You are on page 1of 45

1 Titanium carbide coating to improve surface characteristic, wear and corrosion

2 resistance of spheroidal graphite cast irons

3 Ali GÜNEN1, Betül SOYLU2,3 Özgür KARAKAŞ3

4 1
Department of Metallurgy and Materials Engineering, Faculty of Engineering and Natural
5 Sciences, Iskenderun Technical University, 31200 Hatay, Turkey.
6 2
Department of Metallurgy and Materials Engineering, Institute of Graduate Studies,
7 Iskenderun Technical University, 31200 Hatay, Turkey.
8 3
MMK Metallurgy, Industry, Trade and Port Co., Dortyol, 31400, Hatay, Turkey
9
10 Abstract
11 In this study, titanium carbide (TiC) coatings were grown on the surface of a spheroidal
12 graphite cast iron (SGI) via thermo-reactive diffusion (TRD) using powder-pack processing at
13 800 °C, 850 °C and 900 °C for 4 h. The TiC coatings obtained on the SGI were characterized
14 by scanning electron microscopy (SEM), energy-dispersive X-ray spectroscopy (EDS), X-ray
15 diffraction (XRD), surface profilometry, microhardness, VDI adhesion testing, wear testing,
16 and electrochemical corrosion testing. Depending on the TRD temperature, continuous, crack-
17 free, and smooth TiC coatings of 5–11 µm thickness and 27.96–32.45 GPa hardness were
18 obtained on the surface. The high chemical stability, high hardness, and good adhesion
19 strength of TiC coatings resulted in a reduced friction, high wear resistance, and superior
20 corrosion resistance compared to the untreated sample. Delamination and oxidation assisted
21 abrasive wear transformed into oxidation-assisted adhesive wear in the coated samples.
22 Galvanic corrosion was dominant in untreated SGI, while homogeneous corrosion occurred in
23 the coated samples. TiC coatings grown using TRD have the potential to be used in
24 engineering components exposed to tribo-corrosive conditions.

25

26 Keywords: Spheroidal graphite cast iron, Thermo-reactive diffusion technique, Titanium


27 carbide coating, Wear, Corrosion

28

29

30

31

1 1
2
32 1. Introduction

33 Iron-based metals (such as cast iron and steel) are the most critical materials used
34 commercially. Although steels were prominent when the industrial revolution started to take
35 off, the use of cast iron that can compete with steels regarding their mechanical properties and
36 be obtained at a more affordable cost with the developments in technology has begun to take
37 up in many areas. [1-4]. Cast irons can be defined as Fe-C alloys containing at least 2.1% C
38 and 1-3% Si in their chemical content [1,2,5].In addition to these alloying elements, by adding
39 alloying elements such as Cr, Cu, Nb, Mo and Ni, the mechanical properties and corrosion
40 resistance of cast irons can be improved by changing the chemical composition and cooling
41 behavior of cast irons [3,5-7].

42 Cast irons are gaining widespread use in industries, and the foremost among them is the
43 spheroidal graphite cast iron (SGI), which has the mechanical properties of steel and the
44 advantages of cast iron production [4,8]. In addition to having toughness, ductility, vibration
45 damping, good machinability, mechanical and fatigue strength [9], SGIs also have a lower
46 melting point than steel. These unique properties have empowered SGI in numerous industrial
47 applications such as exhaust manifolds, glass container automotive engines, gas turbine
48 compressor boxes, turbocharger bodies, molds, aluminum and lead melting crucibles, hot
49 rolling mills, burner housings, and furnaces [4,10,11].

50 Although spheroidal graphite nodules are structures that determine the characteristic
51 properties of SGIs such as ductility and toughness, they serve as starting points in damage
52 mechanisms such as corrosion, high-temperature wear, and thermal fatigue [4, 11-16]. This is
53 a factor that reduces the lifetime of SGIs at their place of use. In addition, due to the low
54 surface hardness of SGIs, their wear resistance in many corrosive environments is not
55 satisfactory [14-19]. Therefore, in order to improve the tribo-corrosive resistance of these
56 alloys, alloying [6,7,20-22], coating the surfaces with many coating methods such as PVD
57 [23], CVD [24], laser [11,12], thermal spray [25], and austempering [26,27], boriding [28.29],
58 vanadinizing-chrominizing [4,30,31] etc. studied by many scholars, and improvement to
59 tribo-corrosive condition obtained.

60 The thermo-reactive diffusion technique takes a step forward compared to other


61 coating methods to provide good corrosion, thermal fatigue, and wear resistance and be easily
62 adapted to industrial applications with low installation costs [32-34]. Thermo-reactive
63 diffusion technique is a surface hardening heat treatment based on the principle that the

3 2
4
64 carbon in the chemical content of the material diffuses to the surface with the effect of high
65 temperature (>700°C) and forms carbides with the transition elements (Cr, V, Ti, Nb, etc.)
66 that are desired to be diffused on the surface. [35]. In this method, a thin hard layer in the
67 range of 1500-3800 HV is obtained depending on the chemical composition of the carbide
68 layer formed on the surface, while a gradual decrease in hardness from the surface towards
69 the inner parts ensures that the material remains tough against impacts from the outside
70 without changing the inner parts of the materials. Thus, while increasing the strength and
71 wear resistance of the materials, the decrease in the toughness values is kept at a minimum
72 level. Although TRD applications have been carried out predominantly in a salt bath
73 environment, studies also have fluidized bed and pack cementation techniques [32]. The
74 carbon (C) and nitrogen (N) contents of the alloys to be subjected to TRD treatment are
75 essential [32,35]. When the TRD treatment is applied to alloys containing <0.3 C by weight,
76 metallic coatings are obtained instead of carbide layers [32]. Therefore, it is difficult obtaining
77 hard carbide layers on material groups such as non-metallic alloys, superalloys, cermets,
78 Ni3Al alloy, and high entropy alloys.

79 The properties of the carbide-based layers obtained vary according to the phase/s
80 formed on the surface, depending on the chemical composition of the substrate used,
81 especially the C ratio, the purity of the coating powders used, and the TRD process
82 parameters (temperature, time, TRD environment). Namely: the hardness of the TiC layer,
83 which is the most stable and hardest among the carbide-based layers obtained as a result of
84 the process, is between 3200-3800 kg/mm2, while the other layers obtained by this method
85 are VC 2900 - 3200 kg/mm2, NbC 1800 - 2500 kg/mm2 and Cr23C6 while it is in the range of
86 1600 - 2000 kg/mm2 [34]. In addition to having the highest hardness (3200-3800HV), TiC
87 coatings are thermally stable up to temperatures over 3000 °C and are chemically stable to
88 most molten metals [35,36]. These unique properties make them unmatched in industrial
89 applications.

90 It has been reported by many scientists that TiC coatings are obtained by CVD, PVD,
91 and PLD methods and that TiC improves the hardness and wear properties of many substrate
92 materials [38-42]. However, PVD, CVD, and PLD methods have high initial setup and
93 operating costs. Since these methods require a vacuum and/or protective atmosphere and
94 require qualified personnel. On the other hand, the fact that the adhesion of the coatings
95 obtained with PVD and CVD to the surface is weak, thus causing the substrate to break, takes
96 the TRD methods one step forward.

5 3
6
97 In the literature, it has been reported in many studies that TRD TiC coatings provide a
98 significant life increase in various Fe-based materials. Some of those: Kurt et al. (2018)
99 titanized AISI D2 steel at 900, 1000, 1100°C for 1, 2, 3 h through solid-state box Thermo-
100 reactive diffusion technique. The authors report that i TiC coating improved wear resistance
101 of AISI D2 at all titanizing conditions, but mechanical properties and wear resistance
102 changed depending on uniformity, hardness, elastic modulus, and coating layer's thickness
103 [34]. Zhang et al. (2017) formed a TiC coating layer with a thickness of 7.5 µm and hardness
104 of 1885 HV by using the pure bath TRD method on AISI 1020 steel for 3 h at 900 °C by TRD
105 method. The authors reported that the adhesion forces of the TiC layer to the surface are 52 N,
106 and the TiC coating provided an improving (78%) wear resistance compared with the
107 untreated substrate [43]. In another study, Kurt et al. (2014) reported that they obtained TiC
108 coating layers in the range of 6-23 µm depending on the TRD process parameters with the
109 powder pack method for 2 and 4 h at 950, 1000 1050, 1100 C on AISI D3 steel. They
110 determined that these layers were 1700-2470 HV hard, and the activation energy was 249.42
111 kJ/mol. The authors did not examine the wear and/or corrosion resistance of the coatings [44].
112 Taktak and Ulu (2010) performed the characterization of the TiC coatings they formed on
113 AISI 52100 and AISI 440C steels with the TRD method for 3 h at 950 °C, using Daimler-
114 Benz Rockwell-C, microKnoop adhesion tests, and abrasion tests from room temperature to
115 600 °C. As a result of the study, it was reported that the hardness value of 2400 HK was
116 reached as a result of the formation of Ti6C3.75 phase in both steel types, and it provided better
117 wear resistance from room temperature to 300 °C compared to the untreated material, but at
118 600 °C, the substrate materials showed better wear performance [45].

119 Although there are no publications in the open literature that clearly indicate that TiC
120 coatings can be obtained on cast irons by TRD so far, there are studies in which other carbide
121 coatings are obtained. Soares et al. applied niobizing to Cu-containing and Cu-free spheroidal
122 graphite cast irons at 1000 °C for 4 hours using TRD. As a result of the niobizing treatment,
123 they obtained coatings consisting of NbC with hardness higher than 2000 HV. According to
124 the VDI 3198 adhesion test and Vickers hardness tests, the coatings bonded to the surface
125 with good adhesion. The authors also reported that the obtained NbC coatings exhibited better
126 abrasion performance in the micro-adhesion and micro-abrasion tests than the untreated
127 sample [9]. Mariani et al. produced NbC, NbV and NbVC coatings on vermicular cast irons
128 using the TRD salt bath method at 1000°C for 2 h followed by tempering at 300°C. They
129 reported that the obtained coatings exhibited good adhesion with a thickness of more than 25

7 4
8
130 µm and hardness values of 2300–3050 HV. The austempered carbide coatings exhibited better
131 wear resistance in the range of 10–30 times and better corrosion resistance than uncoated
132 austempered vermicular graphite cast iron [46]. In another study, Mariani et al. (2020) studied
133 the formation of NbC coatings obtained with the powder pack method and the salt bath
134 method for 2 h at 900 °C on gray cast iron using TRD. The authors stated that although
135 coatings with NbC phase structure, 10 µm thickness and 2000 HV hardness values were
136 obtained in both methods, the coatings obtained with the powder pack method were more
137 uniform; there was a transition zone between the coating and the substrate, which resulted in
138 better adhesion. The authors reported that due to the high hardness obtained, the coatings
139 obtained by both methods provided about 12 times higher wear resistance in the adhesive
140 wear tests. However, in the micro adhesive wear tests, the coatings produced by the powder
141 pack method exhibited 8 times higher wear resistance compared to the untreated sample
142 owing to the transition zone, while the NbC coatings produced by the salt bath method
143 (without a transition zone) showed only 2 times higher wear resistance compared to the
144 untreated sample. The authors concluded that the powder pack method wasmore effective for
145 improving wear resistance [47]. Günen et al. deposited Cr-V-C coatings on the spheroidal
146 graphite cast iron surface at 900, 1000 and 1100 °C for 1 h by TRD method. The authors
147 reported increasing thickness and hardness in the Cr-V-C coatings with increasing TRD
148 processing and temperature. The obtained Cr-V-C coatings exhibited higher wear resistance
149 than untreated ductile iron at both room temperature and high temperatures due to their high
150 hardness and stability up to 750°C. The authors also reported that Cr-V-C coatings improved
151 their corrosion resistance in NaCl, H 2SO4 and HNO3 environments compared to untreated
152 spheroidal graphite cast iron due to the Cr-V-C coatings converting the graphite nodules into
153 carbides during their formation [4,31].

154 As a matter of fact, as can be seen in the studies on TiC, the effects of TiC formation
155 on the microstructure, some mechanical properties and wear properties of TiC formation on
156 steel have been studied [34,43-45]. On the other hand the corrosion behavior of TiC coating
157 has been studied less frequently [48]. However, to develop cost-effective systems,
158 components are increasingly being used in harsh environmental conditions, particularly in
159 tribo-corrosion environments. For example, the US Navy investigates materials that exhibit
160 both high wear resistance and corrosion resistance [49]. However, both wear and corrosion
161 resistance may not always be improved as a result of coating processes. Especially since
162 thermochemical methods are carried out at high temperatures, if it causes deterioration in the

9 5
10
163 microstructure of the base material, it may cause deterioration in corrosion resistance [50,51].
164 In addition, the fact that studies on the production and characterization of TiC coatings on cast
165 irons using TRD processes have not yet been included in the literature has triggered the
166 research of this subject. To better understand the possible applications, benefits, and
167 limitations of growing TiC coatings on the surface of ductile cast iron, this study focused on
168 examining the microstructure, phase formation, hardness, adhesion, friction, wear resistance,
169 and corrosion behavior of TiC coatings formed on the SGIs surface.

170 2. Experimental Studies


171 The supplier gave a chemical analysis of spheroidal graphite cast iron samples (SGI-
172 80) as 3.65 wt% C, 1.90 Si, 0.3 Mn, 0.06 P, 0.03 S, .016 Cr, 0.015 Ni, 0.002 Mo, 0.035 Mg
173 and balance Fe. Before the coating process, the specimens were cut with a precision cutting
174 device using a SiC wear disc with 20x20x5 mm. Then the samples were ground with 320-
175 400-600-800 SiC paper, respectively and the residues on the surface were removed with
176 alcohol solution and distilled water. Based on previous studies on steels, a powder blend of
177 45% FeTi + 45% Al2O3 + 10% NH4CI was prepared for the TiC coating process [34,44].
178 Preliminary studies were carried out at 1100 , 1000, 900 and 800 °C, considering the TiC
179 coating temperatures carried out on steels. However, as shown in Figure 1, local spills
180 occurred in the coatings performed at 1000 and 1100 °C. On the other hand, since any spills
181 were observed in the coatings carried out at 800 and 900 C, it was decided to make the final
182 coatings at 800, 850 and 900 C for 4 hours. The crucibles were left to cool in an open-air
183 environment to prevent cracking in the coated samples. After cooling, the surface of the
184 samples was cleaned of dust residues.

11 6
12
185

186 Figure 1. Surface view of samples before and after pre-coating processes

187 The pack TRD technique involves chemical reactions that can form dynamic diffusion
188 atoms and grow diffusion coatings. The chemical reactions forming dynamic Ti atoms during
189 the titanizing process are given below [52].

190 NH4Cl(s) → HCl(g) + NH3(g)  (1)


191 2NH3 → + 3H2(g) + N2(g)  (2)
192 4HCl(g) + Ti(s) → 2H2(g) + TiCl4(g)   (3)
193 2Fe(s)  +  TiCl4(g) → 2FeCl2(g) + Ti (4)
194 TiCl4(g) + 2H2(g) → 4HCl(g) + Ti, (5)
195 and the active atom Ti reacts with C to form TiC.
196
197 After the titanizing process, the samples were cut in appropriate sizes (15×15×5 mm).
198 The cross-sections of the samples were grounded with SiC paper in the range of 320-2500 grit
199 at appropriate times. The polishing process was carried out using 1 µm and 0.25 µm Al 2O3
200 pastes. Finally, etching was applied with 3% Nital for 6 s to reveal the microstructural details
201 of the mirror-shine surfaces. Examinations of cross-sections of TiC coatings were performed
202 in CBS mode using a Thermo Fisher Scientific Apreo S LoVac. (Netherlands) SEM operating
203 at 30 kV equipped with Ultra Dry EDS and Quasor II electron backscatter diffraction (EBSD)
204 detectors. The thickness of the TiC layer was determined by SEM microscope and the
205 microhardness values were determined by both Vickers microhardness and nanoindentation

13 7
14
206 methods. Micro-hardness measurements were determined with a Vickers pyramid-tip Qness
207 GmbH Q10 (Austria) micro-hardness device using a 25 gf load and 10 s dwell
208 time.Nanoindentation hardness measurements were determined using the Hysitron TI-950
209 TriboIndenter device (Germany) with a Berkovich-tip nanoindentation with 10 mN load and a
210 30 s constant loading rate, 15 s dwell time, and 30 s gradually unloading time.. Thickness and
211 hardness values were determined based on the average of 10 different measurements.

212 A computer-controlled Panalytical Malvern Panalytical Empyrean (Netherlands)


213 XRD device was used to determine and analyze XRD patterns with Cu Kα radiation (1.5418A)
214 with scan step size 0.0525211and 2θ angles ranging from 20 to 90°. Surface roughness
215 measurements were made using a 2D profilometer (Hommelwerke T8000 GmbH, Germany)
216 based on 0.25 mm/s speed and 2.5 mm scan length.

217 The adhesion behavior between TiC coatings and substrate SGI was determined by the
218 Daimler-Benz Rockwell-C indentation test (VDI 3198). In this method, the maps of the
219 damage appearances resulting from the test are classified from 1 to 6. According to this map,
220 shapes 1-4 are considered acceptable in terms of adhesion, while 5 and 6 are considered
221 unacceptable (Fig. 2).

222
223 Figure 2. Classification of damage mechanisms according to the VDI 3198 adhesion test [53].
224
225 In order to determine the stability and dry sliding wear behavior of thin TiC coatings (5-11
226 µm thickness) obtained on the spheroidal graphite cast iron surface, 100 m (short distance)
227 and 250 m (mid-distance) wear tests were applied. Thus, friction behavior of TiC coatings
228 during wear, wear volume losses and SEM-EDS characterization of worn surfaces were

15 8
16
229 evaluated. The wear test was performed at room temperature against a 6.3 mm diameter Al2O3
230 ball (14GPa) with 10N load, 200 mm/s speed (477.48 rpm), 8 mm wear track diameter, 100 m
231 and 250 m sliding distance using rotational ball-on disc wear device (Turkyus
232 POD&HT&WT, Turkey). The 100 m abrasion test corresponds to approximately 4000 cycles,
233 and the 250 m abrasion test corresponds to 10,000 cycles. The device automatically recorded
234 the coefficient of friction (COF) during the wear test, and COF graphs were prepared based
235 on the COF data obtained from the test of three samples for each test condition.

236 After the wear process, the widths of the wear marks were determined with an optical
237 microscope and their depths were determined with a 2D profilometer. Since the wear trace
238 resembles half of the ellipse, the wear volume losses were determined as half the volume of
239 the ellipse using the equations detailed in our previous studies as given below [31,50,54].

240

241

242 (6)

243 (7)

244 (8)

245 Where : Length of wear track (mm), : Wear track volume (mm3), : Radius of
246 wear track (mm), : Average wear track width (μm), : Average wear track depth (μm),

247 : Test load (N), : Sliding distance (m) and : Wear rate (mm³/Nm).
248 Finally, to determine the presence of TiC coatings on the surface after wear, wear
249 traces were visualized by SEM, and EDS analysis was taken from some parts of the traces.
250 The corrosion resistance of both un-coated and TiC coated SGI was determined by an
251 electrochemical corrosion test. Open circuit potential test (ASTM G106) and Tafel
252 extrapolation technique (ASTM G5) parameters were used as stated in the previous studies
253 [31,54]. The measured OCP value in the electrolyte gives us the equilibrium potential
254 between cathodic and anodic reactions. When the sample was immersed in the 3.5 wt.%
255 NaCl solution, it was held for 3600 seconds to stabilize. OCP was used to determine
256 corrosion potential, and the Tafel extrapolation method was used to determine corrosion
257 current density and rate. Tafel extrapolation experiments were realized at 0.1 mV/s scan rate
258 from -250 mV to +250 mV vs. OCP. Corrosion solution was chosen as 3.5 wt.% NaCl, which
17 9
18
259 is possible SGIs exposed, solution. CHI 602E model potentiostat and analyzer test device
260 (Shanghai Chenhua Instrument Co., Ltd.)  and conventional three-electrode system were used
261 for all electrochemical experiments in pyrex glass. The surface zone of the samples exposed
262 to the corrosion test was 0.785 cm2. The sample was used as the cathode and a platinum wire
263 was used as the anode. An Ag/AgCI electrode was used as the reference electrode.
264 Experiments were realized at room temperature as described in our previous study [31,54].
265 Experiments were performed in triplicate, and the average of the three measurements was
266 reported as a value. Corrosion potentials (Ecorr) and the corrosion current densities
267 (icorr)were calculated using instantaneous Tafel-type fit CHI 602E corrosion analysis
268 software. After corrosion testing, each sample was then analyzed using a FESEM device
269 equipped with EDS capabilities.
270

271

272

273 3. Result and Discussion


274 Fig. 3 shows the microstructure view of the untreated SGI sample.

275
276 Figure 3. Optical microstructure view of untreated pearlitic SGI etched with 3 wt.% Nital.

19 10
20
277 SEM observation revealed that the ferrite phase passed around the graphite spheres in
278 a ring and a matrix structure consisting of pearlite surrounding this ferrite phase. This
279 structure, which is formed by spherical graphite nodules and ferritic structure, is called "bull's
280 eye" in the literature [26,3,55]. The volume ratios of ferrite and pearlite formed in the matrix
281 structure can vary depending on the cooling rate, the content of alloying elements, and the
282 amount of graphite. It is seen that the calf eye structure does not have a homogeneous
283 appearance throughout the microstructure. This is because there are always segregations in the
284 casting materials, depending on the cooling rate of the casting molds and the purity of the
285 charging materials used, in cases where the casting conditions are not always completely
286 optimal. Charge inputs must be optimally controlled to keep segregations and carbides to a
287 minimum. Thus, it is ensured that the product to be produced has the desired mechanical
288 properties and that the undesired compounds are kept at the lowest level.

289 Fig. 4 illustrates the cross-sectional microstructure of the sample titanized at 800 °C
290 for 4h. It is clear that a coating layer of 4-5 µm thickness is formed on the SGI substrate
291 surface, which is continuous on the surface but has a roughness in places. It is seen that there
292 are no spherical graphite nodules in the coating layer formed as indicated on other carbide
293 coatings obtained by the TRD method on SGI [9,30,31]. EDS line analysis showed that the
294 resulting coating layer consisted of 60% Ti, 20% C and 5% Fe. As seen in the EDS line
295 analysis, it was determined that the C ratio decreased from the surface to the inner parts.
296 Although spherical graphite nodules were not observed in the coating layer, the presence of C
297 in these regions indicates that the spherical graphite nodules spread with the effect of heat and
298 formed a TiC coating layer on the surface.Moreover, It is seen that there is a slight increase in
299 the O ratio from the surface to the transition zone. This situation may have occurred as a
300 result of slight oxidation on the surface of the substrate because the coating processes were
301 carried out in an open-air environment.

21 11
22
302

303 Figure 4. SEM section view and EDS region analysis of SGI sample subjected to titanizing at
304 800 °C for 4 h etched with 3 wt.% Nital.

305 Fig. 5 illustrates the corresponding cross-sectional morphology of the sample titanized
306 at 850 °C for 4h and relevant EDS line analysis. As shown in Fig. 5 the TiC coating is more
307 homogenous, compact and thick according to the sample titanized at 800 °C for 4h.
308 Moreover, spheroidal graphite nodules were not observed in the coating layer as well as in the
309 sample titanized at 800 °C for 4h, and under the coating layer, the spherical graphite nodules
310 dissolved in someplace and/or transformed into polygonal shapes different from the spherical
311 shape. EDS line analysis of Ti, Fe, C concentration depth profile is taken from the surface
312 towards the interior shows that the Ti, C and Fe ratios are ranked at 65%, 25% and 5% levels
313 throughout the coating layer and these ratios continue to be continuous throughout the coating

23 12
24
314 layer. The presence of a transition zone of around 2 µm between the finish line of the coating
315 layer and the substrate, which is not visible in the SEM microstructure pictures, is seen in the
316 EDS line analysis. The formation of the transition zone between the coating layer and the
317 substrate has also been reported in previous studies [43,47]. It has been reported that this
318 transition region is the iron-titanium solid solution region between the coating layer and the
319 substrate, improving the adhesion forces of the carbide layer to the substrate [43, 47,56]. At
320 the end of the transition zone, it is seen that the Fe ratio increases very rapidly, whereas the Ti
321 ratio first decreases gradually and then approaches "0" with a gradual decrease.

322
323 Figure 5. SEM section view and EDS depth line analysis of SGI sample subjected to
324 titanizing at 850 °C for 4 h etched with 3 wt.% Nital.
325

25 13
26
326 When the SEM microstructure of the TiC coating formed by the TRD method at 900
327 °C for 4 h was examined (Fig. 6), a similar coating layer was obtained to the coating layer
328 performed at 850 °C, but it was observed that TiC coating thickness increased slightly. This is
329 an expected result since the titanizing process is a diffusional heat treatment. Because with the
330 increase in temperature, the dissolution of C will spread to the surface and the diffusion rates
331 of Ti atoms on the surface will increase. As a matter of fact, the highest coating layer (around
332 11µm) was reached in this parameter. The critical point in the EDS line analysis is that the Ti
333 ratios in the sample titanized for 4 h at 900 °C, increased, whereas the O ratio in the coating
334 layer decreased, while the C ratio did not change. This point shows that despite the increasing
335 diffusion rate, the C ratio in the structure is limited, so it spreads up to a specific rate, but it
336 continues to spread due to the high Ti concentration in the coating powder content.

337

27 14
28
338 Figure 6. SEM section view and EDS depth line analysis of SGI sample subjected to
339 titanizing at 900 °C for 4 h etched with 3 wt.% Nital.

340
341 Figure 7. XRD analysis taken from the SGI a) untreated, b) 800-4h, c)850-4h and d) 900-4h.
342
343 XRD analyses were performed to support the accuracy of the TiC coatings shown in
344 the SEM pictures and detected by EDS analysis. As seen in Figure 7; Fe phase (ICDD 01-
345 085-1410) was determined in the untreated GG-80 sample at 44.77, 65.00 and 82.46 2-theta
346 deg. On the other hand, in titanized samples, TiC phase was observed at 35.74, 41.52, 60.33,
347 72.20 and 76.08 2-theta deg. The Fe phase was determined in untreated samples. The
348 remarkable point in the XRD analysis is that the intensity of the Fe phase decreased, while the
349 intensity of the TiC phases increased depending on the increase in the coating thickness. This
350 shows that XRD rays act up to a certain distance from the surface. Therefore, it should be
351 considered that XRD rays can detect phases in the substrate as well as the phases in the
352 coating in the thin coatings (below 10 µm). The obtained phases exactly match the phases
353 obtained by Zhang et al. (2019) on AISI 1020 steel, which first carburized and then titanized
354 samples [43].

355 Considering the studies in the literature, although TiC coatings have been studied on
356 steels with the TRD method, it has not yet been studied on cast irons. It has been reported that
357 the best coating layer thickness, hardness and wear resistance of TiC coatings obtained on
358 steels are achieved in titanizing studies performed at 1000 °C and 1100 °C [34,43-45].
359 However, as seen in the preliminary experiments of our study (Fig. 1), although the coatings
360 at 1000 and 1100 C were formed on the surface, they were shed from the surface during the

29 15
30
361 cooling of the samples. The spill of the TiC coating may be occurred due to the thermal
362 expansion coefficient difference between the TiC coating layer and the substrate SGI.
363 Because the thermal expansion coefficient of spheroidal graphite cast iron is 13x106 K1
364 while the thermal expansion coefficient of TiC coatings is 7.43×106 K1. Moreover, SGI,
365 which is named as a natural composite material due to the presence of graphite, ferrite and
366 pearlitic structures in its microstructure, will be exposed to residual stresses due to the
367 different thermal expansion coefficients of the micro components [57] and TiC coating
368 depending on the temperature change during the process. Therefore, it is understood that these
369 residual stresses have exceeded a certain point and the damage has occurred. The targeted
370 coating layers are achieved in spheroidal graphite cast irons at lower temperatures can be
371 attributed to the higher C ratio of these alloys compared to steels and the presence of austenite
372 in the structure at temperatures above 723°C. Because the high C concentration in the
373 structure accelerates the diffusion, thus enabling the formation of TiC coatings even at lower
374 temperatures. On the other hand, converting a part of the structure to austenite above 723 °C
375 increased the C resolution, which accelerated the formation of TiC in the structure. In the
376 coating processes carried out between 800°C-1100°C, the coating layer increases with the
377 increase in temperature, but since the increase in the coating layer increases the thermal
378 expansion difference between the coating layer and the substrate, it is observed that the
379 coating layers carried out at 1100 °C fracture off on a large part of the surface (Fig. 1). The
380 remarkable point in the coating layers obtained in all experimental parameters is that, as in all
381 TRD processes (with vanadium, chroming, niobinizing, etc.) [9,30,31], spherical graphite
382 nodules were not included in the coating layer, unlike boriding [28,29]. This is due to the
383 nature of the thermo-reactive diffusion technique. Because in the Thermo-reactive diffusion
384 technique (unlike boronizing), the C inside the material moves from the material to the
385 surface with the effect of heat [32]. When the carbon and/or nitrogen in the substrate comes
386 into contact with the activator at high temperature, the carbide and nitride forming elements
387 combine with the activators due to their low free energy to form transition metal carbides
388 [32].

31 16
32
389
390 Figure 8. Rockwell C indentations (VDI 3198) on the surfaces of the TiC layers a)800-4h,
391 b)850-4h and c) 900-4h.

392 Fig. 8 shows the optical and SEM surface view of the TiC layer subjected to the VDI
393 3198 Rockwell C adhesion test. Considering the Fig. 8 (2500X) SEM views were examined, it
394 can be seen that the applied load of 160 kgf nucleated some microcracks around the Rocwell
395 C impression and these cracks propagated outward. According to VDI 3198 test standards,
396 Rockwell C supported prints without damages, low number of fine microcracks and layer
397 fracture. Therefore, the damage mechanisms occurring in TiC coatings are similar to the HF1
398 test standard and are at an acceptable level. This showed that the obtained TiC coatings had
399 good adhesion to the substrate, as stated in the previous study for other metal carbide layers
400 [9,30,46].

33 17
34
401 Table 1. Coating layer thickness, microhardness and surface roughness values of the samples
402 subjected to titanizing process.
Naming Temperatur Duratio Thicknes Microhardnes Nanohardnes Surface
e (ºC) n s (µm) s (HV0.025) s roughnes
(h) (GPa) s
(Ra, µm )
800-4h 800 4 5±0.4 - 27.96±2.3 0.49

850-4h 850 4 8±0.3 3093±413 29.57±2.1 0.55

900-4h 900 4 11±0.4 3158±489 32.45±1.8 0.67

Untreate - - - 300±52 3.45±0.4 0.24


d SGI
403
404 It is seen that the hardness values of the obtained coatings are in the range of 27.96-32.45
405 GPa and a hardness value is reached 10 times higher than the hardness value of the untreated
406 GG-80 material (Table 1). Although it has been stated that the TiC hardness value can be
407 obtained in the range of 3200-3800, it has been determined that the hardness value of 1800-
408 3000 HV is generally reached in the studies carried out with TRD in the literature, and
409 therefore the hardness values obtained in the study are by the literature [34,43,44,48]. It was
410 determined that there was an increase between the titanizing time and the surface roughness
411 values in parallel with the increase in the titanizing temperature (Table 1). This is related to
412 the increasing diffusion rate with increasing titanization temperature.
413
414 3.3 Wear response of TiC coatings

415 The friction coefficient graphs obtained during the wear tests performed against a 6.3 mm
416 diameter Al2O3 ball are shown in Figure 9.

417

35 18
36
418
419 Figure 9. Graphs of friction coefficient of untreated GG-80 and TiC coating against Al 2O3
420 ball a)100 m b)250 m through sliding distance.

421 Although the average friction coefficient values of the TiC coated samples were lower
422 than the untreated sample at both 100 m and 250 m slip distance (Table 2), the friction
423 coefficient graphs showed variation in some cases. This shows that COF is dependent on
424 many different parameters such as the mechanical properties of the substrate, the thickness of
425 the coating layer, its hardness, toughness, surface roughness, and exhibits a dynamic change
426 [22,45,58,59]. Considering the friction coefficient graph of 100 m sliding distance (Fig. 9a), it
427 is seen that the friction coefficient of the untreated ductile cast iron sample starts from 0 and
428 reaches 0.4 in 75 seconds, followed by a stable course in the range of 75-100, and then enters
429 a downward trend. In the range of 100-200 seconds, it first went down to 0.2 level with a
430 rapid decrease and rose to 0.4 levels with a rapid increase, and followed a zig-zag course in
431 the range of 0.33-0.55 until the end of the experiment. A similar unstable friction coefficient

37 19
38
432 was observed in the experimental conditions where the sliding distance was increased to 250
433 m (Fig. 9b). The zig-zag coefficient of friction course in the untreated sample can be
434 attributed to 3 distinct regions as graphite nodules, ferrite and pearlitic region in spherical
435 graphite cast iron because it has been reported in many studies that spherical graphite nodules
436 act as a solid lubricant that reduces the friction coefficient by reducing the contact surfaces at
437 temperatures below 100 °C and low sliding distances [55, 59]. However, it has been stated
438 that the solid lubrication effect will be lost due to the delamination of spheroidal graphite
439 nodules at long sliding distances and the oxidation of graphite due to the increase in
440 temperature during wear [14,55,58,60]. In addition, the different mechanical properties of
441 spherical graphite nodules, ferrite and pearlitic phases in the structure caused the friction
442 coefficient graphs to be wavy. When the COF graph of the TiC coated samples at a sliding
443 distance of 100 m is evaluated (Fig. 9a), it is seen that depending on the roughness values on
444 the surface, there is a rapid increase from the beginning of the wear test up to 50 seconds, and
445 then the friction coefficient values have followed a decrease and stable course around 0.2 until
446 the end of the test. It is obvious that the friction coefficient values are lower and more stable
447 due to the homogeneity of the TiC coated samples. TiC coated samples showed a lower COF
448 value than the untreated ductile graphite cast iron sample in the 250 m sliding distance as in
449 the wear tests performed at 100 m sliding distance. However, the COF graphs of TiC coating
450 at 250 m sliding distance are slightly higher and more complex than the tests performed at
451 100 m sliding distance. Namely; While the 800-4h sample followed a constantly increasing
452 trend from the beginning to the end of the wear test, the 850-4h and 900-4h samples followed
453 a near stable course with a friction coefficient then a decreasing trend. The fact that the
454 friction coefficients of the TiC coated specimens first increased and then decreased at the
455 beginning of the wear tests can be explained as follows. Initially, the surface roughness and
456 brittleness of the coating layer are quickly removed from the coating surface. While some of
457 these wastes, high hardness, are removed from the environment with centrifugal force, some
458 get stuck between the ball and the substrate. Due to the high surface hardness of TiC wear
459 debris, it has an abrasive effect on coatings. This increases the coefficient of friction. When
460 the roughness and brittleness on the surface are removed, the debris from the coating layer
461 decreases and some of them are adhered to the surface, resulting in a decrease in the friction
462 coefficient values. As a matter of fact, the fact that the wear rates are higher at 100 m
463 compared to 250 m supports this claim.

39 20
40
464
465 Figure 10. a) volume losses b) wear rates of untreated and TiC coated GG-80 samples against
466 Al2O3 ball.

467 Table 2. Mean COF, wear track width, wear depth, volume loss and wear rate untreated and
468 TiC coated SGI samples subjected to the ball-on-disc wear test in dry sliding condition.

Naming Sliding distance COF Wear track Wear track Volume loss Wear rate
(m) width (µm) depth (µm) (10-3mm3) (10-5mm3/Nm)
800-4h 0.15±0.03 356±14 2.43±0.12 17.37±2.0 1.74±0.08
850-4h 0.23±0.04 515±56 1.40±0.02 14.1±1.3 1.41±0.07
900-4h 0.19±0.04 555±91 1.22±0.03 13.41±1.2 1.34±0.06
100
GG-80 0.33±0.05 467±35 3.13±0.13 28.95±2.4 2.9±0.15
800-4h 0.27±0.03 473±30 3.43±0.12 32.14±2.1 1.29±0.06
850-4h 0.47±0.06 652±36 2.13±0.23 27.43±1.9 1.1±0.06
250
900-4h 0.40±0.04 641±30 1.87±0.12 23.72±1.7 0.95±0.05
GG-80 0.51±0.05 587±55 4.73±0.06 54.17±3.2 2.19±0.09
469
470 As seen in Figure 10a and Table 2, it is seen that the wear volume losses of untreated
471 SGI and TiC coated SGI samples increase with increasing sliding distance. However, as seen
472 in the wear rate graph Figure 10b and Table 2, the wear rates of both the untreated SGI
473 sample and the TiC coated samples decreased with the increase of the sliding distance. This is
474 a result of cyclic loads causing strain hardening in the worn surface due to the increase in the

41 21
42
475 sliding distance. As the hardness values of the worn surface increase, the deformation
476 hardening provides a higher resistance of the material against the penetration of the abrasive
477 ball, resulting in a decrease in the wear rates. In addition, since the high hardness prevents the
478 abrasive ball from sinking inwards, the abrasive ball causes wear on the side surfaces against
479 the applied load. As a matter of fact, in some of the TiC coated samples, the width of the wear
480 trace volume was measured wider than the untreated sample. The fact that the wear trace
481 depths decreased with the increase in the hardness values of the samples supports this claim
482 (Table 2). This is also due to the increased area of the high-hardness TiC coating layer in
483 contact with the surface by eroding the abrasive ball. As a matter of fact, the presence of Al
484 was found in the range of 5%-36% in the EDS analyses of the TiC-coated samples (Fig. 14,
485 Fig. 16 and Fig. 18 pt3 and pt4), while the presence of Al was not found around 0.3% (Fig.
486 12) in the untreated SGI samples. Gül (2010) reported that the pearlite phase provides high
487 adhesive wear resistance due to the increasing sliding distance, as well as the wear volume
488 losses due to the hardening caused by the deformation of the ferrite phase compared to the
489 increasing sliding distance [59]. There is a similar situation in our study. While the wear rate
490 of the untreated SGI sample was 2.9 x10-5 mm3/N.m in the 100 m wear test, this value was
491 measured as 2.19 x10-5 mm3/N.m after the 250 sliding distance. However, Gül (2010) reported
492 that in case of excessive hardening of the ferrite phase, which undergoes strain hardening at
493 higher sliding distances, the wear rates might increase as a result of the delamination
494 mechanism, but the initial wear rates will be lower [59]. On the other hand, the decrease in
495 wear rates of TiC coated samples can be explained as follows. Since the applied TRD
496 methods are diffusion-based, a more disadvantageous part such as surface roughness, oxide
497 residues, and extreme brittleness can be found at the outermost part of the surface compared
498 to the inner parts of the coating. These disadvantageous parts may cause more wear volume
499 losses in the first stage of the wear process. When the main coating area is reached by
500 removing these parts, the response of the coating layers with high hardness and uniform
501 structure against the abrasive ball will be better. As a matter of fact, the wear volume ratios in
502 the 100 m wear tests were in the range of 1.74-1.34x10 -5 mm3/N.m for TiC coated samples,
503 while the wear volume ratios in the 250 m wear tests were 1.29-0.95x10 -5 mm3/N.m.
504 Therefore, the wear rates of both untreated SGI and TiC coated samples decreased with the
505 sliding distance increase.

43 22
44
506
507 Figure 11. Worn surface SEM view a)350 x, b)1000 x and c) EDS analysis of untreated GG-
508 80 sample subjected to wear testing at 100 m sliding distance.

45 23
46
509
510 Figure 12. Worn surface SEM view a)300 x, b) 1000 x, c)2000 x and d) EDS analysis of
511 untreated GG-80 sample subjected to wear testing at 250 m sliding distance.

512 As seen in the SEM image of the wear surface of the SGI samples(Figure 11a and
513 Figure 12a), the width of the wear trace formed on the surface of the untreated SGI sample
514 after the 100m wear test is around 467 µm, while this value is around 587 µm at the sliding
515 distance of 250 m. When the 1000X-2500X magnification SEM views (Fig. 11b and Fig. 12b-
516 c) were examined, it was observed that the damage mechanisms on the surface of the samples
517 mainly occurred in the spherical graphite nodules ferrite regions immediately adjacent to
518 these nodules. This is because the spherical graphite nodules and the ferrite matrix structure
519 (bull's eye structure) surrounding the spherical graphite nodules in the form of rings are of
520 lower hardness than the pearlitic structure of the general microstructure [46,59,61]. As a
521 matter of fact, as a result of the instability of the graphite phase in the ferrite phase and the
522 crack formation towards the surface, delamination occurred in the spherical graphite nodule
523 regions [59], while it is clearly seen that the ferrite phase, which is harder than graphite but
524 softer than pearlite, wears in the form of abrasive wear scratches. In addition, as can be seen
525 from the EDS analysis taken from the spherical graphite nodule regions (Figure 11 Pt2 and
526 Figure 12 Pt2), high O was detected, indicating that the oxidational wear mechanism has
527 occurred. The literature states that stress concentration at the graphite matrix interface is the

47 24
48
528 leading cause of the observed delamination in most wear studies performed on spheroidal
529 graphite cast irons [55,59,60].

530
531 Figure 13. Worn surface SEM view of 800-4h TiC coated sample subjected to wear testing at
532 100 m sliding distance a) 400 x b) 1000 x and c) 2000 x.
533
534

49 25
50
535

536
537 Figure 14. Worn surface SEM view a) 350 x b) 1000 x and c) 2000 x and d) EDS analysis of
538 800-4h TiC coated sample subjected to wear testing at 250 m sliding distance.

539 As seen in the wear traces of 800-4h TiC coated samples (Fig. 13 and Fig. 14), 355 µm
540 wear traces were formed at 100 m sliding distance, while 471 µm wear trace widths were
541 observed at 250 m sliding distance. However, similar wear marks and wear debris were
542 observed at both sliding distances. When the wear marks of the samples at 1000X
543 magnification are examined, it is clearly seen in the wear marks that they have a smoother
544 structure than TiC coatings with a rough structure, but micro-cracks, fracture and ruptures in
545 some parts. This shows that the TiC coating particles detached from the surface during the
546 wear process adhere to the surface due to the pressure force between the substrate and the
547 abrasive ball. It was observed that these structures, which were plastered during the wear
548 testing, were broken in some areas. As a matter of fact, due to this plastering, the TiC coated
549 samples followed a lower friction coefficient course than the untreated SGI samples, both at
550 100 m and 250 m sliding distance. As in the microstructure SEM analysis (Fig. 3), a Ti ratio
551 of around 30% was determined at 4 different EDS points taken from the surface of the 250 m
552 eroded sample. This shows that the TiC coatings maintain their structure in both plastered and
553 fractured areas. Another great point in EDS analysis is that around 45% O 2 was detected. This

51 26
52
554 shows that the wear mechanism occurs as an oxidative assisted by the adhesive wear
555 mechanism [62].

556
557 Figure 15. Worn surface SEM view of 850-4h TiC coated sample subjected to wear testing at
558 100 m sliding distance a) 300 x b) 1000 x and c) 2000 x.

53 27
54
559
560 Figure 16. Worn surface SEM view a) 250 x b) 1000 x and c) 2000 x and d) EDS analysis of
561 850-4h TiC coated sample subjected to wear testing at 250 m sliding distance.
562
563 When the worn surface SEM views of the 850-4h sample are examined (Fig. 15 and Fig.
564 16), it is seen that it exhibits a similar wear mechanism with 800-4h sample. However, it was
565 observed that the fracture zones on the 850-4h sample were smaller than those of the 800-4h
566 sample. Since microcracks do not reach a specific threshold value, it is thought that fracture
567 did not occur. The presence of microcracks in the layers due to cyclic loading can be clearly
568 seen in the sample where the sliding distance is 250 m (Fig. 16b). Fatigue-type wear damage
569 occurs when microcracks that occur due to cyclic loads exceed a certain threshold
570 (progression) [63,64]. However, it is understood that the damage in the samples remained
571 mainly in the form of micro-cracks and did not turn into a fatigue-type wear mechanism. The
572 high O2 ratio in detected in Fig. 16 Pt4 and Pt3 regions also shows that the oxidation wear
573 assists the wear mechanism. Moreover, Al was detected in EDS analysis taken from Pt3 and
574 Pt4 regions. This shows that due to the high hardness of the TiC coatings, it wore the abrasive
575 ball somewhat. As a result of the EDS analysis taken, it means that the TiC coating continues
576 to exist in the structure and a lubricating layer is formed due to the reaction of Ti with O 2 in
577 the environment. The surface of the sample subjected to the 100 m abrasion test is more
578 damaged than the sample subjected to the 250 m abrasion test due to surface roughness,

55 28
56
579 porosity, etc., in the near-surface coating layer. This is because the weakest parts of the
580 coating layer are initially worn at a higher rate due to brittleness. With the increase of the
581 sliding distance, the surface appearance was better due to the removal of these defective areas
582 and the minor abrasion of the more homogeneous coatings in the inner parts of the coating.

583
584 Figure 17. Worn surface SEM view a) 250 x b) 1000 x and c) 2000 x of 900-4h TiC coated
585 sample subjected to wear testing at 100 m sliding distance.
586

57 29
58
587
589 Figure 18. Worn surface SEM view a) 250 x b) 1000 x and c) 2000 x and d) EDS analysis of
590 900-4h TiC coated sample subjected to wear testing at 250 m sliding distance.

591 When the worn surface SEM views of the 900-4h sample are examined, it is seen that it
592 has a different appearance than the 800-4h and 850-4h samples. In the SEM images of the
593 worn surface, a region with three different concentrations of gray, black and brightly is seen.
594 The EDS analysis shows that the gray region (Pt1) contains 52% Ti, 21% C, and 21% O 2, so it
595 is the TiC coating region. In the EDS analysis, the black regions (Pt3 and pt4) were found to
596 be composed of high O2 around 56%, low Ti (1.16-17.20) and high Al (9-36%). The bright
597 region (Pt 2) is composed of 48% Ti, 28.5% O 2 and 16.86% C, so it is thought that this region
598 is a broken TiC coating. Therefore, as a wear mechanism, it can be said that this region is
599 exposed to oxidative-supported adhesive wear, as in the other 800-4h and 850-4h samples.
600 The Al ratio was determined as high as 36% in some regions means that the Al2O3 abrasive
601 ball suffered some damage due to the high hardness of the TiC coating. It has been reported in
602 the literature that the highest wear resistance will be obtained when both oxidation and
603 adhesion wear mechanisms are active [65,66].

604

605

59 30
60
606 3.4 Electrochemical corrosion behavior of TiC coating

607 OCP and potentiodynamic polarization scans were conducted for TiC coatings and
608 untreated SGI in 3.5% NaCl solution. The results are presented in Fig. 19. The corrosion
609 potential Ecorr, corrosion current density Icorr and corrosion rate values were obtained from
610 the polarization curves (Fig. 19b) and are listed in Table 3. OCP measurement is a
611 thermodynamic parameter. Although it does not give information about corrosion resistance
612 but gives information on corrosion susceptibility. Also, more positive OCP values indicate
613 resistance to corrosion [31,67].

614 The OCP curves of the samples tested in 3.5% NaCl in Figure 19a show that the TiC
615 coating samples start at relatively higher OCP values than the untreated SGI sample. When
616 the TiC coated samples are compared among themselves, it is seen that the OCP diagrams
617 shift towards the positive side depending on the increase in the titanizing temperature, and
618 thus the corrosion potentials decrease. This results from the increased coating thickness
619 creating more continuous coating layers on the surface, as seen in the SEM views (Fig. 4 –
620 Fig. 6). Thus, the corrosive liquid is prevented to reach the substrate. The polarization curves
621 in Figure 19b showed that TiC coatings are potentially located on the more positive side and
622 logarithmic currents on the more negative side (Fig. 19b). Compared to untreated SGI, TiC
623 coatings are noted to provide a marked improvement in corrosion resistance. The increase in
624 corrosion resistance can be attributed to the fact that the microstructure of spheroidal graphite
625 cast iron, which consists of ferrite, pearlite and graphite structures with different potentials,
626 has been transformed into a single structure in the form of TiC. These results are consistent
627 with the corrosion trend sequence we obtained previously in OCP measurements. Also
628 corrosion resistance of the TiC coatings increased with the treatment temperature. Therefore,
629 TiC coating showed higher corrosion resistance than the untreated SGI sample, as Table 3.

61 31
62
630
631 Figure 19. The electrochemical response of untreated GG-80 and TiC coated samples in wt.
632 3.5% NaCl solution a) OCP b) Tafel plots.

633 The corrosion rates in Table 3 confirm that the TiC coating improves the corrosion
634 resistance; the corrosion rates in the untreated SGI sample were 12.10x10-6, whereas the
635 corrosion rates for the TiC coated samples range from 3.368x10-6 to 7.74x10-6 . When the
636 corrosion resistance of the TiC coating are compared among themselves, it can be seen that
637 800-4h and 850-h samples showed close I corr values, while 900-4h samples showed lower than
638 800-4h and 850-4h samples. This arises because specific coating properties, such as
639 continuity on the surface, O2 content, coating layer thickness, porosity, and microcracks, are
640 different. Different studies showed that these coating defects cause Cl ions enable to leak to
641 the substrate or coating layer which increased anodic dissolution and pitting susceptibility in
642 corrosion and tribocorrosion conditions [68-69]. Since the corrosion resistance of the coating
643 depends on the ratio of pores and microcracks in the coating, scatter in the corrosion test data
644 is unavoidable [70-72]. Because the penetration of the corrosive liquid into the substrate and
645 the interaction of corrosion products between the substrate and the substrate occur through

63 32
64
646 defects such as cracks and pores on the surface [73]. The best corrosion resistance was
647 obtained in the 900-4h sample, and 3.59 times better corrosion resistance was obtained than
648 the untreated sample. The results indicate that the TiC coating serves as a barrier against
649 corrosion and that increasing the thickness of the TiC coating further improves the corrosion
650 resistance. It has been noted in many thermochemical coating studies that thicker coatings
651 provide better protection unless microcracks occur in coatings due to increased thickness
652 [4,54,72,74]. Since galvanic corrosion formation is prevented by the conversion of spheroidal
653 graphite nodules in the structure to TiC coating, corrosion resistance increased 3.59 times.

654 Table 3. The data obtained from the Tafel curves were recorded at 3.5 wt.% NaCl solution.
Sample Solution Ecorr (V) Icorr (µAcm-2) Corrosion rate (gr/h)
800-4h -0.535 6.131x10-6 6.405x10-6
850-4h -0.480 7.413x 10-6 7.740x10-6
3.5% NaCl
900-4h -0.433 3.224x 10-6 3.368x10-6
Untreated SGI -0.659 11.58x10-6 12.10x10-6
655
656 Figures 20-23 show SEM images and EDS analysis obtained from TiC coating
657 (performed at various temperatures) and untreated SGI samples after being subjected to an
658 electrochemical corrosion test in 3.5 wt.% NaCl solution.

659
660 Figure 20. SEM surface view and EDS analysis of the electrochemical corrosion test area of
661 the untreated GG-80 sample in 3.5% NaCl solution.
662
663 It is verified that the corroded surface of the untreated SGI sample reveals signs of
664 extensive surface damage, signs of pitting and corrosion residues are evident in samples tested

65 33
66
665 in 3.5 wt.% NaCl (Figure 20a). The pre-existing spherical graphite nodules in the structure
666 may have served as starting points for galvanic corrosion [4]. Because the chemical difference
667 between the iron matrix (ferrite and pearlite) and the spheroidal graphite nodules triggers
668 galvanic corrosion in the structure [4,62,75]. EDS analysis of the corroded surface of
669 untreated SGI (yellow rectangular) indicated the presence of Fe, O, C, and Si. As seen in
670 these regions, it is clearly seen that corrosion starts from spheroidal graphite nodules (Pt2). As
671 a matter of fact, in this region (Pt2), 55% O and 17% C ratios were determined, and the Fe
672 ratio was around 25%.

673
674 Figure 21. SEM surface view and EDS analysis of the electrochemical corrosion test area of
675 the titanized 800-4h in 3.5% NaCl solution.

676 On the other hand, SEM views taken from the corroded surfaces of the TiC coated
677 samples showed a less damaged surface appearance than the untreated SGI sample (Fig.21-
678 23). In some parts of the 800-4h sample, it is observed that the coating layer is broken and
679 pitting-type corrosion mechanisms occur in these regions (Fig. 21). However, the presence of
680 Ti close to each other (around 30%) was detected in the EDS analysis taken from the pitting
681 region (Pt3), as well as the entire corroded surface (Pt1) and the (Pt2) the region that
682 maintains the continuity of the TiC coating. This indicates that the pitting formations are
683 within the TiC coating and that the corrosive liquid has not yet reached the substrate.

684
685

67 34
68
686
687 Figure 22. SEM surface view and EDS analysis of the electrochemical corrosion test area of
688 the titanized 850-4h in 3.5% NaCl solution.

689 In the SEM views of the 850-4 TiC coated sample, on the other hand, it has a more un-
690 homogeneous surface appearance. As a matter of fact, it was the most corroded sample among
691 TiC coatings. In EDS analysis, Ti was detected around 26% and 38% in Pt1 and Pt3 regions,
692 respectively. On the other hand, 61% Ti was detected in the Pt2 region. Therefore, it can be
693 argued that the inhomogeneity in the structure increases the corrosion resistance.

694
695 Figure 23. SEM surface view and EDS analysis of the electrochemical corrosion test area of
696 the titanized 900-4h in 3.5% NaCl solution.
697

69 35
70
698 SEM views of 900-4h TiC coating showed minor pitting formation and corrosion
699 residue in the corroded areas on the surfaces. Moreover, It was determined that an increase in
700 titanizing temperature increased the Ti amount and a decrease in the O 2 amount on the
701 corroded surface. As a matter of fact, 68% Ti ratio was determined in the Pt2 region, while
702 the O2 rate in this region was 0%. The high level of Ti and C on the surface after the corrosion
703 tests shows that the TiC coatings remained after the tests for all titanizing conditions.

704 Briefly, TiC coatings showed better corrosion resistance than the untreated SGI sample,
705 as seen from the corrosion tests. As this is shown in figure 3-5, there is a high Ti
706 concentration (up to 60% atm.) on the surface of the TiC coating; this high Ti content retains
707 itself after corrosion and as seen in the EDS analysis taken from the corroded areas.

708 While the corrosion behavior of alloys largely depends on the properties of the surface,
709 in coated alloys, the surface properties of the coatings, as well as the solid-state properties, are
710 also determinative on corrosion resistance. Namely, lattice energy, bond energies, the heat of
711 formation, bandgap etc., of coatings solid-state properties significantly affect the overall
712 dissolution rate during the corrosion process [75]. Structural heterogeneities such as pits,
713 defects, cracks, and other discontinuities in the coating layers also affect the anodic
714 dissolution behavior [76]. Many scholars have reported that layers such as TiO 2, SiO2, Cr2O3
715 and Al2O3, which have very low electronic conductivity or insulation properties, effectively
716 protect the alloys from the corrosive environment [77-79].

717 Although the corrosion resistance of TiC coating has not been studied so far, there are
718 studies in the literature that different carbides improve the corrosion resistance of cast irons.
719 Mariani et al. (2021) to improve wear and corrosion resistance of vermicular cast iron applied
720 Nb-V coating proses by TRD method [46]. The authors analyzed the Nb-V layers employing
721 potentiodynamic polarization in 3.5 wt.% NaCl solution as in our study. The authors
722 concluded that the NbC, NbVC2 and VC layers improved vermicular cast iron corrosion
723 resistance by presenting a greater passive region. In another study, Günen et al. (2020)
724 reported that CrVC coatings formed on the surface of spheroidal graphite cast irons improved
725 the corrosion resistance in corrosive environments of 3.5% NaCl, 5% HCl and 5% H 2SO4 as a
726 result of converting the spheroid graphite nodules in the structure to carbides [4]. As
727 understood from the studies in the literature and our study, the way to increase the corrosion
728 resistance in cast irons is provided by converting the free graphite into carbides and
729 eliminating the phase differences in the microstructure.

71 36
72
730 4. Conclusions
731 TiC coatings were successfully grown on the surface of a spheroidal graphite cast iron.
732 The microstructures, hardness, adhesion, tribological and corrosion behavior of the coatings
733 were investigated. In the light of the results obtained, the following conclusion can be
734 reached:
735 1- Due to the high C content of spheroidal graphite cast irons, unlike steels, TiC coatings
736 were obtained at lower temperatures (800–900 °C).
737 2- Since the TRD process is a diffusion-controlled process, TiC coating thicknesses changed
738 between 5-11 µm thickness and 27.96-32.45 GPa hardness values depending on the
739 increasing TRD temperature. The TiC coatings formed were smooth, continuous, and
740 microcrack-free, with TiC phase structure, and SGI was not observed in the coating layer
741 as spheroidal graphite nodules (C) dissolved during the TRD process and combined with
742 Ti to form TiC.
743 3- TiC coatings exhibited a lower friction coefficient and higher wear resistance than
744 untreated SGI, thanks to their high hardness, good substrate adhesion, and inert structure.
745 Abrasive wear was dominant in the untreated SGI, but signs of delamination and
746 oxidation were also observed. The effective wear mechanism in the TiC coatings was
747 adhesive wear and oxidation.
748 4- Transforming the surface of SGI, which consists of ferrite, pearlite, graphite structures
749 with different potentials, into a single structure in the form of TiC provided an
750 improvement in corrosion resistance.
751 5- The spheroidal graphite nodules, which often serve as the initiation sites for corrosion and
752 wear in spheroidal graphite cast irons, reacted with titanium to form the TiC coatings and
753 the toughness of the core was maintained as it preserved its spheroidal graphite
754 microstructure. It can be argued that the corrosion and wear resistance of SGIs can be
755 improved without compromising ductility and toughness by TRD treatments.
756 6- The great advantage of TiC coatings in terms of adhesion, wear and corrosion indicates
757 them for use in mechanical components subjected to tribo-corrosion environments.
758
759
760
761
762
763

73 37
74
764
765 Acknowledgment
766
767 This article is derived from the results of the master thesis study “Wear and Corrosion
768 Behavior of Titanium Carbide Coatings Grown on the Surface of Ductile Iron” conducted by
769 Betül Soylu under the supervision of Ali GÜNEN and Özgür Karakaş at the Metallurgical and
770 Materials Engineering Department of the Graduate Studies of Iskenderun Technical
771 University.
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797

75 38
76
798
799
800 References
801 [1] ASM international, Ductile Iron Handbook, American Foundry Society, Published: June
802 10, 2010, Pages: 277, ISBN: 978-0-87433-364-0
803 [2] Elorz, J. A. P. S., González, D. F., & Verdeja, L. F. (2018). Physical metallurgy of cast
804 irons. Springer.
805 [3] Labrecque, C., & Gagne, M. (1998). Ductile iron: Fifty years of continuous
806 development. Canadian Metallurgical Quarterly, 37(5), 343-378.
807 [4] Günen, A., Kalkandelen, M., Karahan, İ. H., Kurt, B., Kanca, E., Gök, M. S., & Serdar
808 Karakaş, M. (2020). Properties and Corrosion Behavior of Chromium and Vanadium
809 Carbide Composite Coatings Produced on Ductile Cast Iron by Thermoreactive
810 Diffusion Technique. Journal of Engineering Materials and Technology, 142(4).
811 [5] Angus, H. T. (2013). Cast iron: physical and engineering properties. Elsevier.
812 [6] Zhi, X., Xing, J., Fu, H., & Xiao, B. (2008). Effect of niobium on the as-cast
813 microstructure of hypereutectic high chromium cast iron. Materials letters, 62(6-7), 857-
814 860.
815 [7] Babakr, A. M., Al-Ahmari, A., Al-Jumayiah, K., & Habiby, F. (2008). Sigma phase
816 formation and embrittlement of cast iron-chromium-nickel (Fe-Cr-Ni) alloys. Journal of
817 minerals and materials characterization and engineering, 7(02), 127.
818 [8] Wu, X. (2020). Microstructure-Fatigue Relationships for Cast Irons (No. 2020-01-0187).
819 SAE Technical Paper.
820 [9] Soares, C., Mariani, F. E., Casteletti, L. C., Lombardi, A. N., & Totten, G. E. (2017).
821 Characterization of Niobium Carbide Layers Produced in Ductile Cast Iron Using
822 Thermo-Reactive Treatments. Materials Performance and Characterization, 6(4), 607-
823 616.
824 [10] Zeng, D., Lu, L., Zhang, N., Zhang, Y., & Zhang, J. (2017). Investigation on the scuffing
825 resistance of ductile cast iron as affected by fine particle bombardment to produce
826 surface hardened layer and micro-dimpled surface. Wear, 378, 174-182.
827 [11] Weng, Z., Wang, A., Wang, Y., Xiong, D., & Tang, H. (2016). Diode laser cladding of Fe-
828 based alloy on ductile cast iron and related interfacial behavior. Surface and Coatings
829 Technology, 286, 64-71.

77 39
78
830 [12] Li, Y., Liu, X., Dong, S., Ren, X., Yan, S., & Xu, B. (2021). Influence of laser power on
831 interface characteristics and cracking behavior during laser remanufacturing of nodular
832 cast iron. Engineering Failure Analysis, 122, 105226.
833 [13] Milne, I., Ritchie, R. O., & Karihaloo, B. L. (Eds.). (2003). Comprehensive structural
834 integrity: Cyclic loading and fatigue (Vol. 5). Chapter 5: Thermal–Mechanical Fatigue
835 (Including Thermal Shock), 2003, Elsevier.
836 [14] Menezes, P. L., Nosonovsky, M., Ingole, S. P., Kailas, S. V., & Lovell, M. R. (Eds.).
837 (2013). Tribology for scientists and engineers (pp. 3-4). New York: Springer.
838 [15] Zou, C. L., Pang, J. C., Zhang, M. X., Qiu, Y., Li, S. X., Chen, L. J., ... & Zhang, Z. F.
839 (2018). The high cycle fatigue, deformation and fracture of compacted graphite iron:
840 influence of temperature. Materials Science and Engineering: A, 724, 606-615.
841 [16] Asi, O. (2006). Failure analysis of a crankshaft made from ductile cast iron. Engineering
842 Failure Analysis, 13(8), 1260-1267.
843 [17] Witek, L., Sikora, M., Stachowicz, F., & Trzepiecinski, T. (2017). Stress and failure
844 analysis of the crankshaft of diesel engine. Engineering Failure Analysis, 82, 703-712.
845 [18] Forrest, P. G. (2013). Fatigue of metals. Elsevier.
846 [19] Xu, Z. Y., & Jiang, Y. J. (2012). Study on the Thermal Fatigue Resistance of Ductile Iron.
847 In Advanced Materials Research (Vol. 512, pp. 2093-2096). Trans Tech Publications
848 Ltd.
849 [20] Konca, E., Tur, K., & Koç, E. (2017). Effects of alloying elements (Mo, Ni, and Cu) on
850 the austemperability of GGG-60 ductile cast iron. Metals, 7(8), 320.
851 [21] Çelik, G. A., Tzini, M. I. T., Polat, Ş., Atapek, Ş. H., & Haidemenopoulos, G. N. (2020).
852 Thermal and microstructural characterization of a novel ductile cast iron modified by
853 aluminum addition. International Journal of Minerals, Metallurgy and Materials, 27(2),
854 190-199.
855 [22] Boulifa, I., & Hadji, A. (2021). Study of the influence of alloying elements on the
856 mechanical characteristics and wear behavior of a ductile cast iron. Frattura ed Integrità
857 Strutturale, 15(56), 74-83.
858 [23] Elosegui, I., Alonso, U., & de Lacalle, L. N. L. (2017). PVD coatings for thread tapping
859 of austempered ductile iron. The International Journal of Advanced Manufacturing
860 Technology, 91(5), 2663-2672.
861 [24] Kostylev, A., Pokrovsky, Y., Lumpov, A., & Bryskin, B. (2012). Advanced Chromium
862 Carbide Coatings on Piston Rings by CVD: A Highly Adaptable new method with
863 relatively low cost. Advanced materials & processes, 170(7), 22-27.

79 40
80
864 [25] Ksiazek, M., Boron, L., & Tchorz, A. (2019). Microstructure, Mechanical Properties and
865 Wear Behavior of High-Velocity Oxygen-Fuel (HVOF) Sprayed (Cr3C2-NiCr+ Al)
866 Composite Coating on Ductile Cast Iron. Coatings, 9(12), 840.
867 [26] Panneerselvam, S., Putatunda, S. K., Gundlach, R., & Boileau, J. (2017). Influence of
868 intercritical austempering on the microstructure and mechanical properties of
869 austempered ductile cast iron (ADI). Materials Science and Engineering: A, 694, 72-80.
870 [27] Akinribide, O. J., Akinwamide, S. O., Ajibola, O. O., Obadele, B. A., oluwagbenga
871 Olusunle, S. O., & Olubambi, P. A. (2019). Corrosion behavior of ductile and
872 austempered ductile cast iron in 0.01 M and 0.05 M NaCl Environments. Procedia
873 Manufacturing, 30, 167-172.
874 [28] Azouani, O., Keddam, M., Allaoui, O., & Sehisseh, A. (2017). Characterization of boride
875 coatings on a ductile cast iron. Protection of Metals and Physical Chemistry of
876 Surfaces, 53(2), 306-311.
877 [29] Kayali, Y., & Yalcin, Y. (2011). The effects of boro-tempering heat treatment on
878 microstructural properties of ductile iron. Materials & Design, 32(3), 1414-1419.
879 [30] Su, X., Zhao, S., Sun, H., Yang, X., Zhang, P., & Xie, L. (2019). Chromium carbide
880 coatings produced on ductile cast iron QT600-3 by thermal reactive diffusion in fluoride
881 salt bath: Growth behavior, microstructure evolution and kinetics. Ceramics
882 International, 45(1), 1196-1201.
883 [31] Günen, A., Kalkandelen, M., Karahan, İ. H., Kurt, B., Kanca, E., Gök, M. S., & Serdar
884 Karakaş, M. (2020). Properties and corrosion behavior of chromium and vanadium
885 carbide composite coatings produced on ductile cast Iron by thermoreactive diffusion
886 technique. Journal of Engineering Materials and Technology, 142(4), 041008.
887 [32] Arai, T. (2015). The thermo-reactive deposition and diffusion process for coating steels to
888 improve wear resistance. In Thermochemical surface engineering of steels (pp. 703-
889 735). Woodhead Publishing.
890 [33] Ganji, O., Sajjadi, S. A., Yang, Z. G., Mirjalili, M., & Najari, M. R. (2020). On the
891 formation and properties of chromium carbide and vanadium carbide coatings produced
892 on W1 tool steel through thermal reactive diffusion (TRD). Ceramics
893 International, 46(16), 25320-25329.
894 [34] Kurt, B., Günen, A., Kanca, Y., Koç, V., Gök, M. S., Kırar, E., & Askerov, K. (2018).
895 Properties and tribologic behavior of titanium carbide coatings on AISI D2 steel
896 deposited by thermoreactive diffusion. Jom, 70(11), 2650-2659.

81 41
82
897 [35] Christiansen, T. L., Bottoli, F., Dahl, K., Gammeltoft-Hansen, N. B., Laursen, M. B., &
898 Somers, M. A. (2017). Hard surface layers by pack boriding and gaseous thermo-
899 reactive deposition and diffusion treatments. Materials Performance and
900 Characterization, 6(4), 475-491.
901 [36] Monteleone, C., Poges, S., Petroski, K., Kerns, P., Tobin, Z., Policandriotes, T., & Suib,
902 S. L. (2020). Atmospheric pressure chemical vapor infiltration of a titanium carbide
903 interphase coating on carbon fiber. Ceramics International, 46(10), 15084-15091.
904 [37] Pierson, H. O. (1996). Handbook of refractory carbides and nitrides: properties,
905 characteristics, processing and applications. William Andrew.
906 [38] Rubshtein, A. P., Vladimirov, A. B., Korkh, Y. V., Ponosov, Y. S., & Plotnikov, S. A.
907 (2017). The composition, structure and surface properties of the titanium-carbon
908 coatings prepared by PVD technique. Surface and Coatings Technology, 309, 680-686.
909 [39] Zhang, Y., Wang, Y., Zhang, J., Liu, Y., Yang, X., & Zhang, Q. (2015). Micromachining
910 features of TiC ceramic by femtosecond pulsed laser. Ceramics International, 41(5),
911 6525-6533.
912 [40] Balasubramanyam, N., Prasanthi, S. G., & Yugandhar, M. (2015). Study of coated TiN
913 and TiC on cutting tools for the PVD and CVD coated tungsten carbide by sand blasting
914 pretreatment of nickel and carbon. International Journal of Advanced Science and
915 Technology, 75, 51-58.
916 [41] von Fieandt, L., Johansson, K., Lindahl, E., Larsson, T., Boman, M., & Rehnlund, D.
917 (2018). Corrosion properties of CVD grown Ti (C, N) coatings in 3.5 wt-% NaCl
918 environment. Corrosion Engineering, Science and Technology, 53(4), 316-320.
919 [42]- Mohazzab, B. F., Jaleh, B., Fattah-alhosseini, A., Mahmoudi, F., & Momeni, A. (2020).
920 Laser surface treatment of pure titanium: Microstructural analysis, wear properties, and
921 corrosion behavior of titanium carbide coatings in Hank's physiological
922 solution. Surfaces and Interfaces, 20, 100597.
923 [43] Zhang, J., Li, S., Lu, C., Sun, C., Pu, S., Xue, Q., ... & Huang, M. (2019). Anti-wear
924 titanium carbide coating on low-carbon steel by thermo-reactive diffusion. Surface and
925 Coatings Technology, 364, 265-272.
926 [44] Kurt, B., Sinoplu, O., Carboga, C., & Demirel, B. (2014). The investigation and growth
927 kinetics of TiC coatings on AISI D3 steel produced by thermo-reactive diffusion
928 technique. Practical Metallography, 51(2), 95-106.
929 [45] Taktak, S., & Ulu, S. (2010). Wear behaviour of TRD carbide coatings at elevated
930 temperatures. Industrial lubrication and tribology.

83 42
84
931 [46] Mariani, F. E., Takeya, G. S., Lombardi, A. N., Picone, C. A., & Casteletti, L. C. (2020).
932 Wear and corrosion resistance of Nb-V carbide layers produced in vermicular cast iron
933 using TRD treatments. Surface and Coatings Technology, 397, 126050.
934 [47] Mariani, F. E., Rêgo, G. C., Bonella, P. G., Neto, A. L., Totten, G. E., & Casteletti, L. C.
935 (2020). Wear resistance of niobium carbide layers produced on gray cast iron by
936 thermoreactive treatments. Journal of Materials Engineering and Performance, 1-7.
937 [48] Xue, Q., Li, J., Fan, P., Xin, R., & Zhang, J. (2015, July). Study on corrosion and wear
938 resistance of titanizing coating on steel surface. In The Twenty-fifth International Ocean
939 and Polar Engineering Conference. International Society of Offshore and Polar
940 Engineers.
941 [49] Wu, Q., Li, W., & Zhong, N. (2011). Corrosion behavior of TiC particle-reinforced 304
942 stainless steel. Corrosion science, 53(12), 4258-4264.
943 [50] Çetin, M., Günen, A., Kalkandelen, M., & Karakaş, M. S. (2021). Microstructural, wear
944 and corrosion characteristics of boronized AISI 904L superaustenitic stainless
945 steel. Vacuum, 187, 110145.
946 [51] Deng, D. W., Wang, C. G., Liu, Q. Q., & Niu, T. T. (2015). Effect of standard heat
947 treatment on microstructure and properties of borided Inconel 718. Transactions of
948 Nonferrous Metals Society of China, 25(2), 437-443.
949 [52] Lin, N., Zhao, L., Liu, Q., Zou, J., Xie, R., Yuan, S., ... & Tang, B. (2019). Preparation of
950 titanizing coating on AISI 316 stainless steel by pack cementation to mitigate surface
951 damage: Estimations of corrosion resistance and tribological behavior. Journal of
952 Physics and Chemistry of Solids, 129, 387-400.
953 [53] Verein Deutscher Ingenieure Normen. (1991). VDI 3198.
954 [54] Günen, A. (2020). Properties and corrosion resistance of borided AISI H11 tool
955 steel. Journal of Engineering Materials and Technology, 142(1).
956 [55] Abedi, H. R., Fareghi, A., Saghafian, H., & Kheirandish, S. H. (2010). Sliding wear
957 behavior of a ferritic–pearlitic ductile cast iron with different nodule
958 count. Wear, 268(3-4), 622-628.
959 [56] Emamverdian, A. A., Sun, Y., Cao, C., Pruncu, C., & Wang, Y. (2021). Current failure
960 mechanisms and treatment methods of hot forging tools (dies)-a review. Engineering
961 Failure Analysis, 129, 105678.
962 [57] Rodriguez, F. J., Boccardo, A. D., Dardati, P. M., Celentano, D. J., & Godoy, L. A.
963 (2018). Thermal expansion of a Spheroidal Graphite Iron: A micromechanical
964 approach. Finite Elements in Analysis and Design, 141, 26-36.

85 43
86
965 [58] Günen, A., Kanca, E., Karakaş, M. S., Gök, M. S., Kalkandelen, M., Kurt, B., ... &
966 Karahan, I. H. (2021). Effect of thermal degradation on the properties and wear
967 behavior of Cr− V− C composite coatings grown on ductile iron. Surface and Coatings
968 Technology, 419, 127305.
969 [59] Gul, F. (2010). Effect of boronizing surface treatment on the adhesive wear behavior of
970 ferritic-pearlitic ductile iron, J. Fac. Eng. Arch. Gazi Univ. vol 25, No 2, 389-395.
971 [60] Günen, A., Kalkandelen, M., Gök, M. S., Kanca, E., Kurt, B., Karakaş, M. S., ... & Cetin,
972 M. (2020). Characteristics and high temperature wear behavior of chrome vanadium
973 carbide composite coatings produced by thermo-reactive diffusion. Surface and
974 Coatings Technology, 402, 126402.
975 [61] Gonzaga, R. A. (2013). Influence of ferrite and pearlite content on mechanical properties
976 of ductile cast irons. Materials Science and Engineering: A, 567, 1-8.
977 [62] Sun, C., Xue, Q., Zhang, J., Wan, S., Tieu, A. K., & Tran, B. H. (2018). Growth behavior
978 and mechanical properties of Cr-V composite surface layer on AISI D3 steel by thermal
979 reactive deposition. Vacuum, 148, 158-167.
980 [63] Sosnovskiy, L. A. (2010). Tribo-fatigue: wear-fatigue damage and its prediction. Springer
981 Science & Business Media.
982 [64] Erdogan, A., Yener, T., Doleker, K. M., Korkmaz, M. E., & Gök, M. S. (2021). Low-
983 Temperature Aluminizing Influence On Degradation of Nimonic 80A Surface:
984 Microstructure, Wear and High Temperature Oxidation Behaviors. Surfaces and
985 Interfaces, 101240.
986
987 [65] Kulka, M., Mikolajczak, D., Makuch, N., Dziarski, P., & Miklaszewski, A. (2016). Wear
988 resistance improvement of austenitic 316L steel by laser alloying with boron. Surface
989 and Coatings Technology, 291, 292-313.
990 [66] Moghaddam, P. V., Hardell, J., Vuorinen, E., & Prakash, B. (2020). Effect of retained
991 austenite on adhesion-dominated wear of nanostructured carbide-free bainitic
992 steel. Tribology International, 150, 106348.
993 [67] Karahan, İ. H. (2019). Effect of current density on the corrosion protection performance
994 of polyaniline coated AISI 4140 steel. Transactions of the IMF, 97(1), 48-52.
995 [68] Alkan, S., & Gök, M. S. (2021). Effect of sliding wear and electrochemical potential on
996 tribocorrosion behaviour of AISI 316 stainless steel in seawater. Engineering Science
997 and Technology, an International Journal, 24(2), 524-532.

87 44
88
998 [69] Alkan, S., & Gök, M. S. (2021). Influence of plasma nitriding pre‐treatment on the
999 corrosion and tribocorrosion behaviours of PVD CrN, TiN and AlTiN coated AISI 4140
1000 steel in seawater. Lubrication Science.
1001 [70] Castillejo, F. E., Marulanda, D. M., Olaya, J. J., & Alfonso, J. E. (2014). Wear and
1002 corrosion resistance of niobium–chromium carbide coatings on AISI D2 produced
1003 through TRD. Surface and Coatings Technology, 254, 104-111.
1004 [71] Campos, I., Palomar, M., Amador, A., Ganem, R., Martinez, J. (2006). Evaluation of the
1005 corrosion resistance of iron boride coatings obtained by paste boriding process, Surface
1006 & Coatings Technology, V. 201, p. 2438-2442.
1007 [72] Kayali, Y., & Anaturk, B. (2013). Investigation of electrochemical corrosion behavior in
1008 a 3.5 wt.% NaCl solution of boronized dual-phase steel. Materials & Design, 46, 776-
1009 783.
1010 [73] Attarzadeh, N., Molaei, M., Babaei, K., & Fattahalhosseini, A. (2021). New Promising
1011 Ceramic Coatings for Corrosion and Wear Protection of Steels: A Review. Surfaces and
1012 Interfaces, 100997.
1013 [74] Wang, Q. Y., Behnamian, Y., Luo, H., Wang, X. Z., Leitch, M., Zeng, H., & Luo, J. L.
1014 (2017). Anticorrosion performance of chromized coating prepared by pack cementation
1015 in simulated solution with H2S and CO2. Applied Surface Science, 419, 197-205.
1016 [75] Padhy, N., Kamal, S., Chandra, R., Mudali, U. K., & Raj, B. (2010). Corrosion
1017 performance of TiO2 coated type 304L stainless steel in nitric acid medium. Surface and
1018 Coatings Technology, 204(16-17), 2782-2788.
1019 [76] Courtright, E. L. (1987). The need for microstructural control in coatings exposed to
1020 severe environments. Surface and Coatings Technology, 33, 327-340.
1021 [77] Kazarin, D. A., Volkotrub, N. P., & Prilutskii, M. I. Effect of the disp ersion of charge
1022 materials on aluminothermic p rocesses in melting of ferrotitanium. No. 1 Volume 12
1023 2014, 65.
1024 [78] Pöttgen, R., Huppertz, H., & Hoffmann, R. D. (2008). Structural Chemistry of
1025 Ceramics. Ceramics Science and Technology, Volume 1: Structures, 71.
1026 [79] Tiwari, A., & Hihara, L. H. (2010). Novel silicone ceramer coatings for aluminum
1027 protection. High Performance Coatings for Automotive and Aerospace Industries, Nova
1028 Publishers, New York, 413.

89 45
90

You might also like