Next Article in Journal
Analysis of Severe Spinal Injuries in Korean Elite Female Wrestlers
Previous Article in Journal
Research on an Evaluation Method of Snowdrift Hazard for Railway Subgrades
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Influence of Variable Plasma Welding Parameters on Weld Geometry, Dilution Factor, and Microhardness

by
Sylwia Bazychowska
,
Katarzyna Panasiuk
* and
Robert Starosta
Faculty of Marine Engineering, Gdynia Maritime University, 81-225 Gdynia, Poland
*
Author to whom correspondence should be addressed.
Appl. Sci. 2024, 14(16), 7248; https://doi.org/10.3390/app14167248 (registering DOI)
Submission received: 5 June 2024 / Revised: 28 June 2024 / Accepted: 15 August 2024 / Published: 17 August 2024
(This article belongs to the Section Mechanical Engineering)

Abstract

:
Weld surfacing is the process of applying a layer of metal to the surface of metal objects by simultaneously melting the substrate. As a result of this process, the metal content of the padding weld can be as high as several tens of percents. It is a method used to regenerate machine parts and improve the properties of the surface layer, increasing its resistance to abrasion, corrosion, erosion, and cavitation. It also supports the repair and creation of permanent protective coatings in the engineering, automotive, energy, and aerospace industries. This makes it possible to repair damaged parts instead of completely replacing them, saving time and production costs. Plasma surfacing technology is used for components that require high hardness and corrosion resistance under various environmental conditions. Plasma wire surfacing is not sufficiently presented and described in the current literature, which creates problems in determining the appropriate process parameters. The influence of variable plasma surfacing parameters on steel C45 significantly affects surfacing weld geometry, the dilution factor, and microhardness. Higher currents can increase the dilution factor, integrating more base metal into the weld pool, which may alter the chemical composition and mechanical properties of the weld. Variations in surfacing speed and heat input also affect the microhardness of the surfaced joint, with higher heat inputs potentially leading to softer welds due to slower cooling rates. Optimizing these parameters is essential to achieving desired surfacing weld characteristics and ensuring the structural integrity of C45 steel joints. This paper presents the influence of varying plasma surfacing parameters on the surfacing geometry, the dilution factor, and microhardness. The tests were carried out on a Panasonic TM-1400 GIII automated surfacing machine with CastoMag 45554S solid wire as the filler material. Flat bars of C45 steel were prepared, and then the variable parameters of the surfacing process were developed. Tests were carried out to determine the dilution factor, followed by microhardness measurements. The results showed a significant dependence of the effect of the parameters on the surfacing geometry and the dilution factor.

1. Introduction

Weld surfacing is defined as a method of applying a layer of metal to the surface of metal objects while simultaneously melting a substrate, the content of which in the surfacing can be up to several tens of percents. The advantages of this process are the regeneration of machine parts and the refinement of the surface layer by increasing resistance to abrasion, corrosion, erosion, and cavitation [1]. Surfacing technologies in the global industry are developing in parallel with welding technologies or their modifications. However, cladding technologies require more research effort. There are still no normative studies regarding the quality assurance of welded products, especially in terms of regeneration. In order to increase the durability of machine and equipment parts, and at the same time save costs and, to some extent, protect the environment, surfacing is increasingly used as production and regeneration surfaces in the global industry. The development of research on the functional properties of surface layers and modern welding materials allows the development of production or regeneration technologies, thus contributing to higher efficiency and high economic efficiency [2].
The weld surfacing process has a vital role in modern industries such as mechanical engineering, automotive, power generation, and aerospace. It helps to repair high-performance parts, create durable protective coatings, and improve the performance and durability of components. Using this process allows production to be optimised by repairing damaged parts instead of completely replacing them, leading to time and cost savings [3]. This welding method maximises the functionality of machine part systems, which contributes to significant economic benefits. Plasma surfacing technology is used for components with special properties such as high hardness, abrasion resistance, and corrosion in atmospheric or chemically active environments [4].
Weld surfacing not only makes it possible to restore the dimensions of components altered by wear [5], but at the same time provides the work surface with better wear resistance due to the use of a harder surfacing metal than the base of the component [6]. The higher the hardness of the surfacing metal, the greater the wear resistance of the surface layer and the greater the difficulty of machining. This problem is complicated by the presence of specific irregularities within the surface layer. During machining, this creates impact loads on the cutter, leading to accelerated chipping and puncturing of the cutting edge. It is possible for not only the cutting edge, but the entire replaceable insert to split, i.e., the machining effort then becomes even greater. Another reason limiting the use of surfacing is the tendency to crack. To prevent cracking, preheating of the workpieces before surfacing and heat treatment (tempering) after surfacing are used, which increases the cost of surface work and makes it more difficult to carry out. Therefore, usually the hardness of the surface layer is reduced at the expense of wear resistance [7]. One method of surfacing is wire arc [8] additive manufacturing, (ang. WAAM), which is a popular wire-feed additive manufacturing technology that creates components by depositing material layer by layer [9]. WAAM has emerged as a promising alternative to conventional machining due to its high deposition efficiency, environmental friendliness, and cost competitiveness. The article presents the adaptation of a gantry machine with in situ monitoring and a control system in order to reveal the ability of WAAM technology to produce parts with complex shapes. The machine was retrofitted in several layers called hardware, control, and software layers, respectively. This study demonstrated the feasibility of an additive manufacturing-capable machine as a controlled process [10]. Plasma surfacing with powder additive material requires greater control over temperature and gas parameters to maintain optimal melting conditions, while wire is more resistant to fluctuations in temperature and operating conditions [11]. The use of wire allows for more precise and controlled deposition of thin repair layers [12].
Plasma surfacing by using a plasma jet and a conductive filler wire is carried out using a direct current of simple polarity. The arc burns between the tungsten cathode and the filler wire, applied from the side at right angles to the plasma torch axis. A low intensity pilot arc (15–25 A) also burns continuously between the cathode and the plasma torch nozzle. The arc ensures reliable ignition and stable combustion of the main arc. The parent metal heats up under the heat of the plasma jet and the heat transferred by the secondary metal droplets. The effective heat output of this heat source depends on the arc current and the distance between the wire and the parent metal. By keeping the current constant and varying the current and thus the melting rate of the secondary wire, the power consumed to heat the parent metal can be varied over a relatively wide range. Thus, with plasma jet surfacing, it is possible to control the thermal and diffusion processes at the fusion edge, which determine the depth of penetration of the parent metal and the content of the parent metal in the metal, as well as the length, composition, and properties of the fusion zone [13]. Plasma jet deposition is widely used on ships for the deposition of corrosion-resistant and anti-friction alloys. The deposition of various shafts, equipment rods and other components is carried out using copper alloys with solid supplementary wires or powder wires. Argon is used as a plasma-forming gas and as a shield [14,15].
Steel C45, also known as AISI 1045, is a medium carbon steel that is commonly chosen for plasma welding due to its specific properties that make it suitable for this kind of process. C45 steel has a carbon content of approximately 0.45%, which provides a good balance between hardness and ductility. This makes it easier to weld compared to high-carbon steels, which can be more prone to cracking. C45 can still be welded effectively with the proper techniques and precautions. Preheating and post-weld heat treatment are often employed to prevent issues such as cracking and to improve the overall quality of the weld. Compared to alloy steels or other high-performance materials, C45 steel is relatively cost-effective, making it a practical choice for many industrial applications [16].
The study [17] focused on the regularities of interfacial interactions, intermetallic phase formation (IMP) features, and defects during the surfacing of steels on titanium using four methods: P-MAG, CMT, indirect arc plasma surfacing with a conductive wire, and PAW. A general trend in the occurrence of IMPs in the surfacing of steels on titanium was found with all the methods considered. It was found that the plasma spraying technique using an indirect arc with a conductive wire was less critical with regard to IMP formation. The study showed the possibility of forming, in addition to TiFe2 and TiFe phases, a Ti2Fe phase at low heat input. The indirect arc plasma surfacing technique with a conductive wire minimises the thermal effect on the base metal. When applied at the transition boundary of a titanium-lined steel layer, the phase composition and layer structure, in some cases, approach that of the transition zone of the original ‘titanium–steel’ bimetallic sheet produced by rolling.
In the article [18], the method of multilayer plasma surfacing in an alloyed nitrogen shielding atmosphere was improved. The paper presents a comparative study of the residual stresses developed when surfacing with 3Kh2V8, 4Kh4V10Yu, and R18Yu steel wire. It is shown that the R18Yu-based wire product provides the lowest residual stresses. In addition, an improved method of multilayer plasma surfacing of heat-resistant steels in a nitrogen atmosphere is presented as preventing cracking and increasing surface metal hardness without subsequent heat treatment.
The article [19] is based on the results of fabricating multilayer semi-finished products and laminates, shaping the structure and properties of homogeneous and heterogeneous materials using plasma surfacing with straight and reverse polarity currents. It is shown that reverse-polarity plasma surfacing provides good layer fusion with minimal heat input to the base material and filler. This reduces the thickness of the fusion zone and the mixing of metal in the layers. In addition, cathode cleaning reduces the likelihood of interlayer defects.
The paper [20] presents the results of a study of the structure, phase formation, and properties of chromium–nickel alloys in plasma surfacing of high-alloy steel wire. In order to investigate the possibility of structure modification during argon arc deposition, an additional ultrasonic action was applied to the deposited material using a waveguide connected to the bottom surface of the sample. A comparison of the surface alloy structures obtained by laser deposition showed that arc welding of the alloy combined with ultrasonic action created an additional effect of increasing phase dispersion, which led to an increase in the strength of the alloy at high temperatures.
In the study [21], a medium entropy CoCrFeMnNiW alloy coating on Q235 was produced using plasma spraying technology. The wear resistance of the prepared single-layer coating and the double-layer coating was investigated using a friction and abrasion tester. The microstructure and performance of the CoCrFeMnNiW coating were investigated using optical microscopy, a nano-hardness test, SEM, and a hardness tester. The results showed that the microstructure of the coating was composed of a melt zone, equiaxial dendrites near the fusion zone, coarse co-luminal and near-surface crystals in a specific direction between the melt zone, and near-surface fine equiaxial grains.
What distinguishes this research from the existing literature is its comprehensive analysis of the specific impacts of variable plasma surfacing parameters on C45 steel, particularly focusing on weld geometry, the dilution factor, and microhardness. Unlike previous studies that may have addressed these factors individually or in different contexts, this research provides a holistic examination, integrating these aspects to offer a clearer understanding of their interdependencies. The concrete contribution of this study lies in its detailed investigation of how specific changes in plasma surfacing parameters can be optimized to achieve precise weld characteristics, thereby enhancing weld quality and performance. This work also provides practical guidelines for industry professionals aiming to improve plasma surfacing outcomes for C45 steel, bridging the gap between theoretical research and practical application. In this study, the aim was to investigate the effects of varying plasma surfacing parameters such as arc intensity, arc voltage, deposition speed, feed rate and torch distance from the work surface on the surfacing geometry, the dilution coefficient, and microhardness. The parent material was C45 unalloyed steel, the filler material was 45554S stainless steel wire, and the surfacing was carried out under protective argon gas shielding.

2. Materials and Methods

The first step in the research was to prepare the specimens and develop a set of plasma resurfacing parameters variable on the basis of the initial samples. The samples were then incubated in resin so that metallographic tests and microhardness measurements could be carried out. The metallographic tests are related to the determination of the dilution coefficient, which is important for surfacing technology [22].
The dilution coefficient in welding and surfacing technology refers to the proportion of base material mixed with the filler material in the weld pool. It significantly impacts the chemical composition, mechanical properties, corrosion resistance, microstructure, and wear resistance of the weld or surfaced layer. Controlling the dilution coefficient is crucial for achieving desired material properties and cost efficiency in surfacing applications [22].
The tests were carried out on flat bars made of unalloyed C45 steel with dimensions of 300 × 100 × 6 mm. The surface of the samples was cleaned and degreased before surfacing. The CastoMag 45554S stainless (Gliwice, Polandsteel wire with a diameter of 1.2 mm was used in the process. The wire was used for welding and surfacing difficult-to-weld steels in protective gas Argon N5.0. It is characterised by a high corrosion resistance, resistance to sudden temperature changes, cracking resistance, and the absence of spatter. The chemical composition is shown in Table 1 based on the Castolin Eutetic manufacturer’s (Gliwice, Poland) process card [23].
The surfacing tests were performed on an automated workstation equipped with a Panasonic TM-1400 GIII robot (Gliwice, Poland) (Figure 1), a Castolin Eutectic GAP 2501 DC (Gliwice, Poland) source, and a Castolin Eutectic E52 torch (Gliwice, Poland).
The main variable parameters were the intensity of the main arc, the plasma arc voltage, the surfacing speed, the feeding speed of the additional material, and the distance of the torch from the working surface. Table 2 shows the variables identified through preliminary testing. The constant parameters, which remain consistent throughout the process, are shown in Table 3.
The starting points were obtained after a number of tests to achieve the most stable surfacing possible. All the variable parameters were manipulated so that the surfacing weld stitch was as beneficial as possible. Eighteen samples of five replicates of each whole surfacing stitch were conducted considering the trivalent nature of each variable (Table 4); in order to quantify the influence of the parameter variables, an orthogonal randomised experimental plan as described in the article [24] was used.
Samples for metallographic examination were prepared by cutting the specimens at a predetermined distance from the stitch and then into smaller fragments (Figure 2 and Figure 3) using a Struers Labotom cutter5. Non-continuous surfacing samples were eliminated, and further tests were carried out to refine the final parameters.
The prepared samples were encapsulated in epoxy resin so that their dilution factor could be examined. This is determined from the formula as follows [22]:
    D = B A + B 100 %   [ ]
where
  • D is the dilution factor [−];
  • A is the cross-sectional area of the surfacing [mm2];
  • B is the cross-sectional area of the surfacing mixed with the substrate material [mm2].

3. Results

The results section details the findings from metallographic tests, surfacing parameters analysis, dilution factor assessments, and hardness tests. These analyses provide crucial insights into the microstructural characteristics, optimal surfacing conditions, material mixing ratios, and mechanical properties such as hardness. Together, these results contribute to a comprehensive evaluation of the surfacing process and its impact on the final properties of the welded or surfaced material [25].
Figure 4 shows selected samples during metallographic measurements. It is noticeable that the shape of the stresses varies considerably. Table 5 presents a summary of the parameters of the samples in Figure 4 in order from largest to smallest dilution factor value. The metallographic tests were carried out on a Zeiss Smart Zoom 5 microscope (Warsaw, Poland).
The results show that the same wire feed rate of 3.9 m/min occurred for samples 1, 11, and 18, and that there is also overlap in the first two samples between the arc current of 160 A and the hardfacing speed of 0.50 m/min. The results of the dilution factor for all the samples are shown in Table 6.
The highest dilution factor among all the surfacing tests carried out was achieved by test number 12 (Figure 5) and the lowest by test number 14 (Figure 6). The detailed parameters of these two trials are shown in Table 7.
By comparing the parameters and dilution factor results for test pieces 12 and 14, it can be seen that the arc intensity of the plasma surfacing process has a strong influence on the shape of the padding weld. A higher arc intensity can result in a faster melting of the deposit, which can be advantageous in the case of hard-welded materials such as C45 steel. In addition, as test No. 14 shows, insufficient melting of the material can occur if the surfacing speed is too low, leading to poor coating quality and poor adhesion to the substrate. Excessively high surfacing and supplementary material feed rates can result in uneven material deposition, excessive spatter, and loss of process control, leading to an imperfect coating and reduced process efficiency. Therefore, it is important too maintain appropriate process parameters to ensure the quality and efficiency of plasma surfacing. Hardness measurements were performed on a FM-800 micro hardness tester in accordance with PN-EN ISO 6507-1:2018 [26] and PN-EN ISO-9015-2:2016 [27]. The results are shown in Table 8.
Figure 7 shows a summary of the results obtained and the determination of the influence of parameters on hardness. The variation in hardness results indicates differences in microcrystalline structure. Further investigations should include sample etching in a suitable solution, e.g., nital, microscopic, and mechanical tests to understand the reasons for these differences in hardness [28].
Table 8 contains the results of microhardness measurements of selected samples. Sample 1 shows significant variation in hardness in heat-affected zone 1 (349–417 HV) and stable values in parent materials 1 and 2 (approximately 230–250 HV). For sample 8, there are very high hardness results (507–502 HV) in heat-affected zone 1, indicating clearly higher mechanical properties compared to the other materials. This may be due to the perlite structure encountered. In sample 12, in heat-affected zones 1 and 2, the results show high hardness values, similar to that of padding weld 2, which may suggest good wear resistance. Sample 9 has the highest average hardness value (368 HV), mainly due to very high hardness values in heat-affected zone 2 (608–626 HV) and heat-affected zone 1 (460–537 HV). Sample 7 shows lower hardness values compared to samples 8, 12, and 9, with values in parent materials 1 and 2 having values in the range 215–280 HV.

4. Conclusions

The study of C45 steel in plasma surfacing has limitations such as sensitivity in the heat-affected zone, the need for post-weld heat treatment, and the potential for cracking and distortion. Future research should focus on optimising welding and surfacing parameters, exploring advanced techniques, investigating alternative filler materials, and understanding the long-term performance of welded joints. Additionally, the use of simulations and environmentally friendly practices could enhance welding outcomes for C45 steel.
Research into the plasma surfacing process reveals a significant influence of process parameters on the mechanical properties of the surface layers. Analysis of the results shows a relationship between microhardness and the mixing coefficient. It is observed that with increasing microhardness, the mixing ratio also increases. This trend may be the result of better integration of the surfacing layer with the substrate, resulting in increased durability and mechanical resistance of the resulting structure. Understanding the influence of surfacing parameters on these properties is crucial for various applications where durability and wear resistance are important. This suggests that optimisation of process parameters can be an effective strategy in achieving the desired mechanical properties of surface layers.
The plasma arc intensity has the greatest influence on the geometry of the coating and the mixing ratio. This confirms the importance of controlling this parameter in the process. However, further research is needed to more fully understand this relationship and further implications. Continued experimentation may provide more detailed information on optimal process parameter settings, which could lead to further improvements and innovations in plasma surfacing.
Practical implications include the need for optimized plasma welding parameters and advanced techniques to ensure high-quality welds. Investigating alternative filler materials and understanding long-term performance are crucial for enhancing the reliability and durability of welded joints. Additionally, leveraging simulation tools and environmentally friendly practices can lead to more efficient and sustainable welding processes, ultimately broadening the applications of C45 steel in various industries. These results underscore the critical role of precise surfacing parameters in achieving desirable weld characteristics such as minimised dilution and enhanced hardness. Practical applications could benefit from these insights by optimising welding conditions to improve weld quality and durability in industrial settings. However, remaining questions persist, such as the long-term durability of surfaced materials under varying environmental conditions and the development of more sustainable welding practices. Addressing these gaps in future studies could further refine plasma welding techniques and broaden their applicability across diverse engineering applications.

Author Contributions

Data curation, S.B.; dormal analysis, S.B.; methodology, S.B. and R.S.; resources, S.B. and K.P.; software, S.B. and R.S.; supervision, S.B.; validation, S.B.; visualization, S.B.; writing—original draft, S.B.; writing—review and editing, S.B. and K.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bober, M.; Tobota, K. Badania Istotności Wpływu Podstawowych Parametrów Napawania Plazmowego na Geometrię Napoin; Przegląd Spawalnictwa: Warszawa, Polska, 2015; Volume 87 9/2015, pp. 24–28. [Google Scholar]
  2. Klimpel, A. Industrial surfacing and hardfacing technology, fundamentals and applications. Weld. Technol. Rev. 2019, 91, 33–42. [Google Scholar] [CrossRef]
  3. Cui, W.; Shuang, W.; Song, Z.; Tan, J.; Meng, G.; Kai, L. Influence of welding current on microstructure and properties of nickel-based alloy hardfacing by plasma transferred arc welding. Mater. Transanctions 2017, 54, 2144–2150. [Google Scholar]
  4. Napadałek, W.; Chrzanowski, W.; Woźniak, A. Przyrostowe technologie 3D w odbudowie kształtu zużytych eksploatacyjnie łopat turbin parowych. Autobusy 2017, 12, 1147–1152. [Google Scholar]
  5. Hu, Z.; Qin, X.; Shao, T.; Liu, H. Understanding and overcoming of abnormity at start and end of the weld bead in additive manufacturing with GMAW. Int. J. Adv. Manuf. Technol. 2018, 95, 2357–2368. [Google Scholar] [CrossRef]
  6. Tarng, Y.S.; Juang, S.C.; Chang, C.H. The use of grey-based Taguchi methods to determine submerged arc welding process parameters in hardfacing. J. Mech. Work. Technol. 2002, 128, 1–6. [Google Scholar] [CrossRef]
  7. Korotkov, V.A. Plasma Hardening of a steel 30KhGSA Surfacing Layer. Chem. Pet. Eng. 2015, 51, 319–323. [Google Scholar] [CrossRef]
  8. Fernández-Zabalza, A.; Veiga, F.; Suárez, A.; López, J.R.A. The Use of Virtual Sensors for Bead Size Measurements in Wire-Arc Directed Energy Deposition. Appl. Sci. 2024, 14, 1972. [Google Scholar] [CrossRef]
  9. Veiga, F.; Suárez, A.; Aldalur, E.; Bhujangrao, T. Effect of the Metal Transfer Mode on the Symmetry of Bead Geometry in WAAM Aluminum. Symmetry 2021, 13, 1245. [Google Scholar] [CrossRef]
  10. Artaza, T.; Alberdi, A.; Murua, M.; Gorrotxategi, J.; Frías, J.; Puertas, G.; Melchor, M.A.; Mugica, D.; Suárez, A. Design and integration of WAAM technology and in situ monitoring system in a gantry machine. Procedia Manuf. 2017, 13, 778–785. [Google Scholar] [CrossRef]
  11. Yin, X.; He, G.; Meng, W.; Xu, Z.; Hu, L.; Ma, Q. Comparison Study of Low-Heat-Input Wire Arc-Fabricated Nickel-Based Alloy by Cold Metal Transfer and Plasma Arc. J. Mater. Eng. Perform. 2020, 29, 4222–4232. [Google Scholar] [CrossRef]
  12. Klimpel, A. Technologie Laserowe; Wydawnictwo Politechniki Śląskiej: Gliwice, Poland, 2012. [Google Scholar]
  13. Sanchez-Tovar, R.; Montanes, M.T.; García-Antón, J. Effect of the Micro-plasma Arc Welding Process on the Microstructure and Pitting Corrosion of AISI316L Stainless Steels in Heavy Brines. Corros. Sci. 2011, 53, 2598–2610. [Google Scholar] [CrossRef]
  14. Gladkii, P.V.; Perepletchikov, E.F.; Ryabtsev, I.A. Plasma surfacing. Weld. Int. 2007, 21, 685–693. [Google Scholar] [CrossRef]
  15. Kalenskii, V.K.; Gladkii, P.V. Investigation and development of the method of automatic surfacing outlet valves of vehicles. Avt. Svarka 1963, 1, 15–23. [Google Scholar]
  16. Tanasković, D.; Aranđelović, M.; Đorđević, B.; Jeremić, L.; Sedmak, S.; Gajin, M. Repair attempts of cold crack on forklift made of C45 steel: Case study. Weld. Mater. Test. 2020, 29, 25–28. [Google Scholar]
  17. Korzhyk, V.; Khaskin, V.; Grynyuk, A.; Ganushchak, O.; Peleshenko, S.; Konoreva, O.; Demianov, O.; Shcheretskiy, V.; Fialko, N. Comparing Features in Metallurgical Interaction When Applying Different Techniques of Arc and Plasma Surfacing of Steel Wire on Titanium. East.-Eur. J. Enterp. Technol. 2021, 4, 6–17. [Google Scholar]
  18. Malushin, N.N.; Martyushev, N.V.; Valuev, D.V. i wsp. Wzmacnianie elementów urządzeń hutniczych poprzez napawanie plazmowe w atmosferze azotu. Metalurg 2022, 65, 1468–1475. [Google Scholar] [CrossRef]
  19. Neulybin, S.D.; Schitsyn, Y.D.; Belinin, D.S.; Permyakov, G.L. Prospects of Using Plasma Surfacing to Producing of Layered Materials. Int. J. Emerg. Trends Eng. Res. 2020, 8, 3562–3568. [Google Scholar] [CrossRef]
  20. Krivonosova, E.A.; Schitsin, Y.D.; Trushnikov, D.N.; Myshkina, A.V.; Akulova, S.N.; Neulibin, S.D.; Dushina, A.Y. Structure formation of high-temperature alloy by plasma, laser and TIG surfacing. J. Phys. Conf. Ser. 2018, 1089, 012019. [Google Scholar] [CrossRef]
  21. Hu, Q.; Wang, X.; Miao, J.; Fu, F.; Shen, X. Friction and Wear Performance of CoCrFeMnNiW Medium-Entropy Alloy Coatings by Plasma-Arc Surfacing Welding on Q235 Steel. Coatings 2021, 11, 715. [Google Scholar] [CrossRef]
  22. Blicharski, M. Inżynieria Powierzchni; Wydawnictwo Naukowe PWN: Warszawa, Polska, 2021. [Google Scholar]
  23. Castolin Eutetic Brand Card. Available online: https://issuu.com/castolin_eutectic/docs/castolin-castomag-45554-s-schweisse (accessed on 10 April 2024).
  24. Bazychowska, S.; Starosta, R.; Dudzik, K. Ilościowa ocena wpływu parametrów napawania plazmowego na stopienie metalurgiczne powłoki ze stali austenitycznej z materiałem podłoża wykonanego ze stali C45. J. KONBiN 2022, 52, 27–51. [Google Scholar] [CrossRef]
  25. Uralde, V.; Veiga, F.; Suarez, A.; Aldalur, E.; Ballesteros, T. Symmetry Analysis in Wire Arc Direct Energy Deposition for Overlapping and Oscillatory Strategies in Mild Steel. Symmetry 2023, 15, 1231. [Google Scholar] [CrossRef]
  26. PN-EN ISO 6507-1:2018-05; Metals—Vickers Hardness Measurement—Part 1: Test Method. Polish Committee for Standardization: Warsaw, Poland, 2018.
  27. PN-EN-ISO-9015-2:2016; Destructive Tests on Welds in Metallic Materials—Hardness Testing—Part 2: Microhardness Testing of Welded Joints. Polish Committee for Standardization: Warsaw, Poland, 2013.
  28. Guo, Z.; Ma, T.; Yang, X.; Li, W.; Xu, Q.; Li, Y.; Li, J.; Vairis, A. Comprehensive investigation on linear friction welding a dissimilar material joint between Ti17(α+β) and Ti17(β): Microstructure evolution, failure mechanisms, with simultaneous optimization of tensile and fatigue properties. Mater. Sci. Eng. A 2024, 909, 146818. [Google Scholar] [CrossRef]
Figure 1. Panasonic TM-1400 GIII robotic plasma surfacing station.
Figure 1. Panasonic TM-1400 GIII robotic plasma surfacing station.
Applsci 14 07248 g001
Figure 2. First step in preparing samples for testing.
Figure 2. First step in preparing samples for testing.
Applsci 14 07248 g002
Figure 3. Second stage of sample preparation for testing.
Figure 3. Second stage of sample preparation for testing.
Applsci 14 07248 g003
Figure 4. Metallographic examples of selected samples.
Figure 4. Metallographic examples of selected samples.
Applsci 14 07248 g004
Figure 5. Cross-section of specimen number 12. (The green and red lines are the areas of the surface and the yellow lines are the heights).
Figure 5. Cross-section of specimen number 12. (The green and red lines are the areas of the surface and the yellow lines are the heights).
Applsci 14 07248 g005
Figure 6. Cross-section of specimen number 14. (The green and yellow lines are the surface areas, while the red lines are the heights).
Figure 6. Cross-section of specimen number 14. (The green and yellow lines are the surface areas, while the red lines are the heights).
Applsci 14 07248 g006
Figure 7. Hardness dependence on plasma surfacing parameters.
Figure 7. Hardness dependence on plasma surfacing parameters.
Applsci 14 07248 g007
Table 1. Chemical composition of Castolin Eutetic CastoMag 45554S wire [23].
Table 1. Chemical composition of Castolin Eutetic CastoMag 45554S wire [23].
ElementCSiMnCrNiMoNb
Percentage content0.0350.8001.60019.50011.5002.8000.700
Table 2. Variable surfacing parameters.
Table 2. Variable surfacing parameters.
Arc Intensity [A]Arc Voltage [V]Plasma Surfacing Speed [m/min]Feed Speed of Additional Material [m/min]Torch Distance from Work Surface [m/min]
120280.252.512
140320.503.216
160350.753.920
Table 3. Constant plasma surfacing parameters.
Table 3. Constant plasma surfacing parameters.
ParameterValue
Internal arc current50 A
Generating gas output Plasma Varigon H91 dm3/min
Shielding gas output Argon N5.01.5 dm3/min
Table 4. Plasma surfacing parameters for individual tests.
Table 4. Plasma surfacing parameters for individual tests.
Plasma Surfacing Parameters
Sample No.Arc Intensity [A]Arc Voltage [V]Plasma Surfacing Speed [m/min]Feed Speed of Additional Material [m/min]Torch Distance from Work Surface [m/min]
1160280.503.912
2160320.252.516
3160350.753.220
4120280.502.516
5120320.253.220
6120350.753.912
7140280.253.920
8140320.752.512
9140350.503.216
10160280.753.216
11160320.503.920
12160350.252.512
13120280.253.212
14120320.753.916
15120350.502.520
16140280.752.520
17140320.503.212
18140350.253.916
Table 5. Summary of parameters of selected samples.
Table 5. Summary of parameters of selected samples.
Sample No.Arc Intensity [A]Arc Voltage [V]Plasma Surfacing Speed [m/min]Feed Speed of Additional Material [m/min]Torch Distance from Work Surface [m/min]Dilution Factor D [−]
1160280.503.9120.430
11160320.503.9200.373
18140350.253.9160.303
15120350.502.5200.288
16140280.752.5200.265
6120350.7539120.093
Table 6. Results of measurement of the dilution factor of all samples.
Table 6. Results of measurement of the dilution factor of all samples.
Sample No.D [−]
10.430
20.547
30.343
40.335
50.286
60.093
70.346
80.284
90.276
100.319
110.373
120.623
130.310
140.053
150.288
160.265
170.325
180.303
Table 7. Summary of sample parameters 12 and 14.
Table 7. Summary of sample parameters 12 and 14.
Sample No.Arc Intensity [A]Arc Voltage [V]Plasma Surfacing Speed [m/min]Feed Speed of Additional Material [m/min]Torch Distance from Work Surface [m/min]Dilution Factor D
[−]
12160350.252.5120.623
14120320.753.9160.053
Table 8. Results of microhardness measurements of selected samples.
Table 8. Results of microhardness measurements of selected samples.
HARDNESS HV
Sample No.Parent Mat. 1Parent Mat. 2HAZ1HAZ2Padding Weld 1Padding Weld 2HAZ1HAZ2Parent Mat. 1Parent Mat. 2Average
1243250374278334338361275230232289
231243376294334334359264225225
251238349, 417, 339292, 300, 292349341320, 335, 315270, 282, 254226224
8243235507252250245457237245238296
251240536248255248498254261231
249238502, 491, 501240, 255, 237230251503, 490, 516248, 251, 245259228
12283285434411392439288301275245344
290287389453477476260277267251
287276463, 456, 464464, 451, 460446493291, 280, 282314, 280, 313263272
2282273286284469422286276242250318
291267307313447462289285248249
296260309, 460, 289295, 310, 287531493303, 289, 282295, 275, 297256260
9247303460608210230538483246245368
245380513631231202501376247250
252407537, 588, 591570, 626, 626232225547, 580, 501528, 505, 634247250
7215271282291224236281316264280260
213280272283231234240284260280
228270244, 260, 252341, 257, 286231203276, 238, 270304, 319, 252262266
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bazychowska, S.; Panasiuk, K.; Starosta, R. The Influence of Variable Plasma Welding Parameters on Weld Geometry, Dilution Factor, and Microhardness. Appl. Sci. 2024, 14, 7248. https://doi.org/10.3390/app14167248

AMA Style

Bazychowska S, Panasiuk K, Starosta R. The Influence of Variable Plasma Welding Parameters on Weld Geometry, Dilution Factor, and Microhardness. Applied Sciences. 2024; 14(16):7248. https://doi.org/10.3390/app14167248

Chicago/Turabian Style

Bazychowska, Sylwia, Katarzyna Panasiuk, and Robert Starosta. 2024. "The Influence of Variable Plasma Welding Parameters on Weld Geometry, Dilution Factor, and Microhardness" Applied Sciences 14, no. 16: 7248. https://doi.org/10.3390/app14167248

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop