[go: up one dir, main page]

Next Issue
Volume 2, March
Previous Issue
Volume 1, September
 
 

Corros. Mater. Degrad., Volume 1, Issue 3 (December 2020) – 5 articles

Cover Story (view full-size image): Corrosion is a naturally occurring phenomenon, and the formulation of more protective coatings that can be employed to prevent or minimise corrosion is especially important in the modern world. The opportunities to develop new protective coatings have never been so rich with the emergence of a host of new two-dimensional materials. In this review, the corrosion protection afforded by graphene is described and discussed, highlighting issues with galvanic corrosion and the more encouraging results obtained with functionalised graphene oxide. This is then followed with an account of graphene-like materials, including hexagonal boron nitride and graphitic carbon nitride. View this paper.
  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Select all
Export citation of selected articles as:
35 pages, 6614 KiB  
Review
Modification, Degradation and Evaluation of a Few Organic Coatings for Some Marine Applications
by Guang-Ling Song and Zhenliang Feng
Corros. Mater. Degrad. 2020, 1(3), 408-442; https://doi.org/10.3390/cmd1030019 - 21 Dec 2020
Cited by 16 | Viewed by 5352
Abstract
Organic coatings for marine applications must have great corrosion protection and antifouling performance. This review presents an overview of recent investigations into coating microstructure, corrosion protection performance, antifouling behavior, and evaluation methods, particularly the substrate effect and environmental influence on coating protectiveness, aiming [...] Read more.
Organic coatings for marine applications must have great corrosion protection and antifouling performance. This review presents an overview of recent investigations into coating microstructure, corrosion protection performance, antifouling behavior, and evaluation methods, particularly the substrate effect and environmental influence on coating protectiveness, aiming to improve operational practice in the coating industry. The review indicates that the presence of defects in an organic coating is the root cause of the corrosion damage of the coating. The protection performance of a coating system can be enhanced by proper treatment of the substrate and physical modification of the coating. Environmental factors may synergistically accelerate the coating degradation. The long-term protection performance of a coating system is extremely difficult to predict without coating defect information. Non-fouling coating and self-repairing coatings may be promising antifouling approaches. Based on the review, some important research topics are suggested, such as the exploration of rapid evaluation methods, the development of long-term cost-effective antifouling coatings in real marine environments. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of the cross-sections of epoxy powder coating and epoxy E-coating on a Mg alloy based on [<a href="#B56-cmd-01-00019" class="html-bibr">56</a>]. The defects may be across the entire coating, but present as pores or holes as pointed by arrows on the cross-sections.</p>
Full article ">Figure 2
<p>Schematic illustration of an airflow over an alkyd coating during curing, solvent evaporation distribution and possible micro-crack formation in the curing coating, adapted from [<a href="#B70-cmd-01-00019" class="html-bibr">70</a>].</p>
Full article ">Figure 3
<p>Schematic illustration of local modification of the surface morphology of a semiliquid alkyd film under a DC electric field [<a href="#B75-cmd-01-00019" class="html-bibr">75</a>].</p>
Full article ">Figure 4
<p>The substrate corrosion, coating damage and interface delamination for a coated AZ31 after 1000 h of salt spray, adapted from [<a href="#B56-cmd-01-00019" class="html-bibr">56</a>] (Copyright 2020, with permission from NACE International).</p>
Full article ">Figure 5
<p>Corrosion damage of epoxy coated (<b>a<sub>1</sub></b>,<b>a<sub>2</sub></b>) carbon steel (CS), (<b>b<sub>1</sub></b>,<b>b<sub>2</sub></b>) magnetron-sputtered Mg layer (CS+SM) on carbon steel, (<b>c<sub>1</sub></b>,<b>c<sub>2</sub></b>) magnetron-sputtered Mg layer (SM) on glass and (<b>d<sub>1</sub></b>,<b>d<sub>2</sub></b>) magnetron-sputtered iron layer on magnetron sputtered Mg layer (SM+SI) after salt spray for different periods of time, reprinted from [<a href="#B94-cmd-01-00019" class="html-bibr">94</a>] (Copyright 2020 with permission from Elsevier).</p>
Full article ">Figure 6
<p>Galvanic corrosion current densities of coupled pure Mg and carbon steel immersed in (<b>a</b>) epoxy resin (A), (<b>b</b>) curing agent (B) and (<b>c</b>) their mixture (A<sub>3</sub>B) at different relatively humidity levels at 20 °C, reprinted from [<a href="#B94-cmd-01-00019" class="html-bibr">94</a>] (Copyright 2020, with permission from Elsevier).</p>
Full article ">Figure 7
<p>Surface roughness of epoxy resin after continuous salt water immersion (E<sub>NaCl</sub>), continuous UV irradiation effect (E<sub>UVA</sub>) and alternate UV irradiation and salt water immersion, reprinted from [<a href="#B120-cmd-01-00019" class="html-bibr">120</a>] (Copyright 2020 with permission from Elsevier).</p>
Full article ">Figure 8
<p>Curve fitting of the surface roughness vs. exposure time for epoxy under different weathering conditions, reprinted from [<a href="#B120-cmd-01-00019" class="html-bibr">120</a>] (Copyright 2020, with permission from Elsevier).</p>
Full article ">Figure 9
<p>Schematic illustration of the antifouling mechanism of a serine protease encapsulated sol-gel coating, reprinted from [<a href="#B230-cmd-01-00019" class="html-bibr">230</a>] (Copyright 2020, with permission from American Chemical Society).</p>
Full article ">Figure 10
<p>Schematic illustration of the PEG layers on silicon, reprinted from [<a href="#B243-cmd-01-00019" class="html-bibr">243</a>] (Copyright 2020, with permission from American Chemical Society).</p>
Full article ">Figure 11
<p>Grafted zwitterionic polymer for antifouling, reprinted from [<a href="#B252-cmd-01-00019" class="html-bibr">252</a>] (Copyright 2020, with permission from Royal Society of Chemistry).</p>
Full article ">Figure 12
<p>Schematic illustration of the formation processes of different hydroxyl-rich antibacterial polymers via the ring opening reaction of s-PGMA, adapted from [<a href="#B256-cmd-01-00019" class="html-bibr">256</a>] (RX: Haloalkane, PGMA: poly (glycidyl mehacrylate), PGDED and PGDED-DED: the products after DED and DED/ED being introduced into s-PGMA, QPGDED-R, PGED and QPGDED-DED-R: the products modified by RX and ED, QPGDED-R-Ag, GED-Ag, QPGDED-DED-R-Ag: the products modified by AgNO<sub>3</sub>).</p>
Full article ">Figure 13
<p>Synthesis route of P (H-P-A) and PU films, reprinted from [<a href="#B262-cmd-01-00019" class="html-bibr">262</a>] (Copyright 2020, with permission from American Chemical Society).</p>
Full article ">Figure 14
<p>Schematic illustration of the preparation of a MHMS-based coating (<b>a</b>), SEM images of the microgel spheres (MS) (<b>b</b>) and MHMS (<b>c</b>), size distribution of the MS and MHMS (<b>d</b>), SEM image of the MHMS-based coating (<b>e</b>), the magnified SEM image (<b>f</b>) and cross-sectional SEM image (<b>g</b>) of the MHMS-based coating, reprinted from [<a href="#B16-cmd-01-00019" class="html-bibr">16</a>] (Copyright 2020, with permission from Springer Nature).</p>
Full article ">Figure 15
<p>Schematic illustration of the ingress of corrosive species into coating through the surface layer, the diffusion of corrosive species in the coating body and the interaction of corrosive species with the substrate through the interface layer.</p>
Full article ">
35 pages, 6465 KiB  
Article
Indirect Galvanostatic Pulse in Wenner Configuration: Numerical Insights into Its Physical Aspect and Its Ability to Locate Highly Corroding Areas in Macrocell Corrosion of Steel in Concrete
by Romain Rodrigues, Stéphane Gaboreau, Julien Gance, Ioannis Ignatiadis and Stéphanie Betelu
Corros. Mater. Degrad. 2020, 1(3), 373-407; https://doi.org/10.3390/cmd1030018 - 21 Dec 2020
Cited by 2 | Viewed by 3599
Abstract
The use of indirect electrical techniques is gaining interest for monitoring the corrosion of steel in concrete as they do not require any connection to the rebar. In this paper, we provide insights into the physical aspects of the indirect galvanostatic pulse (GP) [...] Read more.
The use of indirect electrical techniques is gaining interest for monitoring the corrosion of steel in concrete as they do not require any connection to the rebar. In this paper, we provide insights into the physical aspects of the indirect galvanostatic pulse (GP) method in the Wenner configuration. Considering uniform corrosion, the instantaneous ohmic drop is decreased due to the presence of the rebar, which acts as a short-circuit. However, we observed that this phenomenon is independent of the electrochemical parameters of the Butler–Volmer equation. They are, however, responsible for the nonlinear decrease of the current that polarizes the rebar over time, especially for a passive rebar due to its high polarization resistance. This evolution of the resulting potential difference with time is explained by the increase of the potential difference related to concrete resistance and the global decrease of the potential difference related to the polarization resistance of the rebar. The indirect GP technique is then fundamentally different than the conventional one in three-electrode configuration, as here the steady-state potential is not only representative of polarization resistance but also of concrete resistance. Considering non-uniform corrosion, the presence of a small anodic area disturbs the current distribution in the material. This is essentially due to the different capability of anodic and cathodic areas to consume the impressed current, resulting in slowing down the evolution of the transient potential as compared to uniform corrosion. Hence, highly corroding areas have a greater effect on the transient potential than on the steady-state one. The use of this temporal evolution is thus recommended to qualitatively detect anodic areas. For the estimation of their length and position, which is one of the main current problematic issue when performing any measurement on reinforced concrete (RC) structures with conventional techniques, we suggest adjusting the probe spacing to modulate the sensitivity of the technique. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Pictures of the mortar sample with the monitoring device and of the stainless-steel probes. (<b>b</b>) Schematic representation of the sample with the connection to the potentiostat-galvanostat; WE = working electrode, SE = sense electrode, RE = reference electrode, and CE = counter-electrode. (<b>c</b>) Example of typical raw data obtained in this study, showing the evolution of <span class="html-italic">V</span><sub>P1−P2</sub> with time when applying 10 consecutive galvanostatic pulses of 30 s at 500 µA or −500 µA.</p>
Full article ">Figure 2
<p>Geometry and mesh of the finite-element model used in this study (200 × 60 × 13 cm<sup>3</sup>). The two outer electrodes C<sub>1</sub> and C<sub>2</sub> act as current-injection electrodes, while the two inner electrodes P<sub>1</sub> and P<sub>2</sub> act as potential electrodes. Probe spacing <span class="html-italic">a</span> varies between 2.5 and 15 cm. The steel-concrete interface was modeled with a parallel contribution of the exchange current density and the double-layer capacitance using the Butler–Volmer equation.</p>
Full article ">Figure 3
<p>Illustration of the free macrocell current (evaluated at 50 s) that flows spontaneously between anodic and cathodic areas without applying any external polarization (<span class="html-italic">I</span><sub>C1</sub> = 0 µA) for <span class="html-italic">ρ</span> = 100 Ω m. The cover depth <span class="html-italic">e</span> was set at 40 mm and the anode length <span class="html-italic">L</span><sub>a</sub> at 1 cm. The inserts show the temporal evolution of the potential difference between P<sub>1</sub> and P<sub>2</sub>.</p>
Full article ">Figure 4
<p>Experimental results obtained on the mortar sample using the indirect galvanostatic pulse (GP) technique. (<b>a</b>) Influence of the distance of the probe from the rebar on the evolution of <span class="html-italic">V</span><sub>P1−P2</sub> with <span class="html-italic">I</span> = 100 µA and <span class="html-italic">a</span> = 15 cm, all electrodes being parallel to the rebar, (<b>b</b>) influence of the injected current (50–1000 µA) on the evolution of <span class="html-italic">V</span><sub>P1−P2</sub> with <span class="html-italic">a</span> = 15 cm, (<b>c</b>) evolution of <span class="html-italic">V</span><sub>P1−P2</sub> as a function of the impressed current for <span class="html-italic">a</span> = 15 cm at <span class="html-italic">t</span> = 0.1 s and 30 s, (<b>d</b>) influence of the probe spacing on <span class="html-italic">V</span><sub>P1−P2</sub> with <span class="html-italic">I</span> = 200 µA for the passive rebar (reproduced from Reference [<a href="#B2-cmd-01-00018" class="html-bibr">2</a>]), (<b>e</b>) influence of the probe spacing on <span class="html-italic">V</span><sub>P1−P2</sub> with <span class="html-italic">I</span> = 200 µA after the introduction of chlorides from the concrete surface through wetting/drying cycles, and (<b>f</b>) evolution of Δ<span class="html-italic">V</span> with time for each probe spacing, showing the difference between the passive rebar and the rebar after Cl<sup>−</sup> introduction.</p>
Full article ">Figure 5
<p>Evolution of the instantaneous ohmic drop observed at <span class="html-italic">t</span><sub>0</sub> for both active and passive conditions, for <span class="html-italic">I</span><sub>C1</sub> = 100 µA as a function of: (<b>a</b>) probe spacing <span class="html-italic">a</span>, for each concrete resistivity, <span class="html-italic">e</span> = 40 mm and <span class="html-italic">Φ</span> = 12 mm, (<b>b</b>) cover depth <span class="html-italic">e</span>, for each concrete resistivity, <span class="html-italic">a</span> = 5 cm and <span class="html-italic">Φ</span> = 12 mm; (<b>c</b>) rebar diameter <span class="html-italic">Φ</span> for each concrete resistivity, <span class="html-italic">a</span> = 5 cm and <span class="html-italic">e</span> = 40 mm.</p>
Full article ">Figure 6
<p>Comparing input concrete resistivity with the concrete resistivity calculated with Equation (6), considering the accurate geometric factor <span class="html-italic">k</span><sub>COMSOL</sub> for each simulation. The calculated resistivity decreases when increasing <span class="html-italic">a</span>, decreasing <span class="html-italic">e</span> or increasing <span class="html-italic">Φ</span> as more current reaches the rebar.</p>
Full article ">Figure 7
<p>Example of results showing the temporal evolution of <span class="html-italic">I</span><sub>rebar</sub> and <span class="html-italic">V</span><sub>P1−P2</sub> obtained when considering uniform corrosion in both active (<span class="html-italic">i</span><sub>0,a</sub> = 0.1 A m<sup>−2</sup>) or passive states (<span class="html-italic">i</span><sub>0,c</sub> = 10<sup>−5</sup> A m<sup>−2</sup>). (<b>a</b>) <span class="html-italic">a</span> = 5 cm, <span class="html-italic">e</span> = 40 mm and <span class="html-italic">Φ</span> = 12 mm; (<b>b</b>) <span class="html-italic">a</span> = 15 cm, <span class="html-italic">e</span> = 40 mm and <span class="html-italic">Φ</span> = 12 mm; (<b>c</b>) <span class="html-italic">a</span> = 5 cm, <span class="html-italic">e</span> = 60 mm and <span class="html-italic">Φ</span> = 12 mm; (<b>d</b>) <span class="html-italic">a</span> = 5 cm, <span class="html-italic">e</span> = 40 mm and <span class="html-italic">Φ</span> = 25 mm. The complete results from this parametric study are presented in <a href="#app1-cmd-01-00018" class="html-app">Figure S2</a>.</p>
Full article ">Figure 8
<p>Schematic representation of the physical aspect of the indirect GP technique in Wenner configuration. (<b>a</b>) Evolution of the distribution of the current in the concrete and in the rebar; (<b>b</b>) evolution of the potential difference attributed to concrete resistance, polarization resistance, and the global response.</p>
Full article ">Figure 9
<p>Evolution of <span class="html-italic">V</span><sub>P1−P2</sub> and Δ<span class="html-italic">V</span> over time in the presence of a slightly corroding area at different positions relative to the monitoring device. Input parameters: <span class="html-italic">i</span><sub>0,a</sub> = 10<sup>−4</sup> A m<sup>−2</sup>, <span class="html-italic">i</span><sub>0,c</sub> = 10<sup>−5</sup> A m<sup>−2</sup>, <span class="html-italic">L</span><sub>a</sub> = 1 cm, <span class="html-italic">a</span> = 5 cm, <span class="html-italic">e</span> = 40 mm, <span class="html-italic">C</span><sub>dl,a</sub> = <span class="html-italic">C</span><sub>dl,c</sub> = 0.2 F m<sup>−2</sup>, <span class="html-italic">I</span><sub>C1</sub> = 100 µA.</p>
Full article ">Figure 10
<p>Evolution of <span class="html-italic">V</span><sub>P1−P2</sub> and Δ<span class="html-italic">V</span> over time in the presence of a highly corroding area at different positions relative to the monitoring device. Input parameters: <span class="html-italic">i</span><sub>0,a</sub> = 0.1 A m<sup>−2</sup>, <span class="html-italic">i</span><sub>0,c</sub> = 10<sup>−5</sup> A m<sup>−2</sup>, <span class="html-italic">L</span><sub>a</sub> = 1 cm, <span class="html-italic">a</span> = 5 cm, <span class="html-italic">e</span> = 40 mm, <span class="html-italic">C</span><sub>dl,a</sub> = <span class="html-italic">C</span><sub>dl,c</sub> = 0.2 F m<sup>−2</sup>, <span class="html-italic">I</span><sub>C1</sub> = 100 µA.</p>
Full article ">Figure 11
<p>Evolution of <span class="html-italic">V</span><sub>P1−P2</sub> and Δ<span class="html-italic">V</span> over time in the presence of a highly corroding area at different positions relative to the monitoring device. Input parameters: <span class="html-italic">i</span><sub>0,a</sub> = 0.1 A m<sup>−2</sup>, <span class="html-italic">i</span><sub>0,c</sub> = 10<sup>−5</sup> A m<sup>−2</sup>, <span class="html-italic">L</span><sub>a</sub> = 3 cm, <span class="html-italic">a</span> = 5 cm, <span class="html-italic">e</span> = 40 mm, <span class="html-italic">C</span><sub>dl,a</sub> = <span class="html-italic">C</span><sub>dl,c</sub> = 0.2 F m<sup>−2</sup>, <span class="html-italic">I</span><sub>C1</sub> = 100 µA.</p>
Full article ">Figure 12
<p>Evolution of <span class="html-italic">V</span><sub>P1−P2</sub>, <span class="html-italic">I</span><sub>rebar</sub>, <span class="html-italic">I</span><sub>anode</sub>, and <span class="html-italic">I</span><sub>anode</sub>/<span class="html-italic">I</span><sub>rebar</sub> over time when the anode is below C<sub>1</sub>. Input parameters: <span class="html-italic">i</span><sub>0,a</sub> = 0.1 A m<sup>−2</sup>, <span class="html-italic">i</span><sub>0,c</sub> = 10<sup>−5</sup> A m<sup>−2</sup>, <span class="html-italic">L</span><sub>a</sub> = 3 cm, <span class="html-italic">a</span> = 5 cm, <span class="html-italic">e</span> = 40 mm, <span class="html-italic">C</span><sub>dl,a</sub> = <span class="html-italic">C</span><sub>dl,c</sub> = 0.2 F m<sup>−2</sup>, <span class="html-italic">I</span><sub>C1</sub> = 100 µA.</p>
Full article ">Figure 13
<p>Evolution of the current in the different parts of the anode for <span class="html-italic">ρ</span> = 100 Ω m (left) and <span class="html-italic">ρ</span> = 1000 Ω m (right) when the anode is below C<sub>1</sub>. The current probe C<sub>1</sub> is located just above the intersection of Anode 3 and Anode 4. Anode 6 is closest to P<sub>1</sub>.</p>
Full article ">Figure 14
<p>Evolution of <span class="html-italic">V</span><sub>P1−P2</sub> and Δ<span class="html-italic">V</span> over time in the presence of a highly corroding area at different positions relative to the monitoring device. Input parameters: <span class="html-italic">i</span><sub>0,a</sub> = 0.1 A m<sup>−2</sup>, <span class="html-italic">i</span><sub>0,c</sub> = 10<sup>−5</sup> A m<sup>−2</sup>, <span class="html-italic">L</span><sub>a</sub> = 3 cm, <span class="html-italic">a</span> = 2.5 cm, <span class="html-italic">e</span> = 40 mm, <span class="html-italic">C</span><sub>dl,a</sub> = <span class="html-italic">C</span><sub>dl,c</sub> = 0.2 F m<sup>−2</sup>, <span class="html-italic">I</span><sub>C1</sub> = 100 µA.</p>
Full article ">Figure 15
<p>Evolution of <span class="html-italic">V</span><sub>P1−P2</sub> and Δ<span class="html-italic">V</span> over time in the presence of a highly corroding area at different positions relative to the monitoring device. Input parameters: <span class="html-italic">i</span><sub>0,a</sub> = 0.1 A m<sup>−2</sup>, <span class="html-italic">i</span><sub>0,c</sub> = 10<sup>−5</sup> A m<sup>−2</sup>, <span class="html-italic">L</span><sub>a</sub> = 3 cm, <span class="html-italic">a</span> = 15 cm, <span class="html-italic">e</span> = 40 mm, <span class="html-italic">C</span><sub>dl,a</sub> = <span class="html-italic">C</span><sub>dl,c</sub> = 0.2 F m<sup>−2</sup>, <span class="html-italic">I</span><sub>C1</sub> = 100 µA.</p>
Full article ">Figure 16
<p>Evolution of <span class="html-italic">V</span><sub>P1−P2</sub> and Δ<span class="html-italic">V</span> over time in the presence of an anode with a high double-layer capacitance at different positions relative to the monitoring device. Input parameters: <span class="html-italic">i</span><sub>0,a</sub> = <span class="html-italic">i</span><sub>0,c</sub> = 10<sup>−5</sup> A m<sup>−2</sup>, <span class="html-italic">L</span><sub>a</sub> = 3 cm, <span class="html-italic">a</span> = 5 cm, <span class="html-italic">e</span> = 40 mm, <span class="html-italic">C</span><sub>dl,a</sub> = 2 F m<sup>−2</sup>, <span class="html-italic">C</span><sub>dl,c</sub> = 0.2 F m<sup>−2</sup>, <span class="html-italic">I</span><sub>C1</sub> = 100 µA.</p>
Full article ">Figure 17
<p>Evolution of <span class="html-italic">V</span><sub>P1−P2</sub> and Δ<span class="html-italic">V</span> over time in the presence of a highly corroding area with a high double-layer capacitance at different positions relative to the monitoring device. Input parameters: <span class="html-italic">i</span><sub>0,a</sub> = 0.1 A m<sup>−2</sup>, <span class="html-italic">i</span><sub>0,c</sub> = 10<sup>−5</sup> A m<sup>−2</sup>, <span class="html-italic">L</span><sub>a</sub> = 3 cm, <span class="html-italic">a</span> = 5 cm, <span class="html-italic">e</span> = 40 mm, <span class="html-italic">C</span><sub>dl,a</sub> = 2 F m<sup>−2</sup>, <span class="html-italic">C</span><sub>dl,c</sub> = 0.2 F m<sup>−2</sup>, <span class="html-italic">I</span><sub>C1</sub> = 100 µA.</p>
Full article ">Figure 18
<p>Schematic representation of the main scenarios expected when monitoring a rebar with one highly corroding area in concrete with uniform resistivity for <span class="html-italic">e</span> = 40 mm. The figure shows the evolution of Δ<span class="html-italic">V</span><sub>max</sub> values as a function of the measurement position. Depending on the charge transfer coefficients, the two peaks observed in cases A, B, and C can either be symmetric or not. Case A can be expected even if concrete resistivity is not uniform, as the anode effect is almost insignificant irrespective of concrete resistivity when <span class="html-italic">L</span><sub>a</sub> &lt;&lt; <span class="html-italic">a</span>.</p>
Full article ">
28 pages, 7514 KiB  
Review
Application of AFM-Based Techniques in Studies of Corrosion and Corrosion Inhibition of Metallic Alloys
by Kiryl Yasakau
Corros. Mater. Degrad. 2020, 1(3), 345-372; https://doi.org/10.3390/cmd1030017 - 7 Nov 2020
Cited by 30 | Viewed by 6020
Abstract
In this review several scanning probe microscopy techniques are briefly discussed as valuable assets for corrosionists to study corrosion susceptibility and inhibition of metals and alloys at sub-micrometer resolution. At the beginning, the review provides the reader with background of atomic force microscopy [...] Read more.
In this review several scanning probe microscopy techniques are briefly discussed as valuable assets for corrosionists to study corrosion susceptibility and inhibition of metals and alloys at sub-micrometer resolution. At the beginning, the review provides the reader with background of atomic force microscopy (AFM) and related techniques such as scanning Kelvin probe force microscopy (SKPFM) and electrochemical AFM (EC-AFM). Afterwards, the review presents the current state of corrosion research and specific applications of the techniques in studying important metallic materials for the aircraft and automotive industries. Different corrosion mechanisms of metallic materials are addressed emphasizing the role of intermetallic inclusions, grain boundaries, and impurities as focal points for corrosion initiation and development. The presented information demonstrates the importance of localized studies using AFM-based techniques in understanding corrosion mechanisms of metallic materials and developing efficient means of corrosion prevention. Full article
Show Figures

Figure 1

Figure 1
<p>A scheme of tapping mode atomic force microscopy (AFM) (<b>a</b>) and force-distance curve between the AFM tip and sample (<b>b</b>).</p>
Full article ">Figure 1 Cont.
<p>A scheme of tapping mode atomic force microscopy (AFM) (<b>a</b>) and force-distance curve between the AFM tip and sample (<b>b</b>).</p>
Full article ">Figure 2
<p>Electronic states of a probe and a sample metals before connection (<b>a</b>), after electric connection (<b>b</b>), and after applying a DC bias <span class="html-italic">V</span><sub>DC</sub> = <span class="html-italic">V</span><sub>VPD</sub> (<b>c</b>).</p>
Full article ">Figure 3
<p>A scheme of amplitude modulation mode scanning Kelvin probe force microscopy (SKPFM).</p>
Full article ">Figure 4
<p>A scheme of electrochemical AFM (EC-AFM) liquid cell containing main parts such as counter electrode (CE), working electrode (WE), and reference electrode (RE); AFM scanning system containing an AFM probe, a scanner, an optical system, a laser, and a photodetector; the external potentiostat controls the applied voltage between a counter and a working electrode.</p>
Full article ">Figure 5
<p>Topography evolution of AA5083 surface during in situ AFM measurements showing pit initiation and growth: (<b>a</b>) topography after 180 min of immersion in 0.005 M NaCl, (<b>b</b>) pit initiation after 130 min of immersion in 0.5 M NaCl, (<b>c</b>) pit growth after 220 min in 0.5 M NaCl and (<b>d</b>) image after immersion for 450 min in 0.5 M NaCl; the kinetics of the pit depth propagation (<b>e</b>). Reprinted from [<a href="#B47-cmd-01-00017" class="html-bibr">47</a>] Copyright (2020), with permission from Elsevier.</p>
Full article ">Figure 6
<p>AFM images of the alloy surface: (<b>a</b>) before in situ immersion, (<b>b</b>) after 25 min of in situ immersion and (<b>c</b>) after the in situ experiment followed by the removal of the corrosion deposits (performed in air); (<b>d</b>) comparison of the topography profiles across the black line on the AFM images. Reprinted from [<a href="#B47-cmd-01-00017" class="html-bibr">47</a>] Copyright (2020), with permission from Elsevier.</p>
Full article ">Figure 7
<p>AFM images of the alloy surface (<b>a</b>) at the beginning of immersion in 0.5% La(NO<sub>3</sub>)<sub>3</sub> + 0.005 M NaCl, (<b>b</b>) at the end of immersion in 0.5% La(NO<sub>3</sub>)<sub>3</sub> + 0.5 M NaCl; (<b>c</b>) Evolution of topography across the profiles depicted in (<b>a</b>,<b>b</b>) in 0.5% La(NO<sub>3</sub>)<sub>3</sub> solution with 0.005 M or 0.5 M NaCl; (<b>d</b>) Rate of the lanthanum hydroxide precipitation. Reprinted with permission from [<a href="#B11-cmd-01-00017" class="html-bibr">11</a>] Copyright (2020) American Chemical Society.</p>
Full article ">Figure 8
<p>In situ AFM three-dimensional topography images of high carbon martensitic steel surface at (<b>i</b>) 0 min; (<b>ii</b>) 60 min; and (<b>iii</b>) 105 min, the exposure to 0.1 M NaCl solution. Sample A (0.10–0.18 wt% Cr), Sample B (0.60–0.80 wt% Cr), Sample C (0.10–0.18 wt% Cr). Reprinted from [<a href="#B54-cmd-01-00017" class="html-bibr">54</a>].</p>
Full article ">Figure 9
<p>VPD after immersion in 0.5 M NaCl solution vs. corrosion potentials measured in 0.5 M NaCl solution (a), the Figure was adapted from [<a href="#B55-cmd-01-00017" class="html-bibr">55</a>].</p>
Full article ">Figure 10
<p>SKPFM study of the as-cast ZE41 alloy: (<b>a</b>) topography map, (<b>b</b>) surface potential map, and (<b>c</b>) line-profile analysis of relative Volta potential through an α-Mg dendrite; SEM study of the as-cast ZE41 after immersion in 3.5 wt% NaCl solution: (<b>d</b>) 60 min (surface) and (<b>e</b>) 60 min (cross-section). Reprinted from [<a href="#B62-cmd-01-00017" class="html-bibr">62</a>] Copyright (2020), with permission from Elsevier.</p>
Full article ">Figure 11
<p>Topography (<b>a</b>), (<b>b</b>) and Volta potential maps (<b>c</b>), (<b>d</b>) obtained at the same places of 2024 aluminum alloy before (<b>a</b>), (<b>c</b>) and after (<b>b</b>), (<b>d</b>) 5 h of immersion in 0.05 M NaCl solution; topography and VPD profiles (<b>e</b>) taken from maps a,b,c,d. Reprinted from [<a href="#B73-cmd-01-00017" class="html-bibr">73</a>] Copyright (2020), with permission from Elsevier.</p>
Full article ">Figure 12
<p>Topography (<b>a</b>), (<b>b</b>) and Volta potential maps (<b>c</b>), (<b>d</b>) obtained at the same places of 2024 aluminum alloy before (<b>a</b>), (<b>c</b>) and after (<b>b</b>), (<b>d</b>) 5 h of immersion in 0.5 g/L solution of salicylaldoxime with 0.05 M NaCl; topography and VPD profiles (<b>e</b>) taken from maps a,b,c,d. Reprinted from [<a href="#B73-cmd-01-00017" class="html-bibr">73</a>] Copyright (2020), with permission from Elsevier.</p>
Full article ">Figure 13
<p>Example of filament propagation (100 μm × 100 μm) observed on sample coated with 344 nm hexamethyldisiloxane (HMDSO) film at relative humidity around 85%. Part (<b>a</b>) presents observations of the surface from 2 hrs until 4 hrs 15 min, part (<b>b</b>) presents observations of the surface from 8 hrs 30 min until 12 hrs 25 min. Note that the topographic images on the top have a full scale range of 6 μm and the surface potential (SP) maps at the bottom have a full scale of 1.3 V. The contour lines are positioned to show the slight expansion in the borders of the filament when surface potential and topography images are compared. Reprinted from [<a href="#B69-cmd-01-00017" class="html-bibr">69</a>] Copyright (2020), with permission from Elsevier.</p>
Full article ">Figure 14
<p>Schematic illustration of the “jumps” during the propagation of the active head. Note that the nearest intermetallic particles are shown in red. After incorporation into the filament they are shown in ruby. The top sequence shows the propagation of the filigree in the first 4.5 h and the bottom sequence shows the further propagation imaged in the following 17.5 h. Reprinted from [<a href="#B69-cmd-01-00017" class="html-bibr">69</a>] Copyright (2020), with permission from Elsevier.</p>
Full article ">Figure 15
<p>In situ AFM images of corrosion behavior of the sample annealed at 1050 °C in 0.3 M hydrochloric acid solution: (<b>a</b>) <span class="html-italic">t</span> = 0, (<b>b</b>) <span class="html-italic">t</span> = 20 min, (<b>c</b>) <span class="html-italic">t</span> = 50 min, (<b>d</b>) <span class="html-italic">t</span> = 80 min, (<b>e</b>) <span class="html-italic">t</span> = 100 min, (<b>f</b>) t = 120 min; Magnetic (<b>a1</b>,<b>a2</b>) and VPD (<b>b1</b>,<b>b2</b>,<b>c1</b>,<b>c2</b>) measurements and profiles across different ferrite and austenite zones. Reprinted from [<a href="#B76-cmd-01-00017" class="html-bibr">76</a>] Copyright (2020), with permission from Elsevier.</p>
Full article ">Figure 16
<p>AFM topography maps of polished Z galvanized coating before immersion (<b>A</b>) and during immersion in 0.005 M NaCl: 3 min (<b>B</b>), 22 min (<b>C</b>), 35 min (<b>D</b>); Figure (<b>E</b>) presents the evolution of topography across the black lines profiles; Figure (<b>F</b>) presents the cathodic polarization curve on steel and the anodic polarization curves on Zn (Z) and Zn-Al-Mg (ZM) samples. Reprinted from [<a href="#B86-cmd-01-00017" class="html-bibr">86</a>] Copyright (2020), with permission from Elsevier.</p>
Full article ">Figure 17
<p>AFM topography (<b>A</b>) and SKPFM (<b>B</b>) maps of polished ZM galvanized coating before immersion and during immersion in 0.005 M NaCl: 4 min (<b>C</b>), 9 min (<b>D</b>), 22 min (<b>E</b>); Figure (<b>F</b>) presents the evolution of topography across the black lines profiles. Reprinted from [<a href="#B86-cmd-01-00017" class="html-bibr">86</a>] Copyright (2020), with permission from Elsevier.</p>
Full article ">Figure 18
<p>Topographic maps showing a grain boundary (GB) in sensitized AISI 304 stainless steel within 1% aqueous NaCl: (<b>a</b>) before the formation of the intergranular pit, (<b>b</b>) after the formation of the intergranular pit, and (<b>c</b>) the full intergranular pit formed. Reprinted from [<a href="#B87-cmd-01-00017" class="html-bibr">87</a>]—Published by The Royal Society of Chemistry.</p>
Full article ">Figure 19
<p>In-situ EC-AFM topography images of an actively dissolving region of the AA7075 in 0.2 M Na<sub>2</sub>SO<sub>4</sub> solution at (<b>a</b>) open circuit potential (OCP), (<b>b</b>) 1 V, (<b>c</b>) 2 V, (<b>d</b>) 4 V fist scan, (<b>e</b>) 4 V second scan, and (<b>f</b>) 8 V vs. Ag/AgCl, respectively. The inset of each image is the enlargement of the area marked with a square. Reprinted from [<a href="#B89-cmd-01-00017" class="html-bibr">89</a>] Copyright (2020), with permission from Elsevier.</p>
Full article ">
17 pages, 3304 KiB  
Article
Probabilistic Corrosion Initiation Model for Coastal Concrete Structures
by Changkyu Kim, Do-Eun Choe, Pedro Castro-Borges and Homero Castaneda
Corros. Mater. Degrad. 2020, 1(3), 328-344; https://doi.org/10.3390/cmd1030016 - 16 Oct 2020
Cited by 10 | Viewed by 3428
Abstract
Corrosion of the reinforced concrete (RC) structures has been affecting the major infrastructures in U.S. and in other continents, causing the recent several bridge collapses and incidents. While the theoretical understanding is well-established, the reliable prediction of the corrosion process in the RC [...] Read more.
Corrosion of the reinforced concrete (RC) structures has been affecting the major infrastructures in U.S. and in other continents, causing the recent several bridge collapses and incidents. While the theoretical understanding is well-established, the reliable prediction of the corrosion process in the RC structural systems has hardly been successful due to the inherent uncertainties existed in the electrochemical corrosion process and the associated material and environmental conditions. The paper proposes a computational framework to develop evidence-based probabilistic corrosion initiation models for the reinforcing steels in the RC structures, which predicts the corrosion initiation time and quantifies the inherent variances considering various acting parameters. The framework includes: probabilistic modeling with Bayesian updating based on the sets of previously generated experimental data; Bayesian model/parameter selection considering various parameters, such as material properties and environmental conditions; corrosion reliability analyses to predict the probabilities of the corrosion initiation at given time t, structural configurations, and environmental conditions; and sensitivity analyses to measure and to rank the influences of each acting parameter and its uncertainty to the probabilities of the corrosion initiation. Total of 284 sets of experimental data exposed to the coastal atmospheric environments are used for the modeling. The goal of the Bayesian model selection presented in this paper is to obtain the most accurate and unbiased model using the simplest form of expression. The developed example corrosion model is currently limited to the initiation of diffusion-induced corrosion. The model can be updated, improved, or modified upon future available sets of data. The research contributes to the decision making to improve the corrosion reliability, corrosion control, and further the structural reliability of corroding structures. Full article
Show Figures

Figure 1

Figure 1
<p>Initiation and propagation of the corrosion of reinforcing steel in concrete (adapted from Tuutti [<a href="#B20-cmd-01-00016" class="html-bibr">20</a>]).</p>
Full article ">Figure 2
<p>Schematic description of the specimen used in mm scale (adapted from Castro-Borges et al. [<a href="#B7-cmd-01-00016" class="html-bibr">7</a>]).</p>
Full article ">Figure 3
<p>Reference case: (<b>a</b>) model prediction vs. observation and (<b>b</b>) prediction error.</p>
Full article ">Figure 4
<p>Example parameter estimation of correction factor <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mn>1</mn> </msub> </mrow> </semantics></math>: (<b>a</b>) <span class="html-italic">w/c</span> = 0.46, (<b>b</b>) <span class="html-italic">w/c</span> = 0.5, (<b>c</b>) <span class="html-italic">w/c</span> = 0.53, and (<b>d</b>) <span class="html-italic">w/c</span> = 0.7.</p>
Full article ">Figure 5
<p>Model prediction vs. observation and prediction error; Model A: (<b>a</b>,<b>b</b>), Model B: (<b>c</b>,<b>d</b>), and Model C: (<b>e</b>,<b>f</b>).</p>
Full article ">Figure 6
<p>Estimated probability of corrosion initiation: (<b>a</b>) <span class="html-italic">w/c</span> = 0.46, (<b>b</b>) <span class="html-italic">w/c</span> = 0.5, (<b>c</b>) <span class="html-italic">w/c</span> = 0.53, (<b>d</b>) <span class="html-italic">w/c</span> = 0.7, and (<b>e</b>) <span class="html-italic">w/c</span> = 0.76.</p>
Full article ">Figure 7
<p>Example normalized sensitivity measures: (<b>a</b>) <span class="html-italic">w/c</span> = 0.5 and <math display="inline"><semantics> <mrow> <msub> <mi>C</mi> <mi>s</mi> </msub> </mrow> </semantics></math> = 0.14 and (<b>b</b>) <span class="html-italic">w/c</span> = 0.76 and <math display="inline"><semantics> <mrow> <msub> <mi>C</mi> <mi>s</mi> </msub> </mrow> </semantics></math> = 0.14.</p>
Full article ">Figure 8
<p>Example importance measures: (<b>a</b>) <span class="html-italic">w/c</span> = 0.5 and <math display="inline"><semantics> <mrow> <msub> <mi>C</mi> <mi>s</mi> </msub> </mrow> </semantics></math> <span class="html-italic">=</span> 0.14 and (<b>b</b>) <span class="html-italic">w/c</span> = 0.76 and <math display="inline"><semantics> <mrow> <msub> <mi>C</mi> <mi>s</mi> </msub> </mrow> </semantics></math><span class="html-italic">=</span> 0.14.</p>
Full article ">
32 pages, 3381 KiB  
Review
Review of Recent Developments in the Formulation of Graphene-Based Coatings for the Corrosion Protection of Metals and Alloys
by Bronach Healy, Tian Yu, Daniele da Silva Alves and Carmel B. Breslin
Corros. Mater. Degrad. 2020, 1(3), 296-327; https://doi.org/10.3390/cmd1030015 - 25 Sep 2020
Cited by 8 | Viewed by 5219
Abstract
Corrosion is a naturally occurring phenomenon and there is continuous interest in the development of new and more protective coatings or films that can be employed to prevent or minimise corrosion. In this review the corrosion protection afforded by two-dimensional graphene is described [...] Read more.
Corrosion is a naturally occurring phenomenon and there is continuous interest in the development of new and more protective coatings or films that can be employed to prevent or minimise corrosion. In this review the corrosion protection afforded by two-dimensional graphene is described and discussed. Following a short introduction to corrosion, the application of graphene in the formulation of coatings and films is introduced. Initially, reduced graphene oxide (rGO) and metallic like graphene layers are reviewed, highlighting the issues with galvanic corrosion. Then the more successful graphene oxide (GO), functionalised GO and polymer grafted GO-modified coatings are introduced, where the functionalisation and grafting are tailored to optimise dispersion of graphene fillers. This is followed by rGO coupled with zinc rich coatings or conducting polymers, GO combined with sol-gels, layered double hydroxides or metal organic frameworks as protective coatings, where again the dispersion of the graphene sheets becomes important in the design of protective coatings. The role of graphene in the photocathodic protection of metals and alloys is briefly introduced, while graphene-like emerging materials, such as hexagonal boron nitride, h-BN, and graphitic carbon nitride, g-C3N4, are then highlighted. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic representation of pitting attack and micrographs illustrating (<b>b</b>) general-like dissolution and (<b>c</b>) pitting attack.</p>
Full article ">Figure 2
<p>Summary of the number of graphene-based papers published in recent years on corrosion protection, taken from Scopus.</p>
Full article ">Figure 3
<p>Schematic diagram of a Tafel plot illustrating the estimation of <span class="html-italic">j<sub>c</sub></span><sub>orr</sub> and <span class="html-italic">E</span><sub>corr</sub> with (<b>a</b>) depicting a mixed corrosion event and (<b>b</b>) an inhibited corrosion event with lower anodic dissolution currents and (<b>c</b>) Evans diagram, depicting the H<sup>+</sup>|H<sub>2</sub> and Zn<sup>2+</sup>|Zn couples and the mixed corrosion reaction. The dashed traces in (<b>a</b>,<b>b</b>) correspond to measured data, while the solid traces show the linear Tafel regions, fitted to the experimental data.</p>
Full article ">Figure 4
<p>Simulated impedance data. (<b>a</b>) Nyquist diagram illustrating a protective coating (blue) and the onset of substrate dissolution (orange) and (<b>b</b>) Bode plot typical of a protective coating.</p>
Full article ">Figure 5
<p>Schematic illustrating the functionalisation of GO sheets followed by grafting to a polymer, combined with epoxy and cured to give the final coated electrode.</p>
Full article ">Figure 6
<p>Schematic of photocathodic protection, with graphene sheets shuttling the electrons from the semiconductor conduction band to the metal/alloy.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop