Next Article in Journal
On the Problem of the Distillation Separation of Secondary Alloys of Magnesium with Zinc and Magnesium with Cadmium
Previous Article in Journal
Review of In Situ Detection and Ex Situ Characterization of Porosity in Laser Powder Bed Fusion Metal Additive Manufacturing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Surface Growth of Boronize Coatings Studied with Mathematical Models of Diffusion

by
Martín Ortiz-Domínguez
1,*,
Ángel Jesús Morales-Robles
2,
Oscar Armando Gómez-Vargas
3 and
Georgina Moreno-González
4
1
Mechanical Engineering, Escuela Superior de Ciudad Sahagún, Autonomous University of Hidalgo State, Carretera Ciudad Sahagún-Otumba s/n, Zona Industrial, Ciudad Sahagún 43990, Hidalgo, Mexico
2
Academic Area of Earth Sciences and Materials, Institute of Basic Sciences and Engineering, Autonomous University of Hidalgo State, Carretera Pachuca-Tulancingo Km. 4.5 s/n, Colonia Carboneras, Mineral de la Reforma 42184, Hidalgo, Mexico
3
Division of Graduate Studies and Research, Tlalnepantla Institute of Technology, National Technological Institute of Mexico, Avenida Instituto Tecnológico, s/n. Colonia La Comunidad, Tlalnepantla de Baz 54070, Estado de México, Mexico
4
Systems and Computing Department, Tlalnepantla Institute of Technology, National Technological Institute of Mexico, s/n. Colonia La Comunidad, Tlalnepantla de Baz 54070, Estado de México, Mexico
*
Author to whom correspondence should be addressed.
Metals 2024, 14(6), 670; https://doi.org/10.3390/met14060670
Submission received: 1 May 2024 / Revised: 29 May 2024 / Accepted: 4 June 2024 / Published: 5 June 2024

Abstract

:
The following investigation focused on examining the kinetics of Fe2B coating formation on the surface of ASTM A681 steel during the powder-pack boronizing process. The study measured Fe2B coating thicknesses at various temperatures and exposure times to confirm the diffusion-controlled growth mechanism during boronizing. Five distinct mathematical models were devised to determine the boron diffusion coefficients in Fe2B coatings. Understanding the growth kinetics of boronize coatings is imperative as it facilitates the optimization and automation of industrial processes. This ensures the efficient and consistent production of boronize coatings on cutting tools, such as drills and milling cutters, due to their high hardness and wear resistance. The value of the activation energy estimated with five mathematical diffusion models for the Fe2B coating was 209.8 kJ∙mol−1. The X-ray diffraction technique was used to identify the presence of the iron boronize phase. Tribological studies were also performed to evaluate the coefficient of friction (COF) of the boronized (0.256) and untreated (0.781) samples, having a 300% positive effect of the boronize coating on wear resistance. Finally, the models were empirically validated for two supplementary treatment conditions for 1223 K for 3 h and 1273 K for 1.5 h, where the percentage error for both conditions was estimated to be approximately 2.5%.

1. Introduction

Various surface treatments can augment steel’s wear and corrosion resistance. These methods encompass nitriding, boronizing, and aluminizing, as well as physical vapor deposition (PVD) and chemical vapor deposition (CVD). Nevertheless, thermochemical processes are favored due to the superior adhesion properties of the resultant coatings under elevated temperatures. Boronizing is a widely employed thermochemical process, and it is distinguished among these methodologies. This technique engenders the development of a durable surface comprising single- or multi-phase boronizes, achieved through the diffusion of element B onto the surface of the base material. The resultant boronize coatings exhibit exceptional hardness and strength, which are challenging to attain through conventional methods like carburizing, nitriding, and carbonitriding. Consequently, the tribological performance and service life of steels are significantly improved.
The process can be used on ferrous and non-ferrous materials, typically at temperatures between 973 and 1273 K (1300 and 1830 °F) for 2 to 10 h. By doing this, a metal boronize coating of approximately 20–300 µm thickness can be formed. The metal boronize coating that results from boronizing exhibits excellent characteristics such as high hardness, reasonable wear and corrosion resistance, and moderate resistance to high-temperature oxidation. However, it is essential to note that boronizing cannot be applied to some metals and alloys, such as aluminium and magnesium, due to their low melting points. Moreover, copper alloy is incompatible with boron atoms [1]. Notably, a singular Fe2B coating in the industrial sector is typically favored over a dual-phase coating consisting of both FeB and Fe2B due to the potential risk of crack propagation along the (FeB/Fe2B) interface [2]. Solid boronizing has emerged as the most extensively utilized method in the industry for this purpose, owing to its cost-effectiveness, ease of handling, and the capability to modify the chemical composition of the powders as per the requirements.
The academic literature contains numerous documented diffusion models focusing on the kinetics of Fe2B coating formation on various substrates. These simulation techniques have real-world applications that aim to enhance the surface properties of treated steels by optimizing the thickness of the boronize coating. For example, Ortiz-Domínguez et al. [3] utilized two distinct methods to evaluate the growth kinetics of Fe2B coatings on the surface of AISI 1018 steel. The first approach was founded on a diffusion model that employed the mass balance equation at the Fe2B/substrate interface. The second approach involved a dimensional analysis in constructing an equation that could accurately describe the thickness of the boronize coating as a function of process parameters, including treatment time and boronizing temperature. Additionally, a series of experiments were conducted at different temperatures, with an exposure time of 5 h, to facilitate a comparison between the predicted values and the experimental coating thicknesses. Arslan-Kaba et al. [4] conducted a study where they utilized a boronizing process known as pulsed current integrated cathodic reduction/thermal diffusion (PC/CRTD-Bor) on low-carbon steels. The aim was to create Fe2B-type monocoatings in a shorter period. The study utilized a constant duty cycle of 1/4 and various current densities (50, 200, and 700 mA∙cm−2) at a treatment temperature of 1223 K for exposure times of 5, 10, and 30 min. An average diffusion coefficient (ADC) model was utilized to evaluate the boron diffusion coefficients in FeB and Fe2B coatings, an average diffusion coefficient (ADC) model was used, and the calculated coating thicknesses were found to agree with the experimental data. Bouarour et al. [5] studied the kinetics of Fe2B boronize coatings formed on the surface of SAE 1020 steel by considering four mathematical mass transfer models. They also considered the boronize incubation period for the Fe2B coating in the mathematical models. The activation energies of boronize evaluated on SAE 1020 steel were relatively equal and approximately equal to 183 kJ mol−1.
The investigation of boronize needle orientation during the forming of Fe2B coatings on pure Armco iron was carried out by Ramdan et al. [6] through 2D multi-phase field (MPF) simulations. The study aimed to determine the growth behaviour of texture in these boronizes parallel to the substrate surface. The study’s main finding was observing texture growth in these boronizes, which occurs perpendicular to the substrate surface. Notably, the growth behavior was found to be independent of the initial size of the boronize needle. The study conducted by Campos et al. [7] is significant as it provides a comprehensive analysis of the paste-boronizing process of AISI 1045 steel. Fuzzy logic Mamdani and Takagi-Sugeno’s mathematical techniques to simulate the process are particularly interesting due to their ability to handle imprecise or uncertain input data. The formation of Fe2B coatings on the steel’s surface, which is a desirable outcome in the paste-boronizing process, was successfully achieved by varying the thickness of the boron paste. The reported mean errors of 2.61% and 3.62% for the Mamdani and Takagi-Sugeno techniques indicate that the simulation results are highly accurate and reliable. The study’s findings are of practical importance to researchers and practitioners in materials science and engineering as they can optimize the paste boronizing process of AISI 1045 steel.
In a previous study, Campos et al. [8] examined the growth interface between the Fe2B/substrate while accounting for mass conservation. However, the boronize incubation times for pure Armco iron paste boronizing were not taken into consideration. The authors proposed investigating the growth kinetics of boronize coatings by exploring the relationship between the molar volume of the Fe2B phase and the substrate.
Nait Abdellah et al. [9] proposed two mathematical mass transfer models to estimate the coefficients of boronize in the Fe2B coating formed on the surface of ASTM A36 steel in the temperature range from 1173 to 1273 K with treatment times of 2 h, 4 h, 6 h, and 8 h. They also considered the boronize incubation time required to form the primitive coating, which is constant and independent of the boronizing conditions. The first mathematical model was derived from the mass balance equation at the growth interface (Fe2B/substrate), while the second one employed the integral method. The calculated values of boron activation energies for ASTM A36 steel turned out to be very close to each other for both approaches (161.65 kJ·mol−1 and 160.96 y kJ·mol−1).
This research aimed to study the growth kinetics of the boronized coatings (Fe2B) formed on the surface of ASTM A681 steel through five mathematical mass transfer models. Likewise, the models were experimentally validated for two extra boronizing conditions (1223 K for 3 h and 1273 K for 1.5 h). The boronizing treatment was performed using the traditional powder-pack technique. For this reason, four different temperatures and four different durations were selected for the treatment. Furthermore, a powder mixture containing 33.5 wt.% B4C, 5.4 wt.% KBF4, and 61.1 wt.% SiC was used to treat ASTM A681 steel. In this context, the SEM method was used for microstructural characterization, and the XRD and EDS methods were used for chemical characterization of the boronize coatings formed due to the boronizing treatment. The investigation examined the influence of the boronizing treatment on the coefficient of friction (COF) value after the wear test and compared it with the unboronized sample.

2. Materials and Methods

2.1. Properties

During the boronizing process, boron atoms diffuse and adsorb onto the metal lattice of the component’s surface, forming an interstitial boron compound that manifests as a single-phase or polyphase boronize coating. The boronize coating’s morphology, growth, and phase composition are intricately linked to the elements present in the substrate materials and their interactions with the boron atoms. As demonstrated in Table 1, the boronize coating’s characteristics depend on the substrate materials’ constituents [10].
A very interesting and much more complete work is the one presented by Akopov et al. [11], which reviews the advances in the crystal chemistry of the different boronizes in the last 60 years and compiles data on the crystal structure, material properties, and synthetic routes of metal boronizes. Likewise, metal boronizes are chosen because of their rich crystalline chemistry, which gives rise to several properties that offer many application possibilities. The formation of a boronize coating on iron and steel can result in a single-phase or two-phase composition, which is determined by the Fe-B phase diagram (refer to Figure 1) [12].
The following technical information concerns the morphological characteristics of the boronize coating. It has been observed that forming a single-phase coating results in the creation of Fe2B, while a two-phase coating featuring an outer FeB phase and an inner Fe2B phase is formed. The morphology of the boronize coating is characterized by a sawtooth structure, illustrated in Figure 2. This structure is vital in enhancing the mechanical adhesion at Fe2B/substrate interfaces, as documented in reference [13].
It is essential to note that the development of boronize coatings on the substrate surface is subject to a certain period of nucleation or incubation time, as referenced in scholarly literature [14]. During this time, the initial spreading of the coating is limited to specific locations on the substrate surface, where it merges to form a skinny initial coating of iron boronize. At this stage, the initial coating is randomly oriented [15]. However, as the coating grows, it is strongly favored along the preferential direction [002] for FeB and Fe2B [15]. It is important to note that when undergoing boronizing treatment, the formation of the Fe2B phase is preferred over the FeB phase due to the latter being more brittle [2]. The differences in thermal expansion coefficients between the FeB and Fe2B phases and the base material can create a complex stress state, leading to tensile residual stresses in the outer FeB coating and compressive residual stresses in the Fe2B phase [16]. However, the formation of both phases can result in cracks at the FeB/Fe2B interface of the two-phase coating, which may cause spalling and even separation of the coating under mechanical stress or thermal/mechanical shock [2]. Fortunately, the annealing process can help reduce the occurrence of the FeB phase after boronizing treatment, thus mitigating the risk of crack formation [17,18].
The data presented in Figure 2 indicate that the boron concentration at the phase surface is denoted as C u p F e 2 B (=9 wt.%B), while the boron concentration at the Fe2B/substrate growth interface is represented as C l o w F e 2 B (=8.83 wt.%B). Additionally, the boron solubility limit within the substrate is symbolized as C 0 (=35 × 10−4 wt.%B), though it cannot be considered due to its magnitude [12]. It is essential to note that the information presented in Figure 2 is critical for understanding the boron concentration at the phase surface and the Fe2B/substrate growth interface. Therefore, the data can provide valuable insights into the growth process of Fe2B on the substrate.
The unique characteristics of the Fe2B phase are notable [19]:
  • The composition contains approximately 9.0 wt.% boron.
  • The crystal structure of Fe2B is identified as body-centered tetragonal and is characterized by axial lengths a = b = 0.5078 × 10−9 m and c = 0.4249 × 10−9 m (as exemplified in Figure 3).
  • The density of Fe2B is 7.43 × 10−3 kg/cm3.
  • Microhardness ranges from approximately 18–20 × 109 Pa.
  • The Young’s modulus range of Fe2B is between 285–295 × 109 Pa.
  • The thermal expansion coefficient is measured at 7.65 × 10−6/K within the 473–873 K range and 9.2 × 10−6/K within the 373–1073 K range.

2.2. Substrate and Boronizing Process

ASTM A681 is a high-speed tool steel widely acclaimed for its remarkable wear resistance, strength, and toughness. This type of steel is commonly employed in manufacturing cutting tools, dies, and various machinery where its exceptional properties can be highly advantageous. The notable characteristics of this steel make it an ideal choice for applications with high-stress conditions. The material’s high wear resistance maintains its cutting edge even under harsh conditions. Its superior strength and toughness also enable it to withstand the high forces and pressures commonly encountered in industrial applications. The samples to be boronized were cubic and had 10 mm sides. They were sectioned from a square bar using a Buehler-IsoMetTM 1000 precision gravity-fed bench (Lake Bluff, IL, USA). The steel samples were roughened with silicon carbide (SiC) paper grits before undergoing the thermochemical process. The grits ranged from 80 to 2500 and were applied using an EcoMetTM 30 single-hand grinding machine (Lake Bluff, IL, USA). The samples were then cleaned for 20 min in a SONOREX SUPER RK 52 high-performance ultrasonic cleaner (Heinrichstraße, Berlín, Germany). After washing, the samples were placed inside a cylindrical AISI 316L stainless steel container and embedded in a boron-rich powder mixture, as shown in Figure 4.
The chemical mixture has the following weight percentage composition: 33.5% boron carbide (B4C) as the main boronize, 5.4% potassium fluoroborate (KBF4) as the catalyst, and 61.1% silicon carbide (SiC) as the diluent. This specific composition of the boronize powder facilitates the formation of a Fe2B-type boronize monocoating on the treated surface, as cited in [20]. Chemical reactions that form the Fe2B coatings are shown in Figure 5, as referenced by [21].
During the thermochemical treatment process, the surface of the substrate material, ASTM A681 steel, undergoes ionization through the reaction of BF2 gas, resulting in the formation of B2+ and [BF]+ ions. The diffusion of boron (B) atoms further ensues into the interior of the substrate. A TEFIC 1473 K brand furnace (Weiyang District, Shaanxi, China) was employed to execute this process. Different treatment times and temperatures were utilized, while argon gas protected the samples from atmospheric pollution, oxygen, and hydrogen. Argon gas is chemically inert, tasteless, and odorless, making it an ideal choice. The AISI 316L steel container was heated, with the samples embedded in the boron-rich mixture.
According to the phase diagram, the boronizing treatment temperatures (refer to Figure 1) ranged from 1123 K to 1273 K, while the treatment duration ranged from 2 to 8 h. After completion of the boronizing process, the container was removed from the oven and allowed to cool to room temperature (296 K). The sample was then removed from the container and encapsulated in resin using a Buehler SimpliMet 4000 mounting press (Lake Bluff, IL, USA). This encapsulation was carried out to facilitate sample holding during manual polishing. It was chosen based on the shape of the part and the type of analysis to be performed, including edge analysis as part of a surface treatment. The metallographic samples underwent a rigorous two-step polishing process. Firstly, the samples were polished with a Buehler MicroCloth™ (Hong Kong, SAR, China), a soft, versatile, synthetic rayon cloth with a long pile and magnetic backing. The cloth was used with Al2O3 aluminium oxide abrasive (3 and 1 × 10−6 m) to remove any scratches that may have been produced during the roughing operation. Secondly, with each size, the samples were subjected to diamond suspension polishing (0.25 and 0.05 × 10−6 m, Buehler) for 20 min. This allowed for the creation of a specular surface necessary for accurate analysis. Additionally, the samples were cleaned after each step using an ultrasonic bath of acetone for thirty seconds to ensure the highest possible quality. A reagent was employed in immersion with 4% volume Nital to reveal and analyze the samples’ microstructure. The thickness of the boronized coatings was observed using a 4K optical microscope from the VHX-7000 series (Higashi-Yodogawa-ku, Osaka, Japan). The metallographic procedure, including the observation of phases, grain boundaries, impurities, and deformation zones, is presented in Figure 6.
The thicknesses of boronize (Fe2B) coatings were estimated using the specialized software Image-Pro Plus 6.3 (Rockville, MD, USA). An automated process was utilized for this purpose. The growth kinetics study involved six samples, and a replicate was made for each one. To determine the average values of Fe2B coating thickness, one hundred measurements were taken on each of the thirty-two cross sections selected. This approach enabled us to obtain accurate and reliable data. Figure 7 shows a graphical illustration of the selected cross-sections. The phase was examined using the grazing angle X-ray diffraction technique. The peaks of the X-ray diffraction spectrum were analyzed through the use of Match! crystallographic software version 3.15 (Karlsruhe, Baden-Württemberg, Germany). The scanning range was from 25 to 85, and a CoKα-type diffraction radiation with a wavelength of λ = 0.18 × 10−9 m was used. The Intel Equinox 2000 diffractometer (Waltham, MA, USA) was utilized for this purpose.

3. Mathematical Models of Mass Transfer

Correctly understanding growth kinetics optimizes process parameters to reduce production time and costs. Minimizing the processing time required to achieve the desired properties can result in significant savings in energy and resources. Five mathematical models were formulated to study the growth kinetics of boronize (Fe2B) coatings deposited on the surface of ASTM A681 steel via the solid boronizing process. The initial proposed model pertains to the linear mass transfer model, encompassing the mass balance equation at the Fe2B/Fe growth interface ( Δ C ) ( d x / d t u ) x = u = J i n F e 2 B ( x ) x = u J o u t F e ( x ) x = u + d u and the solution of Fick’s second law without time dependence C F e 2 B ( x ) = C 1 x + C 2 (steady state). The second proposition involves utilizing the mass transfer model called the error function. This model considers the mass balance equation at the Fe2B/Fe growth interface, incorporating time dependence, the solution of the second Fick’s law C F e 2 B ( x , t u ) = A + B e r f ( x / 2 ( D F e 2 B t u ) 1 / 2 ) , and the parabolic growth equation of the boronize (Fe2B) coatings u 2 = k t u . For the third model, the integral method of Goodman [22,23] was used, where the mass balance equation at the Fe2B/Fe growth interface with time dependence was considered ( Δ C ) ( C F e 2 B ( x , t u ) / t u / C F e 2 B ( x , t u ) / x x = u = J i n F e 2 B ( x , t u ) x = u J o u t F e ( x , t u ) x = u + d u , a parabolic profile with time dependence C F e 2 B ( x , t u ) = b ( u x ) 2 + a ( u x ) + c and the parabolic growth equation. Only the mass balance equation at the Fe2B/Fe growth interface without time dependence ( Δ C ) ( d x / d t u ) x = u = D F e 2 B ( d C F e 2 B ( x ) / d t u ) ( d t u / d x ) x = u and the chain rule were considered for the optional fourth mass transfer diffusion model. Finally, the fifth model is based on an inverse diffusion problem approach, where the concentration profile is considered known experimental data  C F e 2 B ( x , t u ) , which allows the determination of the diffusion coefficient as a function of the concentration D F e 2 B ( C F e 2 B ( x , t u ) ) . It is based on Fick’s second law but with the consideration that the boron diffusion coefficient is a function of the concentration profile C F e 2 B ( x , t u ) / t u = ( ( D F e 2 B ( C F e 2 B ( x , t u ) ) ) C F e 2 B ( x , t u ) / t u ) / x  and is combined with the similarity variable x / 2 t u 1 / 2 Figure 8 displays the boronize concentration profile for the Fe2B monocoating in a graphical format. The C a d s term mentioned in Figure 8 refers to the adsorbed content of active boron on the substrate surface [24].

3.1. Linear Mass Transfer Model

The boron concentrations in Figure 8 ( C u p F e 2 B , C l o w F e 2 B , and C l o w F e 2 B ) remain consistent and are determined by the phase diagram, as depicted in Figure 1. Under these specific conditions, the conservation of mass equation can be derived at the interface separating the Fe2B phase and the substrate. Notably, the forward velocity of the interface between the Fe2B/substrate phases is directly proportional to the difference between the inflow and outflow rates (as demonstrated in Figure 8).
C u p F e 2 B + C l o w F e 2 B 2 C 0 2 d x d t u x = u = J i n F e 2 B ( x ) x = u J o u t F e ( x ) x = u + d u = D F e 2 B d C F e 2 B x d x x = u D F e d C F e x d x x = u + d u .
The formation time of the Fe2B phase t u = ( t t 0 F e 2 B ) (s) [25], which denotes the treatment time t (s) and t 0 F e 2 B is equivalent to the characteristic time of Fe2B iron boronize [14], has an essential role in material engineering. Additionally, the incoming flux in the Fe2B phase is J i n F e 2 B ( x ) (atoms/m2∙s), and J o u t F e ( x ) corresponds to the outgoing flux in the substrate (atoms/m2∙s). The diffusion coefficient in the Fe2B phase ( D F e 2 B ) and the diffusion coefficient in the Fe phase ( D F e ) are significant parameters that affect the diffusion process. Notably, the diffusion coefficients in Equation (1) are concentration-independent. Therefore, there are direct methods to address the problem of various conditions at the boundary. In particular, there is no boron flux from the Fe2B surface coating to the substrate J o u t F e ( x ) x = u + d u = 0 , as indicated in Figure 8, which illustrates the boron concentration profile for this system C F e 2 B ( x ) . The thickness of the boronize coating is specified as u (m). Lastly, the boundary conditions for no time dependence are as follows:
C F e 2 B x = u 0 0 = C u p F e 2 B = 9 w t . % , for C a d s B > 9 w t . % ,   and   t u > 0 ,
C F e 2 B x = u = C l o w F e 2 B = 8.83 w t . % , for C a d s B < 8.83 w t . % ,   and   t u > 0 .
The parameter ‘ u 0 ’ denotes a fundamental Fe2B (m) coating that terminates at the point of emergence of the first iron boronizes following a characteristic duration specific to boronizes. Fick’s second law, which incorporates time dependence, is expressed as follows:
C F e 2 B x , t u t u = D F e 2 B 2 C F e 2 B x , t u x 2 .
Equation (4) can be rewritten without time dependence as follows:
d 2 C F e 2 B x d x 2 = 0 .
Solving Equation (5) and considering Equations (2) and (3), we arrive at the following:
C F e 2 B x = C l o w F e 2 B C u p F e 2 B u x + C u p F e 2 B .
Substituting Equation (6) into Equation (1) and considering that C F e ( x ) = c t e . , we obtain the following:
u 2 = 4 D F e 2 B C u p F e 2 B C l o w F e 2 B C l o w F e 2 B 2 C 0 + C u p F e 2 B t t 0 F e 2 B = 4 D F e 2 B ε F e 2 B 2 t u .
In Equation (7), it is observed that the coatings obey the law of parabolic growth with
ε F e 2 B 2 = C u p F e 2 B C l o w F e 2 B / C l o w F e 2 B 2 C 0 + C u p F e 2 B = 9.5 × 10 3 ,
where ε F e 2 B 2 is defined as a concentration-dependent adimensional constant.

3.2. Mass Transfer Model of the Error Function

The solution to Fick’s second law, which incorporates time dependence, is typically expressed in terms of the error function. The error function is a mathematical construct extensively used in various academic disciplines and industries, such as finance, engineering, and physics. Specifically, the error function is widely employed in statistical mechanics, quantum mechanics, and probability theory, among other fields. Equation (4) is commonly used to express Fick’s second law regarding the error function. This efficient approach enables accurate modeling of diverse phenomena using a single mathematical construct. As such, the error function is a valuable tool in many research areas, enabling precise calculations and predictions.
C F e 2 B x , t u = A + B e r f x 2 D F e 2 B t u .
The boundary conditions for C F e 2 B ( x , t u ) with time dependence are:
C F e 2 B x = u 0 , t u = t 0 F e 2 B = C u p F e 2 B = 9 w t . % , for C a d s B > 9 w t . % ,   and   t u > 0 .
C F e 2 B x = u , t u = t u = C l o w F e 2 B = 8.83 w t . % , for C a d s B < 8.83 w t . % ,   and   t u > 0 .
To determine constants A and B for Equation (9), the boundary conditions given in Equations (10) and (11) must be considered. Based on these conditions, the values of A and B can be derived accordingly. We must adhere to this process to obtain accurate and reliable results.
C F e 2 B x , t u = C u p F e 2 B + C l o w F e 2 B C u p F e 2 B e r f u 2 D F e 2 B t u e r f x 2 D F e 2 B t u .
Conversely, the equation governing the balance of matter at the interface of coating/substrate growth can be expressed as follows. This equation plays a significant role in understanding the fundamental processes involved in thin coating growth.
C u p F e 2 B + C l o w F e 2 B 2 C 0 2 d x d t u x = u = J F e 2 B ( x , t u ) x = u J F e ( x , t u + d t u ) x = u + d u = D F e 2 B C F e 2 B x , t u x x = u D F e C F e x , t u + d t u x x = u + d u .
The transfer of boron from the Fe2B surface coating to the substrate at J o u t F e ( x , t u + d t u ) x = u + d u = 0 has yet to be considered. This study assumes that the parabolic growth law applies to the boron coatings, similar to the linear model. We obtained the following result by substituting Equation (12) into Equation (13).
C u p F e 2 B + C l o w F e 2 B 2 C 0 / 2 ε F e 2 B = C l o w F e 2 B C u p F e 2 B e x p ε F e 2 B 2 / π e r f ε F e 2 B .
Equation (14)’s numerical values can be obtained through the implementation of the Newton–Raphson numerical method [26]. This method is frequently employed to determine the roots of an equation and can converge rapidly.
ε F e 2 B 2 = 9.6 × 10 3 .

3.3. Goodman Integral Mass Transfer Model

Fe2B coating growth was studied using Goodman’s integral method [22,23]. Specifically, a semi-infinite solid with a parabolic profile was preliminarily defined for this purpose. Implementation of the method allowed for an in-depth examination of the growth process. The results obtained using this method are expected to provide valuable insights into the fundamental mechanisms underlying the growth of Fe2B coatings.
C F e 2 B x , t u = b u x 2 + a u x + c .
Reference [22] indicates that parameters a and b in Equation (16) are time-dependent functions. Upon closer examination, it is observed that Equation can be rewritten in a different form by applying the chain rule on the left-hand side. Parameters a and b are noted as functions that vary with time.
C l o w F e 2 B 2 C 0 + C u p F e 2 B / 2 C F e 2 B ( x , t u ) t u x = u C F e 2 B ( x , t u ) x x = u = D F e 2 B C F e 2 B ( x , t u ) x x = u .
Combining Equations (4) and (17), the following is obtained.
C l o w F e 2 B 2 C 0 + C u p F e 2 B / 2 D F e 2 B x C F e 2 B ( x , t u ) x x = u C F e 2 B ( x , t u ) x x = u = D F e 2 B C F e 2 B ( x , t u ) x x = u .
Replacing Equation (16) with Equation (18), the following occurs:
a 2 = C u p F e 2 B + C l o w F e 2 B 2 C 0 b .
Similarly, by replacing Equation (16) with Equation (13), we obtain the following:
C u p F e 2 B + C l o w F e 2 B 2 C 0 / 2 d x d t u x = u = D F e 2 B a .
The concentration profile described in Equation (16) is defined within a specific range ( 0 x u ). Substitution of the boundary condition stipulated in Equation (11) into Equation (16) may result in a resultant output.
c = C l o w F e 2 B .
With the result obtained from Equation (21), the concentration profile given in Equation (16) can be rewritten as follows:
C F e 2 B x , t u = b u x 2 + a u x + C l o w F e 2 B .
Applying the boundary condition given by Equations (10)–(22) produces the following:
b = C u p F e 2 B C l o w F e 2 B a u / u 2 .
Combining Equation (19) with Equation (23), we obtain the following:
a = C l o w F e 2 B 2 C 0 + C u p F e 2 B 1 + 4 C u p F e 2 B C l o w F e 2 B / C u p F e 2 B + C l o w F e 2 B 2 C 0 1 2 u .
Substituting Equations (7) and (24) in Equation (20), we find the following:
ε F e 2 B 2 = 1 + 4 C u p F e 2 B C l o w F e 2 B / C u p F e 2 B + C l o w F e 2 B 2 C 0 1 2 9.4 × 10 3 .

3.4. Optional Mass Transfer Diffusion Model

This optional diffusion model postulates an examination of the mass balance equation at the Fe2B/substrate growth interface without considering the shape of the boron concentration profile along the Fe2B coating. Equation (1) can be rephrased by applying the chain rule on the right-hand side of equality.
  C u p F e 2 B + C l o w F e 2 B 2 C 0 2 d x d t u x = u = D F e 2 B d C F e 2 B x d t u d t u d x x = u .
The term d C F e 2 B ( x ) / d t u appearing on the right-hand side of Equation (26) is non-zero when considering the differential of the concentration profile, which is expressed as follows:
  d C F e 2 B x = C F e 2 B x x d x .
Deriving both sides with respect to d t u , we obtain the following:
  d C F e 2 B x d t u = C F e 2 B x x d x d t u .
It is observed that the right-hand side of Equation (28) does not cancel, so d C F e 2 B ( x ) / d t u does not cancel either. Equation (26) can also take the following form:
  C u p F e 2 B + C l o w F e 2 B 2 C 0 2 d x d t u 2 d t u x = u = D F e 2 B d C F e 2 B x x = u .
Substituting Equation (7) on the left side of Equation (29), we obtain the following:
  C u p F e 2 B + C l o w F e 2 B 2 C 0 2 ε F e 2 B 2 d t u t u = d C F e 2 B x x = u .
Integrating both sides of the equal of Equation (30) with the corresponding limits, we obtain the following:
  C u p F e 2 B + C l o w F e 2 B 2 C 0 2 ε F e 2 B 2 t u = t 0 F e 2 B t u = t u d t u t u = C F e 2 B ( x = u 0 ) = C u p F e 2 B C F e 2 B ( x = u ) = C l o w F e 2 B d C F e 2 B x x = u .
The left side of Equation (31) reveals a characteristic initial time for forming the first iron boronize crystals t 0 F e 2 B . In other words, the coating does not attain instantaneous growth at t u = 0 . Equation (31) further elucidates that the formation of crystals necessitates a specific time interval. This phenomenon highlights the intricate process involved in the growth of the coating and the formation of crystals.
ε F e 2 B 2 = 2 C u p F e 2 B C l o w F e 2 B / C u p F e 2 B + C l o w F e 2 B 2 C 0 / l n ( t / t 0 F e 2 B ) .
In the case of ASTM A681 steel, experimental data on Fe2B coating thickness indicate that the characteristic incubation time of iron boronizes is roughly t 0 F e 2 B = 1941.9569 s. As a result, Equation (32) can be approximated accordingly.
  ε F e 2 B 2 = 9.7 × 10 3 .

3.5. Boltzmann–Matano Inverse Diffusion Model

The mathematical models discussed previously are limited to scenarios where the diffusion coefficient, denoted by D F e 2 B , remains constant. However, the diffusion coefficient is subject to variability in practical diffusion experiments. To address this, researchers utilize an inverse diffusion problem approach, where the concentration field, denoted by C F e 2 B ( x , t u ) , serves as a known experimental datum. This approach aims to determine the diffusion coefficient as a function of concentration, a technique widely documented in the literature [27]. Additionally, the nonlinear diffusion equation expresses the second Fick’s law for a linear time-dependent flow with a varying diffusion coefficient, denoted by D F e 2 B ( C F e 2 B ( x , t u ) ) .
  C F e 2 B ( x , t u ) t u = ( D F e 2 B ( C F e 2 B ( x , t u ) ) ) C F e 2 B ( x , t u ) t u .
In certain boundary conditions, Boltzmann demonstrated that the proportionate diffusion coefficient, D F e 2 B ( C F e 2 B ( x , t u ) ) , is exclusively a function of concentration. Moreover, C F e 2 B ( x , t u ) can be represented as a single variable called the similarity variable ( x / 2 t u 1 / 2 ). This permits Equation (34) to be reduced to an ordinary differential equation by introducing a new variable, η . The latter variable is expressed as a function of the following:
  η = x / 2 t u 1 / 2 .
Conditions for a semi-infinite medium:
  C F e 2 B x , t u > 0 = C 0 = 35 × 10 4 w t . % , for x = 0 ,
  C F e 2 B x , t u = 0 = 0 , for x > 0 , .
With the similarity variable, the conditions of Equations (36) and (37) are transformed into the following:
  C F e 2 B η = C 0 = 35 × 10 4 w t . % , for η = 0 ,
  C F e 2 B η = 0 , for η = .
With this new variable, we obtain the following:
  C F e 2 B x , t u x = C F e 2 B η η η x = 1 2 t u 1 / 2 d C F e 2 B η d η ,
  C F e 2 B x , t u t u = C F e 2 B η η η t u = x 4 t u 3 / 2 d C F e 2 B η d η ,
x D F e 2 B C F e 2 B x , t u C F e 2 B x , t u x = x D F e 2 B C F e 2 B x , t u 1 2 t u 1 / 2 d C F e 2 B η d η = d 4 t u d η D F e 2 B C F e 2 B η d C F e 2 B η d η .
Substituting Equations (41) and (42) into Equation (34), obtains the following:
2 η d C F e 2 B η d η = d d η D F e 2 B C F e 2 B η d C F e 2 B η d η .
Since Equation (43) contains only total differentials, we can cancel 1 / d η from each side and integrate between C F e 2 B = 0 and C F e 2 B = C F e 2 B , where C F e 2 B is a concentration in the range 0 < C F e 2 B < C 0 .
2 C F e 2 B = 0 C F e 2 B = C F e 2 B η d C F e 2 B η = C F e 2 B = 0 C F e 2 B = C F e 2 B d D F e 2 B C F e 2 B η d C F e 2 B η d η = D F e 2 B C F e 2 B η d C F e 2 B η d η 0 C F e 2 B .
Since D F e 2 B ( C F e 2 B ( η ) ) d C F e 2 B ( η ) / d η = 0 , then C F e 2 B = 0 . Finally, rewriting Equation (42), we obtain the following:
  D F e 2 B C F e 2 B = 2 d x / 2 t u 1 / 2 d C F e 2 B x , t u 0 C F e 2 B x / 2 t u 1 / 2 d C F e 2 B x , t u = 1 2 t u d x d C F e 2 B x , t u 0 C F e 2 B x d C F e 2 B x , t u .
The diffusion coefficient of the dissolved component is represented by D F e 2 B ( C F e 2 B ) . In contrast, d x / d C F e 2 B ( x , t u ) denotes the inverse of the dissolved component C F e 2 B ( x , t u ) concentration profile derivative concerning the diffusion path in the x phase. By utilizing the Ratajski proposal [28], we may rewrite the integral of Equation (45) as follows:
  D F e 2 B C F e 2 B = 1 2 t u d x d C F e 2 B x , t u 0 x i C F e 2 B x , t u d x + x i C F e 2 B x , t u d x .
With
0 x i C F e 2 B x , t u d x + x i C F e 2 B x , t u d x = j = 1 i 1 c ¯ i Δ x j + c ¯ i Δ x i 2 + c ¯ i + c i + 1 2 Δ x i 2 + j = i + 1 n c ¯ j Δ x j , = j = 1 i 1 c ¯ i Δ x j + 1 4 Δ x i 3 c ¯ i + c i + 1 + j = i + 1 n c ¯ j Δ x j .
Substituting Equation (47) into Equation (46), we obtain the following:
  D F e 2 B C F e 2 B = 1 2 t u Δ c i j = 1 n δ i j Δ x i Δ x j .
Δ c i represents the difference of concentrations of the diffusion element of the i-th phase. On the other hand, the following can be obtained:
  δ i j = c ¯ i i > j ( 3 c ¯ i + c i , i + 1 ) / 4 , j = 1 c ¯ j i < j .
The following Equation is presented to transform Equation (48) for a single-phase Fe2B boron coating: c ¯ i represents the average value of boron concentration in the i-th phase, while c i , i + 1 represents the lower limit of boron in the i-th phase, and Δ x i represents the thickness of the boronize coating of the i-th phase, as determined by Equation (7). It is important to note that formal terminology is necessary for clear and concise communication in academic settings.
  D F e 2 B = 1 2 t u Δ c 1 δ 11 Δ x 1 2 = 1 2 t u Δ c 1 δ 11 u 2 = 2 δ 11 4 t u Δ c 1 u 2 .
Equation (50) represents the diffusion coefficient of the Fe2B phase; the parameters on which it depends are expressed as follows: Δ c 1 = C u p F e 2 B C l o w F e 2 B = 0.17 w t . % , δ 11 = ( 3 c ¯ 1 + c ¯ 12 ) / 4 = 8.893 w t . % , c ¯ 1 = ( C u p F e 2 B + C l o w F e 2 B ) / 2 = 8.915 w t . % , and c ¯ 12 = C l o w F e 2 B = 8.83 w t . % , and u 2 = 4 D F e 2 B ε F e 2 B 2 t u . Thus, from Equation (50), we obtain the following:
  ε F e 2 B 2 = 9.6 × 10 3 .

4. Results

4.1. SEM Cross-Sections of the Boronize Coatings

Figure 9 depicts the cross-sectional views of boronize coatings formed on the surface of an ASTM A681 steel, subjected to a treatment temperature of 1273 K over varying periods ranging from 2 to 8 h. The morphology of the boronizes is characterized by an acicular shape, which indicates perpendicular growth relative to the material’s surface. This observation suggests that the boronizes formed through an epitaxial mechanism, in which the development of the crystals is aligned with the underlying crystal structure of the steel. The results of this study have important implications for the design of high-performance materials and coatings, as well as for the development of novel processing techniques that leverage the unique properties of boronize coatings.
The obtained boronize coating thickness ranged from 58 ± 9.61 µm after two hours to 96 ± 15.9 µm after 8 h at a constant temperature of 1273 K. A research study by VillaVelazquez-Mendoza [29] revealed that the process temperature had a significant influence, approximately 67%, compared to the treatment time, approximately 16%, on the kinetics of boronized coatings in AISI 1018 steel, according to ANOVA analysis.
Further, Figure 10 illustrates the energy-dispersive spectroscopy analyses performed on the boronized coatings’ cross-sections generated at 1273 K for 8 h at two distinct locations. The results of an EDS analysis of the surface, as presented in Figure 10a, indicate the presence of several elements, including boron, carbon, manganese, silicon, and iron. The formation of the Fe2B phase under boronizing conditions involves a chemical reaction between boron and the base element, iron. However, due to its characteristic low-intensity peak, EDS analysis cannot reliably identify the boron element. Notably, the distinctive boron peak is absent near the interface in Figure 10b, while other components, such as carbon and silicon, which are insoluble in iron boronize, migrate towards the substrate.

4.2. XRD Analysis

Figure 11 presents the X-ray diffraction patterns of the diffraction peaks obtained from the analyzed surface of ASTM A681 steel boronized after 2 and 8 h at 1273 K for both treatment times. In both images, distinct peaks of the Fe2B phase are evident, and their diffracted intensities are highly similar.
The X-ray diffractogram in Figure 11 provides evidence of the presence of Fe2B coatings in ASTM A681 steel. The most prominent peak in the diffractogram corresponds to the crystallographic plane (002). At the same time, the observed variation in diffracted intensities is attributed to the growth of the coatings with a well-defined texture along the most favorable crystallographic direction [001]. This texture minimizes the growth-induced stress [30]. The literature suggests that the phase composition of boronizing coatings on iron and Armco steels depends on both the amount of boron present in the solid boronizing powder mixture and the type of boronizing agent used in other boronizing methods [15].

4.3. Coefficient of Friction

Figure 12 shows the variation of the coefficient of friction concerning sliding distance on the surface of untreated and boron-treated samples at 1273 K for 8 h. The tribological behavior of the two surfaces under dry conditions was evaluated by pressing a diamond indenter on the tested surfaces. Results indicate a significant difference in the coefficient of friction between the two surfaces, owing to the formation of a hard boronize coating on the boronized sample’s surface. The average coefficient of friction on the boronized surface ranged from 0.485 to 0.256, while that of the untreated substrate was recorded between 0.646 and 0.781. The study further compared the results with previously published literature, and the findings agreed with the data reported by other authors, regardless of the bidding conditions. The survey by Türkmen et al. [31] investigated the evolution of the coefficient of friction (COF) versus sliding distance in the case of steel specimens with and without boronizing treatment.

4.4. Evaluation of the Activation Energy of Boron in Diiron Boronize for ASTM A681 Steel

The present study aimed to calculate boron diffusion coefficients using five different mathematical models, namely Equations (8), (15), (25), (33), and (51), along with the parabolic growth law represented by Equation (7). The calculation process required the determination of experimental parabolic growth constants, as provided in Table 2. The coefficients were computed based on the slopes of the curves that depict the relationship between coating thickness squared and exposure time (refer to Figure 13). Notably, the incubation periods remained consistently similar across the chosen temperature range.
Figure 14 shows the relationship between the thickness of the iron boronize (u) Fe2B coatings with treatment time and temperature. The boronize coating thickness increases with treatment time and temperature [32].
Five distinct models were utilized to evaluate the diffusion coefficients of boron in the Fe2B phase, with a maximal boron concentration of 9 wt.%. The results were fitted with the Arrhenius relation [32], as depicted in Figure 15.
Table 3 presents the grouping of the activation energies of boron in the Fe2B phase and their corresponding pre-exponential factors for ASTM A681 steel based on five distinct models. The energies in question were derived from the slopes of l n D l o w F e 2 B = f ( 1 / T ) , as per the Arrhenius relation [32].
Table 4 compares the values of boron activation energies estimated in the present work with those reported for some boronizing steels [9,31,32,33,34,35,36].
Two supplementary boronizing conditions were employed to verify the validity of five models experimentally. The scanning electron microscope (SEM) cross-sectional images in Figure 16 illustrate the findings. The boronize incubation period required to generate the Fe2B coating on ASTM A681 steel was determined to be 1941.9569 s through analysis of the experimental data fit in Figure 15. Equations (8), (15), (25), (33), and (51), in conjunction with the parabolic growth law (as outlined in Equation (7)), were utilized to calculate the estimated Fe2B coating thickness under conditions of 1223 K for 3 h and 1273 K for 1.5 h. Table 5 and Table 6 provide the results of these estimations.
The study revealed that the estimated values of boronize coating thickness, as determined by Equations (8), (15), (25), (33), and (51), are consistent with the experimental outcomes presented in Table 5 and Table 6, for a Fe2B phase surface boron content of 9 wt.%. Therefore, in industrial applications of this steel type, knowledge of the variables that govern the boronizing process is critical in obtaining the optimal Fe2B coating thickness. The thickness of the boronize coating plays a crucial role in protecting against wear and tear, with thin coatings considered effective against adhesive wear and thicker coatings intended to combat abrasive wear. The optimal thickness of the boronize coating varies according to the steel type used, ranging from 50 to 250 µm for low-carbon and low-alloy steels and from 25 to 76 µm for high-alloy steels [37].

5. Discussion

5.1. Growth Kinetics of the Boronized Coatings

In this section, we discuss the differences between the five mass transfer models that have been used to determine the growth kinetics of the boron coatings formed on the surface of ASTM A681 steel; according to the results obtained, it can be noted that the estimated values with the five mathematical models of the boron activation energy (Q = 209.867 kJ·mol−1) for ASTM A681 steel is the same value. Likewise, the calculated values for the pre-exponential factors of the linear model (D0 = 2.257 × 10−2 m2/s), error function (D0 = 2.233 × 10−2 m2/s), Goodman integral (D0 = 2.280 × 10−2 m2/s), optional diffusion (D0 = 2.210 × 10−2 m2/s), and Boltzmann–Matano inverse diffusion (D0 = 2.233 × 10−2 m2/s) show that there is minimal variation. To determine how this coincidence is possible, we analyze the model of the error function (see Equation (14)), developing in a Taylor series to the parameters e x p ( ε F e 2 B 2 ) and e r f ( ε F e 2 B ) considering only the first order; we obtain the following: e x p ( ε F e 2 B 2 ) 1 and e r f ( ε F e 2 B ) 2 ε F e 2 B / π , in such a way that Equation (14) is transformed into the following: ε F e 2 B 2 = C u p F e 2 B C l o w F e 2 B / C l o w F e 2 B 2 C 0 + C u p F e 2 B , which coincides with Equation (8). Continuing with the analysis of Goodman’s integral model of mass transfer (see Equation (25)), developing in a Taylor series to the term 1 + 4 ( C u p F e 2 B C l o w F e 2 B / C u p F e 2 B + C l o w F e 2 B 2 C 0 ) , considering the first order, the following is obtained: 1 + 4 ( C u p F e 2 B C l o w F e 2 B / C u p F e 2 B + C l o w F e 2 B 2 C 0 1 + 2 ( C u p F e 2 B C l o w F e 2 B / C u p F e 2 B + C l o w F e 2 B 2 C 0 ) , so Equation (25) is expressed as follows: ε F e 2 B 2 = C u p F e 2 B C l o w F e 2 B / C l o w F e 2 B 2 C 0 + C u p F e 2 B , as in the previous case, and coincides with the expression given in Equation (8). For the case of the optional mass transfer diffusion model (see Equation (32)), in the term l n ( t u / t 0 F e 2 B ) with t u = t t 0 F e 2 B and t 0 F e 2 B = 1941.9569 s, for t = 2 h, 4 h, 6 h, and 8 h considered, it can be approximated to be as follows: l n ( t u / t 0 F e 2 B ) ≈ 1.95, so Equation (32) can be expressed as follows: ε F e 2 B 2 = 1.025 ( C u p F e 2 B C l o w F e 2 B / C l o w F e 2 B 2 C 0 + C u p F e 2 B ) . It is also observed that it approximates the expression given in Equation (8). Finally, in the Boltzmann–Matano inverse diffusion model, the term ε F e 2 B 2 = C u p F e 2 B C l o w F e 2 B / ( 3 ( C u p F e 2 B + C l o w F e 2 B ) / 2 + C l o w F e 2 B ) / 2 is equivalent to the term C u p F e 2 B C l o w F e 2 B / C l o w F e 2 B 2 C 0 + C u p F e 2 B . Even though diffusion models are different, we have an equivalence between the five mathematical mass transfer models for the growth kinetics of the boronized coatings. We also determined a common adimensional constant ( ε F e 2 B 2 ) for the five models. Because of its simplicity, the linear model is the most suitable for studying growth kinetics; the study of the growth kinetics of boronized coatings is essential in the field of metallurgy and materials engineering, especially in applications where the wear resistance, hardness, and other mechanical properties of metallic materials are sought to be improved. Understanding the growth kinetics of boronize coatings is essential to optimize and automate surface treatment processes, control the quality of the formed coatings, and design materials with specific properties for industrial applications requiring wear and corrosion resistance.

5.2. Microstructural Characterization of the Boronize Coating (SEM)

The images presented in Figure 9 illustrate the vertical sections of boronized samples obtained through the implementation of scanning electron microscopy (SEM). These visual representations unveil an internal structure characterized by two distinct regions: a surface boronize coating and an underlying matrix region. Upon thoroughly examining the formation of boronize coatings in all samples, it was observed that the coatings exhibited continuous coverage on the material surface with consistent and uniform thickness. Figure 9 depicts the single-phase sawtooth morphology (Fe2B) of boronized coatings. Within the tetragonal lattice of Fe2B, the highest atomic density of boron aligns along the crystallographic direction [001], which is perpendicular to the sample surface. This orientation facilitates boron diffusion within the lattice, establishing it as the favored growth direction for Fe2B crystals. Consequently, the crystals manifest a sawtooth morphology, notably prominent in iron boronize coatings. The development of this morphology is contingent upon various factors, including the presence of alloying elements in the base material, their respective concentrations, process temperature, and duration. Typically, boronized coatings formed during the boronizing processes of unalloyed and low-alloy steels exhibit a sawtooth morphology. This particular structural configuration significantly enhances the interconnection and adhesion between the substrate material and the boronized coating. Upon careful examination of the microstructure photographs of all samples in this study and considering the circumstances, it can be deduced that the boronize coatings are exclusively composed of the Fe2B phase. The XRD analysis presented in Section 4.2 further substantiates this deduction. The determination of the average thickness of boronized coatings on boronized samples involved the use of optical microscope software (Image-Pro Plus 6.3). To ensure precise data in assessing the boronize coating’s thickness, it is vital to conduct a comprehensive examination encompassing a significant number of measurements from various segments of the coating.

5.3. Chemical Characterization (XRD and EDS)

X-ray diffraction analysis was performed on boronized samples to determine the phase or phases forming the boronize coating. Whether the boronize coating has a single-phase or multi-phase structure can affect the tribological, mechanical, and adhesion properties of the boronized material. The XRD patterns obtained from the analysis of two samples boronized at a temperature of 1273 K for two distinct time durations of 2 h and 8 h are depicted in Figure 11. Based on the analysis results, it was determined that the boronize coating had a single-phase (Fe2B) structure. XRD analysis of the relevant samples used a broader range of diffraction angles. The XRD analysis data confirm the idea of a single-phase coating structure formation, as expressed in the microstructural examination results in Section 4.2. The vertical cross-sectional SEM image of the boronized sample and EDS analysis results performed in different regions are shown in Figure 10. The EDS analyses aimed to determine the possible presence of elements such as Fe and B in the boronize coating and their concentrations. Elemental analysis was performed on a total of two regions. Upon evaluation of the results obtained from regions 1 and 2 of the sample (at 1273 K for 8 h), the presence of the B element was determined. It is well established that the boronize coating formed through boronizing unalloyed or low-alloyed iron-based materials may contain a Fe2B or Fe2B+FeB phase. The Fe-B phase diagram shows that the Fe2B and FeB phases consist of approximately 9 wt% and 16.4 wt% B, respectively. The amount of element B in the boronize coatings was determined to be at least 6.03% and 8.15% by weight. Comparing these values with the theoretical weight per cent of element B in the Fe2B phase, it is clear that they are relatively close. It should be noted that the EDS method may not accurately detect elements with low atomic numbers, particularly those below eight. As a result, it was concluded that the Fe2B coating was formed on the surface of ASTM A681 steel by boronizing processes since the determined values were below the amount of boron in the FeB structure and closer to the amount of boron in the Fe2B structure.

5.4. Pin-on-Disc Test (COF)

The graphs depicted in Figure 12 present the variation in COF (coefficient of friction) observed during the wear test of the samples. The COF values were derived from the friction forces recorded about the sliding distance, calculated under steady-state wear conditions (200–800 m). Upon analyzing the initial COF values obtained after the wear test, it is evident that the initial friction coefficient of the unboronized sample is approximately three times higher than that of the boronized sample. The low coefficient of friction (COF) exhibited by boronized steels can be ascribed to the interaction between iron (Fe) and boron (B) within the boronize coating and the oxygen present in the surrounding atmosphere [38]. This observed decrease in friction is postulated to be a result of forming an oxide film. In the study, it was found that the initial COF values of boronized sample were lower than those of unboronized sample.

6. Conclusions

This study focused on the study of the growth kinetics through five mathematical models of mass transfer of boronize coatings formed on the surface of ASTM A681 steel commonly used in the manufacture of cutting tools, dies, and various machinery, where their exceptional properties can be very advantageous, using the powder-pack boronizing process. The process involved the use of a mixture of powders containing 33.5% by weight of B4C (primary boron source), 5.4% by weight of KBF4 (activator), and 61.1% by weight of SiC (diluent). The boronizing processes were carried out at temperatures between 1123 and 1273 K, with 2, 4, 6, and 8 h of treatment. The main results are listed below:
  • As a result of the boronizing processes, the formation of a single-phase boronize coating (Fe2B) with sawtooth morphology, homogeneously distributed on the surface of the substrate material, was observed. The highest boronize coating thickness was obtained in the boronized sample at 1273 K for 8 h.
  • The results indicated that the thickness of the Fe2B boronize coating increased proportionally to the treatment time and temperature.
  • The growth kinetics of the boronize coating were observed to follow the parabolic rate law ( u 2 = 4 D F e 2 B ε F e 2 B 2 t u ).
  • The activation energy for the growth of the boronized coating was determined to be 209.867 kJ·mol−1 for all five mathematical mass transfer models.
  • According to the results presented in Figure 12, the values of the coefficients of friction (COF) obtained after the wear test show that the coefficient of friction of the unboronized sample was approximately three times higher than that of the boronized sample.
  • The growth kinetics of boronized coatings were investigated by proposing five mathematical mass transfer models: linear, error function, Goodman integral, optional diffusion, and inverse Boltzmann–Matano diffusion. The results revealed a dimensionless constant ( ε F e 2 B 2 ) shared by all models, concluding that the models are equivalent.
  • The confirmation of the presence of the Fe2B phase was accomplished through the use of X-ray diffraction (1273 K for 8 h).
  • Two extra processing parameters (1223 K for 3 h and 1273 K for 1.5 h) were employed to validate the five mathematical models of mass transfer empirically. According to Table 5 and Table 6, the predicted coatings’ thicknesses were in line with the experiments.
Consequently, the growth kinetics of the boronized coatings can be studied from the linear model due to its simplicity and equivalence with the other models. The practical importance of the growth kinetics of the boronize coatings is fundamental to understanding and controlling the boronizing process because the growth kinetics influences the thickness of the boronize coating, which in turn affects the final properties of the material, such as wear resistance, corrosion resistance, and surface hardness. The boronizing process has experienced significant advances in recent decades and is expected to continue to develop in several directions; one of them is to understand the growth kinetics with other modeling alternatives, such as artificial neural networks, because it is essential to optimize the boronizing process to maximize efficiency and minimize costs, and thus ensure consistent production. Finally, the proposed mathematical models of mass transfer allow estimating the thickness of the boronize coatings on the surface of the material, which favors the control of factors such as temperature, time, and boriding medium to achieve the desired properties in the surface coating. For example, in the aerospace industry, it is essential to control the thickness of the boronized coating, because wear and corrosion resistance are critical. Components such as turbine shafts, bearings, and structural parts are hardened through boronizing treatment.

Author Contributions

Conceptualization, M.O.-D. and Á.J.M.-R.; formal analysis, M.O.-D., Á.J.M.-R., O.A.G.-V. and G.M.-G.; investigation, M.O.-D., O.A.G.-V. and G.M.-G.; methodology, M.O.-D. and O.A.G.-V.; software, Á.J.M.-R. and G.M.-G.; supervision, M.O.-D. and O.A.G.-V.; validation, O.A.G.-V., Á.J.M.-R. and G.M.-G.; visualization, M.O.-D. and O.A.G.-V.; writing—original draft preparation, M.O.-D. and O.A.G.-V.; writing—review and editing, M.O.-D., O.A.G.-V., Á.J.M.-R. and G.M.-G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Council for The Humanities, Sciences and Technologies, grant number 1034962.

Data Availability Statement

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

Acknowledgments

The authors would like to acknowledge the Metallographic Laboratory of the Escuela Superior de Ciudad Sahagún-UAEH. The authors would also like to acknowledge the support given by Miguel Ángel Abreu González, who is a student at the Colegio de Ciencias y Humanidades Plantel Azcapotzalco UNAM, for the preparation and analysis of the samples.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yıldız, I. Tribological properties and characterization of borided Co–Mg alloys. Open Chem. 2022, 20, 277–286. [Google Scholar] [CrossRef]
  2. Milinović, A.; Marušić, V.; Konjatić, P.; Berić, N. Effect of carbon content and boronizing parameters on growth kinetics of boride layers obtained on carbon steels. Materials 2022, 15, 1858. [Google Scholar] [CrossRef]
  3. Ortiz–Domínguez, M.; Campos–Silva, I.; Hernández–Sánchez, E.; Nava–Sánchez, J.; Martínez–Trinidad, J.; Jiménez–Reyes, M.Y.; Damián–Mejía, O. Estimation of Fe2B growth on low-carbon steel based on two diffusion models. Int. J. Mater. Res. 2011, 102, 429–434. [Google Scholar] [CrossRef]
  4. Arslan-Kaba, M.; Karimzadehkhoei, M.; Keddam, M.; Timur, S.; Sireli, G.K. An experimental and modelling study on pulse current integrated CRTD–Bor process. Mater. Chem. Phys. 2023, 302, 127735. [Google Scholar] [CrossRef]
  5. Bouarour, B.; Keddam, M.; Boumaai, B. Growth kinetics of diiron boride (Fe2B) coating on a carbon steel by four approaches. Corros. Mater. Prot. J. 2022, 66, 1–6. [Google Scholar]
  6. Ramdan, R.D.; Takaki, T.; Yashiro, K.; Tomita, Y. The Effects of structure orientation on the growth of Fe2B boride by multi-phase–field simulation. Mater. Trans. 2010, 51, 62–67. [Google Scholar] [CrossRef]
  7. Campos, I.; Islas, M.; González, E.; Ponce, P.; Ramírez, G. Use of fuzzy logic for modeling the growth of Fe2B boride coatings during boronizing. Surf. Coat. Technol. 2006, 201, 2717–2723. [Google Scholar] [CrossRef]
  8. Campos, I.; Oseguera, J.; Figueroa, U.; García, J.A.; Bautista, O.; Kelemenis, G. Kinetic study of boron diffusion in the paste-boriding process. Mater. Sci. Eng. A 2003, 352, 261–265. [Google Scholar] [CrossRef]
  9. Nait Abdellah, Z.; Keddam, M.; Jurči, P. Simulation of boronizing kinetics of ASTM A36 steel with the alternative kinetic model and the integral method. Corros. Mater. Prot. J. 2021, 65, 33–39. [Google Scholar] [CrossRef]
  10. Morgado González, I. Implementación de Tratamientos Termoquímicos para el Endurecimiento Superficial de un Disco de Sierra Circular y Evaluación de las Propiedades Mecánicas de las Capas Multicomponenciales Formadas. Ph.D. Thesis, University in the State of Hidalgo (UAEH), Pachuca de Soto, Mexico, 12 April 2023. [Google Scholar]
  11. Akopov, G.; Yeung, M.T.; Kaner, R.B. Rediscovering the crystal chemistry of borides. Adv. Mater. 2017, 29, 1604506. [Google Scholar] [CrossRef]
  12. Benyakoub, K.; Keddam, M.; Boumaali, B.; Kulka, M. Kinetic modelling of powder-pack boronized 4Cr5MoSiV1 steel by two distinct approaches. Coatings 2023, 13, 1132. [Google Scholar] [CrossRef]
  13. Kartal, G.; Timur, S.; Sista, V.; Eryilmaz, O.L.; Erdemir, A. The growth of single Fe2B phase on low carbon steel via phase homogenization in electrochemical boriding (PHEB). Surf. Coat. Technol. 2011, 206, 2005–2011. [Google Scholar] [CrossRef]
  14. Mebarek, B.; Keddam, M.; Kulka, M. Simulation of the incubation time for the formation of (FeB/Fe2B) bilayer on pure iron. Corros. Mater. Prot. J. 2021, 65, 49–56. [Google Scholar] [CrossRef]
  15. Guo, P.; Ma, S.; He, X.; Lv, P.; Luo, Y.; Jia, J.; Cui, X.; Xu, L.; Xing, J. Effects of boride orientation and Si content on high-temperature oxidation resistance of directionally solidified Fe–B alloys. Materials 2022, 15, 7819. [Google Scholar] [CrossRef]
  16. Makuch, N.; Dziarski, P. Importance of trimethyl borate temperature used during gas boriding for microstructure, nanomechanical properties and residual stresses distribution on the cross-section of the produced coating. Surf. Coat. Technol. 2021, 405, 126508. [Google Scholar] [CrossRef]
  17. Sabuz, E.H.; Noor-A-Alam, M.; Waseem Haider, W.; Shabib, I. Improving the mechanical and electrochemical performance of additively manufactured 8620 low alloy steel via boriding. Corros. Mater. Degrad. 2023, 4, 623–643. [Google Scholar] [CrossRef]
  18. de Mendonça Ferreira, S.T.; Bacco, A.L.K.; do Nascimento, E.M.; Lepienski, C.M. Mechanical characterization and micro-wear of FeB-Fe2B layers on boriding AISI D2 and AISI 4340 steels. Mater. Sci. Appl. 2021, 12, 330–344. [Google Scholar]
  19. Crisan, O.; Alina Daniela Crisan, A.D.; Randrianantoandro, N. Temperature-dependent phase evolution in FePt-Based nanocomposite multiple-phased magnetic alloys. Nanomaterials 2022, 12, 4122. [Google Scholar] [CrossRef]
  20. Ruiz, M.A.D.; Huitron, D.S.; Bustos, E.D.G.; Suárez, V.J.C.; Perrusquia, N.L. Effect of the boron powder on surface AISI W2 steel: Experiments and modelling. Adv. Mater. Sci. Eng. 2021, 2021, 5548004. [Google Scholar]
  21. Kul, M.; Yilmaz, Y.; Oskay, K.; Kumruoğlu, L.C. Effect of chemical composition of boriding agent on the optimization of surface hardness and layer thickness on AISI 8620 steel by solid and liquid boriding processes. Adv. Mater. Sci. 2022, 22, 14–22. [Google Scholar] [CrossRef]
  22. Rodríguez-Alemán, S.; Hernández-Cooper, E.M.; Pérez-Álvarez, R.; Otero, J.A. Effects of total thermal balance on the thermal energy absorbed or released by a high-temperature phase change material. Molecules 2021, 26, 365. [Google Scholar] [CrossRef]
  23. Santiago-Acosta, R.D.; Hernández-Cooper, E.M.; Pérez-Álvarez, R.; Otero, J.A. Effects of volume changes on the thermal performance of PCM layers subjected to oscillations of the ambient temperature: Transient and steady periodic regimes. Molecules 2022, 27, 2158. [Google Scholar] [CrossRef]
  24. Orihel, P.; Jurči, P.; Keddam, M. Characterizations and kinetic modelling of boride layers on bohler K190 Steel. Coatings 2023, 13, 1000. [Google Scholar] [CrossRef]
  25. Keddam, M.; Peter Jurči, P. Simulating the growth of dual-phase boride layer on AISI M2 steel by two kinetic approaches. Coatings 2021, 11, 433. [Google Scholar] [CrossRef]
  26. Bonilla-Correa, D.M.; Coronado-Hernández, Ó.E.; Fuertes-Miquel, V.S.; Besharat, M.; Ramos, H.M. Application of Newton–Raphson method for computing the final air–water interface location in a pipe water filling. Water 2023, 15, 1304. [Google Scholar] [CrossRef]
  27. Zhang, S.; Zhang, H.; Zhang, H.; Zhao, X.; Li, Y. Study on diffusion kinetics and law of chromium on the surface of low-carbon steel. Coatings 2023, 13, 98. [Google Scholar] [CrossRef]
  28. Ratajski, J. Model of growth kinetics of nitrided coating in the binary Fe–N system. Int. J. Mater. Res. 2004, 95, 823–828. [Google Scholar] [CrossRef]
  29. VillaVelázquez–Mendoza, C.I.; Rodríguez-Mendoza, J.L.; Ibarra–Galván, V.; Hodgkins, R.P.; López–Valdivieso, A.; Serrato–Palacios, L.L.; Leal-Cruz, A.L.; Ibarra-Junquera, V. Effect of substrate roughness, time and temperature on the processing of iron boride coatings: Experimental and statistical approaches. Int. J. Surf. Sci. Eng. 2014, 8, 71–91. [Google Scholar] [CrossRef]
  30. Palombarini, G.; Carbucicchio, M. Growth of boride coatings on iron. J. Mater. Sci. Lett. 1987, 6, 415–416. [Google Scholar] [CrossRef]
  31. Türkmen, I.; Yalamaç, E. Growth of the Fe2B coating on SAE 1020 steel employed a boron source of H3BO3 during the powder-pack boriding method. J. Alloys Compd. 2018, 744, 658–666. [Google Scholar] [CrossRef]
  32. Kayali, Y.; Kara, R. Investigation of wear behavior and diffusion kinetic values of boronized Hardox-450 steel. Prot. Met. Phys. Chem. Surf. 2021, 57, 1025–1033. [Google Scholar] [CrossRef]
  33. Ruiz-Trabolsi, P.A.; Velázquez, J.C.; Orozco-Álvarez, C.; Carrera-Espinoza, R.; Yescas-Hernández, J.A.; González-Arévalo, N.E.; Hernández-Sánchez, E. Kinetics of the boride coatings obtained on AISI 1018 steel by considering the amount of matter involved. Coatings 2021, 11, 259. [Google Scholar] [CrossRef]
  34. Makuch, N.; Dziarski, P.; Kulka, M.; Keddam, M. Growth kinetics and some mechanical properties of plasma paste borided layers produced on Nimonic 80A-Alloy. Materials 2021, 14, 5146. [Google Scholar] [CrossRef]
  35. Zong, X.; Xia, R.; Zhang, Y.; Zhang, Y.; Zhu, Q. Boriding kinetics and mechanical properties of X65Cr14 martensitic stainless steel by pack method. J. Phys. Conf. Ser. 2022, 2368, 012008. [Google Scholar] [CrossRef]
  36. Uslu, I.; Comert, H.; Ipek, M.; Ozdemir, O.; Bindal, C. Evaluation of borides formed on AISI P20 steel. Mater. Des. 2007, 28, 55–61. [Google Scholar] [CrossRef]
  37. Łukaszewicz, G.; Tacikowski, M.; Kulka, M.; Chmielarz, K.; Świątnicki, A.W. The Effect of hybrid treatment combining boriding and nanobainitising on the tribological and mechanical properties of 66SiMnCrMo6-6-4 bearing steel. Materials 2023, 16, 3436. [Google Scholar] [CrossRef]
  38. Ulutan, M.; Yildirim, M.M.; Çelik, O.N.; Buytoz, S. Tribological properties of borided AISI 4140 steel with the powder pack-boriding method. Tribol. Lett. 2010, 38, 231–239. [Google Scholar] [CrossRef]
Figure 1. Iron–boron (Fe-B) phase diagram.
Figure 1. Iron–boron (Fe-B) phase diagram.
Metals 14 00670 g001
Figure 2. Graphical illustration of the Fe2B boronize coating formed on the surface of ASTM A681 steel.
Figure 2. Graphical illustration of the Fe2B boronize coating formed on the surface of ASTM A681 steel.
Metals 14 00670 g002
Figure 3. The Fe2B phase comprises a tetragonal unit cell centered on the body.
Figure 3. The Fe2B phase comprises a tetragonal unit cell centered on the body.
Metals 14 00670 g003
Figure 4. Schematic illustration and cross-sectional view of the AISI 316L stainless steel cylindrical vessel utilized for the pack-boronizing treatment. The cylindrical vessel comprises four distinct components: (1) a lid equipped with a hole for the release of gases generated as a result of the chemical reaction of the mixture, (2) a boron-rich salt mixture consisting of B4C, KBF4, and SiC, (3) an embedded test tube utilized for hardening, and (4) the cylindrical vessel itself. The dimensions of the vessel are specified in millimetres.
Figure 4. Schematic illustration and cross-sectional view of the AISI 316L stainless steel cylindrical vessel utilized for the pack-boronizing treatment. The cylindrical vessel comprises four distinct components: (1) a lid equipped with a hole for the release of gases generated as a result of the chemical reaction of the mixture, (2) a boron-rich salt mixture consisting of B4C, KBF4, and SiC, (3) an embedded test tube utilized for hardening, and (4) the cylindrical vessel itself. The dimensions of the vessel are specified in millimetres.
Metals 14 00670 g004
Figure 5. Schematic diagram of the boron atom diffusion mechanism for forming the phase Fe2B on ASTM A681 steel.
Figure 5. Schematic diagram of the boron atom diffusion mechanism for forming the phase Fe2B on ASTM A681 steel.
Metals 14 00670 g005
Figure 6. Schematic diagram of the boron atom diffusion mechanism for forming the phase Fe2B on ASTM A681 steel. The metallographic preparation involves various steps, including sectioning, grinding, thermochemical treatment, mounting, grinding, polishing, chemical analysis, and transmitted light microscopy.
Figure 6. Schematic diagram of the boron atom diffusion mechanism for forming the phase Fe2B on ASTM A681 steel. The metallographic preparation involves various steps, including sectioning, grinding, thermochemical treatment, mounting, grinding, polishing, chemical analysis, and transmitted light microscopy.
Metals 14 00670 g006
Figure 7. Average Fe2B coating thickness measurement.
Figure 7. Average Fe2B coating thickness measurement.
Metals 14 00670 g007
Figure 8. C F e 2 B ( x ) for boron diffusion-controlled growth on the surface of initially homogeneous ASTM A681 steel. The area ( C l o w F e 2 B C 0 ) d u indicates the amount of solute required to advance the Fe2B phase.
Figure 8. C F e 2 B ( x ) for boron diffusion-controlled growth on the surface of initially homogeneous ASTM A681 steel. The area ( C l o w F e 2 B C 0 ) d u indicates the amount of solute required to advance the Fe2B phase.
Metals 14 00670 g008
Figure 9. SEM images depicting cross-sectional views of boronized coatings formed on ASTM A681 steel, heated to 1273 K for varying lengths. The images correspond to the following treatment durations: (a) 2 h, (b) 4 h, (c) 6 h, and (d) 8 h.
Figure 9. SEM images depicting cross-sectional views of boronized coatings formed on ASTM A681 steel, heated to 1273 K for varying lengths. The images correspond to the following treatment durations: (a) 2 h, (b) 4 h, (c) 6 h, and (d) 8 h.
Metals 14 00670 g009
Figure 10. Electron-dispersive spectroscopy (EDS) analysis was conducted on the cross-sectional view of ASTM A681 steel that underwent boronizing at 1273 K for 8 h. Two distinct locations were subject to analysis, specifically (a) near the sample surface and (b) near the interface.
Figure 10. Electron-dispersive spectroscopy (EDS) analysis was conducted on the cross-sectional view of ASTM A681 steel that underwent boronizing at 1273 K for 8 h. Two distinct locations were subject to analysis, specifically (a) near the sample surface and (b) near the interface.
Metals 14 00670 g010
Figure 11. X-ray diffraction patterns were obtained on the boronized surface of ASTM A681 steel at an elevated temperature of 1273 K for two distinct time durations of (a) 2 h and (b) 8 h. The obtained X-ray diffraction patterns were analyzed to determine the crystalline phase present in the material.
Figure 11. X-ray diffraction patterns were obtained on the boronized surface of ASTM A681 steel at an elevated temperature of 1273 K for two distinct time durations of (a) 2 h and (b) 8 h. The obtained X-ray diffraction patterns were analyzed to determine the crystalline phase present in the material.
Metals 14 00670 g011
Figure 12. Coefficient of friction as a function of sliding distance L in wear tests of ASTM A681 steel: in the initial condition (unboronized) and after 8 h of treatment at the temperature of 1273 K (boronized).
Figure 12. Coefficient of friction as a function of sliding distance L in wear tests of ASTM A681 steel: in the initial condition (unboronized) and after 8 h of treatment at the temperature of 1273 K (boronized).
Metals 14 00670 g012
Figure 13. Insights into the squared thickness u2 of the Fe2B coating formed on the surface of ASTM A681 steel. The graph depicts the correlation between treatment time and different temperatures.
Figure 13. Insights into the squared thickness u2 of the Fe2B coating formed on the surface of ASTM A681 steel. The graph depicts the correlation between treatment time and different temperatures.
Metals 14 00670 g013
Figure 14. The value of boronize coating thickness varies with respect to treatment time and temperature.
Figure 14. The value of boronize coating thickness varies with respect to treatment time and temperature.
Metals 14 00670 g014
Figure 15. Temperature dependence of the evaluated values of boron diffusivities in the Fe2B phase based on five models: (a) linear model, (b) error function model, (c) Goodman integral model, (d) optional diffusion model, and (e) Boltzmann–Matano inverse diffusion model.
Figure 15. Temperature dependence of the evaluated values of boron diffusivities in the Fe2B phase based on five models: (a) linear model, (b) error function model, (c) Goodman integral model, (d) optional diffusion model, and (e) Boltzmann–Matano inverse diffusion model.
Metals 14 00670 g015aMetals 14 00670 g015bMetals 14 00670 g015c
Figure 16. SEM images of cross-sectional views of boronize coatings formed on treated ASTM A681 steel: (a) 1273 K temperature for 1.5 h and (b) 1223 K temperature for 3 h.
Figure 16. SEM images of cross-sectional views of boronize coatings formed on treated ASTM A681 steel: (a) 1273 K temperature for 1.5 h and (b) 1223 K temperature for 3 h.
Metals 14 00670 g016
Table 1. Exhibits significant data on the microhardness values and the constitution of boronize coatings formed on diverse substrates, obtained after the boronizing process.
Table 1. Exhibits significant data on the microhardness values and the constitution of boronize coatings formed on diverse substrates, obtained after the boronizing process.
SubstrateType of Boronize CoatingHardness of Boronize Coating
(HV)
FeFeB1900–2100
Fe2B1800–2000
CoCoB1850
Co2B1550
Co-27·5CrCoB2200
Co2B1550
NiNi4B31600
Ni2B1500
Ni3B900
WW2B2700
WB2700
W2B52700
NbNb2B42600–3000
NbB42600–3000
ZrZr2B2300–2600
ZrB22300–2600
TaTa2B3200–3500
Ta2BTiB22500
TiB2500
Ti-6Al-4VTiB23000
TiB3000
MoMo2B2400–2700
Mo2B52400–2700
ReReB2700–2900
Re2B2700–2900
Table 2. Experimental kinetic constants at the growth interface (Fe2B coating/substrate) with the presence of incubation times.
Table 2. Experimental kinetic constants at the growth interface (Fe2B coating/substrate) with the presence of incubation times.
Experimental Temperature
T
(K)
Experimental Kinetic Constant
4 ε F e 2 B 2 D F e 2 B
(m2⸱s−1)
Incubation Time
t 0 F e 2 B
(s)
11231.4661 × 10−131941.9976
11733.9790 × 10−131941.9725
12239.2007 × 10−131941.9906
12732.0991 × 10−121941.9869
Table 3. Evaluated values of boron diffusivities in Fe2B coating using five diffusion models.
Table 3. Evaluated values of boron diffusivities in Fe2B coating using five diffusion models.
Pre-Exponential Factor
D0
(m2·s−1)
Activation Energy
Q
(kJ·mol−1)
Diffusion Model
2.257 × 10−2209.867Linear
2.233 × 10−2Error function
2.280 × 10−2Goodman integral
2.210 × 10−2Optional diffusion
2.233 × 10−2Boltzmann-Matano inverse diffusion
Table 4. Comparison of boron activation energy values from this work with literature data.
Table 4. Comparison of boron activation energy values from this work with literature data.
SteelActivation Energy
Q
(kJ·mol−1)
MethodReferences
Hardox-450157.9Empirical growth law[32]
ASTM A36161.0Diffusion model[9]
AISI 101891.20–155.2Empirical growth law[33]
Nimonic
80A-Alloy
190.93Diffusion model[34]
X65Cr14
stainless steel
206.53Empirical growth law[35]
AISI P20200.0Empirical growth law[36]
SAE 1020183.15Empirical growth law[31]
ASTM A681209.867Diffusion modelThis study
Diffusion model
Diffusion model
Diffusion model
Diffusion model
Table 5. Comparing the predicted values of the thickness of the diiron boronize coating with the experimental value obtained at 1273 K for 1.5 h.
Table 5. Comparing the predicted values of the thickness of the diiron boronize coating with the experimental value obtained at 1273 K for 1.5 h.
Boronizing ConditionEmpirical Coating Thickness
(µm)
Thickness of Simulated Coating
(µm)
Models
1273 K
For 1.5 h
83.33 ± 10.5685.155Linear
85.602Error function
84.705Goodman integral
86.046Optional diffusion
85.602Boltzmann-Matano inverse diffusion
Table 6. Comparing the predicted values of the thickness of the diiron boronize coating with the experimental value obtained at 1223 K for 3 h.
Table 6. Comparing the predicted values of the thickness of the diiron boronize coating with the experimental value obtained at 1223 K for 3 h.
Boronizing ConditionEmpirical Coating Thickness
(µm)
Thickness of Simulated Coating
(µm)
Models
1223 K
For 3 h
88.88 ± 11.4890.871Linear
91.348Error function
90.392Goodman integral
91.823Optional diffusion
91.348Boltzmann-Matano inverse diffusion
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

Ortiz-Domínguez, M.; Morales-Robles, Á.J.; Gómez-Vargas, O.A.; Moreno-González, G. Surface Growth of Boronize Coatings Studied with Mathematical Models of Diffusion. Metals 2024, 14, 670. https://doi.org/10.3390/met14060670

AMA Style

Ortiz-Domínguez M, Morales-Robles ÁJ, Gómez-Vargas OA, Moreno-González G. Surface Growth of Boronize Coatings Studied with Mathematical Models of Diffusion. Metals. 2024; 14(6):670. https://doi.org/10.3390/met14060670

Chicago/Turabian Style

Ortiz-Domínguez, Martín, Ángel Jesús Morales-Robles, Oscar Armando Gómez-Vargas, and Georgina Moreno-González. 2024. "Surface Growth of Boronize Coatings Studied with Mathematical Models of Diffusion" Metals 14, no. 6: 670. https://doi.org/10.3390/met14060670

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