Você está na página 1de 15

Corrosion Science 136 (2018) 91–105

Contents lists available at ScienceDirect

Corrosion Science
journal homepage: www.elsevier.com/locate/corsci

Inhibition of Cor-Ten steel corrosion by “green” extracts of Brassica T


campestris

Maria Pia Casalettoa, , Viviana Figàa, Antonella Priviteraa, Maurizio Brunob, Assunta Napolitanoc,
Sonia Piacentec
a
Istituto per lo Studio dei Materiali Nanostrutturati, Consiglio Nazionale delle Ricerche, via Ugo La Malfa n. 153, 90146 Palermo, Italy
b
Dipartimento di Scienze e Tecnologie Biologiche, Chimiche e Farmaceutiche, Università di Palermo, viale delle Scienze, Parco d’Orleans II, 90128 Palermo, Italy
c
Dipartimento di Farmacia, Università di Salerno, via Giovanni Paolo II n. 132, 84084 Fisciano, Salerno, Italy

A R T I C LE I N FO A B S T R A C T

Keywords: Extract of Brassica campestris was tested as potential corrosion inhibitor for Cor-Ten steel in NaCl and acidic
Weathering steel solutions, simulating a marine and an urban-industrial environment, respectively. Potentiodynamic polarization
Cor-Ten and electrochemical impedance spectroscopy (EIS) were performed both in absence and in presence of the
‘Green’ corrosion inhibitors extract at room temperature. The surface chemical analysis was investigated by X-ray Photoemission
Brassica campestris
Spectroscopy (XPS), before and after corrosion. Electrochemical results demonstrated that a very small con-
Marine corrosion
Acidic corrosion
centration of Brassica campestris extracts can inhibit Cor-Ten corrosion in NaCl solution (inhibition efficiency of
80–84%) better than in acidic solutions. Surface analysis results confirmed these good results.

1. Introduction et al. investigated a “green” formulation of Nigella sativa L. to protect


iron-based alloys, mimicking historical metallic artefacts, from acidic
Corrosion of metals and alloys is a worldwide big issue that espe- and NaCl corrosion and proved good performances, especially in the
cially influences industrial economy and also infrastructures, buildings, acidic medium [8,9].
ancient and modern artworks. Metallic materials naturally undergo The present paper investigated the corrosion inhibition of Cor-Ten
corrosion, corresponding to the spontaneous return of metals to their steel in adverse environments by means of natural extract of Brassica
ores, due to atmospheric interaction and exposure to aggressive en- campestris. Cor-Ten is a “weathering steel” (WS) widely used for infra-
vironments, such as acidic and NaCl solutions. Generally, metals pro- structures (highway bridges, soundproof and sound adsorbing screens,
tection can be performed by using different methods, such as corrosion buildings), for architectural cladding or urban finishing and for modern
inhibitors, anti-corrosion coatings or cathodic protection. and contemporary artworks, due to its technological and aesthetical
Unfortunately, commercial corrosion inhibitors and coatings are often features. Cor-Ten steel shows better mechanical properties in terms of
toxic, both for human health and for the environment [1]. In the last yield of strength and tensile strenght with respect to many construction
decades, the increasing interest for the environmental preservation steels. Its main characteristic is a good resistance to atmospheric cor-
oriented many research efforts towards eco-friendly corrosion in- rosion, due to the formation of a protective and uniform oxides layer on
hibitors. In particular, attention was focused on natural compounds, the bare steel surface that becomes more compact with exposure times
such as plants extracts with anti-oxidant activity toward different me- avoiding further corrosive processes and the need of painting [10]. The
tals and alloys. Several studies were performed on “green” corrosion rust layer gives also a pleasant aesthetical appearance to structures and
inhibitors as possible substitutes of synthetic products, since they are artefacts and Cor-Ten was initially developed for use in an unpainted
eco-friendly, non-toxic, cheap, easily available in formulations and re- condition. Despite these advantages, there are many limitations to its
newable [2–6]. They resulted efficient for the corrosion protection of employment, since both environmental and surface finish conditions
mild steel, magnesium and aluminium alloys especially in acidic can strongly affect resistance to corrosion [11]. The remarkable per-
medium. Recently, M’hiri et al. studied the effect of orange peel extracts formance of Cor-Ten steel needs some crucial pre-requisites to be sa-
on carbon steel corrosion in acidic medium, demonstrating that the tisfied: (i) regular variation of wet and dry cycles; (ii) low concentration
inhibitive effect was due to both the chemical composition of the ex- of air-borne salts, normally predominant in marine or coastal zones or
tract and the formation of a thin film on the steel surface [7]. Chellouli in cold regions using de-icing salts; (iii) low concentration of


Corresponding author at: Consiglio Nazionale delle Ricerche, Istituto per lo Studio dei Materiali Nanostrutturati, Via Ugo La Malfa 153, Palermo, Italy.
E-mail address: mariapia.casaletto@cnr.it (M.P. Casaletto).

https://doi.org/10.1016/j.corsci.2018.02.059
Received 3 August 2017; Received in revised form 22 February 2018; Accepted 26 February 2018
Available online 27 February 2018
0010-938X/ © 2018 Elsevier Ltd. All rights reserved.
M.P. Casaletto et al. Corrosion Science 136 (2018) 91–105

atmospheric pollutants; (iv) surface orientation and construction de- oxidants exhibited less corrosion resistance than the bare alloy [22].
tails. Coastal areas and harsh-weather environments represent two Additional protection systems, either in the form of corrosion in-
adverse conditions for the use of unpainted Cor-Ten steel, despite its hibitor or as protective coatings, may be needed for broadening the
considerable mechanical properties. Furthermore, dry or wet climates range of applications of Cor-Ten steel and deeply exploiting its me-
or areas that remain permanently damp are not suitable for the for- chanical properties regardless the environmental conditions [1].
mation of the adherent and protective rust layer, since alternating cy- “Painted” Cor-Ten is widely used in Japan in coastal and industrial
cles of wet/dry conditions are requested. The most critical situations for areas for fabricating marine containers, high-speed passengers’ ships,
weathering steels were observed in chloride-rich environments [12], industrial equipment, etc. [23]. Furthermore, the possibility to use
where corrosion rates are similar to those of mild and structural steels corrosion inhibitors can extend the application of Cor-Ten also to in-
[13]. In these aggressive environmental conditions, the formation of the dustrial pipes, seawater-cooling circuits or on board freshwater/central
protective patina is drastically retarded and obstructed by the accu- cooling system circuits. In the literature, despite many works referred to
mulation of hygroscopic chloride ions [14,15]. According to American carbon and stainless steel, just few studies addressed corrosion in-
protocols regarding weathering steels, exposition to levels of chlor- hibitors for Cor-Ten steels [10,22,24–27]. In this work, the extract of
ides > 0.5 mg/100 cm2/day should be avoided [16] and this hinders Brassica campestris was investigated as a potential corrosion inhibitor of
the application of Cor-Ten steel in naval applications and in offshore Cor-Ten A steel in NaCl and acidic solutions, mimicking the aggressive
structures. Many research works were addressed to understanding the marine and urban polluted environments, respectively. Among the
corrosion phenomenon of weathering steels in marine zones in order to different species of natural extracts, the anticorrosion effect of Brassi-
predict the durability of artefacts in such an aggressive environment. cacee plants were tested only in the case of Brassica oleracea [28] and
Melchers proposed a model according to which long-term corrosion loss Brassica oleracea capitate [29]. They showed good anticorrosion activity
is governed not only by oxygen diffusion but also by microbiological both for steel and for zinc in acidic media. In particular, the extract of
activity [17]. Morcillo et al. focalized on the possibility of “designing” Brassica oleracea resulted a mixed-type corrosion inhibitor for pipeline
and developing WS with a better corrosion resistance in marine areas steel in 0.5 M H2SO4. Brassica campestris is a widespread plant in
by studying the effect of elemental components in the alloy and found Southern Europe, easily available and commonly cultivated. The choice
that WS containing Ni (3 wt%) and Cu (0.3 wt%) displayed lower of Brassica campestris derived by the fact that, to the best of our
thickness loss compared to conventional WS [18]. Hao et al. studied the knowledge, this variety of Brassicacee plants was never tested as nat-
evolution of atmospheric corrosion of a kind of weathering steel in ural corrosion inhibitor.
coastal-industrial environment and demonstrated that rust formation is In the present work, the corrosion inhibition was evaluated by
strongly affected by the coexistence of chlorides and sulphates [19]. In means of potentiodynamic polarization and electrochemical impedance
the next future corrosion by air-borne salt particles in Europe will gain spectroscopy. Experiments were performed at room temperature as a
growing concern, since atmospheric corrosion of metals are predicted function of concentration. The extract of Brassica campestris was tested
to be dominated by the effects of chloride deposition in coastal and in solution, since this is the common experimental condition used for a
near-coastal areas [20]. Due to the global warming, these effects are corrosion inhibitor. The worst conditions of Cor-Ten corrosion were
expected to be more pronounced in the future, since a fixed chloride used in this study, since tests were designed for immersion in the cor-
deposition will have a higher impact when the temperature increases. rosive media. This experimental condition can be also useful in the case
When installed in a maritime location, in regular contact with water or of inhibition of atmospheric corrosion. The corrosion behaviour of bare
permanently in a damp environment, additional corrosion protection steel was proven unsatisfactory in stagnant conditions and both the
for Cor-Ten steel becomes essential. corrosion rate and the metal release were found significantly (three
According to the chemical composition of the alloy, three different times) higher than in leaching conditions [11]. In this paper the surface
type of Cor-Ten steels (A, B and C) are commercially available with chemical composition of the Cor-Ten steel substrates was determined
different resistance to atmospheric corrosion and mechanical stresses. by X-ray Photoemission Spectroscopy, both before and after corrosion
In this work, Cor-Ten A alloy was investigated since its chemical in the absence and in the presence of Brassica campestris extract, applied
composition produces an atmospheric corrosion resistance 5–8 times as an additive in small concentrations in the corrosive media. This
higher than that of a common low carbon a steel alloy. Cor-Ten A was surface-sensitive analytical technique can provide extremely useful in-
used in a bare surface finishing as structural material for buildings and formation on the efficacy of protective and/or corrosion inhibiting
civil engineering, due to its very good mechanical and anticorrosion treatments.
properties. When it is installed outdoor in this form, without any in- The originality and novelty of this work concern the application of
tentional pre-patination, in the early years of exposure, the rust layer is “green” extracts of Brassica campestris as a new non-toxic and en-
not well stabilized yet [15] and the environmental impact could be vironmentally friendly inhibitor of Cor-Ten steel corrosion for the
greater. protection of outdoor infrastructures and buildings and for the con-
The corrosion behaviour of Cor-Ten alloy can also be affected by the servation of modern and contemporary artworks and industrial ar-
surface finish commonly used for various applications. Different surface chaeological artefacts. In the latter case, this work represents a first,
finish (bare/pre-patinated/pre-patinated and waxed) of Cor-Ten are preliminary step in the investigation of a possible application as pro-
typically adopted in architecture and in contemporary artworks, since tective coating on both bare and pre-patinated Cor-Ten surfaces.
both the high gloss light grey appearance of the bare metallic surface
and the wide range of colours of its protective patina are particularly 2. Materials and methods
appreciated by architects and artists. Recent studies addressing the
environmental impact of WS as a potential source for metal con- 2.1. Preparation and characterization of the extract of Brassica campestris
tamination in an urban coastal site showed that the surface pre-treat-
ment generating the aesthetically pleasing appearance for outdoor ap- The aerial parts of Brassica campestris L. were collected at the end of
plications resulted not beneficial for corrosion protection [11,21]. The March 2015 from plants at the full flowering stage near Alimena (PA,
pre-patination treatment, even in presence of a beeswax finishing Italy). After washing and air-drying, the dried aerial parts were pul-
coating, worsened WS performances in terms of metal dissolution and verized by using an electric blender. About 100 g of the dried sample
determined a more inhomogeneous behaviour of patinas along time was soaked in 99% ethanol in a 1000 mL volumetric flask. The flask was
[21]. A study on the composition and protective properties of WS ar- covered and allowed to stand for 48 h at room temperature. After fil-
tificial patinas for the conservation of contemporary outdoor sculptures tration, the removal of the excess of solvent was performed by using a
by Ramirez-Barat et al. showed that artificial treatments by different rotary evaporator. The dried residue was stored in an air-sealed

92
M.P. Casaletto et al. Corrosion Science 136 (2018) 91–105

analytical container. In our experiments, a stock solution of Brassica marine environment and labelled as Blank 1. An acidic solution mi-
campestris extract (Brax) was prepared by dissolving 15 mg of plant micking rainwater in an urban polluted atmosphere (Blank 2) was
extract in 5 mL of ethanol. prepared by mixing 0.2 g/L of Na2SO4 (≥99.0% anhydrous, Sigma-
An analytical approach based on high-performance liquid chroma- Aldrich), 0.2 g/L of NaHCO3 (≥99.0%, Fluka) and 0.2 g/L of NaCl
tography coupled to electrospray ionization and multiple-stage linear (≥99.5%, Fluka) in distilled water, acidified with H2SO4 (95–98%,
ion-trap and orbitrap high-resolution mass spectrometry (LC-ESI/ Sigma-Aldrich) to adjust pH to 3.6 [8]. The effects of these different
LTQOrbitrap/MS/MSn) was used for obtaining the metabolite profile of corrosive media on Cor-Ten steel were investigated both by electro-
the ethanol extract of Brassica campestris. Liquid chromatography cou- chemical and surface analysis techniques.
pled to electrospray ionization and mass spectrometry (LC-ESI/MS) is a
powerful tool for the structural characterization and analysis of com- 2.3. Electrochemical measurements
plex mixtures, such as crude and purified plant extracts. Liquid chro-
matography allows to separate complex mixtures and to elute meta- All the electrochemical measurements were performed at room
bolites selectively prior their analysis into the mass spectrometer. Here temperature in an aerated three electrodes cell, consisting of an Ag/
metabolites are converted into positive or negative ions ([M+H]+ or AgCl (3.0 M) reference electrode, a platinum foil counter electrode and
[M−H]−) by electrospray ionization (ESI) technique, analysed and a Cor-Ten steel sample as working electrode. The electrochemical tests
separated according to their mass-to-charge ratio (m/z), and finally were carried out by using as electrolyte a constant volume (20 mL) of
detected to yield a mass spectrum in which the results are displayed in Blank 1 and Blank 2 solutions with addition of different volumes of
the form of a plot of ion abundance versus m/z. By using high resolution Brax. Experiments were performed as a function of Brassica campestris
and high accuracy mass analysers (i.e. Orbitrap), both molecular mass extract concentration by using 15, 40, 80 and 120 ppm Brax. A
and molecular formula of each analysed metabolite can be measured. PGSTAT302N (Metrohm Autolab) potentiostat/galvanostat was used.
Moreover, further structure information can be provided by conven- The open circuit potential (OCP) was stabilized for 30 min before the
tional and multiple-stage tandem mass spectrometry (MS2 and MSn) by electrochemical tests. Potentiodynamic polarization curves were ac-
analysis of the characteristic fragmentation pattern produced by colli- quired at a scan rate of 1 mV/s in the potential range ± 150 mV vs OCP
sion-induced dissociation (CID) of the precursor [M+H]+ or [M-H]− in order to determine kinetic parameters and corrosion inhibition effi-
ions with neutral gas (i.e. helium or argon), yielding product ions ciency. Electrochemical Impedance Spectroscopy (EIS) analysis was
having a m/z value lower than precursor ion. The experimental protocol performed at OCP in the frequency range: 100 kHz ÷ 100 mHz with a
used in this work was based on a LC-ESIHRMS system consisting of an signal amplitude of 10 mV by using a Faraday cage in order to minimize
LTQ Orbitrap XL mass spectrometer with a quaternary Accela 600 the external electrical interferences. EIS spectra were fitted by using
pump and an Accela autosampler (ThermoScientific, San Jose, CA) Nova 1.11 software. The experiments were repeated in triplicate for
operating in negative ion mode. The Orbitrap mass analyser was cali- each Brax concentration.
brated according to the manufacturer’s directions by using a mixture of
caffeine, methionine-arginine-phenylalanine-alanine-acetate (MRFA), 2.4. Surface analysis by X-ray photoelectron spectroscopy (XPS)
and Ultramark 1621 in a solution of acetonitrile, methanol, and acetic
acid. Data were collected and analysed by using the software provided Photoemission spectra were acquired by using a VG Microtech
by the manufacturer. The separation was carried out on a Symmetry ESCA3000 Multilab spectrometer, equipped with a twin anode X-Ray
RP-18 (2.1 × 150 mm, 5 μm; Waters) column at a flow rate of source (Mg and Al) and a five channeltrons detection system. In our
0.2 mL ppm/min. The mobile phase consisted of a combination of A experiment a standard non-monochromatized Al Kα (hν = 1486.6 eV)
(0.1% formic acid in water, v/v) and B (0.1% formic acid in acetoni- excitation source was used. In the ultrahigh vacuum (UHV) chamber
trile, v/v). A linear gradient from 10 to 50% B (v/v) in 20 min was used. the base pressure was lower than 1 × 10−6 Pa during data collection
The following experimental conditions for the ESI source were and the hemispherical analyser operated in the CAE mode. A pass en-
adopted: sheath gas at 15 (arbitrary units); auxiliary gas at 5 (arbitrary ergy of 20 eV was used for acquisition of the single region corre-
units); source voltage at 3.5 kV; capillary temperature at 553 K; capil- sponding to the photoelectron peak and a pass energy of 50 eV for the
lary voltage at −48 V and tube lens at −176.47 V. The mass range registration of the survey spectrum. The binding energy (BE) scale was
extended from m/z = 200 to 800 Da with a resolution of 30000. The calibrated with Au 4f7/2 core level at 83.9 eV. The BE of C 1s peak
first and the second most intense ions were selected for the data-de- (285.1 eV) from the adventitious carbon, which ubiquitously resulted
pendent scan to offer their tandem mass (MS2) product ions with a from the vacuum system of our XPS spectrometer, was used for charge
normalization collision energy at 30%. Liquid chromatography coupled correction of all the samples. The accuracy of the energy measure
to multiple-stage tandem mass experiments (LC-ESI/MS/MSn with is ± 0.1 eV. Photoemission data were collected and processed by using
n = 3) experiments on selected product ions were carried out by the the VGX900 and XPSPEAK4.1 software, respectively. Data analysis was
same collision energy. performed by a nonlinear least square curve-fitting procedure with a
properly weighted sum of Lorentzian and Gaussian component curves
2.2. Steel substrates and corrosion solutions for the shape of Voight function, after Shirley background subtraction
according to Sherwood [30]. Surface relative atomic concentrations
Experiments were carried out by using Cor-Ten steel specimens were calculated by a standard quantification routine, including
(Edilsider Spa, Alcamo, TP, Italy) with the following chemical compo- Wagner’s energy dependence of attenuation length [31] and a standard
sition (wt%): Cr = 1.20%; Cu = 0.40%; C ≤ 0.15%; Mn ≤ 1.00%; set of VG Escalab sensitivity factors. The uncertainty on the atomic
Si ≤ 0.75%; Ni ≤ 0.65%; P ≤ 0.10%; S ≤ 0.03%; Fe balance. Prior to quantitative analysis is about ± 10%. The identification of photoelec-
each experiment, bare commercial substrates were mechanically po- tron signals and peak components was performed according to litera-
lished by metallography papers with decreasing granulometry (P80, ture reference database [32].
P120, P220, P500, P1200 SiC polishing grade), then degreased by
acetone (≥99.0%, Riedel de Haën) and ultra-sonicated in ethanol 2.5. Structural analysis by X-ray diffraction (XRD)
(≥99.8%, Fluka) for 10 min. All the specimens have rectangular
shapes. Samples with flat active areas between 0.10 and 0.16 cm2 were The composition of the corrosion products induced by NaCl and
used for the electrochemical measures and (12 × 10 × 2 mm) samples acidic solutions on Cor-Ten steel substrates was analysed by X-ray
for XPS analysis. An aerated aqueous 3 wt% NaCl (≥99.5%, Fluka) Diffraction (XRD) by using a BRUKER D5000 powder diffractometer,
solution at pH = 6.8 was used in order to simulate corrosive attacks in a equipped with a Ni-filtered radiation of the Cu anode (Kα

93
M.P. Casaletto et al. Corrosion Science 136 (2018) 91–105

Fig. 1. LC-ESI/MS (base peak intensity, negative mode) chromatogram of the ethanol extract of Brassica campestris.

λ = 1.5418 Å; 40KV and 35 mA). Diffraction patterns were registered previous reports [33,34]. Glucosinolates are a group of sulfur- and ni-
in the angular range 10° ≤ 2Θ ≥ 90°, using a step size of 0.05° and a trogen-containing secondary metabolites widely distributed in the
time step of 3 s. The identification of the crystalline phases was based Brassicaceae (syn. Cruciferae) family, especially in Brassica genus.
on the Inorganic Crystal Structure Database (ICSD, FIZ Karlsruhe, Glucosinolates exhibit a common chemical feature, corresponding to a
Germany). C6H11O5S (β-D-1-thioglucopyranose) moiety in the structure, and differ
for their aglycone moieties. They are, generally, classified as alkyl,
3. Results and discussion aliphatic, alkenyl, hydroxyalkenyl, aromatic, or indole derivatives [35].
The neutral loss of 242 Da, originating from the combined loss of sulfur
3.1. Chemical composition of the extract of Brassica campestris trioxide (SO3, 80 Da) and anhydroglucose group (C6H10O5, 162 Da), is
another common fragmentation pathway observed for the glucosino-
The LC-ESI/HRMS analysis of the ethanolic extract of Brassica lates [33]. Analogously, the evidence in LC-ESI/MS/MSn spectra of
campestris was carried out by combining the efficient separation cap- characteristic product ions at m/z = 315, 301, and 285, originated by
ability of HPLC with the great power of structural characterization and neutral loss of one or two hexose units, highlighted the presence of
high sensitivity of mass spectrometry for a sensitive and highly specific isorhamnetin, quercetin, and kaempferol derivatives (2, 4, 6, 8, 9, 11,
levels of metabolite identification. The relative LC-ESI/HRMS (base 12), the most ubiquitous subclasses of flavonoids found in Brassica
peak intensity, negative ion mode) chromatogram is reported in Fig. 1. vegetables [36]. Comparison of MSn data with those reported for re-
The careful analysis of LC-ESI/HRMS and LC-ESI/MS/MSn data allowed ference compounds in literature [36,37] allowed to hypothesize in most
to highlight the presence of flavonoids (2, 4, 6, 8, 9, 11, 12), glucosi- cases the positions and the types of glycosyl units on the basis of the
nolates (3, 7, 10, 14, 16), hydroxycinnamic acid derivatives (5, 13), product ions of the flavonol glycosides, as reported in Table 1.
hydroxy fatty acids (15, 17, 18), and sucrose (1). The retention times, The observation in LC-ESI/MS/MS spectra of compounds 5 and 13
HRMS molecular ions [M−H]− and diagnostic MS2 and MS3 product of product ions at m/z = 223 and 205, corresponding to whole or de-
ions of compounds 1-18 are listed in Table 1. Putative identifications hydrated sinapic acid ion, identified them as hydroxycinnamic acid
were obtained from tandem mass data analysis and confirmed by re- derivatives, well-known secondary metabolites occurring in Brassica
tention times, elution order, and high-resolution mass spectra com- genus. In particular, compound 13 was identified as a hydro-
pared with reference compounds reported in literature. xycinnamoylglycoside containing a gentiobiose unit, according to lit-
In particular, the observation in LC-ESI/MS/MS spectra of product erature data [36,37].
ions at mass-to-charge ratio (m/z) of 195, 241, 259, and 275 (see Finally, on the basis of the negative LC-ESI/HRMS and LC-ESI/MS/
Table 1) was considered indicative of the presence in the plant extract MS, the molecular formulae C16H28O5, C18H32O5 and C18H34O5 were
of glucosinolates 3, 7, 10, 14, 16 in Fig. 1 and this was consistent with assigned, respectively, to compounds 15, 17 and 18, which were

94
M.P. Casaletto et al. Corrosion Science 136 (2018) 91–105

Table 1
LC-ESI/HRMS/MSn data and putative identification of metabolites in Brassica campestris.

peak tR m/z Molecular Formula Δppm Major MS2 ions Major MS3 ions Tentative identification
(min) (m/z) (m/z)

1 1.69 341.1081 C12H22O11 0,827 179 sucrose


2 5.73 625.1387 C27H30O17 −1.817 463, 301 quercetin 3-O -glucoside-7-O -glucoside
3 5.77 372.0416 C11H19O9NS2 −0.158 275, 259, 195 gluconapin
4 6.62 609.1439 C27H30O16 −1.676 489, 447, 285 kaempferol-3-O-glucoside-7-O-glucoside
5 6.71 385.1127 C17H22O10 −0.450 223, 205, 153 sinapic acid hexoside
6 6.79 639.1547 C28H32O17 −1.261 519, 477, 315 357, 315, 314 isorhamnetin diglucoside
7 8.44 386.0572 C12H21O9NS2 −0.515 275, 259, 195, 144 glucobrassicanapin
8 9.15 609.1436 C27H30O16 −2.333 429, 285 kaempferol -7-O-sophoroside
9 10.17 463.0864 C21H20O12 −1.517 301, 300 quercetin glucoside
10 10.70 447.0519 C16H20O9N2S2 −1.517 275, 259 glucobrassicin
11 11.28 447.0914 C21H20O11 −1.628 327, 285, 284, 255 kaempferol-7-O-glucoside
12 11.41 477.1018 C22H22O12 −1.850 357, 314 isorhamnetin-3-O-glucoside
13 11.46 753.2218 C34H42O19 −2.383 529, 223 disinapoyl-gentiobiose
14 12.12 422.0582 C15H21O9NS2 1.993 275, 259, 180, 139 gluconasturtiin
15 13.23 299.1861 C16H28O5 2.606 281, 263, 245, 201, 183, 143 7,10,11-trihydroxyhexadeca-8,13-dienoic acid
16 14.83 477.0623 C17H22O10N2S2 −1.913 446 285, 259 neoglucobrassicin
17 16.74 327.2167 C18H32O5 0.548 309, 291, 273, 229, 211, 171 9,12,13-trihydroxyoctadeca-10,15-dienoic acid
18 17.94 329.2324 C18H34O5 0.545 311, 293, 275, 229, 211, 171 9,12,13-trihydroxyoctadec-10-enoic acid

identified as hydroxy fatty acids, metabolites belonging to the oxylipin


class, already reported in Brassica spp. [36,38]. In particular, LC-ESI/
MS/MS spectra of 17 and 18 showed three main peaks originated by
neutral loss of three water molecules (18 Da), indicative of the presence
of three hydroxyl groups. Moreover, the observation in their tandem
mass spectra of three product ions at m/z = 229, 211 and 171 sug-
gested a common fragmentation pathway that, according to previous
reports, could be tentatively explained locating the hydroxyl groups at
C9, C12, and C13 positions and a double bond at C10 [39]. The mass
difference of 2 Da between 17 and 18 also pointed to indicate an ad-
ditional double bond in 17 as the only structural difference between the
two compounds. This second double bond could be likely located at C15
position, according to the observation that the common product ion at
m/z = 229 originated by the neutral loss of 100 Da from the molecular
ion at m/z = 329 (18) and by the neutral loss of 98 Da from the mo-
lecular ion at m/z = 327 (17) [39,40]. According to these considera-
tions, they were tentatively identified as 9, 12, 13-trihydroxyoctadeca-
10,15-dienoic acid (17) and 9,12,13-trihydroxyoctadec-10-enoic acid
(18), metabolites already reported in Brassica genus [40].
In conclusion, the analysis by LC-ESI/HRMS and LC-ESI/MS/MSn of
the ethanol extract of Brassica campestris allowed ascertaining the pre-
sence of eighteen metabolites mainly representative of flavonoid, glu-
cosinolate, hydroxy fatty acid and hydroxycinnamic acid derivatives. In
particular, to the best of our knowledge, this is the first report of
compounds 5, 8,17 and 18 in Brassica campestris, being compound 5
previously reported only in Brassica oleracea and compounds 8, 17 and
18 in Brassica napus.

3.2. Electrochemical measurements

3.2.1. Open circuit potential variation


Before potentiodynamic polarization and EIS measurements, the
open circuit potential was monitored during the spontaneous corrosion Fig. 2. Variation of the open-circuit potential of Cor-Ten steel in a) NaCl and b) acidic
of the alloy for 30 min, in order to determine the corrosion domain and solutions at different concentrations of Brassica campestris at 293 K.
to define the partial or complete inhibition at different volumes of Brax.
Fig. 2a displays the OCP variation in the time domain in the case of 3 wt
anodic direction was observed in the case of 80 ppm of Brassica cam-
% NaCl solutions (Blank 1) in the absence and in the presence of dif-
pestris extract. Fig. 2b shows the OCP of Cor-Ten steel in acidic solution
ferent concentration of Brassica campestris. The immersion potential
(Blank 2), both in absence and in presence of different volumes of Brax.
was affected by the solution composition. A more negative value was
The immersion potential is more cathodic in the presence of Brax with
found in the case of Cor-Ten steel in Blank 1 solution and it became
respect to the pure Blank 2 solution. OCP decreased with time sug-
more anodic by increasing concentrations of Brassica campestris extract.
gesting that no passivation phenomena occurred. After 30 min of sta-
However, the OCP values decreased with time suggesting the dissolu-
bilization, both the solution with 80 and 120 ppm Brax reached the
tion of the pre-formed oxide films. By comparing the corrosion poten-
most anodic OCP value. In the latter case a weak OCP variation vs time
tial (Ecorr) at different experimental conditions, a shift toward the

95
M.P. Casaletto et al. Corrosion Science 136 (2018) 91–105

and ib are the slopes of the best fitting anodic and cathodic straight
lines; Ecorr,calc is the calculated corrosion potential extrapolated from
the intercept on Tafel plot; Ecorr,obs is the corrosion potential observed
in the experimental plot; icorr is the corrosion current density extra-
polated from the intercept on Tafel plot; IE% ± σ is the average value
and standard deviation of the inhibition efficiency. The inhibition ef-
ficiency was calculated using the following equation (Eq. (1)):
icorr ,0 − icorr , inh
IE (%) = × 100
icorr ,0 (1)

where icorr,0 and icorr,inh are the corrosion current densities in ab-
sence (Blank 1 or Blank 2) and in presence of Brax, respectively.
Corrosion rate (C.R.) was estimated by means of Eq. (2):
e
C .R.=3.27 × icorr ×
d (2)
2
where icorr is the corrosion current density in mA/cm ; e is the
equivalent weight of iron; d is the iron density and 3.27 is a pro-
portionality constant.
In NaCl solution, the fitting parameters revealed a decrease of cor-
rosion current density from 48.0 μA/cm2 in the case of Blank 1 solution
to 9.6 μA/cm2 in presence of 80 ppm of Brassica campestris extract, as
shown in Table 2. By adding 120 ppm Brax, icorr is slightly higher than
in the case of 80 ppm. Despite the small difference of IE% values ob-
tained at different Brax concentrations, data extrapolated from po-
tentiodynamic polarization measurements at 80 ppm Brax showed the
smallest standard deviation. Consequently, the highest value of corro-
sion inhibition efficiency IE ∼ 80% indicates that 80 ppm represents
the ‘optimum’ concentration of Brassica campestris extract for the in-
hibition of Cor-Ten corrosion in 3 wt% NaCl aerated solution in the
investigated range.
As shown in Table 3, the corrosion current density increases by
adding Brax to the acidic solution, except in the case of 80 and
120 ppm, suggesting that in the presence of very low concentration of
Brassica campestris in the acidic environment corrosion of the substrate
Fig. 3. Potentiodynamic polarization curves of Cor-Ten steel in a) NaCl and b) acidic
solutions at different concentrations of Brassica campestris, after ohmic drop compensa-
occurs instead of protection of the alloy surface. The concentration of
tion. 80 ppm Brax had a negligible inhibiting effect corresponding to a very
low inhibition efficiency (∼ 10%). In the range of Brax concentration
investigated in this work, the highest inhibition efficiency was calcu-
was observed with respect to the other solutions, probably due to a
lated for 120 ppm Brax and corresponded to 44%. The addition of
slower dissolution of the native oxide film.
higher Brax concentrations up to 300 ppm did not improve the inhibi-
tion efficiency (results not displayed) and changed the colour of the
3.2.2. Potentiodynamic polarization measurements acidic solution, making it darker.
Dynamic polarization measurements were performed in order to
calculate the kinetic parameters of cathodic and anodic reactions in- 3.2.3. Electrochemical impedance spectroscopy
volving Cor-Ten steel in the corrosive media as a function of different Electrochemical impedance spectra were recorded after the open
concentrations of Brassica campestris extract. The explored potential circuit potential was stabilized for 30 min. In Fig. 4 EIS spectra are
range was Ecorr ± 150 mV starting from the cathodic direction. Fig. 3a reported in NaCl (a) and acidic (b) media, respectively.
shows the dynamic polarization plots in the different chlorides-based Fig. 4a displays the Nyquist diagrams of Cor-Ten steel in 3 wt% NaCl
test solutions, after ohmic drop compensation. The addition of Brax solutions. The plot recorded in absence of the inhibitor (Blank 1) was
volumes shifted the corrosion potential (Ecorr) toward anodic values characterised by lower impedance values than those in the presence of
with respect to Blank 1 solution. The anti-corrosion effect of the green Brax. By adding different Brax concentrations, an increase of im-
inhibitor was observed only in the anodic branch with a decrease of the pedance was recorded, thus demonstrating inhibition effects towards
current density. This result suggested that Brax behaves like an anodic Cor-Ten steel corrosion in NaCl solution. All the EIS spectra recorded in
inhibitor in the presence of chlorine. The same conclusion can be drawn Blank 1 solution and in presence of Brax displayed only one time
in the presence of acidic solution. Fig. 3b displays the dynamic polar- constant, observed from Bode plots (here not represented).
ization plots of the acidic solution (Blank 2) with different Brax con- In order to evaluate the corrosion inhibition efficiency of Brassica
centrations. A current increase in the cathodic branch and a current campestris extract, the EIS spectra of Cor-Ten in NaCl solution were
decrease in the anodic one were recorded as a function of Brax con- fitted by using the electrical circuit represented in Fig. 5a.
centration. The cathodic current weakly increased in the case of Rs represents the solution resistance; Rct is the charge transfer re-
120 ppm. sistance, which is referred to the electron transfer across the electrode/
Electrochemical kinetics parameters associated to the corrosion electrolyte interface [41] and Qdl is the constant phase element, which
processes were estimated from Tafel extrapolation of the dynamic po- represents the distorted double layer capacitance, due to the roughness
larization curves. Results are summarised in Tables 2 and 3, for NaCl of the mechanically treated Cor-Ten. The best fitting parameters are
and acidic solutions, respectively. In the heading of the tables: OCP is summarised in Table 4. IE% ± σ is the average value and standard
the open circuit potential obtained from the open circuit transients; ia deviation of the inhibition efficiency.

96
M.P. Casaletto et al. Corrosion Science 136 (2018) 91–105

Table 2
Kinetic parameters determined from polarization curves in NaCl solution.

Brax Conc. OCP ba bc Ecorr,calc Ecorr,obs icorr Corr. rate IE ± σ


(ppm) (V/AgAgCl) (V/dec) (V/dec) (V/AgAgCl) (V/AgAgCl) (μA/cm2) (mm/year) (%)

0 −0.624 0.24 0.08 −0.70 −0.70 48.0 0.57 –


15 −0.625 0.63 0.12 −0.60 −0.61 16.0 0.15 69.6 ± 3.8
40 −0.638 0.40 0.06 −0.58 −0.60 10.0 0.09 75.3 ± 3.7
80 −0.582 0.48 0.08 −0.58 −0.58 9.6 0.08 79.3 ± 0.9
120 −0.631 0.36 0.07 −0.58 −0.60 11.0 0.10 68.6 ± 7.4

The inhibition efficiency of Brassica campestris extract as a function


of the concentration added to the test solution was calculated by using
the following equation (Eq. (3)):
R ct , inh − R ct ,0
IE (%) = × 100
R ct , inh (3)

where Rct,inh represents the charge transfer resistance in presence of


the inhibitor and Rct,0 is the charge transfer resistance in blank solution.
A value of IE standard deviation ≤3% was found for all the different
Brax concentrations. As reported in Table 4, a concentration of 80 ppm
of Brassica campestris extract showed the most efficient inhibitive effect
of Cor-Ten steel corrosion in NaCl solution. The highest inhibition ef-
ficiency value (IE ∼ 80%) was calculated in the case of 80 ppm Brax in
all the investigated range of concentration. The addition of 120 ppm
induced a decrease in the charge transfer resistance. Despite the con-
centration of Brassica campestris extract, the constant phase element
represents a distorted double layer capacitance. The non-ideal beha-
viour of the capacitor is probably due to the working electrode
roughness, produced by the mechanical cleaning. EIS results are in good
agreement with the kinetic data derived from the dynamic polarization
curves. They confirmed that the best efficiency of Cor-Ten corrosion
inhibition in a 3 wt% NaCl solution in the investigated range was ob-
tained in presence of a concentration of 80 ppm Brax.
In the case of acidic solutions, EIS spectra were recorded in the same
range of Brax concentration (15–120 ppm) as for the NaCl solutions and
a weak inhibition effect was found only in the presence of 120 ppm
Brax. For sake of clarity, electrochemical impedance spectra of Cor-Ten
steel in acidic solution are reported in Fig. 4b in absence (Blank 2) and
in presence of the best performing Brax concentration (120 ppm). By
adding 120 ppm of Brassica campestris extract, an increase in Rct was
detected, confirming the kinetic data previously reported. Impedance
spectra were fitted by using the electrical circuit displayed in Fig. 5b,
where Rs is the solution resistance; Rf is the resistance of the rust layer;
Qf is the constant phase element of the rust layer [42]; Rct is the charge
transfer resistance and Qdl is the constant phase element of the double Fig. 4. Nyquist plots recorded for Cor-Ten steel in a) NaCl and b) acidic solutions at
layer. In the case of acidic solution, two time constants were observed, 293 K.
at variance with NaCl solution. This suggests that the corrosion process
is different in the two different environments, being controlled by
standard deviation of the inhibition efficiency.
charge transfer in NaCl solution and influenced by the rust layer for-
A very low inhibition efficiency (31%) was calculated in the case of
mation in the acidic medium (Fig. 5b). This interpretation is further
acidic solutions containing 120 ppm of Brassica campestris extract, as
confirmed by the results of XPS surface analysis of Cor-Ten steel cor-
reported in Table 5. These data are in good agreement with kinetic
rosion in the two different media (vide infra). The electrical parameters
results. The different values of IE% obtained by EIS and Potentiody-
obtained by fitting the experimental plots with the best fitting electrical
namic Polarization could be ascribed to different immersion times
circuit are summarised in Table 5. IE% ± σ is the average value and

Table 3
Kinetic parameters determined from polarization curves in acidic solution.

Brax Conc. OCP ba bc Ecorr,calc Ecorr,obs icorr Corr. rate IE ± σ


(ppm) (V/AgAgCl) (V/dec) (V/dec) (V/AgAgCl) (V/AgAgCl) (μA/cm2) (mm/year) (%)

0 −0.660 0.31 0.17 −0.708 −0.707 112.18 1.05 –


15 −0.660 0.35 0.21 −0.681 −0.682 169.33 1.58 –
40 −0.666 0.38 0.19 −0.672 −0.673 148.54 1.39 –
80 −0.655 0.51 0.21 −0.664 −0.667 101.26 0.94 11.8 ± 2.2
120 −0.655 0.47 0.18 −0.672 −0.671 62.40 0.58 38.0 ± 6.0

97
M.P. Casaletto et al. Corrosion Science 136 (2018) 91–105

Table 6
XPS relative quantitative analysis of the investigated samples. Elemental concentration is
expressed as atomic percentage (at.%).

DESCRIPTION SAMPLE C 1s O 1s Fe 2p N 1s K 2p

Extract of Brassica campestris 80.2 16.3 – 3.1 0.4

Bare Cor-Ten substrate A 32.7 31.9 35.4


After 1 h corrosion by NaCl B 42.6 33.6 23.8
After 1 h corrosion by NaCl in C 25.7 46.1 27.4 0.8 –
presence of 80 ppm Brax
After 1 h corrosion by NaCl in D 36.8 37.3 24.6 1.3 –
presence of 120 ppm Brax

After 2 h corrosion by NaCl E 32.0 41.6 26.5


After 2 h corrosion by acidic F 32.9 39.2 27.8
solution

3.3. Surface characterization by XPS

X-ray Photoelectron Spectroscopy was used to investigate the sur-


face chemical composition of Cor-Ten specimens, both before and after
corrosion by NaCl solution. Bare Cor-Ten substrate was labelled as
sample A. The surface analysis was performed on the substrate sub-
mitted to corrosion in the presence of a concentration of 80 ppm of
Brassica campestris, according to the best corrosion inhibition revealed
by the electrochemical results. In addition, the effect of the highest
concentration of Brax (120 ppm) was investigated, along with a sample
of pure extract of Brassica campestris. Cor-Ten steel substrates were
immersed in 20 mL of NaCl solution (Blank 1), in the absence and in the
presence of Brassica campestris for one hour, following the same ex-
perimental conditions used for the electrochemical tests. Then, they
Fig. 5. Scheme of the equivalent circuit used to fit EIS data of Cor-Ten steel in a) NaCl and were rinsed with distilled water and air-dried overnight at room tem-
b) acidic solutions, in absence and in presence of different volumes of Brassica campestris
perature. Cor-Ten samples submitted to these treatments were labelled
extract.
B to D, as indicated in Table 6. Moreover, in order to investigate the
early stage of corrosion in pure NaCl and acidic media, bare Cor-Ten
Table 4 steel substrates, cleaned as above described, were immersed for two
Electrochemical data derived from EIS spectra of Cor-Ten steel in NaCl solution (Blank 1)
hours in 20 mL of Blank 1 and in 20 mL of Blank 2 solutions, respec-
at different concentrations of Brassica campestris extract at 293 K.
tively. Samples were, then, washed with distilled water, air-dried
Brax Conc. χ2 Rs Rct Qdl n IE ± σ overnight at room temperature and labelled as E and F, respectively.
(ppm) (Ω cm2) (Ω cm2) (μF/cm2) (%) The description of the investigated samples and their corresponding
surface chemical composition, derived by XPS relative quantitative
0 0.030 4.72 300 408 0.86 –
15 0.012 3.16 1320 772 0.84 73.8 ± 3.0
analysis and expressed as atomic percentage (at.%), are summarised in
40 0.019 2.92 913 954 0.84 68.0 ± 1.4 Table 6.
80 0.012 4.40 1532 575 0.84 82.1 ± 2.0 The analysis of XPS spectra by a nonlinear least-square curve-fitting
120 0.035 3.64 1157 514 0.85 72.0 ± 1.7 procedure allows the identification of elemental species in different
oxidation states and chemical environments on the surface. Iron and
oxygen are the most significant elements on the surface of the in-
required by the experiments. Also in this case, Cor-Ten surface rough-
vestigated samples and their spectral curve fitting procedures could
ness affects the capacitances which are represented by constant phase
help to elucidate quantitative chemical state information. The identi-
elements.
fication of iron surface chemical state speciation by XPS is a very
The different performances of the addition of Brassica campestris to
challenging task, due to the complexity of Fe 2p spectra resulting from
NaCl and acidic corrosion observed by electrochemical measurements
peak asymmetries, complex multiplet splitting, uncertain and over-
can be explained by the different behaviour of a bare Cor-Ten steel
lapping binding energies, shake-up and plasmon loss structure. In the
substrate submitted to these corrosive media. Results of the surface
case of mixed oxide and oxy-hydroxide systems, the curve fitting of Fe
analysis of Cor-Ten steel in NaCl and acidic solutions are discussed in
2p spectra becomes quite complicated for the identification of iron
the following section.
species due to spectral overlaps [43]. The examination of the spectral
curve fitting of the oxygen peak can facilitate the interpretation of XPS
data, giving useful information on the distribution of oxygen species on

Table 5
Electrochemical data derived from EIS spectra of Cor-Ten steel in acidic solution (Blank 2) and in presence of 120 ppm Brassica campestris extract at 293 K.

Brax Conc. χ2 Rs Qf n Rf Qdl n Rct IE ± σ


(ppm) (Ω cm2) (nF/cm2) (Ω cm2) (μF/cm2) (kΩ cm2) (%)

0 0.046 6.93 35.4 0.98 166 842 0.92 1.15 –


120 0.009 7.50 72.0 0.93 181 760 0.89 1.66 27.0 ± 4.0

98
M.P. Casaletto et al. Corrosion Science 136 (2018) 91–105

Fig. 6. XPS curve-fitting of O 1s spectra in the investigated samples.

Table 7 the surface of the investigated samples.


XPS surface distribution of oxygen species in all the investigated samples, as determined
by the curve-fitting of O 1s photoelectron signal (total peak area = 100%). – Curve-fitting of O 1s spectra
BE of O 1s component 529.8 eV 531.2 eV 532.3 eV
Assignment FeO, Fe2O3, Fe3O4 Fe(OH)3, FeOOH Adsorbed H2O The curve-fitting of the O 1s photoelectron spectra of all the in-
vestigated samples (A to D) are reported in Fig. 6.
Sample A 54.6 23.5 21.9 As shown in Fig. 6, XPS O 1s spectra can be deconvoluted by using
Sample B 54.1 32.4 13.4
Sample C 63.6 29.7 6.7
three different components. The first component located at
Sample D 52.8 43.1 4.1 BE = 529.8 eV is assigned to the presence of oxide O2−ions, due to the
presence of iron oxides on the surface (FeO, Fe2O3, Fe3O4). A second
Sample E 52.7 36.1 11.2 component at BE = 531.2 eV is attributed to the presence of hydroxyl
Sample F 61.2 32.6 6.2
groups (OH−) and FeOOH species on the surface. A third component of
the peak, located at BE = 532.3 eV, can be assigned to adsorbed H2O in
the oxide (Fe2O3·nH2O) layer related to the non-uniform porosity

99
M.P. Casaletto et al. Corrosion Science 136 (2018) 91–105

Fig. 7. XPS curve-fitting of Fe 2p3/2 spectra in the investigated samples.

[32,44–47]. The surface distribution of the different oxygen species in Table 8


the investigated samples is reported in Table 7, as resulted from the O XPS surface distribution of iron species in all the investigated samples resulting from the
curve-fitting of Fe 2p3/2 spectra (total peak area = 100%).
1s curve-fittings shown in Fig. 6. Results of these data processing are
expressed as peak area percentage (total photoelectron signal Surface Fe species Fe0 FeO, Fe2O3, Fe3O4 FeOOH Asymmetric envelope
area = 100%).
Sample A 48.4 27.2 19.1 5.4
Sample B 7.8 42.7 38.6 10.9
– Curve-fitting of Fe 2p3/2 spectra
Sample C 11.8 49.5 25.0 13.7
Sample D 9.6 52.1 28.2 10.1
The curve-fitting of the Fe 2p3/2 spectra of all the samples in-
vestigated by XPS (A to D) is reported in Fig. 7 and the results of these Sample E – 15.9 74.3 9.8
data processing are listed in Table 8. Sample F – 39.3 41.2 19.5

As shown in Fig. 7, XPS Fe 2p3/2 spectra can be deconvoluted by

100
M.P. Casaletto et al. Corrosion Science 136 (2018) 91–105

using different components. The peak component at BE = 706.4 eV is


assigned to metallic Fe0 and the component at BE = 707.7 eV can be
attributed to metallic iron strongly interacting with the oxide
[32,48,49]. The peak component at BE = 708.2 eV can be ascribed to
the presence of Fe3O4 oxides; the peak at BE = 709.5 eV is assigned to
the presence of FeO and the component located at BE = 710.0 eV to
Fe2O3 oxides; the component at BE = 711.3 eV can be attributed to iron
oxy-hydroxides as in FeO(OH) [32,48,50]. The highest binding energy
peak around BE ∼ 713 eV originates from the peak asymmetry ob-
served in the Fe oxide band envelope [51].
Thereinafter, the results of XPS analysis will be discussed by con-
sidering separately the two sets of data concerning: a) the corrosion in
NaCl and in acidic solutions, in order to investigate the effect of the
pure corrosive medium on the surface; b) the corrosion in NaCl solution
in the presence of Brassica campestris for supporting the electrochemical
results.

3.3.1. Surface analysis of Cor-Ten steel in pure NaCl and acidic solutions
XPS analysis provided useful information on the very first stage of
the corrosion of Cor-Ten substrates resulting from the two different
environments. The relative quantitative results shown in Table 6 in-
dicated that NaCl and acidic attacks of the Cor-Ten substrate (sample E
and F) induces an increase in the oxygen content and a decrease in the
iron amount on the surface with respect to sample A, due to the for-
mation of corrosion products. These effects are slightly prominent in
the case of steel corrosion by NaCl solution. In such a case, a slightly
higher amount of oxygen and a lower iron content are detected on the
surface of sample E with respect to corrosion in the acidic medium
(sample F). Further differences on the surface chemical composition of
samples submitted to different corrosive solutions can be inferred from
the analysis of XPS spectra by a nonlinear least-square curve-fitting
Fig. 8. Optical images of the Cor-Ten steel submitted to corrosion in a) NaCl, sample E
procedure. The curve-fitting of O 1s spectra of Cor-Ten before (sample
and b) acidic solutions, sample F (Dino-Lite Premier digital microscope, magnifica-
A) and after corrosion in NaCl solution and acidic solutions (sample E tion = 50x; non polarized light).
and F) is reported in Fig. 6(a), (e) and (f), respectively. After corrosion
in the acidic solution, a greater amount of iron lattice oxides species
– Morphology of the corrosion products in pure NaCl and in acidic
was detected on the surface of sample F, as reported by the surface
solutions.
distribution of oxygen species in Table 7. The formation of iron oxides
could be induced by the aerial oxidation of the intermediate complexes
The surface morphology and the structural characterization of Cor-
of iron (II) and (III) [52]. The amount of adsorbed water resulted higher
Ten substrates corroded by pure NaCl and acidic solutions were further
in the case of NaCl solution corrosion (sample E) than for the acidic one
investigated by using optical microscopy and X-ray diffraction. These
(sample F). As reported in Table 7, a very similar distribution of oxygen
supplementary series of surface and bulk analysis were performed in
species was determined on the surface of sample B and E, as expected
order to support the finding of a rust layer contribution in the EIS re-
by the very short difference in the time of immersion in NaCl solutions
sponse in the case of the acidic solution and to give a comprehensive
(one and two hours, respectively).
description of the effect of the corrosive media on the bare Cor-Ten steel
The curve fitting of Fe 2p3/2 spectra of Cor-Ten before (sample A)
substrates.
and after corrosion in NaCl solution (sample E) and acidic (sample F)
The optical images of sample E and F reported in Fig. 8 revealed the
solutions are shown in Fig. 7(a), (e) and (f), respectively and the cor-
presence of a non-homogeneous layer of corrosion products, irregularly
responding results are listed in Table 8. The Fe 2p3/2 spectrum of
covering the surface of the underlying alloy in both cases. A different
sample A revealed the predominant presence of metallic iron on the
surface morphology of the patinas was found. A localized, ulcer/crater
surface of the bare Cor-Ten sample. As reported in Table 8, the highest
shaped corrosion was determined on the surface of sample E (Fig. 8a)
surface concentration of metallic iron was detected in sample A, as
and a localized, pitting shaped corrosion on the surface of sample F
expected for the bare Cor-Ten steel sample. The presence of Fe2+ and
(Fig. 8b) [53].
Fe3+ ions was ascertained on the surface of all the investigated samples
XRD analysis of samples E and F was performed in order to gain
in the mixed form of oxide and hydroxides species. By comparing the
further information about the presence of crystalline phases in the
effects of Cor-Ten corrosion in the two different aggressive environ-
corrosion products layer and the results are shown in Fig. 9. Both XRD
ments, a higher fraction of iron oxides was determined after acidic
patterns are dominated by the iron peaks of the main α-Fe phase of the
corrosion (sample F) than after NaCl attack (sample E), as shown in
underlying Cor-Ten alloy, since a low crystallinity of corrosion products
Table 8. A corroborating evidence of this finding can be gained from the
was found. The presence of γ-FeOOH (Lepidocrocite) was detected as
O 1s peak analysis shown in Table 7. The curve fitting of the O 1s
the only crystalline phase of iron compounds, both for sample E and
spectrum of sample F confirmed that lattice oxide species were pre-
sample F. This result confirmed that lepidocrocite is the main corrosion
dominant on the surface with respect to the fraction of hydroxide-like
product detected on both bare and pre-patinated specimens in the early
species detected (Fig. 6f). This result is in agreement with the electrical
stages of exposure [16,54]. Lepidocrocite proved to be not a protective
circuit used to best fitting the EIS spectra in acidic solution (see Fig. 5b).
phase for the corrosion of weathering steel [22]. Therefore, the feasi-
Furthermore, the presence of iron oxy-hydroxides FeOOH species is
bility of protection systems (corrosion inhibitors or protective coatings)
more prominent on the Cor-Ten surface after NaCl corrosion (sample E)
may be an urgent need both for bare infrastructures exposed to
than that after the acidic attack (sample F).

101
M.P. Casaletto et al. Corrosion Science 136 (2018) 91–105

of samples B, C and D after the corrosion by the NaCl solution. Results


reported in Table 8 showed that on the surface of all the samples cor-
roded by NaCl (samples B, C and D) a significant decrease of metallic
iron and an increase in the amount of both oxides and hydroxide species
occurred with respect to sample A, thus confirming the corrosion at-
tack. After corrosion in NaCl, the fraction of metallic Fe0 drastically
decreased on the surface of sample B with respect to that of the bare
Cor-Ten steel (sample A) and resulted totally absent after a longer im-
mersion time in the corrosive medium (sample E, Fig. 7e). The addition
of 80 ppm Brassica campestris extract produced an increase of the con-
centration of metallic iron and a decrease in the oxy-hydroxides species
on the surface of samples C with respect to sample B, as shown in
Table 8. These evidences confirmed the protective effect towards the
corrosion by NaCl solution. In the case of 120 ppm Brax, a smaller
decrease of metallic iron and an increase of oxy-hydroxides species with
respect to sample C were found on the surface of sample D (see Table 8).
This latter finding is also confirmed by the spectral analysis of O 1s peak
of sample D, as shown in Fig. 6d. These results suggested that, in our
Fig. 9. XRD patterns of sample E and sample F. experimental conditions, the addition of Brax concentration higher than
80 ppm to a NaCl solution did not improve iron corrosion protection, in
aggressive environments and for outdoor weathering steel artworks agreement with the electrochemical data.
with artificial patina. In conclusion, XPS analysis revealed a good performance of Brassica
campestris extract for the protection of the surface of Cor-Ten steel
against the corrosive attack of NaCl solution. The best performances of
3.3.2. Surface analysis of Cor-Ten steel in presence of Brassica campestris the addition of 80 ppm Brax resulted in the highest surface concentra-
extract tion of iron from the XPS relative quantitative analysis and in the
The XPS relative chemical composition reported in Table 6 showed highest fraction of surface metallic Fe0 iron and the lowest fraction of
that, after Cor-Ten corrosion in NaCl solution, a significant decrease in oxy-hydroxides FeO(OH) species, as derived from the O 1s and Fe 2p3/2
the iron concentration was detected on the surface of sample B with spectral analysis.
respect to that on sample A, resulting from the formation of corrosion
products. The presence of the extract of Brassica campestris on the sur- 4. Mechanism of inhibition
face of sample C and D was evidenced by the appearance of the N 1s
photoelectron signal, also detected by XPS in the sample of the pure The chemical composition of the extract of Brassica campestris
extract. A lower amount of iron was found on the surface of sample D proved to be a complex mixture of different metabolites, as described in
than on sample C. After the corrosion treatment in the NaCl solution in more detail above. The presence of such a numerous organic com-
the presence of 80 ppm Brax, a higher concentration of iron was found pounds makes it rather difficult to attempt to assign the observed cor-
on the surface of sample C than that in sample B in the absence of the rosion behaviour to a particular constituent, but still a possible inter-
corrosion inhibitor. From the relative quantitative analysis reported in pretation of our results may be traced out.
Table 6, the presence of 80 ppm Brax in the corrosive solution seems to In the early stage of immersion, the formation of a passive layer
induce a beneficial effect to the surface of the Cor-Ten steel, which was occurred according to the mechanism of polimerization of ferric hy-
less depleted in the iron content. An increase in the surface con- droxo-complex by hydrolysis, olation and oxolation processes [55] and
centration of oxygen species was also detected for sample C. These to the precipitation of amorphous ferric oxy-hydroxide layer [48,56].
evidences could be attributed to the precipitation of insoluble com- This passive layer, whose formation is favoured in acid solution, acts
plexes on the surface, resulting from the interaction of iron ions and the like a self-protective barrier, characterised by abundant water mole-
organic molecules of the Brax extract, as discussed below. After the cules stabilized on the surface of the alloy. In our experimental condi-
corrosion treatment in the NaCl solution in the presence of 120 ppm tions, it did not allow the possibility to other inhibitor agents to show
Brax, a smaller concentration of iron was found on the surface of further protective effect, because the layer resulted adherent and rela-
sample D with respect of that of sample C. The surface chemical com- tively stable, as indicated by the EIS spectral fitting (Fig. 5b). However,
position of sample D was very similar to that of sample B, with the only the resistance of this layer to the pitting formation strongly depends on
exception of nitrogen resulting from the composition of Brassica ex- the presence of some ions in the solution. In NaCl solutions, Cl− ions
tracts. The possible formation of highly soluble organo-metallic com- can replace some water molecules in the amorphous layer and lead to
plexes at higher concentration of Brax that induces the dissolution of the formation of soluble chloride/metal complexes that subtract metal
the metal from the Cor-Ten surface, similarly to the action of chlorides ions from the protective layer by bringing them into solution. A new
ions, can be advanced as discussed in the next paragraph. self-regenerating protective layer immediately replaces the empty site
The curve-fittings of O 1s spectra of samples B, C and D are shown in thus created and the process unevenly advances over the entire surface,
Fig. 6b–d, respectively. The corresponding surface distribution of until chlorine ions are still available, leading to a progressive con-
oxygen species is reported in Table 7. The main difference between Cor- sumption of the alloy [57].
Ten surfaces exposed to the presence of 80 and 120 ppm Brax in NaCl As reported in literature by Okafor et al., the presence of the in-
solutions consisted in a different distribution of oxides and hydroxides hibitor molecules in solution could cause a delay in the anodic dis-
species. The highest concentration of oxides species was detected on the solution of the iron, according to the following mechanism, involving
surface of sample C and the highest amount of hydroxides and FeOOH two adsorbed intermediates, described by Eq. (4) [58]:
species was found on the surface of sample D. This finding confirmed
Fe+H2 O⇌ Fe·H2 Oads (4a)
the better protective action of 80 ppm than 120 ppm Brax.
The Fe 2p3/2 curve-fittings of samples B, C and D are shown in _
Fe·H2 Oads + B⇌ FeOHads + H+ + B (4b)
Fig. 7b–d, respectively and the corresponding results are listed in
Table 8. Very similar Fe 2p spectra were recorded by XPS on the surface Fe·H2 Oads + B⇌ FeBads + H2 O (4c)

102
M.P. Casaletto et al. Corrosion Science 136 (2018) 91–105

Fig. 10. Model of formation of the amorphous ferric oxy-hydroxides layer in pure (a) acidic and (b) NaCl solutions. Possible effects of the addition of 80 ppm Brax (c) and 120 ppm Brax
(d) to NaCl solutions.

_
FeOHads → FeOHads + e− (4d) against chlorides ions for the replacement of some water molecules and
lead to the precipitation of stable and insoluble products on the surface,
FeBads ⇌ FeB+ads + e− (4e) preventing any further progress of corrosion. A similar possible me-
FeOHads+FeB+ads ⇌ FeBads+FeOH+ (4f) chanism was claimed for the inhibition of C-steel and nickel corrosion
by using Lawsonia extract [59]. The addition of higher concentration of
FeOH+ + H+ ⇌ Fe2 + + H2 O (4g) Brassica campestris could lead to the formation of highly soluble organo-
metallic complexes, which induce dissolution of the metal similarly to
where B indicates the inhibitor species.
the chlorides ions. This could explain the worse corrosion behaviour
The replacement/displacement of some water molecules adsorbed
detected in the case of the addition of 120 ppm Brax to the NaCl solu-
on the surface of Cor-Ten steel by the inhibitor species produces a new
tion. The larger number of different chemical compounds may form
adsorbed intermediate FeBads, as in Eq. (4c). Such absorbed inter-
adsorbed intermediates, which may catalyse further metal dissolution
mediate reduces the amount of FeOH− ads species available for the rate-
by means of a chelating action. A direct correlation between the total
determining step in Eq. (4d) [58]. Thus, these two processes are com-
content of phenolics, flavonoids and glucosinolates of a variety of
peting. Depending on the relative solubility, these intermediate organo-
Brassicacea (Brassica olaracea) and the iron chelating activity was al-
metallic complexes could slow down or catalyse the dissolution of the
ready ascertained in the literature [60]. From the observed results it can
metal [58].
be inferred that the insoluble complexes may dominate the adsorbed
The formation of insoluble complex compounds resulting from the
intermediates in the case of 80 ppm of Brax and thus a positive corro-
combination of the metal cations and the organic molecules of the ex-
sion inhibitive effect occurred. At higher Brax concentration (120 ppm),
tract of Brassica campestris adsorbed on the surface is a probable in-
soluble intermediates became predominant in the solution and the re-
terpretation of the inhibition action observed in the case of addition of
moval of iron ions from the surface resulted in a detrimental effect.
80 ppm. In such a case, the molecules of the inhibitor can compete

103
M.P. Casaletto et al. Corrosion Science 136 (2018) 91–105

This hypothesis about the interaction of the extract of Brassica benign Karanj (Pongamia pinnata) seed extract on corrosion of mild steel in hy-
campestris with the Cor-Ten surface in a NaCl solution can be sustained drochloric acid solution, J. Solid State Electrochem. 15 (2011) 1087–1097.
[4] O.K. Abiola, A.O. James, The effects of Aloe vera extract on corrosion and kinetics
by the XPS analysis of the nitrogen photoelectron signal (spectra not of corrosion process of zinc in HCl solution, Corros. Sci. 52 (2010) 661–664.
shown). In the case of the pure extract of Brassica campestris, the N 1s [5] C.A. Loto, O.O. Joseph, R.T. Loto, A.P.I. Popoola, Corrosion inhibitive behaviour of
signal consisted of two components located at BE = 400.1 eV and camellia sinensis on aluminium alloy in H2SO4, Int. J. Electrochem. Sci. 9 (2014)
1221–1231.
401.6 eV, which can be assigned, respectively, to the presence of eN]C [6] N. Chaubey, V.K. Singh, Savita, M.A. Quraishi, E.E. Ebenso, Corrosion inhibition of
bonds and ]NeO groups [48] in the nitrogen-containing organic aluminium alloy in alkaline media by neolamarkia cadamba bark extract as a green
components of Brax. The XPS analysis of the Cor-Ten surface in pre- inhibitor, Int. J. Electrochem. Sci. 10 (2015) 504–518.
[7] N. M’hiri, D. Veys-Renaux, E. Rocca, I. Ioannou, N. Mihoubi Boudhrioua, M. Ghoul,
sence of 80 ppm Brax (sample C) revealed that the N 1s photoelectron Corrosion inhibition of carbon steel in acidic medium by orange peel extract and its
signal was located at BE = 399.3 eV, significantly shifted towards lower main antioxidant compounds, Corros. Sci. 102 (2016) 55–62.
binding energy. This value can be attributed to the presence of iron [8] M. Chellouli, D. Chebabe, A. Dermaj, H. ErramL ppmi, N. Bettach, N. Hajjaji,
M.P. Casaletto, C. Cirrincione, A. Privitera, A. Srhiri, Corrosion inhibition of iron in
organo-metallic complexes on the surface [61]. In the case of addition
acidic solution by a green formulation derived from Nigella sativa L, Electrochim.
of 120 ppm Brax, the N 1s peak was detected on the surface of sample D Acta 204 (2016) 50–59.
as a single component at BE = 399.9 eV that can be assigned to the [9] M. Chellouli, N. Bettach, N. Hajjaji, A. Srhiri, P. Decaro, Application of a for-
presence ofeN]C bonds [48]. This peak corresponds to the first mulation based on oil extracted from the seeds of Nigella Sativa L, Inhibition of
Corrosion of Iron in 3% NaCl, IJERT 3 (4) (2014) 2489–2495.
component of the nitrogen signal found for the pure extract of Brassica [10] L. Damian, R. Fako, Weathering structural steels corrosion in atmospheres of var-
campestris. On the surface of sample D no evidences of ]NeO groups ious degrees of pollution in Romania, Mater. Corros. 51 (2000) 574–578.
were found as a consequence of the interaction of these components of [11] C. Chiavari, E. Bernardi, C. Martini, F. Passarini, A. Motori, M.C. Bignozzi,
Atmospheric corrosion of Cor-Ten steel with different surface finish: accelerated
Brax with iron ions that lead to the formation of soluble intermediates ageing and metal release, Mater. Chem. Phys. 136 (2012) 477–486.
removed from the surface. [12] D.C. Cook, Spectroscopic identification of protective and non-protective corrosion
A pictorial representation of the formation of the amorphous ferric coatings on steel structures in marine environments, Corros. Sci. 47 (2005)
2550–2570.
oxy-hydroxides layer in pure acidic and NaCl solutions is shown in [13] R.E. Melchers, The transition from marine immersion to coastal corrosion for
Fig. 10 (a) and (b). The possible mechanism of interaction of 80 ppm structural steels, Corrosion 63 (2007) 500–514.
and 120 ppm Brax in NaCl solutions is sketched in Fig. 10(c) and (d), [14] Y.S. Choi, J.G. Kim, Aqueous corrosion behaviour of weathering steel and carbon
steel in acid-chloride environment, Corrosion 56 (2000) 1202–1210.
respectively. [15] M. Morcillo, B. Chico, I. Diaz, H. Cano, D. de la Fuente, Atmospheric corrosion data
of weathering steel. A review, Corros. Sci. 77 (2013) 6–24.
5. Conclusions [16] P. Albrecht, A.H. Naeerni, Performance of Weathering Steel in Bridges, National
Cooperative Highway Research Program Report 272, Transportation Research
Board, National Research Council, Washington, DC, 1984 https://trid.trb.org/
In this work electrochemical (OCP, polarization and impedance view/211731 . (Accessed February 2018).
spectroscopy measurements) and surface XPS methods were used to [17] R.E. Melchers, A new interpretation of the corrosion loss processes of weathering
investigate the properties of the ethanolic extracts of Brassica campestris steels in marine atmospheres, Corros. Sci. 50 (2008) 3446–3454.
[18] M. Morcillo, I. Diaz, B. Chico, H. Cano, D. de la Fuente, Weathering steels: from
as a corrosion inhibitor of Cor-Ten steel in NaCl and acidic solutions. empirical development to scientific design. A review, Corros. Sci. 83 (2014) 6–31.
The potentiodynamic results obtained in NaCl solution showed that the [19] L. Hao, S. Zhang, J. Dong, W. Ke, Evolution of atmospheric corrosion of MnCuP
extract of Brassica campestris acts as an anodic inhibitor and the in- weathering steel in a simulated coastal-industrial atmosphere, Corros. Sci. 59
(2012) 270–276.
hibition efficiency increases with increasing Brax concentration. The [20] J. Tidblad, Atmospheric corrosion of metals in 2010–2039 and 2070–2099, Atmos.
best efficiency value of 80–84% was obtained for a concentration of Environ. 55 (2012) 1–6.
80 ppm Brax. The electrochemical impedance measurements confirmed [21] S. Raffo, I. Vassura, C. Chiavari, C. Martini, M.C. Bignozzi, F. Passarini, E. Bernardi,
Weathering steel as a potential source for metal contamination: metal dissolution
the results obtained by the potentiodynamic curves. The XPS surface during 3-year of field exposure in an urban coastal site, Environ. Pollut. 213 (2016)
analysis revealed a good performance of the addition of 80 ppm Brax to 571–574.
the NaCl solution for the protection of the surface of Cor-Ten steel [22] B. Ramírez-Barat, T. Palomar, B. García, D. de la Fuente, E. Cano, et al.,
Composition and protective properties of weathering steel artificial patinas for the
against the corrosive attack. The best performances of the addition of
conservation of contemporary outdoor sculpture, in: R. Menon (Ed.), Metal 2016
80 ppm Brax with respect of that of 120 ppm Brax was confirmed by Proceedings of the Interim Meeting of the ICOM-CC Metals Working Group, New
XPS, as already proven by the best corrosion inhibition derived by the Delhi India, September 26–30 2016, EBook, 2017, pp. 1–6.
[23] COR-TEN™. Nippon Steel & Metal Corporation Catalogue. http://www.nssmc.com/.
electrochemical results. Electrochemical experiments performed in
2008–2017. (Accessed February 2018).
acidic solutions showed a very weak inhibiting effect with the highest [24] J. Aramendia, L. Gomez-Nubla, K. Castro, I. Martinez-Arkarazo, D. Vega, A. Sanz
inhibition efficiency around 31% by addition of 120 ppm Brax. López de Heredia, A. García Ibáñez de Opakuab, J.M. Madariaga, Portable Raman
The results obtained in the investigated experimental conditions study on the conservation state of four CorTen steel-based sculptures by Eduardo
Chillida impacted by urban atmospheres, J. Raman Spectrosc. 43 (2012)
suggest that the extract of Brassica campestris displays much efficacy in 1111–1117.
neutral chlorides environments, encouraging a possible application as [25] H. Mansouri, S.A. Alavi, M. Fotovat, Microbial-influenced corrosion of Cor-Ten steel
Cor-Ten steel corrosion inhibitor in marine context. compared with carbon steel and stainless steel in oily wastewater by Pseudomonas
aeruginosa, JOM 67 (7) (2015) 1594–1600.
[26] A. Crespo, B. Ramírez Barat, E. Cano, Artificial patinas in contemporary weathering
Acknowledgments steel sculpture, Proceeding of the 5th International Conference Youth in
Conservation of Cultural Heritage- YOCOCU 2016, Madrid, 21–23 September,
2016.
This work was supported by the PON03PE_00214_1 TECLA Project: [27] P. Letardi, M. Ramírez-Barat, P. Traverso, E. Cano, Edith Joseph, et al., Copper
“Nanotecnologie e Nanomateriali per i Beni Culturali”, Distretto di Alta alloys and weathering steel used in outdoor monuments: weathering in an urban-
Tecnologia per l’Innovazione nel settore dei Beni Culturali della Marine environment, in: R. Menon (Ed.), Metal 2016 Proceedings of the Interim
Meeting of the ICOM-CC Metals Working Group, New Delhi India, September 26–30
Regione Sicilia. The authors thank Sig. B. Scalici and Dr. F.M. Giordano
2016, EBook, 2017, pp. 320–328.
for technical support. [28] N.C. Ngobiri, E.E. Oguzie, Y. Li, L. Liu, N.C. Oforka, O. Akaranta, Eco-friendly
corrosion inhibition of pipeline steel using brassica oleracea, Int. J. Corros. (2015)
404139.
References
[29] A.J. Chinweuba, Corrosion inhibition potentials of Brassica oleracea capitata ex-
tract on mild steel and zinc in acidic media, Chem. Mater. Res. 6 (3) (2014) 62–67.
[1] E. Almeida, D. Santos, F. Fragata, D. de la Fuente, M. Morcillo, Anticorrosive [30] P.M.A. Sherwood, D. Briggs, M.P. Seah (Eds.), Data Analysis in X-ray Photoelectron
painting for a wide spectrum of marine atmospheres: environmental-friendly versus Spectroscopy in Practical Surface Analysis by Auger and X-ray Photoelectron
traditional paint systems, Prog. Org. Coat. 57 (2006) 11–22. Spectroscopy, Wiley, New York, 1990p. 1.
[2] E.E. Oguzie, Corrosion inhibition of aluminium in acidic and alkaline media by [31] C.D. Wagner, L.E. Davis, W.M. Riggs, Surf. Interface Anal. 2 (1986) 5.
Sansevieria trifasciata extract, Corros. Sci. 49 (2007) 1527–1539. [32] C.D. Wagner, A.V. Naumkin, A. Kraut-Vass, J.W. Allison, C.J. Powell, J.R. Rumble
[3] A. Singh, I. Ahamad, V.K. Singh, M.A. Quraishi, Inhibition effect of environmentally Jr., NIST Standard Reference Database 20, Version 3.4, (2003) http://srdata.nist.

104
M.P. Casaletto et al. Corrosion Science 136 (2018) 91–105

gov/xps/ . (Accessed February 2018). [46] N.S. McIntyre, D.G. Zetaruk, X-ray photoelectron spectroscopic studies of iron
[33] J. Sun, M. Zhang, P. Chen, GLS-finder: a platform for fast profiling of glucosinolates oxides, Anal. Chem. 11 (49) (1977) 1521–1529.
in Brassica vegetables, J. Agric. Food Chem. 64 (2016) 4407–4415. [47] D.T. Harvey, R.W. Linton, Chemical characterization of hydrous ferric oxides by X-
[34] A.M. Ares, J. Bernal, M.J. Nozal, C. Turner, M. Plaza, Fast determination of intact Ray Photoelectron Spectroscopy, Anal. Chem. 53 (1981) 1684–1688.
glucosinolates in broccoli leaf by pressurized liquid extraction and ultra-high per- [48] J.F. Moulder, W.F. Stickle, P.E. Sobol, K.D. Bomben, J. Chastain, R.C. KingJr. (Eds.),
formance liquid chromatography coupled to quadrupole time-of-flight mass spec- Handbook of X-Ray Photoelectron Spectroscopy, Physical Electronics, Inc USA,
trometry, Food Res. Int. 76 (2015) 498–505. Eden Prairie, MN, 1995.
[35] M. Friedman, Applications of the ninhydrin reaction for analysis of amino acids [49] S. Ciampi, V. Di Castro, XPS study of the growth and reactivity of Fe/MnO thin
peptides, and proteins to agricultural and biomedical sciences, J. Agric. Food Chem. films, Surf. Sci. 331–333 (1995) 294–299.
52 (2004) 385–406. [50] B.J. Tan, K.J. Klabunde, P.M.A. Sherwood, X-ray photoelectron spectroscopy stu-
[36] M.A. Farag, M.G. Sharaf Eldin, H. Kassem, M. Abou el Fetouh, Metabolome clas- dies of solvated metal atom dispersed catalysts. Monometallic iron and bimetallic
sification of Brassica napus L. organs via UPLC–QTOF–PDA–MS and their anti- iron-cobalt particles on alumina, Chem. Mater. 2 (1990) 186–191.
oxidant potential, Phytochem. Anal 24 (2013) 277–287. [51] A. Bozzi, T. Yuranova, J. Mielczarski, J. Kiwi, Evidence for immobilized photo-
[37] L.Z. Lin, J. Sun, P. Chen, J. Harnly, UHPLC-PDA-ESI/HRMS/MSn analysis of an- Fenton degradation of organic compounds on structured silica surfaces involving Fe
thocyanins flavonol glycosides, and hydroxycinnamic acid derivatives in red mus- recycling, New J. Chem. 28 (2004) 519–526.
tard greens (Brassica juncea Coss Variety), J. Agric. Food Chem. 59 (2011) [52] T. Misawa, K. Hashimoto, S. Shimodaira, The mechanism of formation of iron oxide
12059–12072. and oxyhydroxides in aqueous solutions at room temperature, Corr. Sci. 14 (1974)
[38] N.Y. Yang, Y.F. Yang, K. Li, Analysis of hydroxy fatty acids from the pollen of 131–149.
Brassica campestris L. var. oleifera DC. by UPLC-MS/MSJ, J. Pharm. (2013) [53] G. Bianchi, F. Mazza, Corrosione e protezione dei metalli, Associazione Italiana di
874875. Metallurgia, AIM, 2005, pp. 10–12.
[39] M.H. Oueslati, H.B. Jannet, Z. Mighri, J. Chriaa, P.M. Abreu, Phytochemical con- [54] J.-R. Park, K.Y. Kim, Passivities of steels weathered for nine years, Corrosion 64
stituents from salsola tetrandra, J. Nat. Prod. 69 (2006) 1366–1369. (2008) 4–14.
[40] D.R. Walters, T. Cowley, H. Weber, N.I. Nashaat, Changes in oxylipins in compatible [55] K. Goto, H. Tamura, M. Nagaya, The mechanism of oxygenation of ferrous ion in
and incompatible interactions between oilseed rape and the downy mildew pa- neutral solution, Inorg. Chem. 9 (1970) 963–964.
thogen Peronospora parasitica, Physiol. Mol. Plant Pathol. 67 (2005) 268–273. [56] T. Misawa, K. Asami, K. Hashimoto, S. Shimodaira, The mechanism of atmospheric
[41] R. Solmaz, G. Kardas, M. Culha, B. Yazici, M. Erbil, Investigation of adsorption and rusting and the protective amorphous rust on low alloy steel, Corros. Sci. 14 (1974)
inhibitive effect of 2-mercaptothiazoline on corrosion of mild steel in hydrochloric 279–289.
acid media, Electrochim. Acta 53 (2008) 5941–5952. [57] G. Okamoto, Passive film of 18-8 stainless steel structure and its function, Corros.
[42] G. Okamoto, Passive film of 18-8 stainless steel structure and its function, Corros. Sci. 13 (1973) 471–489.
Sci. 14 (1973) 471–489. [58] P.C. Okafor, M.E. Ikpi, I.E. Uwaha, E.E. Ebenso, U.J. Ekpe, S.A. Umoren, Inhibitory
[43] M.C. Biesinger, B.P. Payne, A.P. Grosvenor, L.W.M. Lau, A.R. Gerson, R.St.C. Smart, action of Phyllanthus amarus extracts on the corrosion of mild steel in acidic media,
Resolving surface chemical states in XPS analysis of first row transition metals, Corros. Sci. 50 (2008) 2310–2317.
oxides and hydroxides: cr, Mn, Fe, Co and Ni, Appl. Surf. Sci. 257 (2011) [59] A.Y. El-Etre, M. Abdallah, E.Z. El-Tanawy, Corrosion inhibition of some metals
2717–2730. using lawsonia extract, Corros. Sci. 47 (2005) 385–396.
[44] S.F. Lim, Y.M. Zheng, J.P. Chen, Organic arsenic adsorption onto a magnetic sor- [60] A.E.-M.M. Naguib, F.K. El-Baz, Z.A. Salama, H.A. El Baky Hanaa, H.F. Ali,
bent, Langmuir 25 (9) (2009) 4973–4978. A.A. Gaafar, J. Saudi Soc. Agric. Sci. 11 (2012) 135–142.
[45] S. Deng, H. Liu, W. Zhou, J. Huang, G. Yu, Mn–Ce oxide as a high-capacity ad- [61] R.J. Ewen, C.L. Honeybourne, X-ray photoelectron spectroscopy of clean and gas-
sorbent for fluoride removal from water, J. Hazard. Mater. 186 (2011) 1360–1366. doped films of phthalocyanines, J. Phys. Condens. Matter. 3 (1991) S303–S310.

105

Você também pode gostar