Next Article in Journal
A Study on Chloride Corrosion Resistance of Reactive Powder Concrete (RPC) with Copper Slag Replacing Quartz Sand under Freeze–Thaw Conditions
Previous Article in Journal
Growth and Structural Characterization of h-LuMnO3 Thin Films Deposited by Direct MOCVD
Previous Article in Special Issue
Study on the Internal Stress and Thermal Anisotropy in Magnesium Alloys Using a Thermal Elastic Viscoplastic Self-Consistent Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Numerical and Experimental Analysis of Stress–Strain Characteristics in DP 600 and TRIP 400/700 Steel Sheets

1
Department of Automotive Production, Faculty of Mechanical Engineering, Technical University of Košice, Mäsiarska 74, 040 01 Košice, Slovakia
2
USSE Research and Development, U.S. Steel Košice s.r.o., Vstupný Areál U.S. Steel, 044 54 Košice, Slovakia
*
Author to whom correspondence should be addressed.
Materials 2024, 17(1), 210; https://doi.org/10.3390/ma17010210
Submission received: 24 November 2023 / Revised: 27 December 2023 / Accepted: 29 December 2023 / Published: 30 December 2023
(This article belongs to the Special Issue Mechanical Behaviors of Materials: Modelling and Measurement)

Abstract

:
The body constitutes the largest proportion of the total vehicle weight. Recently, increasing efforts have been made towards reducing its weight and improving its crashworthiness. By reducing its weight, fuel consumption will be reduced, and this will also translate into lower CO2 emissions. In terms of safety, vehicle body components use high strength steel which can absorb a substantial amount of impact energy. The present study pays attention to DP 600 and TRIP 400/700 stress–strain characteristics at quasi-static strain rates. The stress–strain characteristics of absorption capacity, stiffness, and deformation resistance force were investigated experimentally by tensile tests, three-point bending tests, and numerical simulations. The results indicate the potential for increasing the absorption capacity, stiffness, and deformation resistance force of the vehicle body’s deformable steel components. The present study verified the possibility of replacing physical testing with numerical simulation. A reasonably satisfactory agreement between the experimentally determined stress–strain characteristics and the numerical simulation was achieved, which can reduce the development time of deformable vehicle body components, reduce costs and optimize the selection of materials. The results extend the state of knowledge on the deformation characteristics of high-strength materials and contribute to the optimization of body components in terms of passive safety and weight.

1. Introduction

In the automotive industry, major efforts are made to develop coherent and balanced car concepts combining safety, emissions and affordability requirements. Such a combination of factors is required to be achievable through realistic, technologically and affordably viable production methods while avoiding an excessive burden on the environment. One of the factors that largely determines the success or failure of a given car concept is its safety. In terms of passive safety, a crucial role belongs to the car body, which consists of a front crumple zone, a safe zone for passengers (safety cage) and a rear crumple zone. In the event of a collision, the deformation structure of the vehicle body’s front and rear sections is designed to absorb as much kinetic energy as possible so that the permissible biomechanical limit of overloading the human body is not exceeded in the event of a collision. The deformation of the structure at the front and rear of the vehicle is controlled. The forces generated by the impact in the bumper cross member are distributed to the longitudinal beams. From the longitudinal beams, the stress is transferred to the tunnel, to the sills, to the floor, to the door structure, and through the door pillar to the roof structure beams. In doing so, the longitudinal beams are subjected to a compressive force and the transverse beams and stiffeners are subjected to a bending force. The consequences of a side impact accident are often more severe than those of a frontal impact because the deformation structure of this part of the vehicle has to absorb the impact energy over a shorter path than in a frontal impact [1,2,3,4,5].
The absorption capacity (EA) of a deformable structure depends on the material used (M), the geometry of the deformable zone components (G) and the nature of the deformation force (FD). Different materials (steel, aluminum alloys, magnesium alloys and composite materials) are used in the body’s construction. The impact energy dissipation mechanisms of metallic and composite structures are significantly different. Structural parts of composite materials are brittle and dissipate energy through various combined fracture mechanisms (fiber breakage, matrix cracking, etc.). In contrast, metal structures allow energy dissipation by controlled elastoplastic deformation of the deformation structure components [6,7,8,9].
As mentioned above, the car body consists of a front crumple zone, a passenger safe zone and a rear crumple zone. Due to their required functionality, the material properties of the components used for each zone also differ. For the front and rear crumple zones, dual and multiphase steels (dual-phase steels—DP, complex-phase steels—CP, transformation induced plasticity steels—TRIP, twinning induced plasticity steels—TWIP) are used, which are equipped with very good EA. For the safe zone (safety cage), martensitic steels—MS, hot formed steels—HF, and dual-phase steels >1000 MPa are used, which are characterized by high strength [1,8,9,10,11].
Dual-phase steels (DP) consist of a fine-grained ferritic matrix in which martensite islands are dispersed. Combining these phases gives the material good fatigue resistance and toughness while maintaining good formability and weldability. DP steels are produced by controlled cooling from the austenitic phase or a two-phase ferritic and austenitic phase. A certain volume fraction of austenite is converted to ferrite, and on sudden cooling, the residual austenite is converted to martensite. The total martensite content in DP steels is in the range of 5–30%. As the proportion of the martensitic phase increases, the strength of the steel increases. High strain hardening exponent n values and high EA of impact energy characterize DP steels [8,12,13]. These steels are produced in standardized grades DP350/600, DP 500/800, and DP 600/980. They are used for floor parts, rear bulkheads and deformation stiffeners, front chassis frame, engine mounts, etc. [12].
High-strength multiphase steels with transformation-induced plasticity (TRIP) exhibit an excellent combination of strength and ductility. Their microstructure consists of islands of residual austenite (7–15%) and bainite (30–35%) dispersed in a soft ferritic matrix (50–55%). In some types of TRIP steels, martensite (1–5%) is present. During plastic deformation, austenite is transformed into martensite (TRIP effect). This effect is not concentrated in a particular area but is uniformly distributed throughout the volume. Plates of martensite formed in the slip planes are an obstacle to the movement of dislocations. A martensitic transformation is indicated at its front if a crack develops in the deformed zone during stressing. Plates of martensite then prevent further crack propagation [8,12,13,14,15,16,17]. TRIP steels are produced in the following standardized grades: TRIP 350/600, TRIP 400/700, TRIP 450/800 and TRIP 600/980. Due to their high EA, these steels are used for vehicle body deformable components (transverse and longitudinal beams, B-pillar stiffeners, bumpers, door sills, etc.) [12].
Designers and material engineers frequently face the dilemma of balancing the conflicting demands of vehicle design, weight and crashworthiness. When selecting an appropriate combination of materials for the body deformation structure components, designers need information on their resistance to deformation under compressive or bending forces (Figure 1) and information on their processability (formability, weldability, etc.)—Figure 1 [4,18,19,20].
This information can be predicted based on the results of the tensile test, three-point bending test, compression test (by squeezing the specimen) and numerical simulation. The present paper aims to draw attention to the possibilities of experimental and numerical methods to predict the stress–strain characteristics of materials such as deformation resistance force—FR or FD, the absorbed energy—EA, the nominal value of deformation resistance force—FMR, the efficiency of deformation resistance force—EFR and the characteristic of the stiffness of the deformation structure—CS. The above characteristics were investigated in the research conducted by tensile test, three-point bending test and numerical simulation of three-point bending [4,21,22,23].

2. Materials and Methods

2.1. Materials

The experimental research of stress–strain characteristics was conducted on sheets of micro-alloyed steel HSLA 240, dual-phase steel DP 600 and multiphase steel TRIP 400/700, which are used in the construction of car bodies. HSLA 240 micro-alloyed steel microstructure is formed by fine equiaxial ferrite grains and small pearlite and tertiary cementite formations—Figure 2a. The microstructure of DP 600 dual-phase steel is a fine-grained polyhedral formed by equiaxial grains of ferrite and very fine grains of martensite—Figure 2b. The TRIP 400/700 microstructure is fine-grained, ferritic, and homogeneous, in which the structural components are arranged in rows. Minority very fine formations of residual austenite and martensite occur in the structure—Figure 2c.
The mechanical properties of the materials were determined by tensile testing on a TIRATEST 2300 testing machine (TIRA Maschinenbau GmbH, Rauenstein, Germany) under the provisions of ISO 6892-1:2019, ISO 10113:2020, and ISO 10275:2020 [24,25,26]. The TIRATES 2300 tensile machine is equipped with a tensometric load cell and extensometer. Five samples were tested in the experiment when the strain rate was 0.002 s−1. Then, average values of mechanical properties and their standard deviation were calculated from five measurements. Figure 3 shows the dependences of both the engineering stress σE and the true stress σT on the engineering strain εE and the true strain φT, which were obtained from the testing machine recordings. The measured values of the mechanical properties: yield strength Re, tensile strength Rm, uniform ductility AU, total ductility A80, material constant K0.1–0.2 and n0.1–0.2 are presented in Table 1.
As the basic deformation characteristic of materials used in body construction, EA expresses the ability to absorb kinetic energy EK into their failure. As it follows from the physio-metallurgical nature of strength and plasticity, EA is closely related to the material’s toughness [27,28]. Toughness can be characterized as the mechanical work related to the unit volume to failure of the material—EA/V0. EAE.t/V0 can be determined from the tensile test record from the area under the σE to εE stress–strain curve—Figure 4, where the dependence of σE on εE is presented by the solid line and the dependence of σT on φT is presented by the dashed line. Then, EAEng.t/V0 was evaluated as follows:
E A E n g . t V 0 = i = 1 n σ E ε i ¯ . ε E
Similarly, the EACTM.t/V0 characteristics can be calculated by trapezoid method from the dependence of σT on φT as follows:
E A C T M , t V 0 = i = 1 n σ T φ i ¯ . φ T
As can be observed in Figure 3, the σT to φT curve is less flattened than the σE to εE dependence curve. This is probably since in the uniform strain area, the cross-section shrinks on one side, but on the other side, there is an intense strain hardening. At the point when σT starts to decrease, the formation of a neck and, consequently, the failure of the specimen occurs. The strain course in the area of neck formation was imaged with the ARAMIS 3-D optical system. The ARAMIS SRX measuring system with camera resolution of 12 Mpix and 8 GB of the internal memory was used for DIC testing. The maximum sensing speed was 335 Hz a few seconds before specimen fracture, while from the beginning the sensing speed was set to 20 Hz. The sample before the test without the speckles is shown on the bottom, and the sample with speckles on the white background after the fracture is shown on the top—Figure 4.
Figure 5, Figure 6 and Figure 7 shows a color map of the strain distribution at the point of deformation before failure of the sample. The strong (deep) blue color shows the undeformed area of the sample, the lighter blue color shows the area in which the localization of plastic deformation has not yet occurred, the green color shows the area of the beginning of the localization of plastic deformation, i.e., local narrowing of the cross-section of the sample begins (forming a neck), the yellow color shows the area of the formation of a gradual local narrowing of the cross-section, and the red shows the area just before the failure of the sample. The amount of plastic strain is expressed by the strains φT1, φT2 and φT3 at the limit points SP, WP, CP—Table 2. As can be seen in Figure 5, Figure 6 and Figure 7, in the blue region, the width of the specimen is the same along the length of the specimen. In the blue region, it can be seen that there is a slight narrowing of the sample width in the deformation localization area, in the yellow region the narrowing of the width is more pronounced than in the green region, and in the red region in the middle of the sample the narrowing is the most pronounced. The strain development at the location of plastic deformation was expressed by the coefficient of normal anisotropy by the Lankford’s anisotropy parameter r, which is given by the ratio of the deformation in the direction of the width φ2 of the sample to the deformation in the direction of the thickness φ3 of the sample:
r = φ 2 φ 3
The r values calculated for individual areas are shown in Table 2.
The deformation work WD respectively EATC.t can be calculated as the area under the dependence curve σTφT from the Equation (4):
W D = E A C T . t = S 0 . L 0 φ = 0.002 φ i σ T i . d φ i
If σT from Krupkovsky Equation (5) is expressed as follows:
σ T i = K . φ 0 + φ i n
Consequently, after substituting σT from Equation (5) into Equation (4), the following is the result:
E A C T , t V 0 = 0.002 φ i F K . φ T i , F n d φ = K . φ T , F n + 1 0.002 n + 1 n + 1
From the measured values of forces and elongations, σTi values were calculated using Equation (7):
σ T i = F T i S 0 . E X P φ T i 1
Calculating φTi strains using the Equation (8):
φ T i = l n 1 + A U , i 100 = l n 1 + L L 0
where FTi is a true instantaneous force; φTi is the true strain value (φTi = ln (1 + ∆Li/L0)); ∆Li is the elongation of the specimen under uniaxial tensile stress; L0 is the specific length of the specimen before deformation; S0 is the initial specimen cross-section; and AU,i is uniform ductility.
The values of K, n obtained by regression analysis at different intervals and the calculated values of EATRA.t/V0 using regression analysis under uniaxial tensile stress are given in Table 3.
As mentioned above, in car collisions, the bumper, cross members, sill, B-pillar of the door and roof reinforcement are subjected to bending force. These parts are connected to the body frame structure. In order to make the bending stresses on the deformation components as close as possible to the actual stresses, the ends of the specimens were mechanically clamped in the jaws—Figure 8. For the experimental research, specimens with a width of 30 mm and a length of L = 300 mm were used. Before the deformation of the specimens, a deformation network was etched on their surface. The specimens were deformed to failure at a testing machine traverse speed of 10 mm.min−1. The bending punch and strain gauge sensor were attached to the moving crossbar of the TIRATEST 2300 testing machine. The machine allows recording the bending force Fb and the traverse distance xFigure 9. Points S0, SZ, WZ and CZ were marked on the curve of the dependence of Fb on the strain path x, which delimit the individual deformable zones. Fb in the range between points X0 and SZ increases linearly as a function of xpl. The area between these points can be referred to as the safe zone. This zone has a sufficient margin of plastic properties until failure. The curve starts to flatten gradually in the area between the SZ and WZ points, and a neck forms. This area can be described as a warning zone. In the area between the WZ and CP points, the curve of the dependence of Fp on xpl is quite flattened. The reserve of plastic properties is exhausted. This area can be referred to as the critical zone. In the area beyond the CP point, the specimen is broken—fracture. The deformation work WD,b or the absorbed energy EAb.CTM was evaluated from the Fb record of the xpl test by three-point bending combined stretching the specimen using Equation (9)—Figure 9:
W D . b . C T M = E A b . C T M = 0 x m a x F b x p i d x i = 1 n F b ( x i ¯ x p l
whereas F b x i ¯ is the average value of the force in the selected interval, x i ¯ is the midpoint of the chosen interval [xi−1,xi], and Δx = xi − xi−1.
The vehicle body consists of a series of deformable zones [29], each designed to resist deformation until a certain stress level on the part is reached. The deformation elements’ resistance force (FR.b = Fb) to plastic deformation is limited by the material properties and geometry. A larger resistance (FR.b) of a deformation element under impact increases the deceleration, which may cause the element to exceed its biomechanical limits. It can be seen from Figure 10 that the area under the curve of the dependence of the bending force Fb on the maximum deformation path xpl. max can be simply calculated as the area of a triangle as follows:
E A b . F = F b m a x . x p l m a 2
A Fbmax can be expressed as a function of the stiffness characteristic:
F b m a x = C S b . F . x p l m a x
If we substitute Expression (11) for Fbmax in Equation (10), we derive the following Equation (12):
E A b . F = 1 2 C S b , F . x p l 2
It is possible to determine the stiffness characteristic more precisely if we use EAb.CTM for EAb.F. Then, from Equation (12), after adjustment, we arrive at Equation (13):
C S b . E A = 2 . E A b . C T M x p l 2
Regression analysis can also determine the stiffness characteristics of CSb,RA. When applying it, the following principles should be observed [30]:
  • the interval of the strain path xpl in linear regression should be chosen as wide as possible,
  • the R2 value of the variance in regression analysis should be greater than >0.95.
Using the stiffness characteristic CSb,RA determined by regression analysis, Equation (14) can be used to calculate EAb,RA/V0 as follows:
E A b . R A = C S b , R A 2 . x p l 2
The determination of the stiffness characteristics CSb.RA for each material by regression analysis is shown in Figure 9. The measured values of Fb = FR.b, xpl, calculated values of stiffness characteristics CSb.RA, CSb.EA and EAb.CTM are presented in Table 4.
The failure of the specimens occurred on one side in the center of the unsupported area of the specimen, and on the other side of the unsupported specimen, there was only a narrowing of the specimen cross-section—Figure 10, Figure 11 and Figure 12. From the deformation grid etched on the specimen, the in-plane φT1 and φT2 along the specimen in 11–12 sections were evaluated by the Argus 3D optical system as it is shown in Figure 10, Figure 11 and Figure 12.
The value of the third true strain φT3, i.e., the thickness strain of the specimen, was calculated based on the validity of the volume constancy law according to the Equation (15):
φ T 1 + φ T 2 + φ T 3 = 0
The Equation (15) after adjusting φT3:
φ T 3 = ( φ T 1 + φ T 2 )
The measured strain values at the boundaries of each zone are shown in Table 5.

2.2. Numerical Simulation of the Three-Point Bending Test

To predict the stress–strain characteristics of the materials used, a numerical simulation of the bending test with fixed ends was also used in the PAM STAMP 2G V2012 software environment. A 3D model of the actual physical bending tool created in Creo 2.0 was used to set up the numerical simulation—Figure 13a. This was saved in *igs format, and the individual functional parts of the tool were imported into the simulation software. A specimen (strip of sheet metal) with dimensions corresponding to the real specimen (with a length of 120 mm, a width of 30 mm and a thickness of 0.75 mm) was also modelled. After importing them into the PAM STAMP 2G V2012 environment, the different parts of the tool were set to a position that corresponded to the real process. Subsequently, the properties such as rigidity, movement, friction, etc. of individual parts of the tool were specified—Figure 13b.
The distribution of stresses and strains depends on the behavior of the material. The authors [31,32] describe three models of the elastic–plastic behavior of the material in bending of the cross member: elastic–perfect plastic, elastic–linear plastic, and power-law hardening. The material properties in our numerical simulations were defined by the Hill-48 plasticity condition in combination with the Krupkovsky isotropic hardening model (Equation (5)). The values of Young’s modulus of elasticity E, Poisson’s number µ, strength constant K, strain hardening exponent n and the plastic strain ratios r were specified in the material card, and the boundary conditions (friction coefficient, loading direction, holding force, punch speed, etc.) were set in the process tree of the simulation software [33]. Surface blank was meshed with square elements of 0.5 mm size and the mesh was checked with no errors.
Measured values of materials—strength constant K, strain hardening exponent n and Lankford’s anisotropy parameter r (Table 4)—were set as input values in the numerical simulation of three-point bending. Subsequently, the boundary conditions corresponding to the conditions of the three-point bending test were set. Before the true strain, the plate strip (specimen) was held at the edge by a force (300 kN) to prevent it from pulling out of the jaw. Subsequently, a bending path of 120 mm was defined. The deformation of the specimen was followed by the actual calculation of the stress–strain characteristics (Fb,xpl, in-plane strains φ1T, φ2T and the thickness strain φ3T) in 11 steps. Figure 14, Figure 15 and Figure 16 show the dependences of the bending force Fs.b on the punch path xpl.s, which were obtained by numerical simulation. The elastic xel and plastic xpl phases of the strain path are indicated in these figures. When evaluating EAS.b, the stress–strain characteristics were shifted to the left by xel (approximately 8 to 10 mm) i.e., by the restitution work (elastic work EAel, etc.)—Figure 14, Figure 15 and Figure 16.
Similar to the above, the deformable zones were marked on the curve of the dependence of the deformation force Fb on the strain path xpl. The dependence of Fs.b on xpl.s is linear in the area between the CP and SP points. This zone can be referred to as the safe zone. In this zone, it is possible to fully exploit the properties of metallic materials in controlling the deformation process of body parts in impact. In the area between the SP and WP points, the dependence of Fs.b on xpl.s begins to flatten and the plastic deformation gradually localizes. The specimen at the point of localization begins to narrow. This zone can be referred to as the warning zone (i.e., edge zone), in which the plastic deformation process of the deformed body parts can also still be controlled. In the critical zone between the WP and CP points, a more pronounced flattening of the Fs.b vs, xpl.s dependence occurs. At the CP point, the plastic properties are exhausted; hence, the possibility of controlling the deformation process of body parts during impact is lost. Table 6 shows the measured values of xpl.s, Fs.b, EAs.b, CSs.b in the plastic deformation area. In evaluating these characteristics, the same procedure was used to evaluate the experimentally determined stress–strain characteristics of the materials used.

3. Results and Discussion

The basic deformation characteristic of the materials used in the construction of the body is the ability to absorb (EA) kinetic energy (EK) upon impact to material failure. As follows from Equation (1), the lower predictive ability of EAEng.t is related to the fact that the initial S0 cross-section and not the actual ST cross-section is used in its determination. The predictive ability of the dependence σT on φT using the Krupkovsky equation was also verified outside the interval recommended by ISO 10275:2020 [26], namely in the following strain intervals: φT from 0.002 to the uniform elongation AU (pre-break φT,F), φT,F from 0.5 to the uniform elongation AU (pre-break φT,F), and φT from 0.1 to the uniform elongation AU (fracture or pre-break φT,F)—Figure 17, Figure 18 and Figure 19.
The designer’s goal during the process of car body design is to reduce its weight and increase resistance to impact. Criteria for optimization for the impact resistance of body components are energy absorption, stiffness, etc. The absorption capacity EACTM.t and EAb.CTM of the investigated materials was determined from the record of the tensile test and the three-point bending test using the trapezoidal method using Equations (2) and (9). The results obtained by the trapezoidal method were considered as a reference when comparing several assessment methods. The Figure 20 shows a comparison of the absorption capacity of micro-alloyed steel HSLA 240, two-phase steel DP 600, and multiphase steel TRIP 400/700 from the tensile test record.
From the comparison of EACTM.t DP 600 against micro-alloyed steel HSLA 240, it can be seen that this characteristic has increased by 30% and compared to TRIP 400/700 by up to 79%. The increase in absorption capacity of DP 600 and TRIP 400/700 steel compared to micro-alloyed steel is related to the strength coefficient K and strain hardening exponent n. In the case of DP 600, the value of K was greater by 47% and in the case of n by 13%. In the case of TRIP 400/700, the value of K was greater by up to 204% and n was greater by 61%. From the mentioned comparison, it follows that materials with a two- and multiphase structure strengthen more intensively than materials with a single-phase ferritic structure. The evaluation of EACTM.t by the trapezoidal method is accurate but time-consuming, therefore the Relation (6) was derived. From the comparison of EACTM.t and EACT.t values calculated according to Relation (6) shows that agreement between these characteristics was achieved in the range from 1% to 6%. When qualitatively evaluating the dependence model σT on φT on Figure 21, we observe that R2 for the dependence of EARA.t/V0 and the measured values of EACTM.t/V0 takes the value of 0.9998. Similarly with the dependence of EACT.t/V0 on EACTM.t R2 was 0.9932. The dependence of EACT.t/V0 on EACTM.t shows a tendency that can be expressed as follows:
E A C T . t / V 0 = 1.0212 . E A C T M T . t / V 0 + 0.931
If this tendency is confirmed or, if it could be, refined, then after substituting for EACT.t/V0 the expression from the right side of Equation (6) we derive:
K . φ T , F n + 1 0.002 n + 1 n + 1 = 1.0212 . E A C T M T . t / V 0 + 0.931
Deformable parts of the body (bumper, B-pillar, door and roof reinforcements, etc.) are exposed to a bending force during an impact. In this case, the results of EAb.CTM evaluated by the trapezoidal method were considered as a reference when comparing. The absorption capacity (EAb.CTM) of the materials used was evaluated from the recording of the dependence of the bending force (Fb) on the deformation path x. The Figure 22 shows a comparison of EAb.CTM of the investigated materials. From the comparison, it can be seen that the material DP 600 shows 1.3 times greater absorption potential and the material TRIP 400/700 approx. 1.9 times greater impact energy absorption potential than HSLA 240 material.
Another deformation characteristic is stiffness (CS). The CSb.EA value calculated according to Relation (13) was considered as a reference value when comparing the stiffness of the investigated materials. It can be seen from the CSb.EA comparison shown in the Figure 23, that DP 600 material exhibits 1.54 times greater stiffness and TRIP 400/700 approximately 1.74 times greater stiffness than HSLA 240. Based on the above, it can be assumed that DP 600 materials and TRIP 400/700 have the potential to reduce the weight of body parts (using thinner wall thickness components) while maintaining or even improving impact resistance properties compared to HSLA 240.
Furthermore, the possibility of using regression analysis in determining the characteristics of CSb.RA was verified. A comparison of CSb.RA and CSb.EA characteristics show that the agreement between these characteristics ranges from 86% to 95%. The possibility of predicting EA and CS was also verified based on the results of numerical simulation. For EA, concordance with EAs.b ranged from 86% to 90%, and for CS.s.b ranged from 64% to 67%. The issue of verifying the results of numerical simulation will be addressed in further research.
The characteristics of EA and CS can guide designers when selecting materials for deformable body parts. For example, if Fb loads the deformation element, the deceleration can be expressed from the energy balance condition:
E K = W D , b = E A b . C T M
Assuming that a vehicle of weight m moves on a path xpl with acceleration a acquires kinetic energy:
E K = 0 x p l m . a . d x = m . a . x p l
This EK on impact is consumed by the WD or absorbed by the deforming structure in the xpl path. The vehicle moves on this path with a deceleration a. When substituted into the energy equilibrium condition (Equation (19), we obtain:
m . a . x p l = C S b . E A . x p l 2 2
After adjusting the deceleration a is as follows:
a = C S b . E A . x p l 2 . m
The effect of material substitution on deceleration can be expressed by the deceleration change index Ia:
I a = a i a r e f
whereas ai is a deceleration in the use of innovative material, aref is the deceleration when using the reference material.
If we assume that on the same vehicle (of the same weight), the reference material of micro-alloyed HSLA 240 steel will be replaced with DP 600 dual-phase steel, then after substituting for ai in Equation (23) and adjusting, the result is as follows:
I a   D P 600 = C S b . E A   D P 600 C S b . E A   H S L A 240 = 0.441 0.338 = 1.30
When replacing DP 600 with TRIP 400/700 steel:
I a   T R I P 400 / 700 = C S b . E A   T R I P 400 / 700 C S b . E A   H S L A 240 = 0.571 0.338 = 1.69
After substitution for aRA.b:
I a R A , b   D P 600 = C S R A . b   D P 600 C S R A . b   H S L A 240 = 0.482 0.361 = 1.33
Similarly, when replacing DP 600 with TRIP 400/700 steel:
I a R A , b   T R I P 400 / 700 = C S R A . b   T R I P 400 / 700 C S R A . b   H S L A 240 = 0.607 0.361 = 1.68
The comparison of the Ia indices calculated using the CSb.EA Equation (13) and the regression analysis CSb.RA indicates that the differences are very small, ranging from −1% to 3%. When substituting HSLA 240 with DP 600, a change in the deceleration index of up to 30% can be expected, and when substituting HSLA 240 with TRIP 400/700, a change in deceleration of up to 69% can be expected.
Another important characteristic is the strain path xpl, at which controlled deformation of the deformation structure occurs. From the comparison of the experimentally measured xpl values and the calculated xpl,s shown in Table 6, it can be observed that there was a difference of 4% for the HSLA 240 material with a single-phase structure and a difference of 27% to 32% for the DP 600 material with a two-phase structure. The largest difference between xpl and xpl,s was observed for the TRIP 400/700 material with a three-phase structure (30% to 43%)—Table 6. We assume that these differences are related to the model used for prediction in the program file. Although plastic deformation is localized in the warning zone between the SP and WP points (the beginning of the neck formation was observed on the specimen), it is still possible to control the body parts’ deformation processes at impact effectively.
For predicting deformation states can be used FLC [34,35,36]. These curves allow the separation of the different transition zones from the safe deformation behavior of the material (with a certain margin) to the states where plastic deformation is localized, i.e., neck formation and subsequent failure. In FLC design, it is important to know the value of FLC0 i.e., the state when at φ2T = 0 the value of φ1T (FLC0 = φ1T) is determined. The example is shown in Figure 24 for DP 600 material. Based on the results of [34], the value of FLC0 for HSLA 260 material is about 0.3, based on [35], the value of FLC0 for DP 600 material is about 0.2 and based on [36] for TRIP 400/600 material the value is about 0.2. These FLC0CP values for separating the critical zone from the warning zone (FLC0WP) are recommended to be shifted by 10%—Table 7.
The curve that separates the critical zone from the warning zone is indicated in the FLD by a dashed line—Figure 24. A comparison of the numerical simulation results presented in Table 7 with the experimental results published in [32,33,34] shows for the HSLA 240 material, with the assumed strain φ2WP = 0, the FLD corresponds to FLC0WP.s = 0. 27; for the DP 600 material the value FLC0WP.s = 0.18 [33]; and for the TRIP 400/700 material the value FLC0WP.s = 0.18 [34]. The above comparison shows that the differences of FLC0WP.s obtained by simulation for the DP 600 and TRIP 400/700 material are 50% as large as the data published in [35,36].
Based on the strain measurements φ1T and φ2T.b using ARAMIS on the tensile test specimens, the FLC0WP value for the warning zone was evaluated—Table 7. The WP point defined the strain at the interface between the warning zone. At point WP, the measured strain values φ1T and φ2T started to deviate from the directive of their interdependence (R2 > 0.99), or the difference of the values of the normal anisotropy coefficient r was greater than 0.01. This difference then increased more and more—Figure 25, Figure 26 and Figure 27. From the combination of the values φ1T and φ2T, the value of FLC0WP was calculated at the defined point WP from the well-known equation [37]:
φ 1 T i = F L C 0 W P φ 2 T i
Calculated values of FLC0WP are presented in Table 7.

4. Conclusions

The deformation properties were analyzed for HSLA 240, DP 600 and TRIP 400/700. From the results obtained, it was found that for the EA prediction of the materials used, it is appropriate to use the material constants K and n determined using the Krupkovsky regression model, which showed a 99.98% agreement between the measured EACTM.t values and the calculated EARA.t values. The results indicate that EARA.t, EAb.RA and CSb.RA strongly correlate with the material constant K and n. Further, it can be concluded that DP 600 and TRIP 400/700 materials can potentially reduce intrusion into the passenger compartment. It should also be noted that when replacing (material upgrading) the micro-alloyed HSLA 240 steel with DP 600 or TRIP 400/700 dual-phase steel, the human body will be overloaded during impact. The results indicate that the EARA.t, EAb.RA and CSb.RA characteristics depend on the microstructure state (single-phase, dual-phase or multiphase, phase transformations, etc.) occurring during deformation. Future research will focus on the effect of strain rate, the nature of the deformation force, the geometry of the deforming components and the application of composite materials on the stress–strain characteristics of metallic and non-metallic materials. From a numerical simulation point of view, the focus will be on material models and the detection of input data on material properties (K, n, FLC, friction, etc.).
The benefit of this article lies in the proposal methodology for predicting the impact characteristics of body parts loaded with bending force based on the mechanical properties of materials determined by tensile testing.

Author Contributions

Conceptualization and methodology, E.E.; software, M.T. and S.N.; validation, E.E., M.T. and S.N.; formal analysis, E.E.; data curation, E.E.; writing—original draft preparation, E.E. and M.T.; writing—review and editing, E.E. and M.T.; project administration, E.E.; funding acquisition, E.E. All authors have read and agreed to the published version of the manuscript.

Funding

This research and the APC was funded by Vedecká grantová agentúra MŠVVaŠ SR a SAV, grant number VEGA 1-0238-23.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data available on request due to restrictions of funder.

Acknowledgments

The work was accomplished under the grant project VEGA 1-0238-23 “Implementation of CAx systems and virtual engineering techniques in the redesign of car-body parts for deformation zones”.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Baroutaji, A.; Sajjia, M.; Olabi, A.G. On the crashworthiness performance of thin-walled energy absorbers: Recent advances and future developments. Thin-Walled Struct. 2017, 118, 137–163. [Google Scholar] [CrossRef]
  2. Santos, K.; Silva, N.M.; Dias, J.P.; Amado, C. A methodology for crash investigation of motorcycle-cars collisions combining accident reconstruction, finite elements, and experimental tests. Eng. Fail. Anal. 2023, 152, 107505. [Google Scholar] [CrossRef]
  3. Yoon, T.; Kim, H.; Heo, C. An experiment and FE simulation for the development of a SPFC1180 AHSS one-body door impact beam about a car side impact collision. Int. J. Precis. Eng. Manuf. 2016, 17, 81–89. [Google Scholar] [CrossRef]
  4. Bouaziz, O.; Zurob, H.; Huang, M. Driving Force and Logic of Development of Advanced High Strength Steels for Automotive Applications. Steel Res. Int. 2013, 84, 937–947. [Google Scholar] [CrossRef]
  5. Kiran More, K.C.; Patil, G.M.; Belkhede, A.A. Design and analysis of side door intrusion beam for automotive safety. Thin–Walled Struct. 2020, 53, 106788. [Google Scholar] [CrossRef]
  6. Beer, F.; Johnston, R.; Dewolf, J.; Mazurek, D. Mechanics of Materials, 8th ed.; McGraw-Hill: New York, NY, USA, 2020; p. 823. ISBN 9781260569971. [Google Scholar]
  7. Schmitt, J.H.; Iung, T. New developments of advanced high-strength steels for automotive applications Nouveaux développements dans le domaine des aciers à très haute résistance pour les applications automobiles. Comptes Rendus Phys. 2018, 19, 641–656. [Google Scholar] [CrossRef]
  8. Xu, P.; Yang, C.; Peng, Y.; Yao, S.; Zhang, D.; Li, B. Crash performance and multi-objective optimization of a gradual energy absorbing structure for subway vehicles. Int. J. Mech. Sci. 2016, 107, 1–12. [Google Scholar] [CrossRef]
  9. Baroutaji, A.; Morris, E.; Olabi, A.G. Quasi-static response and multi-objective crashworthiness optimization of oblong tube under lateral loading. Thin-Walled Struct. 2014, 82, 262–267. [Google Scholar] [CrossRef]
  10. Burdzik, R.; Folęga, P.; Konieczny, L.; Łazarz, B.; Stanik, Z.; Warczek, J. Analysis of material deformation work measures in determination of a vehicle’s collision speed. Arch. Mater. Sci. Eng. 2012, 58, 13–21. [Google Scholar]
  11. Miklós, T. Development of Advanced High Strength Automotive Steels. Acta Mater. Transylvanica 2012, 4, 9–17. [Google Scholar]
  12. Eroğlu, M. Advanced High Strength Steels (AHSSs): Production and Applications. In Proceedings of the UDCS’19 Fourth International Iron and Steel Symposium, Karabuk University, Turey, Karabuk, 4–6 April 2019. [Google Scholar]
  13. Larn, R.H.; Yang, J.R. The effect of compressive deformation of austenite on the bainitic ferrite transformation in Fe-Mn-Si-C steels. Mater. Sci. Eng. 2000, 278, 278–291. [Google Scholar] [CrossRef]
  14. Basuki, A.; Aernoudt, E. Influence of rolling of TRIP steel in the intercritical region on the stability of retained austenite. J. Mater. Process. Technol. 1999, 89–90, 37–43. [Google Scholar] [CrossRef]
  15. Jacoques, P.J.; Ladriere, J.; Delannay, F. On the influence of Interactions between Phases on the Mechanical Stability of Retained Austenite in Transformation- Induced Plasticity Multiphase Steels. Metall. Mater. Trans. A 2001, 32, 2759–2768. [Google Scholar] [CrossRef]
  16. De, A.K.; Speer, J.G.; Matlock, D.K. Color tint-etching for Multiphase Steels. Adv. Mater. Process. 2003, 161, 27. [Google Scholar]
  17. Girault, E.; Jacques, P.; Hamlet, P.; Mols, K.; Van Humbeeck, J.; Aernoudt, E.; Delnnay, F. Metallographic Methods for Revealing the Multiphase Microstructure of TRIP-Assisted Steels. Mater. Charact. 1998, 40, 111–118. [Google Scholar] [CrossRef]
  18. Liu, Y.; Hu, H.; Zhao, L.; Long, H. Compression behavior of warp-knitted spacer fabrics for cushioning applications. Text. Res. J. 2015, 82/1, 11–20. [Google Scholar] [CrossRef]
  19. Kumar, A.; Vijay Gautam, V. Formability of tailor welded blanks of high strength steel: A review. Mater. Today 2021, 46, 6547–6551. [Google Scholar] [CrossRef]
  20. Čada, R.; Pektor, T. Design of Stamping Drawing Technology from Thin Sheet. In Proceedings of the 31st International Conference on Metallurgy and Materials METAL 2022, Brno, Czech Republic, 18–19 May 2022; Tanger: Ostrava, Czech Republic, 2022; pp. 262–267, ISBN 978-80-88365-06-8. [Google Scholar] [CrossRef]
  21. Li, B.; Tian, Y.; Tian, J.; Zuo, S.; Chen, S. Bending-damaged fracture criterion and numerical implementation of limestone beam under three-point bending test. Front. Earth Sci. 2023, 11, 1147129. [Google Scholar] [CrossRef]
  22. Cao, P.; Cao, L.; Chen, G.; Shi, F.; Zhou, C.; Wang, J. Experimental and Simulation Study of the Fracture Instability Behavior in Polypropylene Fiber-Reinforced Concrete. Materials 2023, 16, 4729. [Google Scholar] [CrossRef]
  23. Wójcik, M.; Skrzat, A. Identification of Chaboche–Lemaitre combined isotropic–kinematic hardening model parameters assisted by the fuzzy logic analysis. Acta Mechanica. 2021, 232, 685–708. [Google Scholar] [CrossRef]
  24. ISO 6892-1:2019; Metallic Materials—Tensile Testing—Part 1: Method of Test at Room Temperature. International Organization for Standardization: Geneva, Switzerland, 2019.
  25. ISO 10113:2020; Metallic Materials—Sheet and Strip—Determination of Plastic Strain Ratio. International Organization for Standardization: Geneva, Switzerland, 2020.
  26. ISO 10275:2020; Metallic Materials—Sheet and Strip—Determination of Tensile Strain Hardening Exponent. International Organization for Standardization: Geneva, Switzerland, 2020.
  27. Čada, R. Testing of Strain in Stampings by Embossed Grids. Tech. Gaz. 2003, 10, 9–13. [Google Scholar]
  28. Buršák, M.; Bacso, J. Skúšanie, Kontrola a Hodnotenie Materiálov, 1st ed.; Eminia: Košice, Slovakia, 2008. [Google Scholar]
  29. Lukoševicius, V.; Juodvalkis, D.; Kersys, A.; Makaras, R. Investigation of Functionality of Vehicle Crumple Zones Recovered after a Traffic Accident. Appl. Sci. 2023, 13, 1686. [Google Scholar] [CrossRef]
  30. Summers, S.M.; Prasad, A.K.; Hollowell, W.T. NHTSA’s Compatibility Research Program Update; SAE Technical Paper 2001-01-1167; SAE International: Warrendale, PA, USA, 2001. [Google Scholar] [CrossRef]
  31. Joudaki, J.; Sedighi, M. Effect of material’s behavior on residual stress distribution in elastic–plastic beam bending: An analytical solution. Proc. Inst. Mech. Eng. Part L J. Mater. Des. Appl. 2017, 231, 361–372. [Google Scholar] [CrossRef]
  32. De Angelis, F.; Cancellara, D.; Grassia, L.; Alberto, D. The influence of loading rates on hardening effects in elasto/viscoplastic strain-hardening materials. Mech Time-Depend Mater. 2018, 22, 533–551. [Google Scholar] [CrossRef]
  33. Mulidrán, P.; Spišák, E.; Tomáš, M.; Majerníková, J.; Varga, J. The Effect of Material Models In The Fem Simulation on The Springback Prediction of The Trip Steel. Acta Metall. Slovaca 2021, 27, 103–108. [Google Scholar] [CrossRef]
  34. Martínez-Palmeth, L.H.; Martínez-Donaire, A.J.; Vallellano, C. Formability limits of high-strength H240LA steel sheets under stress/strain gradients. Mech. Mater. 2019, 132, 47–56. [Google Scholar] [CrossRef]
  35. Basak, S.; Panda, S.K.; Zhou, Y.N. Formability Assessment of Prestrained Automotive Grade Steel Sheets Using Stress Based and Polar Effective Plastic Strain-Forming Limit Diagram. J. Eng. Mater. Technol. Trans. ASME 2015, 137, 041006. [Google Scholar] [CrossRef]
  36. Habibi, N.; Sundararaghavan, V.; Prahl, U.; Ramazani, A. Experimental and Numerical Investigations into the Failure Mechanisms of TRIP700 Steel Sheets. Metals 2018, 8, 1073. [Google Scholar] [CrossRef]
  37. Paul, S.K. Controlling factors of forming limit curve: A review. Adv. Ind. Manuf. Eng. 2021, 2, 100033. [Google Scholar] [CrossRef]
Figure 1. Scheme of the three-point bending test and compression test.
Figure 1. Scheme of the three-point bending test and compression test.
Materials 17 00210 g001
Figure 2. Microstructure of the investigated materials: (a) HSLA 240; (b) DP600; (c) TRIP 400/700.
Figure 2. Microstructure of the investigated materials: (a) HSLA 240; (b) DP600; (c) TRIP 400/700.
Materials 17 00210 g002
Figure 3. Dependence of engineering σE and true σT stress on true φT strain.
Figure 3. Dependence of engineering σE and true σT stress on true φT strain.
Materials 17 00210 g003
Figure 4. The sample before and after the tensile test when recorded by ARAMIS SRX optical system.
Figure 4. The sample before and after the tensile test when recorded by ARAMIS SRX optical system.
Materials 17 00210 g004
Figure 5. Deformation record by the ARAMIS system along the HSLA 240 specimen during the tensile test.
Figure 5. Deformation record by the ARAMIS system along the HSLA 240 specimen during the tensile test.
Materials 17 00210 g005
Figure 6. Deformation record by the ARAMIS system along the DP 600 specimen during the tensile test.
Figure 6. Deformation record by the ARAMIS system along the DP 600 specimen during the tensile test.
Materials 17 00210 g006
Figure 7. Deformation record by the ARAMIS system along the TRIP 400/700 specimen during the tensile test.
Figure 7. Deformation record by the ARAMIS system along the TRIP 400/700 specimen during the tensile test.
Materials 17 00210 g007
Figure 8. Principle scheme of the three-point bending test.
Figure 8. Principle scheme of the three-point bending test.
Materials 17 00210 g008
Figure 9. Record of the bending force on the punch path.
Figure 9. Record of the bending force on the punch path.
Materials 17 00210 g009
Figure 10. Record from the strain evaluation by the 3D optical system Argus—material HSLA 240. Note: The measured strain φ1 values are indicated by a thin solid black line and φ3 value are indicated by a thin solid blue line.
Figure 10. Record from the strain evaluation by the 3D optical system Argus—material HSLA 240. Note: The measured strain φ1 values are indicated by a thin solid black line and φ3 value are indicated by a thin solid blue line.
Materials 17 00210 g010
Figure 11. Record from the strain evaluation by the 3D optical system Argus—material DP 600. Note: The measured strain φ1 values are indicated by a thin solid black line and φ3 value are indicated by a thin solid blue line.
Figure 11. Record from the strain evaluation by the 3D optical system Argus—material DP 600. Note: The measured strain φ1 values are indicated by a thin solid black line and φ3 value are indicated by a thin solid blue line.
Materials 17 00210 g011
Figure 12. Record from the strain evaluation by the 3D optical system Argus—material TRIP 400/700. Note: The measured strain φ1 values are indicated by a thin solid black line and φ3 value are indicated by a thin solid blue line.
Figure 12. Record from the strain evaluation by the 3D optical system Argus—material TRIP 400/700. Note: The measured strain φ1 values are indicated by a thin solid black line and φ3 value are indicated by a thin solid blue line.
Materials 17 00210 g012
Figure 13. Three-point bending test with fixed ends: (a) 3-D model created in Creo 2.0; (b) simulation model created in Pam Stamp software.
Figure 13. Three-point bending test with fixed ends: (a) 3-D model created in Creo 2.0; (b) simulation model created in Pam Stamp software.
Materials 17 00210 g013
Figure 14. Comparison of simulation and experiment results for material HSLA 240.
Figure 14. Comparison of simulation and experiment results for material HSLA 240.
Materials 17 00210 g014
Figure 15. Comparison of simulation and experiment results for material DP 600.
Figure 15. Comparison of simulation and experiment results for material DP 600.
Materials 17 00210 g015
Figure 16. Comparison of simulation and experiment results for material TRIP 400/700.
Figure 16. Comparison of simulation and experiment results for material TRIP 400/700.
Materials 17 00210 g016
Figure 17. Dependence of the true stress on the true strain within the strain interval <0.002; φT,F>.
Figure 17. Dependence of the true stress on the true strain within the strain interval <0.002; φT,F>.
Materials 17 00210 g017
Figure 18. Dependence of the true stress on the true strain within the strain interval <0.05; φT,F>.
Figure 18. Dependence of the true stress on the true strain within the strain interval <0.05; φT,F>.
Materials 17 00210 g018
Figure 19. Dependence of the true stress on the true strain within the strain interval <0.1; φT,F>.
Figure 19. Dependence of the true stress on the true strain within the strain interval <0.1; φT,F>.
Materials 17 00210 g019
Figure 20. Comparison of the absorption energy of the investigated materials on the basis of applied models—tensile test.
Figure 20. Comparison of the absorption energy of the investigated materials on the basis of applied models—tensile test.
Materials 17 00210 g020
Figure 21. Dependence of the energy absorbing calculated from the regression model, according to Equation (6) models and according to Equation (1) on the energy absorbing according to Equation (2).
Figure 21. Dependence of the energy absorbing calculated from the regression model, according to Equation (6) models and according to Equation (1) on the energy absorbing according to Equation (2).
Materials 17 00210 g021
Figure 22. Comparison of the absorption energy of the investigated materials on the basis of applied models—three-point bending test.
Figure 22. Comparison of the absorption energy of the investigated materials on the basis of applied models—three-point bending test.
Materials 17 00210 g022
Figure 23. Comparison of the stiffness characteristics of the investigated materials based on the applied models—three-point bending test.
Figure 23. Comparison of the stiffness characteristics of the investigated materials based on the applied models—three-point bending test.
Materials 17 00210 g023
Figure 24. Determination of FLC0 by numerical simulation—material DP 600.
Figure 24. Determination of FLC0 by numerical simulation—material DP 600.
Materials 17 00210 g024
Figure 25. Dependence of φ1T on φ2T for material HSLA 240.
Figure 25. Dependence of φ1T on φ2T for material HSLA 240.
Materials 17 00210 g025
Figure 26. Dependence of φ1T on φ2T for material DP 600.
Figure 26. Dependence of φ1T on φ2T for material DP 600.
Materials 17 00210 g026
Figure 27. Dependence of φ1T on φ2T for material TRIP 400/700.
Figure 27. Dependence of φ1T on φ2T for material TRIP 400/700.
Materials 17 00210 g027
Table 1. Measured values of mechanical properties of the materials used.
Table 1. Measured values of mechanical properties of the materials used.
Mechanical Properties
ReRmAuA80K0.1–0.2n0.1–0.2EACTM.t/V0EACT.t/V0
Material(MPa)(MPa)(%)(%)(MPa)(-)(N.mm)
Acc. to Equation (2)
(N.mm)
Acc. to Equation (6)
HSLA 240388 ± 2449 ± 117 ± 129 ± 1719 ± 30.178 ± 0.002104.6104.7
DP 600376 ± 4632 ± 219 ± 0.528 ± 0.51059 ± 20.202 ± 0.001144.4151.1
TRIP 400/700440 ± 3764 ± 225 ± 130 ± 21468 ± 60.287 ± 0.002187.2198.7
Note: yield strength Re, tensile strength Rm, uniform ductility AU, total ductility A80, Kinterval—strength coefficient, ninterval—strain hardening exponent, 0.1–0.2—interval 0.1 within to 0.2 strain, strength coefficient K0.1–0.2, n0.1–0.2 strain hardening exponent, EACTM.t/V0 absorbed energy calculated by trapezoidal method according to Equation (2) from the tensile test record from the area under the σT to φT true stress–true strain curve, EACT.t/V0 absorbed energy calculated according to Equation (6) from the tensile test record from the area under the σT to φT true stress–true strain curve, V0 unit volume.
Table 2. The strain values measured with the Aramis optical system.
Table 2. The strain values measured with the Aramis optical system.
Strain ZoneStrainMaterialLankford’s Anisotropy
Parameter r (-)
HSLA 240DP 600TRIP 400/700
Safe point
SP (green)
Major φT1 0.3650.3700.310
Minor φT2 (-)−0.166−0.184−0.1430.840.980.86
Thickness φT3 (-)−0.299−0.187−0.167
Warning point
WP (yellow)
Major φT1 (-)0.5400.4040.380
Minor φT2 (-)−0.245−0.200−0.1800.830.980.86
Thickness φT3 (-)−0.295−0.204−0.210
Critical point
CP (red)
Major φT1 (-)0.6400.6350.490
Minor φT2 (-)−0.260−0.275−0.2100.680.860.75
Thickness φT3 (-)−0.380−0.360−0.280
Table 3. Measured values of n, K and energy absorbing at different intervals of true strain.
Table 3. Measured values of n, K and energy absorbing at different intervals of true strain.
MaterialK0.002–F
(MPa)
n0.002–F
(-)
K0.05–F
(MPa)
n0.05–F
(-)
K0.1–F
(MPa)
n0.1–F
(-)
EATRA0.002–F
(N.m)
EATRA0.05–F
(N.m)
EATRA0.1–F
(N.m)
HSLA 2407190.1787180.1777140.174104.5104.6104.6
DP 60010800.21310730.20910510.198144.4144.8145.5
TRIP 400/70013860.26014600.28614320.275187.5186.7187.5
Note: Kinteraval—strength coefficient; n.interval—strain hardening exponent; EATRAinterval—energy absorption evaluated by regression analysis; 0.002–F—interval within 0.002 strain to fracture strain; 0.05–F—interval within 0.05 strain to fracture strain; 0.1–F—interval within 0.1 strain to fracture strain.
Table 4. Measured and calculated values of the deformation characteristics by the three-point bending test.
Table 4. Measured and calculated values of the deformation characteristics by the three-point bending test.
Measured and Calculated Data
xplmaxEAb.CTMEAb.RAFbmaxCSb.EACSb.RACSb.RA/CSb.EA
Material(mm)(Nm)
Acc. to
Equation (9)
(Nm)
Acc. to
Equation (14)
(kN)(kN/m)
Acc. to
Equation (13)
(kN/m)(-)
HSLA 24038.5285.4279.013.89385.1377.40.980
DP 60035.5374.4308.217.76594.2489.10.823
TRIP 400/70040.0535.1589.824.45668.9612.20.915
Table 5. The strain values measured by the 3D optical system Argus.
Table 5. The strain values measured by the 3D optical system Argus.
Strain ZoneStrainMaterial
HSLA 240DP 600TRIP 400/700
Safe zone
SP
Major φT1 (-)0.26 ± 0.020.15 ± 0.020.16 ± 0.02
Minor φT2 (-)−0.12 ± 0.01−0.07 ± 0.01−0.08 ± 0.01
Thickness φT3 (-)−0.14 ± 0.01−0.08 ± 0.01−0.08 ± 0.01
Warning zone
WP
Major φT1 (-)0.33 ± 0.020.23 ± 0.010.21 ± 0.02
Minor φT2 (-)−0.16 ± 0.01−0.11 ± 0.01−0.12 ± 0.01
Thickness φT3 (-)−0.17 ± 0.01−0.12 ± 0.01−0.09 ± 0.01
Table 6. Deformation characteristics determined by simulation of the three-point bending test.
Table 6. Deformation characteristics determined by simulation of the three-point bending test.
MaterialMeasured and Calculated Data
xpl.sEAs.bEAs.b/EAb.CTMFs.bCSs.bCSs.b/CSb.EAxpl.s/xpl.b
(mm)(Nm)(-)(kN)(Nm)(-)(-)
HSLA 2400.002–F40222.610.7812.372950.7661.04
HSLA 2400.05–F4012.39
HSLA 2400.1–F4012.23
DP 6000.002–F47340.700.9117.62369.80.6221.32
DP 6000.05–F4517.641.30
DP 6000.1–F4617.081.27
TRIP 400/7000.002–F52470.890.8824.635090.7611.30
TRIP 400/7000.05–F5726.371.43
TRIP 400/7000.1–F5625.981.40
Table 7. Comparison of values FLC0WP.
Table 7. Comparison of values FLC0WP.
Data fromHSLA 240DP 600TRIP 400/700
Measured0.220.2050.2
Simulation0.210.2750.275
Acc.to [34,35,36]0.27 [34]0.18 [35]0.18 [36]
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

Evin, E.; Tomáš, M.; Németh, S. Numerical and Experimental Analysis of Stress–Strain Characteristics in DP 600 and TRIP 400/700 Steel Sheets. Materials 2024, 17, 210. https://doi.org/10.3390/ma17010210

AMA Style

Evin E, Tomáš M, Németh S. Numerical and Experimental Analysis of Stress–Strain Characteristics in DP 600 and TRIP 400/700 Steel Sheets. Materials. 2024; 17(1):210. https://doi.org/10.3390/ma17010210

Chicago/Turabian Style

Evin, Emil, Miroslav Tomáš, and Stanislav Németh. 2024. "Numerical and Experimental Analysis of Stress–Strain Characteristics in DP 600 and TRIP 400/700 Steel Sheets" Materials 17, no. 1: 210. https://doi.org/10.3390/ma17010210

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