Home Physical Sciences Investigation of microstructure, machinability, and mechanical properties of new-generation hybrid lead-free brass alloys
Article Open Access

Investigation of microstructure, machinability, and mechanical properties of new-generation hybrid lead-free brass alloys

  • Özbey Semih EMAIL logo and Artir Recep
Published/Copyright: February 27, 2023

Abstract

In this study, hybrid alloys were obtained by casting method with alloy elements and additive such as Si and MoS2, which can be used instead of lead, and compared with Ecobrass and free cutting brass samples used in the market in terms of microstructure, mechanical, and machinability properties. The microstructures of lead-free hybridized brass consists of alpha, beta, and intermetallic compound which were confirmed by the results of X-Ray Diffraction analysis and Scanning Electron Microscopy-Energy Dispersive Spectroscopy. The hardness values of the beta phase in the microstructure are between 180 and 220 Vickers hardness. It has been observed that increasing the amount of beta prime phase also increases the hardness. The machinability of samples was evaluated in terms of surface roughness and chip formation. Chips obtained from samples after machining process were categorized according to ISO 6385-G1 standard. Chip morphologies were examined under optic microscope and scanning electron microscope. The surface roughness value of samples with MoS2 additives was found to be the lowest due to its lubricity effect. Moreover, morphologies, distribution of phases, and intermetallic compounds in the microstructure are found to have a great impact on the machinability and ultimate tensile strength.

1 Introduction

Brass, whose main alloys are made of copper and zinc, is the preferred non-ferrous material in the industry due to its good mechanical properties, relatively high corrosion resistance, and excellent machinability.

Lead is another alloying element added to brass alloys to improve machinability and is particularly found in the brass alloys which are generally used in the production of fittings for the transportation of sanitary water [1]. Furthermore, it acts as a chip breaker and leads to decrease in tool wear by virtue of lubricative properties [2]. Since lead is one of the very hazardous heavy metals having toxic properties,the presence of it in the composition of some metallic alloys has to be avoided even at very low trace amounts [3]. The contain of lead in metal alloys has been restricted by community service organization such as Restriction of Hazardous Substances 2011/55/EU [RoHS] Directive up to max 0.1%wt. and Safe Drinking Water Act, USA Directive 2000/53/EC [ELV] up to max 0.25%wt. Thus, the development of new lead-free brass alloy to comply with new legislation is required [4]. Brass, whose main alloys are copper and zinc, is the preferred non-ferrous material in the industry due to good mechanical properties, relatively high corrosion resistance and excellent machinability. Lead is non-soluble in brass matrix and generally precipitates with globular form with a diameter of 1–10 µm at grain boundaries [4,5,6].

There are several alternatives in the substitution of lead such as bismuth, silicon, and selenium to enhance the machinability of brass [7,8,9].

Gr particle is one of the most prominent candidates because they are cheap and environmentally friendly and acts as solid lubricant; however, they are hardly distributed in brass matrix by casting method due to high density differences [10]. Bismuth which is another alternative element that can replace lead has similar properties, but it is considered a rare element and is more expensive than lead; therefore, manufacture of brass with bismuth is very costly. Moreover, it segregates to the grain boundaries, which results into the brittleness of the part.

Titanium (Ti) has been used as reinforcement to produce graphite–brass composites. Thus, it has been observed that the interface bonding between Gr and brass matrix is improved by interfacial reaction of Ti with Cu and Gr to form intermetallic compounds, and also increasing addition of nano titanium has enhanced the mechanical properties of brass. It has been determined that although Gr added to the brass alloy matrix with Ti improves the cutting performance, it deteriorates the mechanical properties considerably [11].

Silicon improves machinability of brass by promoting formation of secondary hard phases in the microstructure. On the other hand, it raises the cost of machinability by accelerating tool wear rate [12]. Manufacturing process with selenium of brass is very complex and cost-producibility is very high; therefore, it is not appropriate for mass production with selenium [6].

There are many studies about silicon (Si) and molybdenum disulfide (MoS2) as less toxic additives than lead, but no study has considered the production of brass by combining them. Each alternative has some negative aspects. The aim of this study was to develop a new generation of lead-free brass alloys through the hybridization of Si and MoS2. Both Si and MoS2 are widely available commercially. The new-generation brass alloy was produced with lower-toxicity elements, which could lead to the enhancement of machinability. The low-cost production and easily available additives are other positive aspects. The new alloy could be used in faucet system components and shows similar mechanical properties to other alternative lead-free alloys used in the industry.

2 Experimental process

In order to prepare the hybrid brass alloy, Cu–Si master alloys which had been produced by casting method were placed in a Gr crucible and then commercial pure Cu, and Zn was added into the preheated Gr crucible in an appropriate order. The chemical composition of the main alloying elements supplied for brass production is shown in Table 1. The molten process was carried out at about 1,120°C temperature by induction furnace, and followed by addition of different amount of MoS2 to the molten brass. The crucible was drilled on the edge to measure the desirable temperature with K-type thermocouple. Borax is used as a commercial flux to insulate from the atmosphere and trace amount of high purity aluminum was added into molten brass to minimize zinc evaporation. The molten alloys were poured into ingot mold which made by hot steel tool according to ISODISC w303. The dimensions of the ingot mold are 20 mm in diameter and 150 mm in length.

Table 1

Chemical composition of supplied main alloying elements

Element (wt%) Cu Zn Sn Pb Si
Zinc wire 0.005 99.995 0.0029 0.0017
Copper sheet 94.712 5.852 0.0065 0.2720

The chemical composition of the master alloys and the prepared sample are shown in Table 2. As-received Ecobrass and C3600 leaded brass were used as reference samples.

Table 2

Chemical compositions of samples

Sample Cu Zn Sn S Mo Pb Si
Master alloys (Cu–Si) 80.49 0.321 0.021 0.196 18.24
FCB 57.38 38.98 0.292 2.9
Ecobrass 75.14 21.81 0.047 0.093 2.81
MoS2-1 63.33 32.04 3.46 0.155 0.212 0.541
MoS2-2 62.45 31.53 3.62 0.25 0.345 1.324
MoS2-3 63.55 30.83 2.73 0.628 0.421 1.712

Components produced for faucet system by casting method are directly used after machining process. In order to provide the comparison conditions for the same level, reference samples taken commercially were reproduced by casting method under the same conditions. In this article, C3600 leaded brass is named free cutting brass and will be abbreviated as F.C.B from now on.

The lower parts of casting alloys were cut into slices with 10 mm thickness and then ground by the range of 120 and 2,500 grind paper for metallographic examination, successively. It was polished by solid lard to obtain a flat surface. So as to reveal the microstructure of each sample, electrolytic etching was performed at 2 V for 20 s with an aqueous solution containing 60% phosphoric acid. General microstructure was examined using an optical microscope. The morphology of the prepared samples was determined by scanning electron microscopy (SEM, FEI Sirion XL 30). Phase composition was determined by X-Ray Diffraction (XRD) techniques (Bruker D8 Advanced). The chemical composition of prepared samples was identified by the energy-dispersive X-ray spectrometer (EDS) connected to the SEM instrument, and X-ray fluorescence spectroscopy equipment (Thermo Scientific Niton XL2 Gold).

The machinability of the prepared and reference samples was evaluated in terms of surface roughness and chip formation. Tool wear, which is one of the determining parameters of machinability, could not be examined due to the need of too many samples and taking too much time. Machinability parameters include very different combinations. While choosing these parameters, the previous article about brass alloys were taken as a basis [13]. Machinability test was performed using commercial lathe instrument which has 2 kW spindle power and highest spindle speed of the machine was 2,500 rpm. Machining operation condition is presented in Table 3. In order to eliminate the different structures and defects on the surface of the prepared samples during casting, machining process was performed from the surface prior to main machining operation.

Table 3

Machining operation conditions

Geometrical features of cutting tool Cutting velocity Feed rate Depth of cut
−6° rake angle, 7° clearance angle, nose radius: 0.8 mm (Vc) 50 m·min−1 (f) 0.2 mm·rev−1 (doc) 1.0 mm
1,500 rpm* (D = 15 mm)

The significance of “*” is maximum speed at revolutions per minute.

The samples were formed in accordance with the EN-ISO 6892 standard to carry out a tensile test with a Devotrans DVT GP universal testing machine. Hardness test of transverse cross-section of casting samples was carried out with HMV-2T, Shimadzu microhardness tester instrument under 0.02 kgf for 15 s at room temperature. Measurement was performed five times starting from side of sample to the center of sample.

Most alloying elements exist as a solid solution in the brass matrix. Therefore, the copper-zinc diagram partially loses its validity. In order to determine the phases in the phase diagram of the brass alloy, the zinc equivalent was formulated by equation (1).

This equation can just estimate the phases on the copper-zinc diagram by basing on fictitious zinc equivalent. The formula is described with X% symbol as follows [14]:

(1) X % = C Zn + C i K i C Zn + C Cu + C i K i × 100 ,

where C Cu is the Cu content in brass, C Zn is the Zn content in brass, C i is the sum of additives content in brass, and K i is the fictitious zinc equivalent value of additives.

(2) V α = Total values of alpha phase intensities ( Total values of alpha phase intensities + Total values of beta prime phase peak intensity ) .

The fictitious zinc equivalent value of Si, Al, and Sn elements significantly influences the reaction on binary Cu–Zn diagram and existence of phases on microstructure are 10, 6, and 2, respectively [15]. In this study, zinc equivalent calculations were taken into account while determining the chemical components and phase distributions.

3 Results and discussion

3.1 XRD analysis

Figure 1 shows the result of XRD analysis for the prepared samples. All samples consist of alpha and beta phases, including reference samples. The alpha phase exists as a solid solution in the beta prime phase, ordered intermetallic compound, which transforms from beta phase below 450°C. The peaks at 42.36, 49.64, and 72.24° detected on the diffraction pattern of both the MoS2 samples and reference samples was associated with alpha phase. In addition, the peaks appearing at 43.28, 62.87, and 79.39° on diffraction pattern was corresponded to beta prime phase. The lower peaks detected at 31.37, 36.34, and 62.30° on the diffraction pattern of F.C.B brass was associated with lead element. Undefined peaks were detected at 39.97 and 45.59°, which are considered to belong to к phase by the result of diffraction pattern of Ecobrass. In the study conducted with the same composition, the presence of silicon-rich kappa and gamma phases dispersed in the matrix was detected in differential thermal analysis analysis [16]. In another study with the same composition, it was suggested that this peak, which was not identified because of the analysis performed with a different XRD source, belonged to the kappa phase [17].

Figure 1 
                  XRD pattern results of samples on spectrum.
Figure 1

XRD pattern results of samples on spectrum.

When the diffraction patterns of MoS2 samples were compared, the peak intensities of the beta prime phase increase whereas the peak intensities of the alpha prime phase decrease with the increase in the silicon contents. This case indicated that the increase in the silicon content in microstructure promotes the beta prime phase formation. This result is in agreement with the results of Scanning Electron Microscopy-Energy Dispersive Spectroscopy.

Furthermore, volume fractions of beta prime phase and alpha phase were calculated by using equation (2), where Vα is the volume fraction of alpha phase [18].

Using the same equation, the volume fraction of the beta phase is calculated. Volume fraction of beta prime phase and alpha phase for all the samples are presented in Table 4. The increase in the volume fraction of the beta prime phase in the microstructures of MoS2 samples can be attributed to zinc equivalent value. Most of the additives except nickel additive promote the ratio of beta phase formation [19].

Table 4

Volume fraction of alpha and beta prime phases for all samples

Samples Alpha phase (vol%) Beta prime phase (vol%)
F.C.B 0.625425 0.384575
Ecobrass 0.538664 0.461336
MoS2-1 0.808646 0.191354
MoS2-2 0.305895 0.694105
MoS2-3 0.182758 0.817242

Figure 2 shows from the result of XRD analysis scanned in a narrow angle on diffraction pattern of F.C.B and Ecobrass samples. The presence of silicon atoms, which is a substitutional atom in unit cell in brass matrix, brings about micro-strain due to different radii of silicon and zinc atoms [13]. Since the radius of the silicon atom (1.11 Å) is smaller than the radius of the zinc atom (1.34 Å), it causes the formation of tensile stress on the lattice. The peaks shifted to the right in the result of XRD analysis because the measured d-spacing is smaller than normal due to the tensile sites. This is related to the Poisson ratio.

Figure 2 
                  XRD patterns comparison of samples F.C.B and Ecobrass with narrow scan.
Figure 2

XRD patterns comparison of samples F.C.B and Ecobrass with narrow scan.

3.2 Microstructure examination

The microstructure of the prepared and reference samples obtained from optic microscope is shown in Figure 3. In general, there are significant difference appearing in the samples of microstructure after casting process. The microstructure of Ecobrass is a typical casting structure consists of dendritic structure with coarse grains. Microstructure of F.C.B has fine grains with agglomerated lead particles which are located at alfa–beta interfaces due to the insolubility of lead in alfa and beta prime phases. Dark spots appear on the microstructure of F.C.B (Figure 3a), which was already reported in pervious study [20]. Figure 3c shows Widmanstätten structures in MoS2-1 sample which are separated from each other by grain boundaries. During solidification, the alpha phase preferentially precipitates in the entire beta matrix to form the Widmanstätten structure that is compatible with the hyper-peritectic (L + α → β) composition in the Cu–Zn phase diagram [21]. The formation mechanisms of Widmanstätten structure have been reported in studies with similar composition [20]. The Widmanstätten structure precipitated at the beta phase grain boundary regions as small needle like shapes is shown in Figure 3d. Grains appeared in the microstructure of MoS2-3 sample due to the increase in the addition of silicon amount, thus the dominant phase has become beta prime phase in the matrix. This case has been confirmed by the result of XRD analysis (Figure 1).

Figure 3 
                  Microstructure on longitudinal cross-section of samples taken by optical microscopy (a) F.C.B; (b) Ecobrass; (c) MoS2-1; (d) MoS2-2; and (e) MoS2-3.
Figure 3

Microstructure on longitudinal cross-section of samples taken by optical microscopy (a) F.C.B; (b) Ecobrass; (c) MoS2-1; (d) MoS2-2; and (e) MoS2-3.

The size of the grain in the microstructure of MoS2-3 is smaller than that in the microstructure of MoS2-2 because of the high content of silicon, which results in grain refinement. During solidification, many additions, including alloying elements, form a nuclei to form a new grain at the liquid–solid interface [23]. Furthermore, the alloying element causes inhibition of the driving force of α-phase formation [24].

The longitudinal cross-sectional SEM images of all samples are presented in Figure 4. From SEM and EDS studies, it was found that the microstructure of F.C.B. alloy contains a typical alpha-beta dual phase. Lead particles appear in bright contrast in the brass matrix and are homogenously distributed at grain boundaries (Figure 5a). Since the lead element has low melting point, it maintains the liquid form during solidification and thus it can be seen in Figure 4a as lead particles have agglomerated form at the alpha–beta interface. The SEM micrograph image of the Ecobrass reference sample shows alpha and beta phases with dominant beta phase. The SEM images of MoS2-1 have two Widmanstätten morphologies, one is needle-like formation, while the other is plate-like formation. The plate formation of Widmanstätten α phase is related to Zn atoms constantly diffusing out from the Widmanstätten phase due to the fast mobile regions like grain boundaries. On the other hand, slow mobile Zn atoms pave the way for the formation of needle-like Widmanstätten structure due to the hindering diffusivity from alloy elements. As can be seen from Figure 4c, the island formation is the same morphology as Widmanstätten structure but its orientation is different. The reason for different appearance of Widmanstätten structure is random orientation during solidification and the information about this appearance of the Widmanstätten structure is given in detail in the previous study [25].

Figure 4 
                  Micrographs of SEM under secondary electron images mode: (a) F.C.B; (b) Ecobrass; (c) MoS2-1; (d) MoS2-2; and (e) MoS2-3.
Figure 4

Micrographs of SEM under secondary electron images mode: (a) F.C.B; (b) Ecobrass; (c) MoS2-1; (d) MoS2-2; and (e) MoS2-3.

Figure 5 
                  The results of EDS analysis from regions labeled (a–c) in the SEM image of F.C.B.
Figure 5

The results of EDS analysis from regions labeled (a–c) in the SEM image of F.C.B.

Figure 5 shows the EDS analysis results of the F.C.B sample. EDS analysis was performed to determine the presence of alpha, beta, and secondary particles and their amounts. In the EDS analysis, a selected areas mode was adopted. As shown in Figure 5a, the intensity of lead particles is high from the EDS taken from aggregated white particles, thus it is understood that they have represented lead-rich particles. Since the alpha phase is rich in copper, the EDS analysis in Figure 5b represents the alpha phase. With the increasing amount of zinc in the alloy, the microstructure turns from alpha phase to beta phase and the crystal structure transforms from face-centered cubic to a body-centered cubic crystal system. The EDS analysis in Figure 6c represents the beta phase because the intensity of zinc has been augmented.

Figure 6 
                  The results of EDS analysis from regions labeled (a–c) in the SEM image of MoS2-3.
Figure 6

The results of EDS analysis from regions labeled (a–c) in the SEM image of MoS2-3.

Figure 6 shows the results of EDS analysis taken from two different selected areas and one-point on the microstructure of the MoS2-3 sample. From the result of the EDS analysis of MoS2-3 sample, the peaks of Si and Sn elements on spectrum in Figure 6a have high intensity. Considering the zinc equivalent, this region probably represents the beta phase. The intensities of the peaks of the silicon and tin were augmented, and the existence of the sulfur peak was detected in Figure 6b. This was probably due to the complex intermetallic compounds occurring at grain boundaries.

With the increase of Si content, the morphology has turned into beta grains with large size and Widmanstätten α phase precipitated in acicular form at mostly grain boundaries as shown in (Figure 5d). The microstructure of the MoS2-3 sample consists entirely of beta grains and intermetalic compounds (IMC) particles located at the grain boundaries of beta grains. When the SEM image was examined more closely, it was seen that snowflake-shaped complex compounds precipitated at grain boundaries and their surroundings. These complex compounds were formed with the increase in the amount of MoS2. From the result of the EDS analysis shown in Figure 6c, it is quite difficult to give information about the composition of the IMC in this region. MoS2 can decompose and react with the other alloying elements. In previous studies, it was emphasized that the addition of MoS2 could thermodynamically turn into Cu2S and ZnS compounds at high temperatures, but no information was given about the kinetics of the process [26,27]. From the results of EDS analysis, it is thought that complex compounds with partial sulfides and partial oxides are formed due to the presence of oxygen in the environment and high temperature effect.

3.3 Machinability

3.3.1 Chip formation and morphology

The chip formation and morphology are the fundamentals of machinability characterization. Chip formation types are categorized into two main groups, continuous and discontinuous. Since all of the samples have good chip breakage features, they form discontinuous chip shapes during the machining process.

The obtained chips after machining process were collected and to evaluate their formation and dimensions, as shown in Figure 7. The chip formation of all samples was according to the ISO 6385-G1 standard using photographs taken after the machining process. Disposable chips are light and have small chip dimension for all samples. According to the reports of Copper Development Association Incorporation, the brass material has both higher machinability and results in lower tool-tip wear. In addition, it has superior surface finish quality in terms of aesthetics. Therefore, the power and robustness of the latest high-speed machine tools and shops have the capacity to increase the rates of extracting brass workpiece materials in milling, drilling, and turning operations.

Figure 7 
                     Macrographs of chip formation for (a) F.C.B; (b) Ecobrass; (c) MoS2-1; (d) MoS2-2; and (e) MoS2-3.
Figure 7

Macrographs of chip formation for (a) F.C.B; (b) Ecobrass; (c) MoS2-1; (d) MoS2-2; and (e) MoS2-3.

The chip formation in F.C.B alloy is of elemental type, as shown in Figure 8a. A research report presented by the Copper Development Association Incorporation also confirmed previously the elemental type chip formation. Lead causes a low coefficient of friction and significantly affects the chip form during machining [28]. This case tends to form chips in smaller form compared to other brass alloys; however, it is not to be preferred in industrial environments because one of the main alloying elements is lead which has toxic properties. In addition, fine chip particles scatter around during the chip removal process. Scattering of chips into the various machine components, especially electrical ones, causes dysfunction of equipment over time, such as electrical short circuits. The chips formation of Ecobrass, MoS2-2, and MoS2-3 samples correspond to the arc chip with loose form because of the presence of dominant beta phase. The chips formation of MoS2-1 sample correspond to the short tubular chip form due to alpha Widmanstätten structure. MoS2-1 sample exhibits lower chip breakage performance amongst other samples due to the high percentage of ductile alpha phase [6]. Microstructural changes, phase transformations, and the presence of alloying elements can significantly affect the chip formation and size during the process [29]. Both the phase dispersion and presence of intermetallics in the brass matrix have enormous effect on chip size and shape [30].

Figure 8 
                     Optical microscopy images of chip formation for (a) F.C.B; (b) Ecobrass; (c) MoS2-1; (d) MoS2-2; and (e) MoS2.
Figure 8

Optical microscopy images of chip formation for (a) F.C.B; (b) Ecobrass; (c) MoS2-1; (d) MoS2-2; and (e) MoS2.

Chip morphology substantially affects chip size and formation. Thus, optical and SEM images were used to examine the chip morphology, as shown in Figures 7 and 8. The occurrence and mechanism of chip formation of the samples are different from each other. The chips of FCB have good chip breakage features, and exhibits fragmented form. Often in the literature, this appearance of the sawdust is called “Sawtooth” or “closed-split sawtooth.” The reason for this appearance is that a great deal of fragmentation appears in the FCB chips due to lower plastic deformation in the cutting region. Lead considerably reduces the friction and chip thickness ratio in cutting zone leading to a split saw-tooth type appearance of MoS2-3’s chip shape [31]. The addition of additives has paved the way for an inhomogeneous brass microstructure due to the precipitation of ultrafine IMCs around the grain boundaries. Thus, it can facilitate to be fragmented chips by virtue of existence of shear/slip bands. The more separated each tooth is in the serration, the highest chip fragmentation takes place eventually. Thus, the serrated chips turned into the fragmented chips. Additives with lubricating properties in the brass matrix cause low friction forces and low tool temperatures at the contact length between the tool and the chip, so the chip forms preferentially result in much less serrated formation than saw tooth formation. In addition, the tool life of cutting tools substantially improves due to the low thermal deformation [32].

The chip formation of Ecobrass is completely compatible with saw-tooth formation and not separated like F.C.B and MoS2-3 samples. The chip formation of Ecobrass mainly due to the beta phase, which has higher Si content in the microstructure [14]. The chip formation for MoS2-1 and MoS2-2 samples has a serrated or toothlike appearance. Plastic deformation is more difficult in the cutting region compared to the chip formation of FCB and MoS2-3 samples.

SEM micrographs of broken chips were investigated to determine the mechanism of chip morphology. The chip morphology of Ecobrass sample, which resulted from inhomogeneous deformation due to existence of dominant beta prime phase, has many cracks and traces of crack propagation on the contact surface. In materials prone to brittleness, it is possible to have such cracks on chip surfaces under tension. As shown in Figure 9b, the chip shapes have a serrated appearance in separate segments with a thickness of about 10 µm in the F.C.B. A flat smooth surface which takes place by shear motion of two neighbor fractured surface on chip surface during machining process are noticeable, and related to the ductile fracture surface features. It appears that serrated components appear distinctively on a fragmented chip form for MoS2-3 sample, as shown in Figure 9c, while slip surface and undeformed surface appear on chip surface formed with lamella formation. On the other hand, high-density cracks are abundant on the chip surface of both F.C.B and Ecobrass, bringing about traces of concaveness and a cavity, hence increasing the surface roughness values. Lower plasticity resulted in smooth regions on the shear surface and fracture surface during the machining process. The presence of secondary phases, inclusions, and other irregularities (many interfaces in matrix) induces stress concentration in the microstructure [33].

Figure 9 
                     SEM images of chip surface morphology: (a) Ecobrass, (b) F.C.B, and (c) MoS2-3.
Figure 9

SEM images of chip surface morphology: (a) Ecobrass, (b) F.C.B, and (c) MoS2-3.

3.3.2 Surface roughness

Surface quality is a measurement of surface roughness, which is classified differently in various industries. It is also a desirable characteristic for aesthetics. For this reason, surface roughness is an important criterion for machinability. The outer layers of brass samples called the “chill zone”, were removed at a certain thickness to obtain a relatively homogeneous microstructure. Since this “chill zone” does not reflect the real mechanical behavior of the samples, the part of this zone was removed as chip by turning operation and then machinability test was performed. Among the average surface roughness values of the samples shown in Figure 10, the MoS2-3 sample, had an average surface value of 0.63 µm, which was the lowest value. The F.C.B sample had the highest value with an average surface roughness value of 2.25 μm. However, it has been determined that the lubricating effect of the additives significantly affects the surface roughness. Ecobrass and F.C.B samples contain different ratios of alpha and beta phases and have their Ra values close to each other. The dispersion of Pb within the brass matrix is not homogenous. The cause of which is the preferential accumulation of Pb in globular form at alfa/beta interface [31]. The surface roughness value of MoS2-1 sample is 1.36 μm, showing a better value than the reference samples. In previous studies, the ductility of materials, the presence of secondary phases, and strain hardening significantly affect the surface roughness after machining. Surface irregularities differ depending on the conditions during processing [34]. Compared to the reference samples, the surface roughness values were quite low due to the lubricating property of the MoS2 addition. Its lubricating property is based on the fact that it is structurally composed of a multi-layered structure [35]. In addition, the increased Si alloy element content in brass caused a higher beta volume fraction in the microstructure and caused a decrease in the surface roughness due to the structure passing from the dual phase to the mono phase system.

Figure 10 
                     Average surface roughness values of the samples.
Figure 10

Average surface roughness values of the samples.

3.4 Mechanical properties

3.4.1 Hardness

Figure 11, shows the average values of the hardness samples measured from five different regions of the microstructure. The hardness value of Ecobrass, F.C.B, and MoS2-3 samples exhibited nearly the same behavior, while MoS2-3 and MoS2-1 samples have lower hardness values. The high Si amount in the Ecobrass alloy promotes the formation of the beta prime phase, which has higher hardness than the alpha phase. Fine grains in the FCB sample and the existence of IMCs in MoS2-3, which act as obstacles against dislocation movement, are also factors that cause an increase in the hardness [36]. The high hardness values of the MoS2-3 sample can be attributed to oxide precipitates at the grain boundaries. It has been reported that oxidized intermetallic precipitates formed by chemical reaction increase crack formation and locally increase the microhardness values of these precipitated regions [37].

Figure 11 
                     Average hardness values of the samples.
Figure 11

Average hardness values of the samples.

3.4.2 Ultimate tensile strength

Figure 12 shows both the ultimate tensile strength and elongation of all tested samples. The sample of F.C.B exhibits moderately high tensile strength and excellent elongation properties since microstructure of F.C.B consists of fine grains. Ecobrass sample has the highest ultimate tensile strength due to the presence of dominant beta prime phase. MoS2-1 sample has a moderate percentage elongation and the lowest tensile strength because it has its Widmanstätten structure in the brass matrix. In previous studies, it has been reported that the microstructure wholly consisting of intergranular Widmanstätten structure can promote the formation of cracks, so the microstructure becomes brittle and decreases the mechanical properties [38,39]. The MoS2-3 sample had the lowest percentage of elongation since it had IMCs that precipitated at grain boundaries in the microstructure. Since the diameter of Sn and Zn atoms are larger than the diameter of a Cu atom, it is attracted by the edge dislocations that occur during tensile stress which reduces the elastic energy [40]. Thus, the material gradually begins to lose its ductility. The promising results of the MoS2-2 alloy show desirable mechanical properties due to the small number of acicular Widmanstätten structures that precipitated at the grain boundaries. A small number of Widmanstätten structures in the brass matrix significantly reduces the deterioration of mechanical properties.

Figure 12 
                     Comparison of the ultimate tensile strength and elongation of the samples.
Figure 12

Comparison of the ultimate tensile strength and elongation of the samples.

4 Conclusion

In this study, the microstructure, machinability, and mechanical properties of hybridized new generation brass alloys were investigated. The major findings of this study were as follows:

Characterization:

  1. The increase in the content of Si and MoS2 additives promoted beta phase formation; besides, resulting in variation in morphology formations in brass alloys.

  2. The microstructure and XRD analysis revealed that peritectic reaction has been shifted to lower zinc content by addition of Si in Cu–Zn phase diagram. Thus, the microstructure was changed from duplex (α + β) phases to a single beta phase with the increase in the Si and Sn contents in the brass matrix.

  3. Peaks shifted on the XRD pattern when Sn and Si elements located as substitutional atoms in the crystal lattice induced stress rising fields in the brass matrix.

  4. SEM images revealed that due to the lower atomic mobility during solidification, Widmanstätten α precipitated and grew with acicular morphology inside the grains and at grain boundaries in the microstructure of MoS2-2.

Machinability:

  1. The presence of hard and brittle phases in the brass matrix affected the formation, distribution, and morphology of chips.

  2. The chip fracture mechanism was determined by the presence of hard brittle phases, while the distribution and size of IMC’s influenced the shear band formation and micro voids.

  3. The samples with MoS2 additives had the lowest surface roughness values due to its lubricant effect.

  4. One of the chip formations had type “C” shapes (loose), which are generally found in samples with high beta fraction in the microstructure.

Mechanical properties:

  1. The addition of additives promotes the formation of beta prime phase; thus, bringing about increased hardness values of samples.

  2. Intergranular Widmanstätten structures deteriorate the mechanical properties by promoting crack initiation. On the other hand, Widmanstätten structure with needle-like shape that mostly formed at grain boundaries enhanced the mechanical properties by hindering slip that plane movement.

  3. MoS2-1 sample had low hardness, low mechanical properties, and continuous chip form compared to other samples due to Widmanstätten morphology in the whole microstructure. Variation in the chip size and shape depending on the microstructure, as well as the mechanical and hardness values were greatly affected by the microstructure morphology. In some cases, the presence of soft and hard phases in the microstructure had more effect on the machinability rather than mechanical properties.

Future studies should consider investigating the machinability performance after heat treatment, as well as the corrosion behavior of new-generation hybridized brass alloys.

Acknowledgement

The authors would like to thank Marmara University Technology and Engineering Faculties for their contributions to this study.

  1. Funding information: The authors state no funding involved.

  2. Author contributions: ÖZBEY SEMIH: writing – reviewing and editing the manuscript, performing the experimental studies, and interpretation of data analyses ARTIR RECEP: contributed to the conception of the study.

  3. Conflict of interest: The authors state no conflict of interest.

References

[1] Kuyucak, S. and M. Sahoo. A review of the machinability of copper-base alloys. Canadian Metallurgical Quarterly, Vol. 35, No. 1, Jan. 1996, pp. 1–15. 10.1179/cmq.1996.35.1.1.Search in Google Scholar

[2] Imai, H., S. Li, K. Kondoh, Y. Kosaka, A. Kojima, H. Atsumi, et al. Effect of chromium precipitation on machinability of sintered brass alloys dispersed with graphite particles. Materials Transactions, Vol. 52, No. 7, 2011, pp. 1426–1430. 10.2320/matertrans.M2011051.Search in Google Scholar

[3] Chen, X., A. Hu, M. Li, and D. Mao. Study on the properties of Sn–9Zn–xCr lead-free solder. Journal of Alloys and Compounds, Vol. 460, No. 1–2, 2008, pp. 478–484.10.1016/j.jallcom.2007.05.087Search in Google Scholar

[4] Estelle, A. Drinking water lead regulations: impact on the brass value chain. Materials Science and Technology, Vol. 32, No. 17, 2016, pp. 1763–1770.10.1080/02670836.2016.1220906Search in Google Scholar

[5] Trent, E. M. and P. K. Wright. Metal cutting, Butterworth-Heinemann, Wildwood Avenue, Woburn, MA, USA, 2000.10.1016/B978-075067069-2/50007-3Search in Google Scholar

[6] Nobel, C., F. Klocke, D. Lung, and S. Wolf. Machinability enhancement of lead-free brass alloys. Procedia CIRP, Vol. 14, 2014, pp. 95–100. 10.1016/j.procir.2014.03.018.Search in Google Scholar

[7] Whiting, L., P. Newcombe, and M. Sahoo. Casting characteristics of red brass containing bismuth and selenium. Transactions of the American Foundrymen’s Society, Vol. 103, 1995, pp. 683–691.Search in Google Scholar

[8] Vilarinho, C., J. P. Davim, D. Soares, F. Castro, and J. Barbosa. Influence of the chemical composition on the machinability of brasses. Journal of Materials Processing Technology, Vol. 170, No. 1–2, Dec. 2005, pp. 441–447. 10.1016/j.jmatprotec.2005.05.035.Search in Google Scholar

[9] Saigal, A. and P. Rohatgi. Machinability of cast lead-free yellow brass containing graphite particles. Transactions of the American Foundrymen’s Society, Vol. 104, 1996, pp. 225–228.Search in Google Scholar

[10] Kobayashi, T., T. Maruyama, K. Nakao, and Y. Murakami. Structure and machinability on hot extrusion material of Pb free Cu–Zn system alloy which dispersed fine graphite particles. Journal of Japan Research Institute for Advanced Copper-Base Materials and Technologies, Vol. 42, 2003, pp. 223–228.Search in Google Scholar

[11] Zhang, X., C. Ma, S. Li, D. Pan, and F. Zheng. Interface design of lead-free free-cutting titanium reinforced graphite brass composites and its effect on mechanical properties and cutting performance. Materials Science & Engineering, A: Structural Materials: Properties, Microstructure and Processing, Vol. 774, 2020, id. 138909.10.1016/j.msea.2020.138909Search in Google Scholar

[12] Schultheiss, F., D. Johansson, M. Linde, P. L. Tam, V. Bushlya, J. Zhou, et al. Machinability of CuZn21Si3P brass. Materials Science and Technology, Vol. 32, No. 17, 2016, pp. 1744–1750.10.1080/02670836.2016.1189199Search in Google Scholar

[13] Jayasankar, K., A. Pandey, B. Mishra, and S. Das. Evaluation of microstructural parameters of nanocrystalline Y2O3 by X-ray diffraction peak broadening analysis. Materials Chemistry and Physics, Vol. 171, 2016, pp. 195–200.10.1016/j.matchemphys.2016.01.005Search in Google Scholar

[14] Yang, C., Z. Ding, Q. C. Tao, L. Liang, Y. F. Ding, W. W. Zhang, et al. High-strength and free-cutting silicon brasses designed via the zinc equivalent rule. Materials Science & Engineering, A: Structural Materials: Properties, Microstructure and Processing, Vol. 723, Apr. 2018, pp. 296–305. 10.1016/j.msea.2018.03.055 Search in Google Scholar

[15] Vander Voort, G. F., S. R. Lampman, B. R. Sanders, G. J. Anton, C. Polakowski, J. Kinson, et al. ASM handbook. Metallography and Microstructures, Vol. 9, 2004, id. 44073–0002.10.31399/asm.hb.v09.9781627081771Search in Google Scholar

[16] Taha, M., N. E‐Mahallawy, T. Mousa, R. Hamouda, and A. Yousef. Microstructure and castability of lead‐free silicon brass alloys. Materialwissenschaft und Werkstofftechnik, Vol. 43, No. 8, 2012, pp. 699–704.10.1002/mawe.201200932Search in Google Scholar

[17] Tam, P.L., F. Schultheiss, J.-E. Ståhl, and L. Nyborg. Residual stress analysis of machined lead-free and lead-containing brasses. Materials Science and Technology, Vol. 32, No. 17, 2016, pp. 1789–1793.10.1080/02670836.2016.1223266Search in Google Scholar

[18] Li, S., H. Imai, H. Atsumi, K. Kondoh, A. Kojima, Y. Kosaka, et al. The effects of Ti and Sn alloying elements on precipitation strengthened Cu40Zn brass using powder metallurgy and hot extrusion. Materials Science & Engineering, A: Structural Materials: Properties, Microstructure and Processing, Vol. 535, Feb. 2012, pp. 22–31. 10.1016/j.msea.2011.12.036.Search in Google Scholar

[19] Handbook, M. Properties and selection: Nonferrous alloys and special-purpose materials, Vol. 2, 1990, p. 102.Search in Google Scholar

[20] Pantazopoulos, G. and A. Vazdirvanidis. Characterization of the microstructural aspects of machinable alpha-beta phase brass. Microscopy and Analysis. Vol. 22, No. 5, 2008, pp. 13–16.Search in Google Scholar

[21] Scott, D. A. Metallography and microstructure in ancient and historic metals, Getty Publications, Singapore, 1992.Search in Google Scholar

[22] Rajabi, Z. and H. Doostmohammadi. Effect of addition of tin on the microstructure and machinability of α-brass. Materials Science and Technology, Vol. 34, No. 10, Jul. 2018, pp. 1218–1227. 10.1080/02670836.2018.1435484.Search in Google Scholar

[23] Wang, F., Z.-L. Liu, D. Qiu, J. A. Taylor, M. A. Easton, and M.-X. Zhang. The influence of the effect of solute on the thermodynamic driving force on grain refinement of Al alloys. Metallurgical and Materials Transactions A: Physical Metallurgy and Materials Science, Vol. 46, No. 1, 2015, pp. 505–515.10.1007/s11661-014-2599-0Search in Google Scholar

[24] Alirezaei, M. and H. Doostmohammadi. Microstructure evolution in cast and equilibrium heat-treated CuZn30-(Si) alloys. International Journal of Cast Metals Research, Vol. 29, No. 4, 2016, pp. 222–227.10.1080/13640461.2015.1126430Search in Google Scholar

[25] Hsieh, C.-C., J.-S. Wang, P. T.-Y. Wu, and W. Wu. Microstructural development of brass alloys with various Bi and Pb additions. Metals And Materials International, Vol. 19, No. 6, Nov. 2013, pp. 1173–1179. 10.1007/s12540-013-6002-2.Search in Google Scholar

[26] Sato, T., Y. Hirai, and T. Kobayashi. Development of lead-free bronze with sulfide dispersion for sliding applications. International Journal of Metalcasting, Vol. 11, No. 1, Jan. 2017, pp. 148–154. 10.1007/s40962-016-0099-2.Search in Google Scholar

[27] Sueyoshi, H., K. Inoue, Y. Yamano, Y. Maeda, and K. Yamada. Machinability of copper sulfide-dispersed lead-free bronze. Materials Transactions, Vol. 50, No. 4, 2009, pp. 847–852.10.2320/matertrans.D-MRA2008840Search in Google Scholar

[28] Johansson, J., P. Alm, R. M’Saoubi, P. Malmberg, J.-E. Ståhl, and V. Bushlya. On the function of lead (Pb) in machining brass alloys. Manufacturing Technology, Vol. 120, No. 11, 2022, pp. 7263–7275.10.1007/s00170-022-09205-0Search in Google Scholar

[29] Konečná, R. and S. Fintová. Copper and copper alloys: casting, classification and characteristic microstructures. Copper alloys-early applications and current performance-enhancing processes, InTech, 2012 Accessed: Sep. 28, 2017. [Online]. Available https://www.intechopen.com/download/pdf/30472.10.5772/39014Search in Google Scholar

[30] Atsumi, H., H. Imai, S. Li, K. Kondoh, Y. Kousaka, and A. Kojima. High-strength, lead-free machinable α–β duplex phase brass Cu–40Zn–Cr–Fe–Sn–Bi alloys. Materials Science & Engineering, A: Structural Materials: Properties, Microstructure and Processing, Vol. 529, Nov. 2011, pp. 275–281. 10.1016/j.msea.2011.09.029.Search in Google Scholar

[31] Taha, M. A., N. A. El-Mahallawy, R. M. Hammouda, T. M. Moussa, and M. H. Gheith. Machinability characteristics of lead free-silicon brass alloys as correlated with microstructure and mechanical properties. Ain Shams Engineering Journal, Vol. 3, No. 4, Dec. 2012, pp. 383–392. 10.1016/j.asej.2012.05.004.Search in Google Scholar

[32] Johansson, J., P. Alm, R. M’Saoubi, P. Malmberg, J.-E. Ståhl, and V. Bushlya. On the function of lead (Pb) in machining brass alloys. The International Journal of Advanced Manufacturing Technology, Vol. 120, No. 11, 2022, pp. 7263–7275.10.1007/s00170-022-09205-0Search in Google Scholar

[33] Pantazopoulos, G. and A. Vazdirvanidis. Failure analysis of a fractured leaded-brass (CuZn39Pb3) extruded hexagonal rod. Journal of Failure Analysis and Prevention, Vol. 8, No. 3, 2008, pp. 218–222.10.1007/s11668-007-9084-7Search in Google Scholar

[34] Schultheiss, F., S. Hägglund, V. Bushlya, J. Zhou, and J.-E. Ståhl. Influence of the minimum chip thickness on the obtained surface roughness during turning operations. Procedia Cirp, Vol. 13, 2014, pp. 67–71.10.1016/j.procir.2014.04.012Search in Google Scholar

[35] Donnet, C., J. Martin, T. Le Mogne, and M. Belin. Super-low friction of MoS2 coatings in various environments. Tribology International, Vol. 29, No. 2, 1996, pp. 123–128.10.1016/0301-679X(95)00094-KSearch in Google Scholar

[36] Pelleg, J. Plastic deformation. Mechanical properties of materials, Springer, Dordrecht, 2013, pp. 147–194.10.1007/978-94-007-4342-7_3Search in Google Scholar

[37] Kunčická, L., M. Jambor, A. Weiser, and J. Dvořák. Structural factors inducing cracking of brass fittings. Materials, Vol. 14, No. 12, 2021, id. 3255.10.3390/ma14123255Search in Google Scholar PubMed PubMed Central

[38] Gil, F., M. Ginebra, J. Manero, and J. Planell. Formation of α-Widmanstätten structure: effects of grain size and cooling rate on the Widmanstätten morphologies and on the mechanical properties in Ti6Al4V alloy. Journal of Alloys and Compounds, Vol. 329, No. 1–2, 2001, pp. 142–152.10.1016/S0925-8388(01)01571-7Search in Google Scholar

[39] Liu, Y. Research on the formation mechanism of widmanstatten ferrite in 30Mn2 steel. Emerging computation and information technologies for education, Springer, Berlin, Heidelberg, 2012, pp. 695–699.10.1007/978-3-642-28466-3_94Search in Google Scholar

[40] Choucri, J., A. Balbo, F. Zanotto, V. Grassi, M. E. Touhami, I. Mansouri, et al. Corrosion behavior and susceptibility to stress corrosion cracking of leaded and lead-free brasses in simulated drinking water. Materials, Vol. 15, No. 1, 2021, id. 144.10.3390/ma15010144Search in Google Scholar PubMed PubMed Central

Received: 2022-09-29
Revised: 2022-12-06
Accepted: 2022-12-12
Published Online: 2023-02-27

© 2023 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Research Articles
  2. First-principles investigation of phase stability and elastic properties of Laves phase TaCr2 by ruthenium alloying
  3. Improvement and prediction on high temperature melting characteristics of coal ash
  4. First-principles calculations to investigate the thermal response of the ZrC(1−x)Nx ceramics at extreme conditions
  5. Study on the cladding path during the solidification process of multi-layer cladding of large steel ingots
  6. Thermodynamic analysis of vanadium distribution behavior in blast furnaces and basic oxygen furnaces
  7. Comparison of data-driven prediction methods for comprehensive coke ratio of blast furnace
  8. Effect of different isothermal times on the microstructure and mechanical properties of high-strength rebar
  9. Analysis of the evolution law of oxide inclusions in U75V heavy rail steel during the LF–RH refining process
  10. Simultaneous extraction of uranium and niobium from a low-grade natural betafite ore
  11. Transfer and transformation mechanism of chromium in stainless steel slag in pedosphere
  12. Effect of tool traverse speed on joint line remnant and mechanical properties of friction stir welded 2195-T8 Al–Li alloy joints
  13. Technology and analysis of 08Cr9W3Co3VNbCuBN steel large diameter thick wall pipe welding process
  14. Influence of shielding gas on machining and wear aspects of AISI 310–AISI 2205 dissimilar stainless steel joints
  15. Effect of post-weld heat treatment on 6156 aluminum alloy joint formed by electron beam welding
  16. Ash melting behavior and mechanism of high-calcium bituminous coal in the process of blast furnace pulverized coal injection
  17. Effect of high temperature tempering on the phase composition and structure of steelmaking slag
  18. Numerical simulation of shrinkage porosity defect in billet continuous casting
  19. Influence of submerged entry nozzle on funnel mold surface velocity
  20. Effect of cold-rolling deformation and rare earth yttrium on microstructure and texture of oriented silicon steel
  21. Investigation of microstructure, machinability, and mechanical properties of new-generation hybrid lead-free brass alloys
  22. Soft sensor method of multimode BOF steelmaking endpoint carbon content and temperature based on vMF-WSAE dynamic deep learning
  23. Mechanical properties and nugget evolution in resistance spot welding of Zn–Al–Mg galvanized DC51D steel
  24. Research on the behaviour and mechanism of void welding based on multiple scales
  25. Preparation of CaO–SiO2–Al2O3 inorganic fibers from melting-separated red mud
  26. Study on diffusion kinetics of chromium and nickel electrochemical co-deposition in a NaCl–KCl–NaF–Cr2O3–NiO molten salt
  27. Enhancing the efficiency of polytetrafluoroethylene-modified silica hydrosols coated solar panels by using artificial neural network and response surface methodology
  28. High-temperature corrosion behaviours of nickel–iron-based alloys with different molybdenum and tungsten contents in a coal ash/flue gas environment
  29. Characteristics and purification of Himalayan salt by high temperature melting
  30. Temperature uniformity optimization with power-frequency coordinated variation in multi-source microwave based on sequential quadratic programming
  31. A novel method for CO2 injection direct smelting vanadium steel: Dephosphorization and vanadium retention
  32. A study of the void surface healing mechanism in 316LN steel
  33. Effect of chemical composition and heat treatment on intergranular corrosion and strength of AlMgSiCu alloys
  34. Soft sensor method for endpoint carbon content and temperature of BOF based on multi-cluster dynamic adaptive selection ensemble learning
  35. Evaluating thermal properties and activation energy of phthalonitrile using sulfur-containing curing agents
  36. Investigation of the liquidus temperature calculation method for medium manganese steel
  37. High-temperature corrosion model of Incoloy 800H alloy connected with Ni-201 in MgCl2–KCl heat transfer fluid
  38. Investigation of the microstructure and mechanical properties of Mg–Al–Zn alloy joints formed by different laser welding processes
  39. Effect of refining slag compositions on its melting property and desulphurization
  40. Effect of P and Ti on the agglomeration behavior of Al2O3 inclusions in Fe–P–Ti alloys
  41. Cation-doping effects on the conductivities of the mayenite Ca12Al14O33
  42. Modification of Al2O3 inclusions in SWRH82B steel by La/Y rare-earth element treatment
  43. Possibility of metallic cobalt formation in the oxide scale during high-temperature oxidation of Co-27Cr-6Mo alloy in air
  44. Multi-source microwave heating temperature uniformity study based on adaptive dynamic programming
  45. Round-robin measurement of surface tension of high-temperature liquid platinum free of oxygen adsorption by oscillating droplet method using levitation techniques
  46. High-temperature production of AlN in Mg alloys with ammonia gas
  47. Review Article
  48. Advances in ultrasonic welding of lightweight alloys: A review
  49. Topical Issue on High-temperature Phase Change Materials for Energy Storage
  50. Compositional and thermophysical study of Al–Si- and Zn–Al–Mg-based eutectic alloys for latent heat storage
  51. Corrosion behavior of a Co−Cr−Mo−Si alloy in pure Al and Al−Si melt
  52. Al–Si–Fe alloy-based phase change material for high-temperature thermal energy storage
  53. Density and surface tension measurements of molten Al–Si based alloys
  54. Graphite crucible interaction with Fe–Si–B phase change material in pilot-scale experiments
  55. Topical Issue on Nuclear Energy Application Materials
  56. Dry synthesis of brannerite (UTi2O6) by mechanochemical treatment
  57. Special Issue on Polymer and Composite Materials (PCM) and Graphene and Novel Nanomaterials - Part I
  58. Heat management of LED-based Cu2O deposits on the optimal structure of heat sink
  59. Special Issue on Recent Developments in 3D Printed Carbon Materials - Part I
  60. Porous metal foam flow field and heat evaluation in PEMFC: A review
  61. Special Issue on Advancements in Solar Energy Technologies and Systems
  62. Research on electric energy measurement system based on intelligent sensor data in artificial intelligence environment
  63. Study of photovoltaic integrated prefabricated components for assembled buildings based on sensing technology supported by solar energy
  64. Topical Issue on Focus of Hot Deformation of Metaland High Entropy Alloys - Part I
  65. Performance optimization and investigation of metal-cored filler wires for high-strength steel during gas metal arc welding
  66. Three-dimensional transient heat transfer analysis of micro-plasma arc welding process using volumetric heat source models
Downloaded on 12.12.2025 from https://www.degruyterbrill.com/document/doi/10.1515/htmp-2022-0263/html
Scroll to top button