corrosion inhibitors for the preservation of metallic...

50
Corrosion inhibitors for the preservation of metallic heritage artefacts E. Cano, D. Lafuente Centro Nacional de Investigaciones Metalúrgicas (CENIM)- Consejo Superior de Investigaciones Científicas (CSIC) Avda. Gregorio del Amo 8, 28040 Madrid [email protected] Publicado originalmente en: EFC book nº65 “Corrosion and conservation of cultural heritage artefacts” P.Dillmann, A. Adriens, E. Angelini and D. Watkinson (eds.) WoodHead Publishing. European Federation of Corrosion.2013. pp. 570-594 ISBN 978-1-78242-154-2(Print) 978-1-78242-157-3 (Online) D.O.I. 10.1533/9781782421573 INTRODUCTION With few exceptions, all metals are subject to degradation by chemical reaction of the metal with its environment, that is, corrosion. This includes, of course, metals that make up or are part of cultural heritage assets. While

Upload: dangmien

Post on 07-Apr-2018

219 views

Category:

Documents


3 download

TRANSCRIPT

Corrosion inhibitors for the preservation of metallic heritage

artefacts

E. Cano, D. Lafuente

Centro Nacional de Investigaciones Metalúrgicas (CENIM)-Consejo Superior de

Investigaciones Científicas (CSIC)

Avda. Gregorio del Amo 8, 28040 Madrid

[email protected]

Publicado originalmente en: EFC book nº65 “Corrosion and conservation

of cultural heritage artefacts” P.Dillmann, A. Adriens, E. Angelini and D.

Watkinson (eds.) WoodHead Publishing. European Federation of

Corrosion.2013. pp. 570-594

ISBN 978-1-78242-154-2(Print) 978-1-78242-157-3 (Online)

D.O.I. 10.1533/9781782421573

INTRODUCTION

With few exceptions, all metals are subject to degradation by chemical reaction of the

metal with its environment, that is, corrosion. This includes, of course, metals that make

up or are part of cultural heritage assets. While corrosion of metals in the industrial field

can be, in many circumstances, expressed in economic terms, due to the economic

losses caused by this process, due to the costs involved in the maintenance of metallic

objects or their replacement, in the case of cultural heritage, every object is unique and

therefore any loss is irreplaceable.

Corrosion is defined by the IUPAC as[1]:

“An irreversible interfacial reaction of a material (metal, ceramic, polymer) with its

environment which results in consumption of the material or in dissolution into the

material of a component of the environment. Often, but not necessarily, corrosion

results in effects detrimental to the usage of the material considered. Exclusively

physical or mechanical processes such as melting or evaporation, abrasion or

mechanical fracture are not included in the term corrosion”.

ISO 8044 Standard defines it as [2]:

“Physicochemical interaction between a metal and its environment that results in

changes in the properties of the metal, and which may lead to significant impairment of

the function of the metal, the environment, or the technical system, of which these form

a part”

The corrosion effects1 in the case of artistic or historic artefacts can be seen as positive,

for instance, producing a patina which is considered aesthetically pleasant. However, in

most cases, it produces a damage. In the case of heritage artefacts, the impairment of the

function produced by corrosion is related with the loss of some specific values (artistic,

historic, scientific, social, etc.) of that object.

These definitions of corrosion give us some clues about different strategies that can be

used to prevent or reduce corrosion. First of all, being a reaction of a material with its

environment, the first choice could be to change either the material or the environment.

While the first is not applicable to cultural heritage, since the physical nature of the

object cannot be changed, the modification of the environment is probably the first

choice: preventive conservation2 strategies involve no action on the object itself, and are

therefore preferable from the point of view of current conservation ethics. This strategy

is more easily applied in indoor environments, such as museums, where the relative

humidity and pollution can be controlled. For outdoor environments, this approach is

1 See ISO 8044 [2] ISO 8044:1999, 'Corrosion of metals and alloys -- Basic terms and definitions'.for the definition of “corrosion effect” and “corrosion damage” .2Preventive conservation is defined as “all measures and actions aimed at avoiding and minimizingfuture deterioration or loss. They are carried out within the context or on the surroundingsof an item, but more often a group of items, whatever their age and condition. Thesemeasures and actions are indirect – they do not interfere with the materials and structuresof the items. They do not modify their appearance.” [3]

ICOM-CC: 'Terminology to characterize the conservation of tangible cultural heritage', Resolution adopted by the ICOM-CC membership at the 15th Triennial Conference, New Delhi, 2008.

more difficult to implement: atmospheric humidity cannot be controlled –although

covering some artefacts to protect them from direct precipitation is sometimes feasible-

and reducing pollution involves large scale actions, such as reducing the traffic around

some monuments, usually with a limited impact. Even in indoor environments, in many

cases it is not economically or practically feasible to act on the environment, but the

interfacial character of the reaction, as pointed out the IUPAC’s definition, gives us the

option for a different strategy: acting on the metal surface to avoid its contact with the

environment or reduce the electrochemical reaction rates.

Many corrosion prevention treatments fall in this category, being the most usual organic

coatings, such as paints and varnishes, were a polymeric material is applied on the

metal; or coatings with inorganic materials, such as metals (usually nobler than the base

metal) or ceramics (applied by sol-gel, PVD, CVD, etc.). Passivation by formation of a

protective and homogeneous layer of corrosion products on the surface of the metal

(either naturally or artificially) also produces the isolation of the metal from the

environment. Many corrosion inhibitors can also be included in this category, since they

form a protective layer (of molecular thickness) that avoids the reaction of the metal

with the environment, as it will be shown later.

Some requirements should be considered when choosing a corrosion protection

treatment for cultural heritage objects3: they should produce no or very little change in

the surface appearance; should be as reversible as possible, that is, it should be possible

to remove them and return the object to its original state; should not modify the material

of the original artefact, including, in most cases, the modifications suffered during the

history of the object, such as patinas or corrosion layers (as far as they do not threaten

the object conservation and its legibility); they need to have long term efficiency, since

heritage artefacts are intended to be preserved for a long time (as long as possible); and

finally, it is desirable for them to have an easy maintenance, because any treatment will

eventually need to be renewed.

Corrosion inhibitors fulfil to a large extent some of these requirements. Some of them

reduce corrosion settling adsorbed layers of the inhibitor molecules on the surface of the

metal. In most cases, the low thickness of the inhibitor protective layers makes them

invisible (in other cases, however, the inhibitors produces visible changes). These layers 3 For terms and definitions related with conservation of cultural heritage, refer to EN 15898 Standard. [4]

EN 15898:2011, 'Conservation of cultural property. Main general terms and definitions'.

are chemically stable in the environment in which they are formed. Due to their low

thickness, they are not resistant to mechanical removal but if the inhibitor is present in

the environment (and replenished when it is consumed), the layer will eventually be

formed again. Another advantage of inhibitors is that they can be used in many cases –

as it is common practice in metallic heritage conservation– in combination with

protective coatings, increasing the protective function of the whole system.

TYPES AND MECHANISMS OF CORROSION INHIBITORS

Corrosion inhibitors are defined by ISO 8044 as “a chemical substance that decreases

the corrosion rate when present in the corrosion system at suitable concentration,

without significantly changing the concentration of any other corrosion agent” [2]. As

we will see later, the use of corrosion inhibitor for metallic heritage conservation is, in

many cases, in the limits of this definition and closer to the coating or conversion

coating ones, that is: “a substance layer that, on the metal surface, decreases corrosion

rate”

Metals corrosion, specifically which affects cultural heritage, is in a vast majority of

cases an electrochemical reaction, involving an anodic reaction, typically:

[1]and a cathodic process,

[2][3]

The mechanism of inhibition involves the reduction of the anodic, the cathodic, or both

reactions rates. Accordingly, a first typical classification of corrosion inhibitors is made

in anodic inhibitors (for those inhibiting the anodic reaction), cathodic inhibitors (those

inhibiting the cathodic reaction) or mixed type inhibitors (acting on both anodic and

cathodic reactions). Depending on the type of inhibitor, the corrosion potential (Ecorr) of

the system is modified in a positive (anodic inhibitor) or negative direction (cathodic

inhibitor), or remains unaltered (mixed inhibitor).

Many other classifications can be found in the literature, attending to the chemical

composition (organic, inorganic, surfactants…), the type of corrosive media in which

they are effective (inhibitors for acid, neutral or alkaline solutions, for chloride-

containing solutions, vapor phase inhibitors…), or the field of application (for cooling

systems, for drinking water systems, for reinforced concrete or, as in our case, for

cultural heritage).

Specific requirements and needs for corrosion inhibitors in conservation

treatments

Some inhibitors have been and are currently being used extensively in conservation and

restoration treatments. Under the European project PROMET, a survey was made

amongst conservators-restorers of ten Mediterranean countries to determine the type of

coatings and corrosion inhibitors used for conservation treatments of copper, iron and

silver alloys [5]. The results showed that most of conservators used ethanol solutions of

benzotriazole (BTA) for copper alloys, applied to the objects by brushing, immersion or

spraying. For iron alloys, the use of corrosion inhibitors was not so popular, being

tannic acid and BTA the preferred inhibitors. For silver, the use of inhibitors was scarce

but again BTA was the selected product. A summary of corrosion inhibitors used for

conservation-restoration treatments of different metals reviewed in this chapter is

presented in Table 1.

As opposed to industrial applications, in the metal conservation field, the main way of

using inhibitors is not adding the substance to the corrosive liquid media, since the

majority of objects are exposed to atmospheric conditions. On the contrary, inhibitors

are used to produce surface modifications or films by adsorption of the inhibitor on the

metal surface, by means of the metal immersion on a non-corrosive inhibitor solution

for a given time [6], followed by drying and, in many cases, a top layer with a varnish

or wax coating [5]. This different way of use is significant for the researches on the use

of inhibitors in the cultural heritage field. While in immersion tests the competitive

adsorption of the ions of the solution, the water molecules and the inhibitor molecules

have a key role in the inhibition process and its efficiency, in their application as films

the physical and chemical resistance of the formed film is a key factor affecting the

efficiency of the inhibitor. The different application methods might produce differences

in the inhibition properties, as has been shown for instance by Mansfeld et al., who

reported that the BTA was a good inhibitor for Cu immersed in 5% NaCl, but not when

it was pre-coated by BTA and then exposed to the 5% NaCl solution [7]; or by Kosec et

al, that showed that the inhibition properties of BTA and 1-(p-tolyl)-4-methyl imidazole

were different when brushed onto patinated bronze and when the patinatedbronze was

immersed in a solution containing the dissolved inhibitor[8].

Another key difference is that, while in basic and industrial-orientedresearches the

inhibitors are applied to the clean metal, in heritage conservation they are applied in

most cases over pre-existing corrosion products (or patinas) that have to be preserved.

Therefore, the testing of these products for this application requires the use of a specific

methodology adapted to the particular needs and conditions of their use. For instance,

the PROMET project combined accelerated and electrochemical laboratory tests on

artificially and naturally corroded coupons, simulating the condition of historic

artefacts, with natural exposure tests in real conditions, both for coupons and real

objects[9]. Most of the inhibitors’ studies are made on clean metal, but some recent

papers have dedicated some attention to the recreation of surfaces similar to the ancient

objects ones. Faltermeier, in 1999, pointed out that studies published on inhibitors did

not dealt with heterogeneous corrosion layers and/or ternary alloys such as Cu-Zn-Pb

commonly found in archaeological artefacts[10]. He proposed a standard methodology

for testing inhibitors including the formation of a cupric chloride patina on the samples

prior to inhibitor application, and the evaluation of the inhibitor efficiency using

gravimetric methods. In recent years, many researchers have carried out studies of

corrosion inhibitors using different types of patinas on bronze alloys, trying to simulate

as close as possible the real conditions of inhibitors application in conservation-

restoration treatments [9, 11-14]. Kosec et al. demonstrated the relevance of the patina

composition in a recent paper, which showed significant differences in the inhibitor

efficiency depending on the patina on which they are applied: they found out

thatinvestigated inhibitors (BTA and 1-(p-tolyl)-4-methyl imidazol) inhibited the

corrosion of both electrochemically formed and chloride-based patina, but were

ineffective in the case of a nitrate-based patina[8].

Some papers have specifically studied the reaction of the inhibitors with the corrosion

products. Brostoff studied the reaction of BTA with different Cu corrosion products and

demonstrated that the presence of copper chloride have a great effect on the Cu-BTA

reactions, predominating Cu(I)-BTA complexes in reactions with cuprite and copper

powder, and Cu(II)-BTA complexes in reactions with chloride containing minerals

(nantokite, atacamite and paratacamite)[15]. Rahmouni et al. studied the

electrochemical behaviour of different natural and artificial patinas in presence of some

inhibitors: BTA, amino-triazole (ATA) and bi-triazole, showing that the inhibitive

properties of the compounds are diverse for each patinas[16]. The specific reaction of5-

amino-2-mercapto-1,2,4-thiadiazole (AMT) with typical corrosion products in heritage

artefacts, namely paratacamite, malachite and brochantite, has been recently studied by

D’Ars et al., who concluded that AMT reacts with copper salts forming an AMT-Cu(II),

and that brochantite suffers a partial alteration after its reaction with AMT [17].

In some cases, tests are carried out using real objects, to test them in actual-life

conditions[9, 16, 18, 19]. The use of real objects have many disadvantages, mainly the

low reproducibility, due to the reduced number of available samples and the huge

variability in their composition, conditions, etc., but the historic materials behaviour

could be in some cases very different to the modern ones. For instance, Bastidas and

Otero demonstrated that the behaviour of copper from ancient chalcographic plates in

acid cleaning baths with inhibitors was be very different to the modern copper samples,

due to the presence of numerous inclusions in the ancient ones which can act as

preferential sitesfor pitting[20, 21].

Inhibitors can also be used as vapour corrosion inhibitors (VCI), also known as vapour

phase inhibitors (VPI). VCIs are substances with a low vapour pressure that have the

ability to vaporise and condense on the metal surface, forming an adsorbed layer that

protects the metal from the corrosive environment [22]. This makes them suitable for

metal protection in enclosed spaces, such as display cases or packages, and the use of

VCI for protection of metallic heritage has been proposed in some cases [23, 24].The

need of closed spaces and, especially, the possible health hazards for the conservation

professionals or visitors of the museum make conservators reluctant to use this kind of

inhibitors[25], even though some recent works claim the safety of VCI in packaging

materials [26].

INHIBITORS EVALUATION

The parameter commonly used to quantify the corrosion inhibition properties of a

substance is the inhibitor efficiency (IE), defined as:

[4]whereCRabs and CRpre are the metal corrosion rate in the absence and presence of

inhibitor, respectively. Corrosion rates can be obtained by different ways, being the

most used gravimetric and electrochemical techniques. Gravimetric measurements are

usually carried out using coupons and measuring the weight before and after exposure

to the corrosive environment, without the inhibitor and with the inhibitor added to the

solution or the metal pre-treated. Electrochemical techniques, such as polarization

resistance, calculation of Tafel slopes from voltametries and electrochemical impedance

spectroscopy (EIS), allow for an indirect calculation of corrosion rates and have the

advantage of providing information on the mechanisms of the corrosion and inhibition

processes [27, 28].

It should be noticed that, in the case of heritage artefacts, the weight measurements (as a

direct measure of the chemical reaction rate) might not reflect the damage suffered by

the object, which is far more complex and it is related with the notion of “loss of value”

(that could be aesthetic, symbolic, historic, socioeconomic, scientific, technologic, etc.)

[4]. In some cases, an small corrosion effect might imply a significant loss of value, e.g.

in the case of silver tarnishing; while in others, a stronger corrosion effect might be

considered acceptable, such as in the formation of a patina in an outdoor sculpture.

Some attempts have been made to quantify this loss of value in the case of damage

caused by pollution, using concepts such as “non observed adverse effect level”

(NOAEL) or “lowest observed effect level” (LOAEL) [29]. For this reason and due to

the indefinite life expectancy of a heritage object, it is not feasible to establish a target

efficiency value for a corrosion inhibitor for this application, which should, in principle,

be “as high as possible”.

Since the adsorption of the inhibitor molecules on the metal surface is a fundamental

step in the inhibition process, the study of the adsorption process can also provide useful

information of the inhibition mechanisms. The study of the adsorption isotherms is a

classical method for studying this process and they have the general form:

[5]

were k is the equilibrium binding constant of the adsorption reaction; c is the inhibitor

concentration; g(,) is the configurational term parameter, in which is the number of

water molecules replaced by one molecule of organic inhibitor and is the degree of

coverage of the metallic surface; and f is the interaction term parameter (f> 0 lateral

attraction, and f< 0 lateral repulsion between the adsorbed inhibitor molecules) [30-35].

These models assume that: (i) the adsorption sites on the metal surface are homogeneous,

(ii) a mono-layerinhibitor adsorption is formed, and (iii) corrosionis uniform and no

localised attack takes place[30], which is not always the case, especially in heritage

objects. They also consider a thermodynamic equilibrium between the inhibitors in the

environment and the adsorbed layer, thus in those cases where the concentration of the

inhibitor changes or they are applied in solvents and then exposed to a different

environment, these models are not useful.

The use of quantum chemical calculations for the evaluation of inhibition properties is

not a new tool, but has gained a huge popularity in the last years due to the

improvement of the calculation capabilities of personal computers. A recent review of

the use of this techniques has been recently published by Gece[36]. This calculations

allow to correlate inhibitor efficiencies with molecular properties such as orbital

energies (mainly highest occupied molecular orbital energy, EHOMO, and lowest

unoccupied molecular orbital energy, ELUMO), dipole moment, charge density, heat of

formation and ionization potential [37]. Quantum chemistry calculations can be very

helpful to study fundamental inhibition mechanisms and have shown a good correlation

with experimental data in some cases, for simple corrosion systems. However, in others,

the correlation is not so clear, since the assumptions and simplifications needed to allow

the computing of the models might neglect important factors in the corrosion inhibition

process [36]. This is especially true in the case of very complex systems such as

metallic heritage artefacts, which typically have inhomogeneous surfaces, usually

covered by corrosion products.

Surface analysis techniques have also been extensively used for the characterization of

the inhibitor layers formed on the metals [38-43]. The use of this techniques allows the

study of the layer composition formed on metals following the procedures used by

conservators-restorers, and the study after exposure of the coated metals to the

atmospheric environment, closer to the real life of the objectsthan the immersion tests

necessary for electrochemical measurements [42]. The main disadvantage is that

sometimes the efficiencies observed by electrochemical or gravimetric measurements

are difficult to correlate with the surface characterization results [44].

For readers interested in further details on evaluation methods for corrosion inhibitors,

a comprehensive review can be found in the book by V.S. Sastri “Corosion Inhibitors.

Principles and Applications” [45] or the recent one by the same author on “Green

Corrosion Inhibitors: Theory and Practice” [46].

CORROSION INHIBITORS USED IN CONSERVATION TREATMENTS

Inhibitors for Copper and its alloys

BTA is by far the most used and most studied corrosion inhibitor for copper and its

alloys, both for industrial applications and for heritage conservation uses. Dugdale and

Cotton’s pioneering work in 1963 reported that BTA was at that time already in use for

industrial applications since many years, but this is considered to be the first scientific

study of its inhibition mechanism [47]. This first work demonstrated that BTA forms a

polymeric complex with copper that acts as a barrier for corrosion, and that this film is

very thin and chemically resistant. It was also proposed that BTA protects copper in

aqueous and gaseous environments polluted with sulphur dioxide, hydrogen sulphide

and salt mist [24]. Cotton and Scholes also suggested different forms for BTA

application, such as its addition to an aqueous solution, its incorporation into lacquers

and polishes, or its use as VCI by exposing the metal to benzotriazole-impregnated

paper [24].

Scientific literature regarding the use, applications and mechanisms of BTA as a copper

corrosion inhibitor is huge, in all types of corrosive solutions: acid, neutral and alkaline,

oxidizing or reducing, containing different ions (chlorides, sulphates, ammonia…), etc.

BTA is known to be chemically adsorbed on the metal, displacing the adsorbed water

and forming different complexes with Cu(I) and Cu(II) [30, 32, 34, 42]. The

composition, structure and orientation of the adsorbed layer have also been matter of

different studies, but a complete agreement has not been reached. The review of all this

topics is out of the scope of this text, but interested readers can find a detailed review on

a recent publication by FinšgarandMilošev[43].

BTA first application for heritage conservation was contemporary to Dugdale and

Cotton’s research. Madsen [48] proposed BTA as a treatment for bronze disease in

1967, and suggested its application by objects immersion in a 3% BTA solution in

ethanol in a vacuum chamber at room temperature; or, if the vacuum was not available,

in a 3% water solution at 60º C. Ethanol solution was considered to be better because

the low surface tension of the solution allows it to reach deep pores in the corrosion

layer. BTA is also less soluble in water and treatment with the water solution might also

produce an undesirable white deposit on the surface of the patina [49]. The current

standard practise in conservation follows the original recommendation by Madsen,

most of the times without vacuum, and for immersion times of up to 24 h [50].

Madsen also suggested using BTA treatments as a first layer, with an additional coating

with Incralac -an acrylic based varnish containing BTA itself-. Even though BTA was

originally included in the Incralac formulation as an UV stabilizer, not as a corrosion

inhibitor [50], Madsen suggested that it might migrate to the metal surface and

contribute to its inhibition.

BTA became a very popular treatment for antiquities in the following years. Sease, in a

review made in 1978about the BTA use for antiquities, recognized this treatment as a

popular one, and pointed out that the exact mechanism of inhibition was not completely

understood, since the formation of a polymeric layer on the copper surface would not

satisfactorily explain the inhibition mechanism in heavily corroded objects such as

antiquities [51]. She also raised some concerns about its efficiency to arrest bronze

disease, since the cupric chloride-BTA complex layer formed upon treatment would

only be superficial and therefore subject to eventual disruption and reactivation of the

corrosion process [51]. Indeed, reactivation of bronze disease after BTA treatments has

been reported in many occasions [52]. The low pH present in pits in active bronze

disease might be responsible for the lower efficiency of inhibition in these cases [53],

since at low pH the adsorption of molecules, rather than complex formation, is favoured

[43].

The efficiency of BTA for treatment of copper alloys seems to be lower than for pure

copper. Using electrochemical techniques, Brunoro et al. demonstrated that the

inhibition of some BTA derivatives was lower in complex multiphasic bronze alloys,

which was attributed to the alloying elements (Sn, Zn and Pb), forming a weaker metal–

triazole bond[54]. However, results obtained by Sharma et al. using leaded bronzes

treated with a neutral BTA aqueous solution, showed the formation of a lead complex

film, so this treatment was proposed for leaded bronzes[55]. More recent studies by

Galtayries et al, have tried to elucidate these differences using surface analysis

techniques. Their results, however, did not demonstrate a straightforward relationship

between the inhibitor efficiency and the amount of adsorbed inhibitor. The

discrepancies found with Sharma’s work were attributed to differences in both the

concentrations of the inhibitor used and the samples pre-treatments [44].

Quite early in its application some concerns were raised about the toxicity of the BTA.

Oddy and Sease recommended some cautions in its use [51, 56]. The environmental and

health hazards of this compound is still a controversial issue: while some authors claim

that it is carcinogenic and causes toxic effects on flora and fauna [6], others classify it

only as slightly toxic [43]. Some recent toxicologic studies reported BTA as having a

low acute toxicity, and non presenting evidence of antiestrogenic activity in the in vivo

assays [57]. However, suspects about its safety remain and the search for “safe”

alternatives to BTA has been a constant in corrosion inhibitor studies in the last 30

years, both in corrosion science in general and in the metal conservation literature.

Since the first BTA researches, alternatives to this molecule have been sought. Cotton

and Schonlesstudied other triazoles: indazole, benzimidazole, indole and methyl-

benzotriazole, and found that just indazoleinduced resistance to tarnishing on copper

strips exposed to salt spray, but the layer was less resistant to organic solvents than

BTA [24]. Ganorkar et al. proposed the use of AMT as a complexing agent which is

able to remove cuprous chloride from corroded bronzes and forming a polymeric layer

on the surface capable of inhibiting the corrosion process [58].

Faltermeier studied several nitrogen based and sulphur based alternatives: BTA and

AMT, 2-aminopyrimidine, 5,6-dimethylbenzimidazole, 2-mercaptobenzimidadole, 2-

mercaptobenzoxazole, 2-mercaptobenzothiazole and 2-

mercaptopyrimidine.Nevertheless, none of them yielded a better efficiency than BTA.

Besides, some of them caused unacceptable colour changes on the coupons surface,and

thereforethe conclusion of his work was that none of them could be recommended for

conservation treatments of chloride containing archaeological artefacts[59]. AMT and

BTA were also compared as inhibitors for their use in acid cleaning of historic

chalcographic plates, resulting in a better efficiency of AMT, especially in citric acid

[20, 21].

Balbo et al. also tested some thiadiazole and imidazole derivatives and 3-

mercaptopropyl-trimetoxysilane (PropS-SH) as alternatives to BTA for inhibition of

cast bronze exposed to concentrated acid rain and in 3.5 wt.%NaCl solution. The best

results were yielded by AMT and especially by PropS-SHwhen enough curing times

were used[6]. As an alternative to BTA, Thachli et al. proposed the use of

electropolimerizedATA for copper protection, reporting an efficiency of 99% and

remaining efficient even after one month of immersion in 0.5 M NaCl solution [60].

These tests were made on pure clean copper, but Rahnouni and co-workers tested the

same compound and bi-triazole in artificial patinas -simulating ancient ones- and real

coins. In this work, ATA yielded good results, but not as good as BTA [16]. Other

researchers have also reached good results with other azoles: 4-methyl-1-(p-tolyl)-

imidazole (TMI), 1-phenyl 4-methyl-imidazole (PMI), 2-mercapto 5-R-acetylamino-

1,3,4-thiadiazole (MAcT), 2-mercapto 5-R-amino-1,3,4-thiadiazole (MAT), were

studied by Muresan et al. for their use on the inhibition ofartificially patinated bronze.

Again, results showed that TMI and MAcT were efficient inhibitors but with a lower

performance than BTA [12]. TMI has also been studied by Marušić et al. on three

different artificial patinas on bronze [11].

Brunoro and co-workers tested 5-methyl-1,2,3-benzotriazole; 5-hexyl-1,2,3-

benzotriazole; 5-octyl-1,2,3-benzotriazole; 5-methoxy-1,2,3-benzotriazole; 5-

(piridinethoxycarbonyl)-1,2,3-benzotriazole chloride; 2-chloroethyl-1,2,3-benzotriazol-

5-carboxylate; and 5-mercapto-1-phenyltetrazole as inhibitors for different cast bronzes,

containing Sn, Zn and Pb[54].

Dermaj et al. proposed the use of 3-phenyl-1,2,4-triazole-5-thione (PTS) as inhibitor for

bronzes [61, 62]. This compound has also been tested on pre-corroded bronze artifacts,

simulating archaeological ones, with promising results[18]; and on real objects,

providing a good protection except in Ag containing objects, in which the sulphur in the

inhibitor caused a darkening on theobjects surface [9].

A very active area of research in corrosion inhibitors in last years has been the use of

the so-called “green inhibitors”, i.e., inhibitors obtained from natural plant extracts [63].

These products present the advantage of their biodegradability, easy availability and

non-toxic nature, but also disadvantages such as the very complex nature and high

variability of the extracts (depending on the exact origin of the plants). Their application

to cultural heritage is, so far, quite limited, even though efforts are being made to use

this kind of products as substitutes for unhealthy compounds. Hammouch et al. have

done some tests of Opuntiaficusindica extract (OTH) for bronze and iron based

artefacts[64], but in the case of bronze no long-term or real object tests have been made

[9].

Great attention has been given in the last years to the application of saturated linear

carboxylic acids and their sodium salts to the protection of heritage metals. Sodium

heptanoate (CH3(CH2)5COONa) was first tested as inhibitor for copper in immersion,

showing a good inhibition of the corrosion attributable to the formation of a copper

heptanoate layer [65, 66]. Tests in real objects using sodium heptanoate have also given

satisfactory results [67]. Sodium decanoate and other carboxylation treatments (a

solution containing the carboxylic acid and an oxidant, either sodium perborate or

hydrogen peroxide) have also been studied [14, 68]. While the carboxylation solutions

formed a thicker layer when applied on copper [14], they were not suitable for

application on brass since they caused the appearance of white stains on the surface of

the metal[68]. Alternative methods for the layer deposition have been evaluated with

different results: Elia et al. tested the deposition fromethanolic solutions of heptanoic,

decanoic and docecanoic acids, in an attempt to improve the resistance of the coating

and to avoid the use of water (as authors consider that it can promote corrosion), but the

results have not been satisfactory[69]; on the other side, the deposition of the same

compounds using cyclic voltammetry showed promising results, saving treatment time

and allowing a good control of the deposited layer [70].

Since inhibition treatments are not completely effective in some cases, even with BTA

which is usually the best inhibitor, a very interesting approach is to take advantage of

the synergetic effects of the combination of various inhibitors. Golfomitsouet al. have

carried out a very interesting research on the topic, evaluating the efficiency and

mechanisms of the combination of BTA with other inhibitors and additives: AMT, 1-

phenyl-5-mercapto-tetrazole, ethanolamine, benzylamine, potassium ethyl xantate and

potassium iodide [53, 71, 72]. She found that the combination of AMT 0.01M and BTA

0.1 M in distilled water increased the efficiency of the treatment, with the advantageof

requiring a lower concentration of the inhibitor. The lower concentrations of the

inhibitors and the use of water as solvent also result –according to the author– in both a

safer and cheaper application of the treatment, and less alteration of the visual aspect of

the objects. On the other side, it was found that some combinations were ineffective, in

some cases even increasing the corrosion rate[53].

Inhibitors for Iron and its alloys

While inhibitors for iron and steel are widely used in industrial applications, their use

for conservation purposes is not so widely spread as in the case of copper alloys.

Interestingly, references to their use in the conservation of cultural heritage are earlier

than for copper. Plenderleith, in his book “The Conservation of Antiquities and Works of

Art: Treatment, Repair and Restoration” mentioned the use of commercial rust cleaning

products with the additional advantage of working as corrosion inhibitors[73].

Evidences have been found of the use of products including chromate based inhibitors

for iron conservation at the beginning of the XX century [25].

The most popular corrosion inhibitors for conservation of iron and steel heritage,

nowadays, are tannins [5, 67]. The name tannin is applied to a wide group of

polyphenolic compounds extracted from leaves, bark or fruits from different plants,

with an exact composition depending on their origin. The main advantages of tannins as

corrosion inhibitors are that, being natural extract from plants, they are non-toxic,

inexpensive and can be applied on pre-rusted metal, without the need of a cleanun-

corroded surface [74]. Tannins protect the metal by formation of a conversion layer of

iron (III) tannate complex which protects the metal, but produces a significant and, in

many cases, unacceptable change in the colour of the rust from reddish to blue-black

[74, 75]. While in some cases very good results are reported[76] the efficiency attained

in other cases is not very high [74] and they can even act as corrosion accelerators for

bare steel in neutral solutions [77].

As it has already been mentioned, the application of OTH has also been studied for

protection of historic steel exposed to atmospheric conditions, obtaining good results

with pre-corroded coupons, especially when the formulation was applied by brushing

[64, 78]. PTS has also been tested for this application, showing also promising results in

tests using artificially corroded coupons [79]. However, the behaviour of these

protection systems in long-term and real object tests was not so good, and PTS and

OTH provided less protection than other protection systems studied, being only

recommendable as short term protection [9, 80].

Long-chain carboxylic acids and sodium carboxylate solutions have also been applied to

protect iron and steel for conservation purposes. Sodium decanoate provided a slightly

better protection than sodium heptanoate when applied on iron [67]. In the case of iron,

the carboxylation layer formed by immersion in sodium decanoate is very thin -

nanometric scale-, but even so it provides better protection than other traditional

inhibitors used for iron such as tannins or phosphates[81]. When the carboxylation

solutions (composed of decanoic acid with the addition of an oxidant) are used, thicker

layers can be obtained[14, 82], but their performance is controversial: while Hollner et

al. have obtained better results than with sodium decanoate[82, 83], Rapp et al. obtained

poor results [68].

As opposed to the main use of inhibitors as final protection layer for copper based

artefacts, inhibitors for iron are used in many other steps of the conservation process.

One of these applications is their use to arrest iron corrosion in desalinization processes

of archaeological iron artefacts. Amines are traditional inhibitors for iron [33, 84], and

some of this compounds have also been used for these purposes. Immersion of iron

objects in a 5% ethylenediamine (EN) solution has been used as part of the stabilization

treatment for archaeological iron. While effective in many cases, it was proven to

stimulate corrosion in some other cases due to its ability to form soluble iron (II)

complexes, what summed to its toxicity turned out to reduce its applicability [85, 86].

A major problem for conservation of waterlogged iron-wood composite objects is the

iron corrosion in the traditional impregnation treatment for wood using polyethylene

glycol (PEG) solutions [25]. To prevent such problem, some commercial corrosion

inhibitors have been proposed. Hostacor KS1 (a triethanolamine salt of an

arylsulphonamido carboxylic acid) has been studied as an additive to the PEG 400

solution to supress the corrosion of iron [87, 88]. It works by reacting with the dissolved

oxygen of the solution and forming a passivating layer on the iron surface [89].

Bobichon et al. demonstrated that in 20% PEG 400 solutions it was more efficient than

triethanolamine[90]. The application of this product faced one of the main challenges of

using commercial products: its production was discontinued and was substituted by

Hostacor IT (a triethanolamine salt of an acrylamido carboxylic acid). Consequently,

new studies were made to assess the applicability of the new product to the iron

corrosion inhibition in PEG solutions, both in laboratory tests using electrochemical

techniques [89] and in real objects treatments [91], showing that Hostacor IT is effective

in slowing down the corrosion of exposed iron in PEG solutions. Phosphates have also

been tested as inhibitors for PEG solutions:Gourbeyre et al. studied the layers formed

on iron samples exposed to a 20% PEG solution with Na2HPO4 added as a corrosion

inhibitor and found that, for concentrations below 5 × 10–3 M, the layer was composed

of iron oxides and phosphates, while for higher concentrations corrosion was inhibited

by a layer of PEG and phosphates resulting from segregation of the former one [92]. A

subsequent paper demonstrated that phosphates acted as an anodic inhibitor, since the

PEG/phosphate complex was preferentially deposited on metal’s anodic sites, inhibiting

the iron dissolution[93].

Inhibitors have also been tested as additives to traditional and innovative coatings for

iron protection. Under the PROMET project, several commercial corrosion inhibitors4

were tested as additives for traditional coatings (Paraloid B-72 and Rennaissance wax)

and innovative coatings (Poligen ES910095)[9]. These inhibitors were already in use for

protecting industrial objects but had never been tested in the conservation field. They

were tested on clean and pre-corroded steel analogues, using accelerated ageing in

climate chambers[94],electrochemical tests (Rp and EIS) [95], as well as in real objects

[96]. Results showed no clear improvement in the coating’s behaviour with the

corrosion inhibitor additives and, in some cases, the effect was clearly deleterious. The

4 The additives were commercial products by Cortec Corporation: M435 (a blend of triazoles), M370 (ammonium salt of tricarboxylic acid) and M109 (calcium sulphonate); and by Dow Chemical: Alkaterge-T (bis-oxazoline). 5Poligen ES91009 is a synthetic polyethylene wax manufactured by BASF presented as a dispersion in water (ready-to-use).

reasons for the failure of these inhibitors were not investigated, but authors considered

that might be related with the interaction of the additives with the coating itself,

affecting the crosslinking of the polymer or the wettingproperties of the binder emulsion

with respect to the substrate, hence impairing the barrier effect of the coating.

Industrial heritage conservation has a significant space for the application of corrosion

inhibitors. Most of this heritage includes a significant part made of iron or steel, and

conservators face unusual problems that can, in some cases, be solved by the application

of corrosion inhibitors. The requirements of industrial heritage are not the same of the

industrial machines: even when kept in operating conditions, these artefacts are only

occasionally operated, so the corrosion problems might be different. One example is the

conservation of collections of vehicles, in which corrosion of the hydraulic braking

systems can be a problem. While in operating vehicles the main reason for decay of the

braking fluid –and loss of its anticorrosion properties– is thermal cycling, in these

infrequently used systems the degradation is mainly attributed to water uptake by

hygroscopic braking fluids. Hedditch et al. studied the inhibition efficiency of different

commercial hydraulic braking fluids and the effect of the addition of dodecanedioic

acid, finding a good performance in different fluids up to 5% water content[97].

Commercial tannate based inhibitors were also tested for the conservation of a boiler

from an operational paddle steamer, using gravimetric and electrochemical techniques,

and it was showed that the addition of these inhibitors would reduce corrosion to

negligible rates even under the boat’s operation [76]. In this case, it was important to

evaluate the corrosion under the exact operation conditions, such as the specific water

used for the steam system.

Other metals

The use of corrosion inhibitors for silver conservation treatments is by far less usual

than for copper or iron. Some compounds such as morpholine, BTA, chlorophyl,

pyridine or cysteamine can be found in the literature as corrosion inhibitor for silver,

[98, 99]but they have not been studied in detail and their application is not widespread

in the conservation community .

Notoyaet al. studied the structure of the polymeric layer formed on Cu, Ag, Au, Cr, Ni,

Fe and Zn using time-of-flight secondary ion mass spectroscopy (ToF SIMS) of BTA-

pretreated metals [100]. They found that the corrosion inhibition in a NaCl solution was

closely related with the degree of polymerization, following the order

Cu>>Ag>>Zn>Ni, Fe. Comparing these results with the survey made by

Argyropoulos[5], it seems that the use of BTA for silver might make any sense, but not

for iron.

Self-assembled monolayers (SAMs) have in recent years attracted some attention as

corrosion protection systems, and some studies have been made on their application on

silver objects with conservation purposes. Burleigh et al. described the procedure to

produce a SAM of alkanethioles on silver by immersion, applicable to coins or

jewellery [101]. They found tetradecanethiol (C14) and hexadecanethiol to be the most

effective in preventing silver tarnishing. Evesque et al. studied the corrosion protection

using electrochemical techniques and electrochemical quartz crystal microbalance

(EQCM), concluding that the protective film was composed of an inner self-assembled

layer of one or two monolayers, plus an outer one of about 10 monolayers [102].

Bernard et al. compared an electrodeposited film of poly(amino-triazole) and a SAM of

hexadecane-thiol, and found the later to be more effective in preventing silver tarnishing

exposed to a sulphur containing solution[103]. Recently, Liang et al. have tested

octadecanethiol from aqueous solution (OSA), octadecanethiol from organic solvent

(OSO), phytic acid (IP6) and silicon tungstemic acid (STA) monolayers for protection of

silver coins, and found that OSA was the most effective treatment, which is especially

interesting for conservation treatments since it avoids the use of organic solvents[104].

The use of inhibitors for lead is even scarcer, since lead is usually considered to be

resistant to atmospheric corrosion. However, when exposed to acetic or formic acid

vapours it undergoes a catastrophic corrosion process [105, 106]. These pollutants can

reach very high concentrations in display cases and storage boxes in museums and

inside organs pipes; for that reason, lead protection treatments (including inhibitors) had

focused on acetic acid environments.

The use of BTA as corrosion inhibitor for lead was studied by Sharmaet al. [107]. BTA

has been reported to form a Pb-BTA complex, creating a layered structure

Pb-BTA/PbO/Pb capable of preventing lead corrosion, so it was proposed as an

inhibitor for lead objects and leaded bronzes. The proposed treatment involved

immersion or brushing with an aqueous solution of BTA neutralized with 1 g of pure

calcium carbonate to obtain a pH ~ 7.

Sankarapapavinasam et al. studied the inhibitive properties of hydrazine (Hy) and

substituted hydrazines (phenyl hydrazine (PHy), 2,4dinitrophenyl hydrazine (2,4-

DNPHy), 4-nitrobenzoyl hydrazine (4-NBHy), and tosyl hydrazine (THy)) in acetic

acid solutions. THy and 4-NBHy were found to be the most efficient inhibitors, and

their mechanism of inhibition was explained by the blockage of anodic dissolution sites

[108].

The most studied type of lead inhibitor in acetic acid environments has been the group

of carboxylates. A first paper by Rocca and Steinmetz studied carboxylates with

different carbon chain length (from C7 to C11) and found the best inhibition

performance for long-chain molecules, being attributable the protection to a metallic

soap formation on the lead surface [109]. Sodium decanoate was selected for the best

compromise between protection and solubility, and it was tested in atmospheric

conditions with presence of acetic acid vapours, yielding good results, as well as in real

heritage objects [67, 110].

The most comprehensive study of corrosion inhibitors for lead exposed to acetic acid

environments was made under the COLLAPSE project6, applying these protection

systems to lead alloy organ pipes [111]. Sodium dodecanoate and undecanoate were

tested and compared with thiourea, phosphatising and sulphatising treatments,

exhibiting the best efficiency of all tested treatments [112]. Carboxylates also performed

very well in comparison with Paraloid B72 and microcrystalline wax, but none of the

protective treatments was completely effective in the long term; therefore, these surface

protection systems were not recommended for the treatment of organ pipes, and only the

change in the environment (reducing the concentration of acetic acid vapours) was

considered to be a good system[111, 113].

Dowsett et al. studied the formation of lead decanoate layers using in-situ

spectroelectrochemical techniques[28]. They showed that there were significant

differences in the formation and protective properties of decanoate layers that strongly

depend on the solution’s preparation conditions, and that solution’s pH alone is not a

6 COLLAPSE “Corrosion of Lead and Lead-Tin Alloys of Organ Pipes in Europe” EC 5 FP: Energy, Environment and Sustainable Development. EVK4-CT-2002-00088

good indicator for the preparation of good quality layers. Electrochemical deposition of

lead dodecanoate has also been explored by De Wael et al., having the advantage of a

shorter treatment time [114]. A later work by same authors have demonstrated that

immersion treatments produce a better quality layer than electrochemically assisted

ones, especially if an electrochemical pre-reduction conditioning treatment is applied.

[115]

Regarding zinc (and also brass), the benefits of BTA as an inhibitor have been

demonstrated as well [116, 117]. Carboxylation treatments have also been tested on it,

giving very good results and turning out the longer the C- chain length, the better

corrosion resistant [118], as well as the application of PEG, in alkaline media, where

PEG400 had fair effective performance but the research on this field is still

developing[119].

CONCLUSIONS

Corrosion inhibitors are one of the different methods that conservation-restoration

professionals have available to protect and prolong the life of metallic cultural heritage.

The scientific literature on corrosion inhibitors is huge, but the vast majority of it deals

with fundamental studies of corrosion inhibition or industrial applications. Protection of

heritage metals has specific needs and requirements, therefore, scientific studies on

corrosion inhibitors for this application should address and follow these specificities.

Since the first papers dealing with the BTA application for copper in the late sixties of

last century, the number and quality of scientific studies on the application of corrosion

inhibitors for metallic heritage conservation have been increasing, especially in the last

15 years. Inhibitors for copper and their alloys, mainly bronze, have attracted most of

the attention; on the other hand, those for iron, silver and lead have been less studied.

While many alternatives have been sought, BTA is probably still the best inhibitor for

copper and its alloys. The combination of different inhibitors, in order to take advantage

of their synergetic effects seems a promising way to increase the protection while

reducing the dosage of inhibitors and, therefore, the possible health or environmental

risks. Most studies have pursued not only better inhibitors, but also safer (for the people

and environment) and easier to apply ones. In this respect, the use of natural plant

extracts as corrosion inhibitors is a very up-to-date trend. They are easily available and

are usually innocuous, but they have a complex nature, which added to the complexity

of the heritage metals, make in many cases difficult to understand the mechanisms of

protection or the reasons for their failure.

SAMs are probably the state-of-the-art system for silver, showing a good protection

against tarnishing in sulphur-containing environments, but they are not easy to apply so

their applicability in real life conservation practice is not simple.

Finally, it is worth noting the great success obtained with carboxylates treatments for

different metals (copper, iron and mainly lead), which also fulfil the reversibility

requirements of modern conservation-restoration ethics. Nevertheless, there is still a

vast field to research on this topic in the future.

REFERENCES

[1] IUPAC: 'Compendium of Chemical Terminology (The "Gold Book")', 2006; http://goldbook.iupac.org/C01351.html, Accessed 2012/04/09.

[2] ISO 8044:1999, 'Corrosion of metals and alloys -- Basic terms and definitions'.

[3] ICOM-CC: 'Terminology to characterize the conservation of tangible cultural heritage', Resolution adopted by the ICOM-CC membership at the 15th Triennial Conference, New Delhi, 2008.

[4] EN 15898:2011, 'Conservation of cultural property. Main general terms and definitions'.

[5] V. Argyropoulos, M. Giannoulaki, G. P. Michalakakos and A. Siatou: in 'Strategies for Saving our Cultural Heritage', (eds. V. Argyropoulos et al.), 166-170; 2007, Cairo, Athens, TEI of Athens.

[6] A. Balbo, C. Chiavari, C. Martini and C. Monticelli: Corrosion Science, 2012, 59, (0), 204-212.

[7] F. Mansfeld, T. Smith and E. P. Parry: Corrosion, 1971, 28, (7), 289-294.

[8] T. Kosec, H. O. Ćurković and A. Legat: Electrochimica Acta, 2010, 56, (2), 722-731.

[9] C. Degrigny: in 'Metals and Museums in the Mediterranean. Protecting, Preserving and Interpreting', (ed. V. Argyropoulos), 179-235; 2008, Athens, TEI of Athens-PROMET project.

[10] R. B. Faltermeier: Stud. Conserv., 1999, 44, (2), 121-128.

[11] K. Marušić, H. Otmačić-Ćurković, S. Horvat-Kurbegović, H. Takenouti and E. Stupnišek-Lisac: Electrochimica Acta, 2009, 54, (27), 7106-7113.

[12] L. Muresan et al.: Electrochimica Acta, 2007, 52, (27 SPEC. ISS.), 7770-7779.

[13] E. Stupnišek-Lisac, H. Otmačić, K. Tadić, A. D. Manee and H. Takenouti, Vol. 2, 31-42; 2007, Denver, CO.

[14] S. Hollner, F. Mirambet, A. Texier, E. Rocca and J. Steinmetz: in 'Strategies for Saving our Cultural Heritage', (eds. V. Argyropoulos et al.), 156-161; 2007, Cairo, Athens, TEI of Athens.

[15] L. B. Brostoff: in 'Metal 95. Proceedings of the International Conference on Metals Conservation', (eds. I. D. MacLeod et al.), 99-108; 1995, Semur-en-Auxois, France, James and James.

[16] K. Rahmouni, H. Takenouti, N. Hajjaji, A. Srhiri and L. Robbiola: Electrochim. Acta, 2009, 54, (22), 5206-5215.

[17] J. C. D'Ars F. Jr, V. M. De Bellis, V. F. C. Lins and L. A. C. Souza: Studies in Conservation, 2007, 52, (2), 147-153.

[18] N. Hajjaji et al.: in 'Strategies for Saving our Cultural Heritage', (eds. V. Argyropoulos et al.), 138-142; 2007, Cairo, Athens, TEI of Athens.

[19] W. A. Mohamed, N. M. Rateb and A. A. Shakour: in 'Metal 04: Proceedings of the International Conference on Metals Conservation', (eds. J. Ashton et al.), 369-378; 2004, Canberra, Australia, National Museum of Australia.

[20] J. M. Bastidas and E. Otero: Materials and Corrosion - Werkstoffe und Korrosion, 1996, 47, (6), 333-337.

[21] E. Otero and J. M. Bastidas: Materials and Corrosion - Werkstoffe und Korrosion, 1996, 47, (3), 133-138.

[22] D. M. Bastidas, E. Cano and E. M. Mora: Anti-Corrosion Methods and Materials, 2005, 52, (2), 71-77.

[23] Y. M. Ding, C. C. Xu and J. L. Wang: Beijing Huagong Daxue Xuebao (Ziran Kexueban)/Journal of Beijing University of Chemical Technology (Natural Science Edition), 2006, 33, (2), 68-71.

[24] J. B. Cotton and I. R. Scholes: British Corrosion Journal, 1967, 2, (1), 1-5.

[25] D. A. Scott and G. Eggert: 'Iron and steel in art: Corrosion, colorants, conservation'; 2009, London, Archetype Publications.

[26] S. Eibl and D. Reiner: Materials and Corrosion, 2011, 62, (8), 745-752.

[27] E. Cano, D. Lafuente and D. M. Bastidas: J. Solid State Electrochem., 2010, 14, (3), 381-391.

[28] M. Dowsett, A. Adriaens, B. Schotte, G. Jones and L. Bouchenoire: Surface and Interface Analysis, 2009, 41, (7), 565-572.

[29] J. Tétreault: 'Airborne Pollutants in Museums, Galleries, and Archives: Risk Assessment, Control Strategies and Preservation Management.'; 2003, Ottawa, Canada., Canadian Conservation Institute.

[30] D. M. Bastidas, P. P. Gómez and E. Cano: Revista de Metalurgia (Madrid), 2005, 41, (2), 98-106.

[31] J. M. Bastidas, P. Pinilla, E. Cano, J. L. Polo and S. Miguel: Corros. Sci., 2002, 45, (2), 427-449.

[32] J. M. Bastidas, P. Pinilla, J. L. Polo and E. Cano: Corrosion, 2002, 58, (11), 922-931.

[33] J. M. Bastidas, J. L. Polo and E. Cano: J. Appl. Electrochem., 2000, 30, (10), 1173-1177.

[34] E. Cano, J. L. Polo, A. L. A. Iglesia and J. M. Bastidas: Adsorption, 2004, 10, (3), 219-225.

[35] J. L. Polo, P. Pinilla, E. Cano and J. M. Bastidas: Corrosion, 2003, 59, (5), 414-423.

[36] G. Gece: Corrosion Science, 2008, 50, (11), 2981-2992.

[37] N. Khalil: Electrochimica Acta, 2003, 48, (18), 2635-2640.

[38] B. S. Fang, C. G. Olson and D. W. Lynch: Surface Science, 1986, 176, (3), 476-490.

[39] T. Hashemi and C. A. Hogarth: Electrochimica Acta, 1988, 33, (8), 1123-1127.

[40] M. Metikoš-Huković, R. Babić and A. Marinović: Journal of the Electrochemical Society, 1998, 145, (12), 4045-4051.

[41] J. O. Nilsson, C. Törnkvist and B. Liedberg: Applied Surface Science, 1989, 37, (3), 306-326.

[42] E. Cano, D. M. Bastidas, J. Simancas, J. A. López-Caballero and J. M. Bastidas: Adsorption Science and Technology, 2004, 22, (7), 565-582.

[43] M. Finšgar and I. Milošev: Corrosion Science, 2010, 52, (9), 2737-2749.

[44] A. Galtayries, A. Mongiatti, P. Marcus and C. Chiavari: in 'Corrosion of metallic heritage artefacts', (eds. P. Dillman et al.), Vol. , 335-351; 2007, Cambridge, England, Woodhead Publishing.

[45] V. S. Sastri: 'Corrosion Inhibitors. Principles and applications'; 1998, New York, Wiley.

[46] V. S. Sastri: 'Green Corrosion Inhibitors: Theory and Practice', Wiley Series in Corrosion; 2011, Wiley.

[47] I. Dugdale and J. B. Cotton: Corrosion Science, 1963, 3, (2), 69-74.

[48] H. B. Madsen: Studies in Conservation, 1967, 12, 163-167.

[49] S. Angelucci, P. Fiorentino, J. Kosinkova and M. Marabelli: Studies in Conservation, 1978, 23, 147-156.

[50] D. A. Scott: 'Copper and Bronze in Art. Corrosion, colorants, conservation'; 2002, Los Angeles, Getty Conservation Institute

[51] C. Sease: Studies in Conservation, 1978, 23, 76-85.

[52] B. Blanco, S. Díaz and E. Willke: in 'MetalEspaña '08', (eds. J. Barrio et al.), 153-160; 2008, Madrid.

[53] S. Golfomitsou and J. F. Merkel: in 'Metal 04: Proceedings of the International Conference on Metals Conservation', (eds. J. Ashton et al.), 344-368; 2004, Canberra, Australia, National Museum of Australia.

[54] G. Brunoro, A. Frignani, A. Colledan and C. Chiavari: Corros. Sci., 2003, 45, (10), 2219-2231.

[55] V. C. Sharma, U. Shankar Lal and T. Singh: Studies in Conservation, 2003, 48, (3), 203-209.

[56] W. A. Oddy: Studies in Conservation, 1974, 19, 188-189.

[57] C. A. Harris et al.: Environmental Toxicology and Chemistry, 2007, 26, (11), 2367-2372.

[58] M. C. Ganorkar, V. Pandit Rao, P. Gayathri and T. A. Sreenivasa Rao: Studies in Conservation, 1988, 33, 97-101.

[59] R. B. Faltermeier: Studies in Conservation, 1999, 44, 121-128.

[60] B. Trachli, M. Keddam, H. Takenouti and A. Srhiri: Corrosion Science, 2002, 44, (5), 997-1008.

[61] A. Dermaj et al.: Electrochimica Acta, 2007, 52, (14), 4654-4662.

[62] A. Dermaj et al.: in 'Metal 07: Interim meeting of the ICOM-CC Metal Working Group', (eds. C. Degrigny et al.), Vol. 5, 119; 2007, Amsterdam, Rijksmuseum Amsterdam.

[63] P. B. Raja and M. G. Sethuraman: Materials Letters, 2008, 62, (1), 113-116.

[64] H. Hammouch, A. Dermaj, M. Goursa, N. Hajjaji and A. Srhiri: in 'Strategies for Saving our Cultural Heritage', (eds. V. Argyropoulos et al.), 149-155; 2007, Cairo, Athens, TEI of Athens.

[65] G. Bertrand et al.: Journal of Electroanalytical Chemistry, 2000, 489, (1–2), 38-45.

[66] E. Rocca, G. Bertrand, C. Rapin and J. C. Labrune: Journal of Electroanalytical Chemistry, 2001, 503, (1-2), 133-140.

[67] E. Rocca and F. Mirambet: in 'Corrosion of metallic heritage artefacts', (eds. P. Dillman et al.), Vol. , 308-334; 2007, Cambridge, England, Woodhead Publishing.

[68] G. Rapp, C. Degrigny, F. Mirambet, S. Ramseyer and A. Tarchini: in 'Metal 2010. Proceedings of the Interim Meeting of the ICOM-CC Metal WG', (eds. P. Mardikian et al.), 185-192; 2010, Charleston, S.C., U.S.A., Clemson University.

[69] A. Elia, M. Dowsett and A. Adriaens: in 'Metal 2010. Proceedings of the Interim Meeting of the ICOM-CC Metal WG', (eds. P. Mardikian et al.), 193-200; 2010, Charleston, S.C., U.S.A., Clemson University.

[70] A. Elia, K. De Wael, M. Dowsett and A. Adriaens: Journal of Solid State Electrochemistry, 2012, 16, (1), 143-148.

[71] S. Golfomitsou: 'Synergistic effects of additives to benzotriazole solutions applied as corrosion inhibitors to archaeological copper and copper alloy artefacts', PhD Thesis, UCL, 2006.

[72] S. Golfomitsou and J. F. Merkel: in 'Metal 07: Interim meeting of the ICOM-CC Metal Working Group', (eds. C. Degrigny et al.), Vol. 5, 38-43; 2007, Amsterdam, Rijksmuseum Amsterdam.

[73] H. J. Plenderleith and A. E. A. Werner: 'The Conservation of Antiquities and Works of Art: Treatment, Repair and Restoration'; 1956, Oxford, Oxford University Press.

[74] A. A. Rahim, E. Rocca, J. Steinmetz and M. Jain Kassim: Corrosion Science, 2008, 50, (6), 1546-1550.

[75] L. S. Selwyn: 'Metals and Corrosion: A handbook for the conservation professional', 223; 2004, Ottawa, Canada, Canadian Conservation Institute.

[76] A. Greiner, D. Hallam, D. Thurrowgood and D. Creagh: in 'Metal 07: Interim meeting of the ICOM-CC Metal Working Group', (eds. C. Degrigny et al.), Vol. 5, 51-55; 2007, Amsterdam, Rijksmuseum Amsterdam.

[77] S. Martinez and I. Stern: Applied Surface Science, 2002, 199, (1–4), 83-89.

[78] H. Hammouch, A. Dermaj, N. Hajjaji, C. Degrigny and A. Srhiri: in 'Metal 07: Interim meeting of the ICOM-CC Metal Working Group', (eds. C. Degrigny et al.), Vol. 5, 56-63; 2007, Amsterdam, Rijksmuseum Amsterdam.

[79] A. Dermaj et al.: in 'Strategies for Saving our Cultural Heritage', (eds. V. Argyropoulos et al.), 143-148; 2007, Cairo, Athens, TEI of Athens.

[80] C. Degrigny et al.: in 'Metal 07: Interim meeting of the ICOM-CC Metal Working Group', (eds. C. Degrigny et al.), Vol. 5, 31-37; 2007, Amsterdam, Rijksmuseum Amsterdam.

[81] S. Hollner, F. Mirambet, E. Rocca and J. Steinmetz: in 'Metal 07: Interim meeting of the ICOM-CC Metal Working Group', (eds. C. Degrigny et al.), Vol. 5, 64-70; 2007, Amsterdam, Rijksmuseum Amsterdam.

[82] S. Hollner, F. Mirambet, E. Rocca and S. Reguer: Corrosion Engineering Science and Technology, 2010, 45, (5), 362-366.

[83] S. Hollner, F. Mirambet, E. Rocca and S. Reguer: in 'Metal 2010. Proceedings of the Interim Meeting of the ICOM-CC Metal WG', (eds. P. Mardikian et al.), 160-166; 2010, Charleston, S.C., U.S.A., Clemson University.

[84] J. M. Bastidas, J. L. Polo, E. Cano and C. L. Torres: J. Mater. Sci., 2000, 35, (11), 2637-2642.

[85] L. S. Selwyn: in 'Metal 04: Proceedings of the International Conference on Metals Conservation', (eds. J. Ashton et al.), 294-306; 2004, Canberra, Australia, National Museum of Australia.

[86] L. S. Selwyn and V. Argyropoulos: Studies in Conservation, 2005, 50, 81-100.

[87] M. R. Gilbert, D. W. Grattan and D. Rennie: in 'Conservation of Wet Wood and Metal: Proceedings of the ICOM Conservation Working Groups on Wet Organic Archaeological Materials and Metals, Fremantle 1987', (eds. I. D. MacLeod et al.), 265-270; 1989, Western Australian Museum.

[88] L. S. Selwyn, D. A. Rennie-Bisaillion and N. E. Binnie: Studies in Conservation, 1993, 38, 180-197.

[89] V. Argyropoulos, J.-J. Rameau, F. Dalard and C. Degrigny: Studies in Conservation, 1999, 44, 49-57.

[90] C. Bobichon, C. Degrigny, F. Dalard and Q. K. Tran: Stud. Conserv., 2000, 45, (3), 145-153.

[91] V. Argyropoulos, C. Degrigny and E. Guilminot: Studies in Conservation, 2000, 45, (4), 253-264.

[92] Y. Gourbeyre, E. Guilminot and F. Dalard: Journal of Materials Science, 2003, 38, (6), 1307-1313.

[93] E. Guilminot, F. Dalard and C. Degrigny: Corrosion, 2005, 61, (5), 437-443.

[94] A. Siatou, V. Argyropoulos, D. Charalambous, K. Polikreti and A. Kaminari: in 'Strategies for Saving our Cultural Heritage', (eds. V. Argyropoulos et al.), 115-120; 2007, Cairo, Athens, TEI of Athens.

[95] E. Cano et al.: J. Solid State Electrochem., 2010, 14, (3), 453-463.

[96] V. Argyropoulos et al.: in 'Metal 07: Interim meeting of the ICOM-CC Metal Working Group', (eds. C. Degrigny et al.), Vol. 5, 10-15; 2007, Amsterdam, Rijksmuseum Amsterdam.

[97] C. Hedditch, A. Greiner, D. Hallam and D. Thurrowgood: in 'Metal 2010. Proceedings of the Interim Meeting of the ICOM-CC Metal WG', (eds. P. Mardikian et al.), 152-159; 2010, Charleston, S.C., U.S.A., Clemson University.

[98] V. Costa: Reviews in Conservation, 2001, 2, 18-34.

[99] M. Marabelli: 'Conservazione e restauro dei metalli d'arte'; 2007, Rome, Italy, Bardi Editore.

[100] T. Notoya et al.: Journal of Corrosion Science and Engineering, 2003, 6.

[101] T. D. Burleigh, Y. Gu, G. Donahey, M. Vida and D. H. Waldeck: Corrosion, 2001, 57, (12), 1066-1074.

[102] M. Evesque, M. Keddam and H. Takenouti: Electrochimica Acta, 2004, 49, (17–18), 2937-2943.

[103] M. C. Bernard, E. Dauvergne, M. Evesque, M. Keddam and H. Takenouti: Corros. Sci., 2005, 47, (3 SPEC. ISS.), 663-679.

[104] C. H. Liang, C. J. Yang, N. B. Huang and B. Wu: Surface Engineering, 2011, 27, (3), 199-204.

[105] J. Tétreault et al.: Studies in Conservation, 2004, 48, (4), 237-250.

[106] J. Tétreault, J. Sirois and E. Stamatopoulou: Studies in Conservation, 1998, 43, (1), 17-32.

[107] V. C. Sharma, T. Singh and U. Shankar Lal: Studies in Conservation, 2003, 48, 203-209.

[108] S. Sankarapapavinasam, F. Pushpanaden and M. F. Ahmed: British Corrosion Journal, 1989, 24, (1), 39-42.

[109] E. Rocca and J. Steinmetz: Corrosion Science, 2001, 43, (5), 891-902.

[110] E. Rocca, C. Rapin and F. Mirambet: Corros. Sci., 2004, 46, (3), 653-665.

[111] C. Martini, C. Chiavari and D. Prandstraller: in 'The Collapse project. Corrosion of organ pipes - causes and recomendations', (ed. J. Speerstra); 2011, Gothenburg, European Commision.

[112] C. Chiavari, C. Martini, G. Poli and D. Prandstraller: in 'Metal 04: Proceedings of the International Conference on Metals Conservation', (eds. J. Ashton et al.), 281-293; 2004, Canberra, Australia, National Museum of Australia.

[113] A. Aslund et al.: in 'Metal 07: Interim meeting of the ICOM-CC Metal Working Group', (eds. C. Degrigny et al.), Vol. 5, 16-22; 2007, Amsterdam, Rijksmuseum Amsterdam.

[114] K. De Wael et al.: Journal of Solid State Electrochemistry, 2010, 14, (3), 407-413.

[115] M. De Keersmaecker, K. De Wael and A. Adriaens: Progress in Organic Coatings, 2012, 74, (1), 1-7.

[116] K. Wang, H. W. Pickering and K. G. Weil: Journal of the Electrochemical Society, 2003, 150, (4), B176-B180.

[117] V. Sirtori, F. Zambon and L. Lombardi: Journal of Electronic Materials, 2000, 29, (4), 463-467.

[118] J. Peultier, E. Rocca and J. Steinmetz: Corrosion Science, 2003, 45, (8), 1703-1716.

[119] J. Dobryszycki and S. Biallozor: Corrosion Science, 2001, 43, (7), 1309-1319.

TABLE

Table 1: Summary of inhibitors used for conservation-restoration treatments of different

metals.

Inhibitor Metal/alloy References

Benzotriazole (BTA)

Cu [5, 7-10, 15, 19-21, 24, 30, 32, 34,

38-43, 47, 49-51, 53, 67, 72, 100]

Ag [5, 9, 99, 100]

Fe [5, 9, 100]

Bronze [6, 9, 16, 44, 48, 50, 52, 54, 55]

Lead [55]

Zn [100, 117]

2-aminopyrimidine Cu [10]

2-amino-5-mercapto-1,3,4-

thiadiazole (AMT)

Cu [10, 19-21, 50, 53, 58, 67, 72]

Bronze [17, 50, 52]

Benzylamine (BZA) Cu [53, 72]

Ethanolamine (ETH) Cu [53, 72]

1-Phenyl-5-Mercapto-Tetrazole

(PMT)

Cu [53, 72]

Potassium Ethyl Zanthate (KEX) Cu [53, 72]

Potassium Iodide (KI) Cu [53, 72]

5,6-dimethylbenzimidazole (DB) Cu [10, 67]

2-mercaptobenzimidazole (MBI) Cu [10, 67]

2-mercaptobenzoxazole (MBO) Cu [10, 67]

2-mercaptopyrimidine (MP) Cu [10, 67]

2-mercaptobenzothiazole (MBTS) Cu [10, 19, 67]

4-methyl-1-(p-tolyl) imidazole

(TMI)

Bronze [11, 12]

1-phenyl 4-methyl-imidazole (PMI) Bronze [12, 13]

2-mercapto 5-R-acetylamino-1,3,4-

thiadiazole (MAcT)

Bronze [12]

2-mercapto 5-R-amino-1,3,4-

thiadiazole (MAT)

Bronze [12]

1-(p-tolyl)-4-methylimidazole Bronze [8, 13]

Tributylamine Steel [84]

Ethylenediamine (EN) Fe [85, 86]

Hexylamine Steel [33]

Triphenylmethane derivatives Cu [31, 35]

Dodecylamine Steel [33]

Carboxylates

Fe [14, 25, 67, 68, 82]

Cu [14, 65-70, 82]

Zn [67, 118]

Lead [28, 67, 109-115]

3-phenyl 1,2,4-triazole-5 thione

(PTS)

Bronze [18, 61, 62]

Steel [79]

Amino-triazole (ATA) Bronze [16]

Poly-amino 1,2,4-triazole (pATA) Cu [60]

Bi-triazole (BiTA) Bronze [16]

VCI/VPI

Cu [22]

Bronze [23]

Ag [67]

Fe [22, 25]

Zn [22]

TanninsFe [5, 67, 75, 76]

Steel [74, 77]

Chromate based Fe [25]

Thiourea Fe [25]

Siderophores Fe [25]

PEGFe [87, 89, 90, 92, 93]

Zn [119]

Hostacor KS1 Fe [87, 89, 90]

(triethanolamine salt of an

arylsulphonamido carboxylic acid)

Hostacor IT

(triethanolamine salt of an

acrylamido carboxylic acid)

Fe [89, 91]

Phosphates Fe [25, 81, 92, 93]

Morpholine Ag [99]

Chlorophyl Ag [98]

Pyridine Ag [98]

Cysteamine Ag [98]

Opuntiaficusindica (OTH) extractBronze [64, 80]

Fe [64, 78, 80, 81]

Self-assembled monolayers (SAMs) Ag [101-104]

Hydrazine (Hy) Pb [108]

Phenyl hydrazine (PHy) Pb [108]

2,4 dinitrophenyl hydrazine (2,4-

DNPHy)

Pb [108]

4-nitrobenzoyl hydrazine (4-NBHy) Pb [108]

Tosyl hydrazine (THy) Pb [108]