, Kyoto, Japan.
[5] Ido S and Shimrit P-F (2015) Blue is the new green—Ecological enhancement of concrete
based coastal and marine infrastructure. Ecological Engineering 84: 260–271.
[6] Wiltsie EA, Brown WC and Al-Shafei K (1984) Rock borer attack on Juaymah trestle
concrete piles. pp. 767–772. In: First International Conference on Case Histories in
Geotechnical Engineering, Missouri University of Science and Technology, Rolla, MO,
USA.
[7] Scott PJB, Moser KA and Risk MJ (1988) Bioerosion of concrete and limestone by marine
organisms: A 13 year experiment from Jamaica. Marine Pollution Bulletin 19(5): 219–222.
[8] Atwood WG and Johnson AA (1924) Marine Structures—their Deterioration and
Preservation. National Research Council, Washington DC, USA.
[9] Castagna M (1973) Shipworms and other marine borers. Marine Fisheries Review 35(8):
7–12.
[10] Jaccarini V, Bannister WH and Micallef H (1968) The pallial glands and rock boring in
Lithophaga lithophaga (Lamellibranchia, Mytilidae). Journal of Zoology 154(4): 397–401.
[11] Warme JE and Marshall NF (1969) Marine borers in calcareous terrigenous rocks of the
pacific coast. American Zoologist 9(3): 765–774.
[12] Moreira J (2006) Patterns of occurrence of grazing molluscs on sandstone and concrete
seawalls in Sydney Harbour (Australia). Molluscan Research 26(1): 51–60.
[13] Tucker JS and Giese AC (1959) Shell repair in chitons. Biological Bulletin 116(2): 318–322.
[14] Trudgill ST (1983) Preliminary estimates of intertidal limestone erosion, One Tree Island,
southern Great Barrier Reef, Australia. Earth Surface Processes and Landforms 8(2):
189–193.
[15] Barbosa SS, Byrne M and Kelaher BP (2008) Bioerosion caused by foraging of the tropical
chiton Acanthopleura gemmata at one tree reef, southern Great Barrier Reef. Coral Reefs
27(3): 635–639.
[16] Kazmer M and Taborosi D (2012) Bioerosion on the small scale—examples from the
tropical and subtropical littoral. Hantkeniana 7(7): 37–94.
[17] Holland RB, Kurtis KE, Moser RD, Kahn LF, Aguayo F and Singh PM (2014) Multiple
deterioration mechanisms in coastal concrete piles: A forensic case study. Concrete
International 36(7): 45–51.
[18] Nava H and Carballo JL (2008) Chemical and mechanical bioerosion of boring sponges
from Mexican Pacific coral reefs. The Journal of Experimental Biology 211(17): 2827–2831.
[19] Duckworth AR and Peterson BJ (2013) Effects of seawater temperature and pH on the
boring rates of the sponge Cliona celata in scallop shells. Marine Biology 160(1): 27–35.
Damage to Concrete from Animal Activity 277
[20] Hinton S (1987) Seashore Life of Southern California: An Introduction to the Animal Life
of California Beaches South of Santa Barbara. University of California Press, Oakland,
CA.
[21] Dorsett DA (1961) The behaviour of Polydora ciliata (Johnst.). Tube-building and
burrowing. Journal of the Marine Biological Association of the United Kingdom 41(3):
577–590.
[22] Haigler SA (1969) Boring mechanism of Polydora websteri inhabiting Crassostrea virginica.
American Zoologist 9(3): 821–828.
[23] Building Research Establishment (1996) BRE Digest 418: Bird, Bee and Plant Damage
to Buildings. Building Research Establishment, Watford, UK.
[24] Maiuolo J, Oppedisano F, Gratteri S, Muscoli C and Mollace V (2016) Regulation of uric
acid metabolism and excretion. International Journal of Cardiology 213: 8–14.
[25] Iwata H, Nishio S, Yokoyama M, Matsumoto A and Takeuchi M (1989) Solubility of
uric acid and supersaturation of monosodium urate: why is uric acid so highly soluble
in urine? Journal of Urology 142(4): 1095–8.
[26] Grases F, Villacampa AI, Costa-Bauzá A and Söhnel O (1999) Uric acid calculi. Scanning
Microscopy 13(2-3): 223–234.
[27] Dawson RMC, Elliott DC, Elliott WH and Jones KM (1986) Data for Biochemical Research
(3rd Ed.) Oxford University Press, Oxford, UK.
[28] Babić-Ivančić V, Füredi-Milhofer H and Brničević N (1992) Precipitation and solubility
of calcium hydrogenurate hexahydrate. Journal of Research of the National Institute
of Standards and Technology 97(3): 365–372.
[29] Presores JB, Cromer KE, Capacci-Daniel C and Swift JA (2013) Calcium urate
hexahydrate. Crystal Growth and Design 13(12): 5162−5164.
[30] Huang CP and Lavenburg G (2011) Impacts of Bird Droppings and Deicing Salts
on Highway Structures: Monitoring, Diagnosis, Prevention. Delaware Center for
Transportation, University of Delaware, Newark, Delaware, USA.
[31] Bassi M and Chiatante D (1976) The role of pigeon excrement in stone biodeterioration.
International Biodeterioration Bulletin 12(3): 73–79.
[32] Vera-Ponce de León A, Sanchez-Flores A, Rosenblueth M and Martínez-Romero E (2016)
Fungal community associated with Dactylopius (Hemiptera: Coccoidea: Dactylopiidae)
and its role in uric acid metabolism. Frontiers in Microbiology 9 Article 954.
Index
carbonation 3, 83, 84, 87, 92, 93, 121, 160, citric acid 3, 8, 9, 66–69, 71, 115, 157, 163,
162, 225, 249 175, 179–182, 184, 190, 195, 229, 230,
carbonic acid 30–38, 87, 88, 122, 125, 126 235, 256
carboxylic acid 34, 39, 42, 66, 127 Cladonia 166
Candelaria 177, 178 Cladosporium 165, 172, 174, 194, 204, 205
Candelariella 177, 195 Clauzadea 177
caoxite 48 Clematis vitalba 223
caper 222, 224 cleaning 4, 134, 135, 198, 200, 201, 204, 253,
Capparis 222, 224 254, 256
Carditamera 268 Climacium 237
Catapyrenium 177 clinker 16, 17, 89, 105
Catillaria 177 coastal protection 245, 246
cellulose 127, 153, 155, 156, 175, 196, 212, coatings 137, 139–141, 203, 251–253
213, 215, 231 Coccobotrys 233
Celtis australis 223 Coccomyxa 233
cement 1, 3, 5, 7, 15–18, 20–22, 24–26, 28, colemanite 202
29, 31–33, 35, 39, 42, 45, 47, 49, 55, 58, Collema 177
62–68, 70, 71, 77, 83–90, 92–96, 98–105, Colletotrichum 170
107–121, 125–129, 136–139, 142, 143, Cololejeunea 238
160, 162, 163, 165, 168–171, 179–196, common ion effect 13
199–205, 225, 229, 230, 231, 245, 246, complex 3, 13–15, 20, 21, 24, 28, 29, 31, 32,
247, 249, 250, 252, 254, 256, 259, 269, 34, 35, 38, 39, 41, 44, 45, 47, 50, 60, 63,
272–275 64–67, 71, 86, 89, 105, 136, 154, 159, 164,
calcium aluminate 18, 20, 22, 33, 70, 165, 217, 274
93, 100, 101, 105, 111, 116, 120, 121, complexolysis 70, 71
183, 190, 200, 249, 256 Coniosporium 160, 172
calcium sulphoaluminate 101, 104, Conocephalum 238
111, 113, 181–185 copper 133, 139, 163, 201, 204, 258
cement content 88, 98–100, 109, 111, Coprinus 159
119, 121, 136–138 coral 270, 272
Portland 16–18, 20–22, 24–26, 29, Corella 266
31–33, 35, 39, 42, 45, 47, 62–66, 68, corrosion 1–3, 24, 98, 105, 109, 111, 123, 127,
70, 85, 88, 89, 93, 94, 99–101, 107–111, 128, 141, 198, 246, 247, 254, 267, 274
115, 116, 118, 125, 127, 138, 143, 160, cracks 71, 86, 96, 138, 141, 205, 219–221, 235,
162, 170, 171, 181–189, 191, 192, 201, 242, 245, 254, 256
202, 246, 247, 249, 250, 256, 275 Croton bonplandianus 224
Centranthus ruber 222 Cryptomonas 225
Ceratium 233 CSH gel – see calcium silicate hydrate gel
Ceratodon 237 Ctenidium 238
Chaetomium 174, 199 Curvularia 172
Chaetophoma 172 cyanobacteria 85, 88, 90, 124, 155, 157, 171,
chelation 15 212, 214, 228, 231, 250
chemotrophic 78 Cyanosarcina 124
Chernobyl 171, 174, 175 cypress 243
Chiloscyphus 237
chiton 270, 271 D
Chlamydomonas 230
dairy industry 115
Chlorella 230, 233, 249–252
Dalbergia sissoo 224
chlorine 132
Damiria 272
chlorophyll 197, 212–214, 250
Davies equation 11, 12
Chlorosarcinopsis 233
Debye-Hückel equation 11
Choricystis 233
decalcification 22, 26, 70, 108, 116, 138
Chroococcus 124
deprotonation 8, 24, 43, 44, 46, 50, 53, 54,
Chrysosporium 174
57, 66
Ciona 266
282 Biodeterioration of Concrete
Deschampsia flexuosia 229 fly ash 17, 88, 89, 100–103, 111, 112, 116, 117,
Desmococcus 233 119, 121, 127, 137, 138, 143, 181–184,
Desulfovibrio 90, 135, 136 199, 200, 246
Dicranella 238 Fonsecaea 173
Dicranum 238 Fontinalis 238
Didymon 238 formate 34, 35, 38, 39
Dirina 177 formic acid 34, 38, 39, 40, 230
Dirinaria 177 foundation 84, 243, 259
dissociation ratio 8 fragmentation 3, 71, 100, 158, 163, 181, 182,
Dittrichia viscosa 222 184, 185, 189, 190
Doratomyces 174 Fraxinus excelsior 223
Dreponacladus 238 Fucus 233, 245
Dryopteris filix-mas 223 Fulgensia 177
dry rot 175 fumarate 64
fumaric acid 57, 64, 229, 241
E fumarprotocetraric acid 167
Funaria 238
Eh 19, 20
fungi 3–5, 72, 153–163, 165, 169–175, 179,
elderberry 222
180, 188, 190–201, 203, 249, 265, 275
elm 223, 224, 243
calcicolous 160
Encalypta 238
silicicolous 160, 180
Endocarpon 177
corrosion of post-tensioned
Enterococcus 115
reinforcement 210
enzymes 82, 88, 127, 155, 156, 159, 196
fungicide 201, 202
Epicoccum 173
Fusarium 173, 174, 191, 193, 198
Erysimum cheiri 222
ettringite 17, 18, 24–26, 33, 93–95, 98, 100, G
122, 136, 183
Eucalyptus 245 gastropods 271
Euonymus europaeus 222 Geminella 233
Eurhynchium 238 geochemical modelling 71, 94, 95, 101, 104,
Evernia 166 107, 108, 111, 113, 183–185, 187–190,
Exophiala 160, 173 274, 275
expansive reaction products 71 geopolymer 246, 247
extracellular polymeric 82, 107, 157, 165, Geotrichum 173, 174
168, 170, 219 GGBS – see ground granulated blastfurnace
slag
F gibbsite 21, 22, 26, 39, 44, 50, 57, 66, 70, 71,
111, 183
FA – see fly ash
Gloeocapsa 124
Fallopia japonica 222, 255
Gloeocapsa-Chroococcus 124
false acacia 222
Gloeothece 124
fern 218, 223
glycolate 46, 47
male 217, 218, 223
glycolic acid 46–48
ferrihydrite 10, 21, 22, 26, 46, 50, 66, 70, 89,
glyoxylic acid 275
95, 188
Grimmia 238
fibres 82, 137, 139, 141, 195, 203, 204, 246,
ground granulated blastfurnace slag 17, 18,
258
20, 100–102, 120–122, 126, 199, 250
Ficus 223, 224
growth 3, 5, 77, 79–81, 83–85, 93, 101, 107,
fig 223
108, 113, 114, 122–125, 127, 130–132,
Fissidens 238
135, 138, 154–156, 158–161, 165, 175,
flagella 82, 238
176, 178, 192–195, 197, 199–205, 216–
Flavobacterium 123
218, 220, 221, 225, 226, 228, 231, 232,
243–245, 247–253, 255, 256, 258–260
Index 283