You are on page 1of 44

Journal Pre-proof

Effect of tempering conditions on adiabatic shear banding during dynamic


compression and ballistic impact tests of ultra-high-strength armor steel

Min Cheol Jo, Selim Kim, Dong Woo Suh, Sung Suk Hong, Hong Kyu Kim, Seok Su
Sohn, Sunghak Lee
PII: S0921-5093(20)30891-1
DOI: https://doi.org/10.1016/j.msea.2020.139818
Reference: MSA 139818

To appear in: Materials Science & Engineering A

Received Date: 6 April 2020


Revised Date: 18 June 2020
Accepted Date: 22 June 2020

Please cite this article as: M.C. Jo, S. Kim, D.W. Suh, S.S. Hong, H.K. Kim, S.S. Sohn, S. Lee,
Effect of tempering conditions on adiabatic shear banding during dynamic compression and ballistic
impact tests of ultra-high-strength armor steel, Materials Science & Engineering A (2020), doi: https://
doi.org/10.1016/j.msea.2020.139818.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2020 Published by Elsevier B.V.


Author contributions

Min Cheol Jo: Conceptualization, Methodology, Formal analysis, Investigation, Data Curation,

Writing-Original Draft, Visualization

Selim Kim: Methodology, Data Curation, Validation

Dong Woo Suh: Conceptualization, Formal analysis

Sung Suk Hong: Resources, Funding acquisition

Hong Kyu Kim: Resources, Funding acquisition

Seok Su Sohn: Conceptualization, Visualization, Funding acquisition, Writing - Review & Editing,

Supervision

Sunghak Lee : Project administration, Writing - Review & Editing, Supervision


Effect of tempering conditions on adiabatic shear banding during dynamic compression

and ballistic impact tests of ultra-high-strength armor steel

Min Cheol Jo a, Selim Kim a, Dong Woo Suh b, Sung Suk Hong c, Hong Kyu Kim c, Seok

Su Sohn d,*, Sunghak Lee a,**


a
Center for Advanced Aerospace Materials

Pohang University of Science and Technology, Pohang 790-784, South Korea


b
Graduate Institute of Ferrous Technology

Pohang University of Science and Technology, Pohang 37673, South Korea


c
Materials Directorate

Agency for Defense Development, Daejeon 305-600, South Korea


d
Department of Materials Science and Engineering

Korea University, Seoul 02841, South Korea

Abstract

In this study, roles of strain-hardening rate on susceptibility of adiabatic shear band (ASB)

formation and subsequent cracking were investigated in two ultra-high-strength armor steel

plates heat-treated differently. The quenched and tempered steel contained ~2% of retained

austenite in the tempered martensitic matrix, while the quenched and austempered steel

contained ~4% retained austenite in the bainitic matrix partly with the tempered martensite.

The actual ballistic impact test results revealed the lower sensitivity of ASB formation in the

austempered steel than in the tempered steel, which corresponded well to the higher critical

strain for ASB formation in the dynamic compressive test using a laboratory-scale split

Hopkinson pressure bar (SHPB). The austempered steel caused the higher internal stress

among various constituents, and all the retained austenite transformed into martensite
-1-
during the deformation, thereby leading to transformation-induced plasticity (TRIP)

effect. The higher strain-hardening rate induced by these higher internal stress and TRIP

effect increased resistance to ASB formation, which was confirmed by a calculation of

ASB susceptibility. Thus, the austempered steel was much less susceptible to the ASB

formation during the ultra-high-speed deformation. Consequently, the increased resistance

to ASB formation retarded the initiation and propagation of ASBs and cracks.

Keywords: Adiabatic shear band (ASB); Tempering; Split Hopkinson pressure bar (SHPB);

Ballistic impact; Strain-hardening rate;

*Corresponding author

**Corresponding author

E-mail address: sssohn@korea.ac.kr (S.S. Sohn), shlee@postech.ac.kr (S. Lee)

-2-
1. Introduction

Excellent protective capability, maneuverability, and transportability are required in

armored vehicles and tanks for achieving the ensured survival of combat crew members [1,2].

Intensive efforts on designing and modifying armored materials and structures have been

made to obtain the excellent protective capability against ballistic impacts. When the armor

materials are ballistically impacted, adiabatic shear bands (ASBs) can be generated as the

heavily localized deformation occurs within a very narrow region before the sufficient

emission of thermal energy [2–6]. At a region ballistically penetrated by a projectile, a

number of ASBs often form by a thermomechanical instability induced from the heavy

plastic flow and abrupt temperature rise [7–13], thereby reducing a load-carrying capability

near the ASB [14–19] to initiate and propagate cracks.

To enhance the ballistic performance of armor materials as well as basic dynamic

properties, the investigation of ASB formation behavior during the high-strain-rate loading is

essential for preventing or minimizing ASBs. Ballistic impact test is the most direct method

for evaluating the performance, but is time-consuming and costly. In addition, the test has

limitations in ASB analyses because of difficulties of sample collection from fragmented or

penetrated target specimens. In this respect, several studies have been conducted on the

behavior of ASB formation in the ultra-high-speed deformation using a split Hopkinson

pressure bar (SHPB) [20–29]. However, very few attempts have been made to investigate the

ASB formation behavior in actual ballistic impacts in relation to the SHPB test results.

Studies on microstructural controls by varying tempering conditions have been actively

conducted in ultra-high-strength martensitic or bainitic steels. The tempering treatment alters

volume fraction, morphology, and chemistry of various microstructures and constituent

phases, which significantly affects the mechanical properties such as hardness, strength,

-3-
ductility, strain-hardening rate, and impact toughness [30–39]. Appropriate tempering

treatments can provide an important idea for preventing or minimizing the ASB formation

during the dynamic loading condition. Thus, dynamic deformation processes related to the

formation of ASBs and cracks, which vary with tempering treatments, should be clearly

understood, from which high-ballistic-performance armor materials can be successfully

developed.

In the present study, two ultra-high-strength armor steel plates heat-treated differently were

fabricated by controlling the tempering treatment, and their ASB formation and cracking

behavior was evaluated by the laboratory-scale SHPB. Susceptibility of ASB formation was

examined in detail by utilizing interrupted dynamic compressive tests to define a critical

strain for ASB formation and subsequent quantitative analyses of ballistic impacted samples.

The ASB formation processes were then related to basic microstructures and dynamic

compressive behavior so that they could be used for the enhancement of actual ballistic

performance by suppressing the ASB formation.

2. Experimental

An ultra-high-strength steel (nominal composition; 0.5C-0.7Mn-1.5Si-0.2V-0.6Cr-0.25Mo

(wt.%)) was fabricated by a vacuum induction melting for the armor application. A steel

ingot was homogenized at 1200 °C for 4 h and hot-rolled at 1100–900 °C to produce 12-mm-

thick plates. In order to obtain different constituent phases, one of the plates was austenitized

at 900 °C for 1/2 h and oil-quenched to room temperature. And then, the plates were

tempered at 200 °C for 4 h. Another one was austenitized at 900 °C for 1/2 h followed by the

austempering in a salt bath at 250 °C for 1/2 h. For convenience, the former and latter plates

are referred to as “QT” and “AT”, respectively.

-4-
1%-nital-etched microstructures of longitudinal-short-transverse(L-S) plane were observed

by using an optical or a scanning electron microscope (SEM). For the transmission electron

microscope (TEM) observation, the specimens were mechanically polished to a thickness of

50 μm, punched to prepare disk specimens (diameter; 3 mm) by a disk cutter, and then

electro-polished in a 90% CH3COOH +10% HClO4 solution by a twin-jet polisher (Tenupol-5,

Struers, Denmark) to make thin foils. They were observed by a TEM (JEM-2100F, Jeol,

Japan) at 200 kV. Main microstructures of tempered martensite, bainite, and retained

austenite were examined by using an X-ray diffraction (XRD, Cu Kα radiation) analysis, and

their volume fractions were measured by a direct comparison method [40]. EBSD analysis

(0.06 μm step size) was performed on electro-polished steel specimens. In order to observe a

detailed ASB microstructure, thin foils were prepared by a focused ion beam (FIB, Helios

600, FEI Company, USA), and were observed by the TEM at 200 kV. Vickers hardness test

was conducted under a load of 500 g. Plate-type tensile specimens (width 5 mm; thickness

1.4 mm; gage length 12.5 mm; longitudinal direction) were tested at room temperature at a

strain rate of 10-3 s-1 by a 100 kN-capacity universal testing machine (8801, INSTRON, USA).

Quasi-static compressive tests were conducted on cylindrical specimens (5 mm diameter; 4

mm height) in the temperature range of 25–500 °C at a strain rate of 2x10-3 s-1 by a 600 kN-

capacity universal testing machine (5589, INSTRON, USA).

The SHPB was used for dynamic compressive tests [24]. The specimen (5 mm diameter; 4

mm height) situated between incident and transmission bars (19 mm diameter; 1430 mm

length) was compressed by a striker bar (19 mm diameter) firing an air pressure of 0.2 or 0.3

MPa. Velocities of the striker bar measured by using an ultra-high-speed camera were 19.9

and 24.3 m/s for the air pressures of 0.2 and 0.3 MPa, respectively, from which the impact

momentums were measured to be 12.4 and 15.2 kg∙m∙s-1, respectively. Incident, reflective,

and transmitted waves were detected at strain gages during the dynamic compression, and
-5-
were recorded at an oscilloscope. All tests were carried out three times to ensure the data

reliability.

Ballistic impact tests were conducted in accordance with the MIL-DTL-46100E standard

[41]. Target plate specimens (300×400×12 mm size) were collected from the QT and AT

steel plates, and were ballistically tested at a 30-degree oblique by using a 0.50-caliber APM2

projectile. After the ballistic test at a velocity of about 840 m/s, penetrated target regions

were half-sectioned, polished, and 1%-nital-etched to observe ballistically-impacted

microstructures.

3. Results

3.1. Microstructures

Optical and SEM micrographs of the QT and AT steels are shown in Fig. 1a–d. The QT and

AT steels consisted mainly of tempered martensite and bainite, respectively, which cannot be

clearly differentiated in these micrographs. Band structures were weakly developed along the

rolling direction (Fig. 1a,b). The band forms by the segregation of substitutional alloying

elements like C, Mn, Ni, and Mo during the casting of conventional high-alloyed steels, and

is hardly removed by the sufficient homogenization [42,43]. Martensite and bainite start

temperatures (Ms and Bs, respectively) were calculated by the equations of Andrews et al.

[44] and Steven et al. [45], respectively, and the data are 297 and 569 °C, respectively.

Considering the calculated Ms temperature is 297 °C, the present austempering is expected to

lead to a small fraction of tempered martensite within a lower bainitic matrix and the

increased fraction of retained austenite compared to the QT steel.

Fig. 2a–d shows TEM bright field (BF) and dark field (DF) images with selected area

diffraction (SAD) patterns of the QT and AT steels. The QT steel consisted of the tempered
-6-
martensitic matrix along with a very small amount of austenite (~0.1 μm in size) (Fig. 2a),

which was identified by the SAD pattern and DF image (Fig. 2b). In the AT steel, the

austenite of ~0.5 μm in size existed inside the bainitic matrix partly with the tempered

martensite (Fig. 2c,d). The microstructures obtained from the isothermal treatments below Ms

are mainly composed of bainite, tempered martensite, and retained austenite, although the

characterization of tempered martensite and bainite is quite difficult because of their

morphological similarities in the TEM analyses [46–48]. Several studies [48–50] have

characterized the two microstructures via SEM, EBSD, and TEM, and explained that the

isothermally-treated tempered martensite shows wavy boundary characteristics, unlike the

bainite. However, the characteristics depends on the chemical composition and heat treatment

condition, according to Kim et al. [50], the nature of those isothermal laths is neither purely

martensitic nor purely bainitic.

EBSD inverse pole figure (IPF) and phase maps of the QT and AT steels are shown in Fig.

3a–d. The QT and AT steels consisted mainly of lath-type tempered martensite and bainite,

respectively. The austenite was observed at lath boundaries, and its fraction (Vγ) was

measured to be 0.8 and 2.7% in the QT and AT steels, respectively. The XRD analyses were

performed to accurately measure the Vγ in the wide area as well as to confirm the EBSD data.

In Fig. 3e, peaks of ferrite and austenite were observed in both steels. The measured Vγ was

1.9% and 3.9% in the QT and AT steels, respectively, which corresponded to the EBSD data,

although the Vγ was slightly underestimated for the EBSD analyses.

3.2. Hardness and tensile properties

The room-temperature hardness and tensile test data are shown in Table 1. The hardness of

the QT steel was higher by 20 HV than that of the AT steel, which reflects the harder

tempered martensite than the bainite. Fig. 4a shows engineering tensile stress-strain curves of

the QT and AT steels. The yield and tensile strengths of the QT steel were higher than those
-7-
of the AT steel, which corresponded to the hardness data, while the elongation was somewhat

lower in the QT steel. The strain-hardening rate was higher in the AT steel than in the QT

steel. The yield ratio (yield strength/tensile strength) of the AT steel was lower than that of

the QT steel (0.58 vs. 0.78), which reflects the higher strain-hardening rate in the AT steel.

3.3. Quasi-static and dynamic compressive properties

Fig. 4b shows quasi-static true compressive stress-strain curves, and the measured

compressive properties are listed in Table 2. The QT steel exhibited the higher yield strength

than the AT steel, which corresponded to the quasi-static tensile data (Table 1), while the

maximum compressive strength was similar in both steels. Fig. 4c,d shows dynamic true

compressive stress-strain and strain-hardening-rate (dσ/dε) curves obtained from the dynamic

compressive test under an impact momentum of 12.4 or 15.2 kg∙m∙s-1 (air pressure; 0.2 or 0.3

MPa, respectively). Under the impact momentums of 12.4 and 15.2 kg∙m∙s-1, compressive

strain rates were 1780 and 3030 s-1 in the QT steel, respectively, and 1870 and 3100 s-1 in the

AT steel, respectively. The 23% increase in impact momentum resulted in the 66–70%

increase in strain rate. The amount of impact momentum does not continuously maintain

during the dynamic compression, but drastically reduces. Thus, the strain rate does not

increase in proportion to the initial impact momentum. This is comparable to the results by

Odeshi et al. [28] exhibiting about 54% increase of strain rate even in ~20% increase of

impact momentum in a martensitic steel. Since both steel specimens were not fractured under

12.4 kg∙m∙s-1, the total strain implies the strain induced from one hit of striker bar. When the

dynamic compressive test was conducted under 15.2 kg∙m∙s-1, both steel specimens were

completely separated, and the resultant stress-strain curves and dynamic compressive

properties are shown in Fig. 4c and Table 2, respectively. The AT steel exhibited a lower

yield strength than the QT steel, but a similar maximum compressive strength level to the QT

steel, like the quasi-static compressive-property trend (Fig. 4b). The AT steel exhibited the
-8-
higher strain-hardening rate than the QT steel under both impact momentum conditions (Fig.

4d).

Low-magnification photographs of the dynamically compressed cylindrical specimens are

shown in Fig. 5a,b. Under 12.4 kg∙m∙s-1, the specimens were deformed until the strain of

0.17–0.18 without the cracking in both steels. When the impact momentum increased to 15.2

kg∙m∙s-1, the deformation initiated from the specimen surface developed into cracks in a shear

mode along the about 45-degree direction to reach the complete failure.

3.4. Dynamic compressive deformation mechanisms

Fig. 6a,b shows optical micrographs of the half-sectioned area of the dynamically

compressed specimens under 12.4 kg∙m∙s-1. All the specimens were not fractured, but shear

localized zones having narrow white-etched characteristics, i.e., adiabatic shear bands (ASBs),

formed at specimen edge areas of the QT steel (Fig. 6a), whereas they were not found in the

AT steel (Fig. 6b).

In order to define the critical strain for starting the ASB formation in the QT and AT steels,

the specimens were dynamically compressed under 15.2 kg∙m∙s-1 after a stopper ring was

inserted between incident and transmitter bars to obtain sequential dynamic compressive

strains at a 2.5% interval. ASBs did not form in both steels until the strain of 17.5%, and

started to form at the strain of 20% (true strain of 0.22) in the QT steel, while they started to

appear at 22.5% (true strain of 0.26) in the AT steel. Fig. 6c,d shows optical micrographs of

the specimens compressed at 20% and 22.5% in the QT and AT steels, respectively. ASBs

formed at specimen edges of both steels. Thus, the critical strain for ASB formation was 20%

and 22.5% in the QT and AT steels, respectively.

3.5. Martensitic transformation during dynamic compression

The XRD analyses were performed on the dynamically compressed specimens at the true
-9-
strains of 0, 0.11, and 0.16 for the QT and AT steels. Fig. 7a shows the volume fraction of

martensitic transformation as a function of compressive true strain. In the QT steel (initial Vγ;

1.9%, Fig. 3e), only a little austenite (0.4%) transformed at the strain of 0.16. In the AT steel,

on the other hand, most of austenite (initial Vγ; 3.9%, Fig. 3e) transformed to martensite

(3.0%) at 0.11, and all the austenite transformed at 0.16. This indicates that the martensitic

transformation activates in the dynamically compressed AT steel specimen. In order to

confirm the transformation occurred in the AT steel, the EBSD analyses were carried out, and

the phase maps are shown in Fig. 7b–d. At the strain of 0.11, most of austenite transformed to

the martensite (Fig. 7c), which well corresponded to the XRD data (Fig. 7a).

3.6. Observation of ballistically-impacted specimens

Fig. 8a,b shows montages of nital-etched high-magnification optical micrographs of

the half-sectioned area near the penetrated region of the ballistically-impacted target

specimen for the QT and AT steels. In the QT steel, the penetrated surface was covered

with ASBs (yellow-arrow marks in Fig. 8a), which indicates that the main crack

propagates along these ASBs formed on the penetrated surface. A number of ASBs also

formed in a narrow band shape, and most of them were found within the depth of 200 μm

from the penetrated surface, while some ASBs were branched off into several ASBs. A

few cracks propagated along these embrittled ASBs, as indicated by blue arrows, and

were opened by high shear stresses [51,52]. In the AT steel, ASBs formed mainly near the

penetrated surface, but their length and width were shorter than those of the QT steel

while surface cracks were not found (Fig. 8b). From the above optical micrographs of the

ballistically-impacted specimens, the average length (per unit area of 600 µm distant from the

penetrated surface) and width of ASBs were counted, and the results are presented in Table 3.

The average length and width of ASBs of the QT steel were larger than those of the AT steel.

- 10 -
3.7. TEM observation of ASB

In order to clearly reveal the microstructures of ASBs after the dynamic compression

and ballistic impact tests, TEM analyses were conducted on the ASB of the QT and AT

steels, and the results are shown in Fig. 9a–d and 10a–d. Fig. 9a–d shows TEM BF

images with SAD patterns of the interior and exterior areas of ASB after the dynamic

compression under the impact momentum of 15.2 kg∙m∙s-1. Very fine equiaxed grains

were observed inside the ASB as shown in Fig. 9a,c. The grain size was similar at about

75 nm in both steels. The SAD patterns displayed ring-like patterns (insets of Fig. 9a,c),

which indicates that multiple crystalline orientations of very fine grains exist within such

a narrow ASB area. In the exterior areas near the ASBs (Fig. 9b,d), grains were

elongated to a width range of 65–135 nm along the shear direction in both QT and AT

steels. According to Yuan et al. [53], shear strains are more experienced at the central area of

ASB than at its surroundings, thereby resulting in the different microstructural evolution in

the interior and exterior of ASB.

Fig. 10a–d shows TEM BF images with SAD patterns of the interior and exterior areas

of ASB after the ballistic impact test. Overall microstructures were similar to those of the

dynamic compression. Very fine equiaxed grains of ~140 nm in average size were

observed in the interior of ASBs (Fig. 10a,c). This average grain size was about twice

larger than that of the dynamic compression. The equiaxed grains exhibited a body-

centered-cubic (BCC) crystal structure, as shown in the inset of Fig. 10a,c. In the exterior

areas near the ASBs, elongated grains having a width of 50–125 nm along the shear

direction were found in both QT and AT steels (Fig. 10b,d).

Several previous studies [54,55] have shown that the fine equiaxed grains within the

ASB form by a rotational dynamic recrystallization (RDR) mechanism. According to the

RDR mechanism, initial grains are elongated along the shear direction, and are
- 11 -
subsequently split into subgrains. As the deformation proceeds, fine equiaxed subgrains

form by the sub-boundary rotation of 30°. In order for the sub-boundary rotation of 30°

within a short time, a high temperature rise above half of the melting temperature is required,

adiabatic heating occurring during high-speed deformation such as dynamic compression and

ballistic impact makes it possible [54].

4. Discussion

Various cracking and penetrating phenomena occurring during the ballistic impact of ultra-

high-strength armor steels are closely related to the ASB formation. Although efforts for

suppressing ASBs have been constantly made, most of armor steel plates are still vulnerable

to the ASB formation and subsequent cracking. This cracking leads to a serious deterioration

of ballistic performance. This study, thus, investigated the effect of strain-hardening

capability on the ASB formation to restrain the cracking phenomena. Detailed ASB

formation phenomena were quantitatively evaluated by the laboratory-scale SHPB, and the

susceptibility of ASB formation was then related to the ballistic behavior by examining ASB

characteristics of ballistically-impacted specimens.

4.1. Higher strain-hardening behavior of AT steel compared to QT steel

As shown in Fig. 4b–d, the AT steel exhibited the higher strain-hardening rate in both

quasi-static and dynamic compressive tests than the QT steel. This difference is greatly

dependent upon the configuration of constituent phases, resulting in variations of

compressive internal stress in the matrix according to mechanical properties of each

phase. During the compressive deformation, the strain is concentrated at a relatively soft

phase, and an internal stress due to the strain partitioning between those phases

contributes to the strain hardening [56–60]. The AT steel consisted of 3.9% of retained

- 12 -
austenite with bainite and tempered martensite, while the QT steel was composed of 1.9%

of retained austenite in the tempered martensite matrix (Fig. 3e). Thus, the internal stress

caused by various kinds of phases in the AT steel affects significantly the strain-

hardening rate. According to Sugimoto et al. [56], the retained austenite in the bainitic

matrix leads to a long-range compressive internal stress, this contributes to a high strain

hardening both in an early stage of strain and at high strains.

As well as the role of retained austenite in the strain partitioning, the deformation-

induced transformation enables the austenite to enhance the strain-hardening rate. The

QT steel contained a little initial retained austenite (1.9%, Fig. 3e), and the martensitic

transformation hardly occurred even in the dynamic compressive deformation. On the

other hand, the AT contained 3.9 vol.% of retained austenite (Fig. 3e), and all the

retained austenite transformed into the martensite during the dynamic deformation (Fig.

7a–d). According to the previous research results of Garcia-Mateo et al. [61], the

deformation-induced transformation of retained austenite in bainitic steels leads to

transformation-induced plasticity (TRIP), which significantly increases the strain-

hardening rate and thus favorably contributes to tensile properties [35,62–64]. This

increase of strain hardening is attributed to a volume expansion of the transforming

region, which results in additional plastic deformation of the surrounding area [65].

Therefore, the AT steel exhibited the higher strain-hardening rate than the QT steel due

to the internal stress occurring during the deformation among various constituents and

due to the TRIP effect by retained austenite.

4.2. Comparison of susceptibility of ASB formation between QT and AT steels

According to the observations of the dynamically-compressed or ballistically-impacted

specimens (Fig. 6a–d and 8a,b), ASBs play a critical role as cracking and failure paths.

These ASBs essentially result from the thermomechanical instability during the dynamic
- 13 -
compression or ballistic impact [7–13]. During the high-strain-rate deformation, the

temperature rises significantly due to little time for the heat to dissipate, which results in the

severely localized deformation [3–5]. It is well accepted in researches on ASB formation

[66–71] that materials having high density, specific heat capacity, strain-hardening rate, and

strain-rate sensitivity, along with low thermal softening parameter, have the lower

susceptibility of ASB formation. Xue et al. [70] reported that a 304 stainless steel exhibiting

the much lower work-hardening rate than the other steels displays the significantly higher

propensity for shear localization because the work hardening plays a key role in the initiation

of shear localization. In this work, thus, an attempt was made to quantitatively calculate the

ASB susceptibility of the QT and AT steels by using the following analytical equation.

According to Wright et al. [66,67], the ASB susceptibility (χSB) is estimated by:

/
= min 1, (1)

where m, n, and α are the strain-rate sensitivity, strain-hardening exponent, and non-

), respectively. The higher


/
dimensional thermal softening parameter (defined as α =

the χSB value, the greater the ASB susceptibility. The density and specific heat capacity

measured by using an electronic densimeter and a plastic-differential scanning calorimetry,

respectively, are 7.644 g/cm3 and 0.447 J/g∙K, respectively, in the QT steel, and 7.789 g/cm3

and 0.443 J/g∙K, respectively, in the AT steel. The strain-rate sensitivity (m) was calculated by

the following formula:

= (2)

where ! and " are the true stress and strain rate, respectively. The m values were calculated

by using flow stress values measured at the true strain of 0.1 (at strain rates of 2x10-3 s-1 and

- 14 -
3030–3100 s-1), and the results are 0.0044 and 0.0045 in the QT and AT steels, respectively.

The n values were calculated from the quasi-static true compressive stress-strain curves (Fig.

11a). The AT steel had the about twice higher n value than the QT steel (0.070 vs. 0.038).

The thermal softening parameters were calculated from quasi-static compressive stress-

strain curves at elevated temperatures. Fig. 11b shows quasi-static true compressive stress-

strain curves of the QT and AT steels in the temperature range of 200–500 °C. Overall stress

level was similar for each temperature in the QT and AT steels. Fig. 11c shows variations of

true flow stress at the strain of 0.1 as a function of temperature. The curve slopes, i.e, ∂σ/ ∂T,

were -5.74 and -5.37 MPa/K in the QT and AT steels, respectively, from which the α values

were calculated to be 1.68 and 1.56, respectively. Consequently, the estimated ASB

susceptibility (χSB) values were calculated to be 32.98 and 17.74 in the QT and AT steels,

respectively. This χSB result indicates that the AT steel is much less susceptible to the ASB

formation than the QT steel. The lower χSB of the AT steel is attributed to the higher n value

because the m and α values of the QT and AT steels are not varied much. This implies that the

χSB decreases in the AT steel by the effect of the higher strain hardening, from which the

results of the less heavily formed ASBs in the AT steel than in the QT steel during the

dynamic compression and ballistic impact (Fig. 6a–d and 8a,b) are plausibly interpreted.

4.3. Role of strain-hardening rate on dynamic compressive and ballistic impact properties

Since the ASB formation deteriorates the resistance to ballistic impact, it should be

prevented or minimized for improving the ballistic performance. In this study, ballistic

impact tests of the QT and AT steels were conducted with the ballistic impact speed of

~840 m/s, and the generated ASBs were quantitatively analyzed near the penetrated

region (Fig. 8a,b and Table 3). In the AT steel, the overall length and width of ASBs

were smaller than those of the QT steel. This result was in good agreement with the

- 15 -
dynamic compression results of Fig. 6a–d. In the case of the ballistic impact test, the

strain rate is approximately 105–3×105 [72,73], which is two orders of magnitude greater

than the dynamic compression strain rate (1780–3100 s-1). However, the ASBs formed in

both ballistic impact and dynamic compression tests, provided crack initiation and

propagation sites. In particular, the results of dynamic compression test showed that the

ASB does not form in the AT steel at impact momentum 12.4 kg∙m∙s-1, whereas it

formed in the QT steel. In addition, at impact momentum 15.2 kg∙m∙s-1, the critical strain

for ASB formation was higher in the AT steel than in the QT steel (22.5 vs. 20%). These

ASB analyses for the dynamic compression and ballistic impact indicate that the

resistance to the ASB formation in the AT steel is higher than that of the QT steel during

the ultra-high-speed deformation, which is attributed to the higher strain-hardening rate

of the AT steel. Since this higher strain-hardening rate reduced the ASB susceptibility

(χSB), the AT steel was much less susceptible to the ASB formation than the QT steel during

the ultra-high-speed deformation. Consequently, the increased resistance to ASB

formation retards the crack initiation and propagation of ASBs and cracks, which can

enhance the ballistic performance.

The susceptibility of ASB formation was analyzed in this study by the microstructural

evolution and dynamic compressive characteristics in the ultra-high-strength armor steels,

and was correlated well with the ballistic impact data. Since dynamic compressive

characteristics as well as the ASB formation provide the excellent data in improving the

ballistic performance, further studies on ultra-high-strength steel design and microstructural

and process parameters should be proceeded for preventing or minimizing the ASB formation

and cracking. Furthermore, the strain hardening occurring during the dynamic compression

should be positively utilized because it beneficially influences the suppression of ASBs. The

present laboratory-scale SHPB effectively evaluates the ASB formation behavior, and
- 16 -
provides an important idea to predict the susceptibility of ASB formation during the ballistic

impact.

5. Conclusions

In the present study, two ultra-high-strength armor steel plates heat-treated differently were

fabricated by controlling the tempering condition, and their dynamic deformation and ASB

formation behavior were evaluated. The results are summarized as follows.

1) The AT steel consisted of 3.9% of retained austenite with bainite and tempered

martensite, while the QT steel contained 1.9% of retained austenite in the tempered

martensite matrix. The calculated martensite and bainite start temperatures were 297

and 569 °C, respectively. Thus, the AT steel consisted partly of tempered martensite in the

bainitic matrix because the austempering was carried out at the lower temperature

(250 °C) than the calculated Ms (297 °C).

2) The tensile yield and tensile strengths of the QT steel were higher than those of the AT

steel, while the elongation was somewhat lower in the QT steel. The yield ratio of the AT

steel was 0.58, which was lower than that of the QT steel (0.78). The quasi-static and

dynamic compression results showed that the QT steel exhibited the higher yield strength

than the AT steel, which corresponded to the quasi-static tensile data, while the maximum

compressive strength was similar in both steels. In particular, the AT steel exhibited the

higher strain-hardening rate than the QT steel under both quasi-static and dynamic

deformation.

3) The higher strain-hardening rate in the AT steel was attributed to the configuration of

constituent phases, resulting in variations of compressive internal stress in the matrix

according to the mechanical properties of each phase. As well as the internal stress,
- 17 -
the TRIP effect due to deformation-induced transformation of retained austenite

played a significant role in increasing the strain-hardening rate.

4) The higher strain-hardening rate in the AT steel reduced the ASB susceptibility (χSB).

The χSB was quantitatively estimated to be 32.98 and 17.74 in the QT and AT steels,

respectively. This χSB result indicate that the AT steel is much less sensitive to the ASB

formation than the QT steel.

5) The dynamic compression test results showed that the ASB did not form in the AT

steel at the impact momentum 12.4 kg∙m∙s-1, whereas it formed only in the QT steel.

In addition, at the impact momentum 15.2 kg∙m∙s-1, the critical strain for ASB

formation was higher in the AT steel than in the QT steel (22.5 vs. 20%). The

ballistic impact test results showed that the overall length and width of ASBs in the

AT steel were smaller than those of the QT steel, which corresponded well to the

dynamic compression results. Therefore, the high strain-hardening rate of the AT

steel led to the higher resistance to the ASB formation than that of the QT steel.

Acknowledgments

This work was supported by the Korea University Grant for S.S. Sohn, by Agency for

Defense Development (grant No.; UE161030GD), by Korea Institute for Advancement of

Technology (KIAT) grant funded by the Korea Government (MOTIE) (P0002019, The

Competency Development Program for Industry Specialist), and by the Brain Korea 21

PLUS Project for Center for Creative Industrial Materials.

Data Availability Statement

- 18 -
The data that support the findings of this study are available from the corresponding author

on request.

References

[1] T. Demir, M. Übeyli, R.O. Yıldırım, Investigation on the ballistic impact behavior of

various alloys against 7.62 mm armor piercing projectile, Mater. Des. 29 (2008) 2009–

2016.

[2] H. Lee, J.H. Choi, M.C. Jo, I. Jo, S.-K. Lee, S. Lee, Effects of strain rate on compressive

properties in bimodal 7075 Al–SiCP composite, Met. Mater. Int. 24 (2018) 894–903.

[3] B. Mishra, P.K. Jena, B. Ramakrishna, V. Madhu, T.B. Bhat, N.K. Gupta, Effect of

tempering temperature, plate thickness and presence of holes on ballistic impact behavior

and ASB formation of a high strength steel, Int. J. Impact Eng. 44 (2012) 17–28.

[4] S.H. Atapek, S. Karagoz, Ballistic impact behavior of a tempered bainitic steel against

7.62 mm armour piercing projectile, Def. Sci. J. 61 (2011) 81–87.

[5] K. Sun, X. Yu, C Tan, H. Ma, F. Wang, H. Cai, Effect of microstructure on adiabatic

shear band bifurcation in Ti–6Al–4V alloys under ballistic impact, Mater. Sci. Eng. A 595

(2014) 247–256.

[6] A.G. Odeshi, S. Al-ameeri, M.N. Bassim, Effect of high strain rate on plastic deformation

of a low alloy steel subjected to ballistic impact, J. Mater. Process. Technol. 162–163

(2005) 385–391.

[7] F. Martinez, L.E. Murr, A. Ramirez, M.I. Lopez, S.M. Gaytan, Dynamic deformation and

adiabatic shear microstructures associated with ballistic plug formation and fracture in

Ti–6Al–4V targets, Mater. Sci. Eng. A 454–455 (2007) 581–589.

- 19 -
[8] C. Zheng, F. Wang, X. Cheng, K. Fu, J. Liu, Y. Wang, T. Liu, Z. Zhu, Effect of

microstructures on ballistic impact property of Ti-6Al-4V targets, Mater. Sci. Eng. A 608

(2014) 53–62.

[9] P.K. Jena, B. Mishra, K.S. Kumar, T.B. Bhat, An experimental study on the ballistic

impact behavior of some metallic armour materials against 7.62 mm deformable

projectile, Mater. Des. 31 (2010) 3308–3316.

[10]L.E. Murr, A.C. Ramirez, S.M. Gaytan, M.I. Lopez, E.Y. Martinez, D.H. Hernandez, E.

Martinez, Microstructure evolution associated with adiabatic shear bands and shear band

failure in ballistic plug formation in Ti–6Al–4V targets, Mater. Sci. Eng. A 516 (2009)

205–216.

[11]K.-M. Cho, S. Lee, S.R. Nutt, J. Duffy, Adiabatic shear band formation during dynamic

torsional deformation of an HY-100 steel, Acta Metall. Mater. 41 (1993) 923–932.

[12]A. Mishra, M. Martin, N.N. Thadhani, B.K. Kad, E.A. Kenik, M.A. Meyers, High-strain-

rate response of ultra-fine-grained copper, Acta Mater. 56 (2008) 2770–2783.

[13]A. Marchand, J. Duffy, An experimental study of the formation process of adiabatic shear

bands in a structural steel, J. Mech. Phys. Solids 36 (1988) 251–283.

[14]Y. Guo, Y. Li, A Novel approach to testing the dynamic shear response of Ti-6Al-4V,

Acta Mech. Solida Sin. 25 (2012) 299–311.

[15]M.N. Bassim, A.G. Odeshi, Shear strain localisation and fracture in high strength

structural materials, Arch. Mater. Sic. Eng. 31 (2008) 69–74.

[16]S.E. Schoenfeld, T.W. Wright, A failure criterion based on material instability, Int. J.

Solids. Struct. 40 (2003) 3021–3037.

[17]A.G. Odeshi, S. Al-ameeri, S. Mirfakhraei, F. Yazdani, M.N. Bassim, Deformation and

failure mechanism in AISI 4340 steel under ballistic impact. Theor. Appl. Fract. Mech. 45

(2006) 18–24.

- 20 -
[18]J. Peirs, P. Verleysen, J. Degrieck, F. Coghe, The use of hat-shaped specimens to study

the high strain rate shear behaviour of Ti–6Al–4V, Int. J. Impact Eng. 37 (2010) 703–714.

[19]B.B. Singh, G. Sukumar, A. Bhattacharjee, K.S. Kumar, T.B. Bhat, A.K. Gogia, Effect of

heat treatment on ballistic impact behavior of Ti–6Al–4V against 7.62 mm deformable

projectile, Mater. Des. 36 (2012) 640–649.

[20]W.H. Yin, F. Xu, O. Ertorer, Z. Pan, X.Y. Zhang, L.J. Kecskes, E.J. Lavernia, Q. Wei,

Mechanical behavior of microstructure engineered multi-length-scale titanium over a

wide range of strain rates, Mater. Sci. Eng. A 61 (2013) 3781–3798.

[21]A. Azimi, G.M. Owolabi, H. Fallahdoost, N. Kumar, G. Warner, Dynamic failure

investigation in ultrafine grained AA2219: mechanical and microstructural analysis, Met.

Mater. Int. 25 (2019) 900–911.

[22]L. Qiang, X. Yongbo, M.N. Bassim, Dynamic mechanical properties in relation to

adiabatic shear band formation in titanium alloy-Ti17, Mater. Sci. Eng. A 358 (2003)

128–133.

[23]X. Liu, C. Tan, J. Zhang, F. Wang, H. Cai, Correlation of adiabatic shearing behavior

with fracture in Ti-6Al-4V alloys with different microstructures, Int. J. Impact Eng. 36

(2009) 1143–1149.

[24]M.C. Jo, S. Kim, D.W. Suh, S.S. Hong, H.K. Kim, S.S. Sohn, S. Lee, Role of retained

austenite on adiabatic shear band formation during high strain rate loading in high-

strength bainitic steels, Mater. Sci. Eng. A 778 (2020) 139118.

[25]Z. Tianfeng, W. Junjie, C. Jiangtao, W. Ying, W. Xibin, Dynamic shear characteristics of

titanium alloy Ti-6Al-4V at large strain rates by the split Hopkinson pressure bar test, Int.

J. Impact Eng. 109 (2017) 167–177.

- 21 -
[26]C. Zheng, F. Wang, X. Cheng, J. Liu, T. Liu, Z. Zhu, K. Yang, M. Peng, D. Jin, Capturing

of the propagating processes of adiabatic shear band in Ti–6Al–4V alloys under dynamic

compression, Mater. Sci. Eng. A 658 (2016) 60–67.

[27]C.K. Syn, D.R. Lesuer, O.D. Sherby, Microstructure in adiabatic shear bands in a

pearlitic ultrahigh carbon steel, Mater. Sci. Technol. 21 (2005) 317–324.

[28]A.G. Odeshi, M.N. Bassim, Evolution of adiabatic shear bands in a dual-phase steel at

very high strain rates, Mater. Sci. Eng. A 488 (2008) 235–240.

[29]A.G. Odeshi, M.N. Bassim, S. Al-ameeri, Effect of heat treatment on adiabatic shear

bands in a high-strength low alloy steel, Mater. Sci. Eng. A 419 (2006) 69–75.

[30]H.S. Hasan, M.J. Peet, M-N. Avettand-Fènoël, H.K.D.H. Bhadeshia, Effect of tempering

upon the tensile properties of a nanostructured bainitic steel, Mater. Sci. Eng. A 615

(2014) 340–347.

[31]F.G. Caballero, H.K.D.H. Bhadeshia, K.J.A. Mawella, D.G. Jones, P. Brown, Very strong

low temperature bainite, Mater. Sci. Technol. 18 (2002) 279–284.

[32]C. Garcia-Mateo, M. Peet, F.G. Caballero, H.K.D.H. Bhadeshia, Tempering of hard

mixture of bainitic ferrite and austenite, Mater. Sci. Technol. 20 (2004) 814–818.

[33]H.F. Lan, L.X. Du, Q. Li, C.L. Qiu, J.P. Li, R.D.K. Misra, Improvement of strength-

toughness combination in austempered low carbon bainitic steel: The key role of refining

prior austenite grain size, J. Alloy. Comp. 710 (2017) 702–710.

[34]J. Kang, F.C. Zhang, X.W. Yang, B. Lv, K.M. Wu, Effect of tempering on the

microstructure and mechanical properties of a medium carbon bainitic steel, Mater. Sci.

Eng. A 686 (2017) 150–159.

[35]C. Hofer, F. Winkelhofer, H. Clemens, S. Primig, Morphology change of retained

austenite during austempering of carbide-free bainitic steel, Mater. Sci. Eng. A 664 (2016)

236–246.

- 22 -
[36]J. Meng, Y. Feng, Q. Zhou, L. Zhao, F. Zhang, L. Qian, Effects of austempering

temperature on strength, ductility and toughness of low-C high-Al/Si carbide-free bainitic

steel, J. Mater. Eng. Perform. 24 (2015) 3068–3076.

[37]B. Avishan, S. Yazdani, S.H. Nedjad, Toughness variations in nanostructured bainitic

steels, Mater. Sci. Eng. A 548 (2012) 106–111.

[38]F.G. Caballero, H.K.D.H. Bhadeshia, K.J.A. Mawella, D.G. Jones, P. Brown, Design of

novel high strength bainitic steels: part 1, Mater. Sci. Technol. 17 (2001) 512–516.

[39]F.G. Caballero, H.K.D.H. Bhadeshia, K.J.A. Mawella, D.G. Jones, P. Brown, Design of

novel high strength bainitic steels: part 2, Mater. Sci. Technol. 17 (2001) 517–522.

[40]A.P. Bentley, G.C. Smith, Phase transformation of austenitic stainless steels as a result of

cathodic hydrogen charging, Metall. Trans. A 17 (1986) 1593–1600.

[41]MIL-DTL-46100E (MR), Armor plate, steel, wrought, high-hardness (2008).

[42]G. Krauss, Solidification, segregation, and banding in carbon and alloy steels, Metall.

Mater. Trans. B 34 (2003) 781–792.

[43]M.A. Mohtadi-Bonab, H. Ghesmati-Kucheki, Important factors on the failure of pipeline

steels with focus on hydrogen induced cracks and improvement of their resistance:

review paper, Met. Mater. Int. 25 (2019) 1109–1134.

[44]K.W. Andrews, Empirical formulae for the calculation of some transformation

temperatures, J. Iron Steel Inst. 203 (1965) 721–727.

[45]W. Steven, A.G. Haynes, The temperature of formation of martensite and bainite in low-

alloy steels, J. Iron Steel Inst. 183 (1956) 349–359.

[46]S.M.C. van Bohemen, M.J. Santofimia, J. Sietsma, Experimental evidence for bainite

formation below Ms in Fe–0.66C, Scr. Mater. 58 (2008) 488–491.

- 23 -
[47]L. Zhao, L. Qian, J. Meng, Q. Zhou, F. Zhang, Below-Ms austempering to obtain refined

bainitic structure and enhanced mechanical properties in low-C high-Si/Al steels, Scr.

Mater. 112 (2016) 96–100.

[48]A. Navarro-López, J. Hidalgo, J. Sietsma, M.J. Santofimia, Characterization of

bainitic/martensitic structures formed in isothermal treatments below the Ms temperature,

Mater. Charact. 128 (2017) 248–256.

[49]D. Kim, J.G. Speer, B.C. De Cooman, Isothermal transformation of a CMnSi steel below

the Ms temperature, Metall. Mater. Trans. A 42 (2011) 1575–1585.

[50]D. Kim, S.J. Lee, B.C. De Cooman, Microstructure of low carbon steel isothermally

transformed in the Ms to Mf temperature range, Metall. Mater. Trans. A 43 (2012) 4967–

4983.

[51]R. Vogt, Z. Zhang, E. Huskins, B. Ahn, S. Nutt, K.T. Ramesh, E.J. Lavernia, J.M.

Schoenung, High strain rate deformation and resultant damage mechanisms in ultrafine-

grained aluminum matrix composites, Mater. Sci. Eng. A 527 (2010) 5990–5996.

[52]Y. Bai, B. Dodd, Adiabatic shear localization: Occurrence, theories, and applications,

Pergamon Press, Oxford (1992).

[53]F. Yuan, X. Bian, P. Jiang, M. Yang, X. Wu, Dynamic shear response and evolution

mechanisms of adiabatic shear band in an ultrafine-grained austenite–ferrite duplex steel,

Mech. Mater. 89 (2015) 47–58.

[54]M.A. Meyers, Y.B. Xu, Q. Xue, M.T. Pérez-Prado, T.R. McNelley, Microstructural

evolution in adiabatic shear localization in stainless steel, Acta Mater. 51 (2003) 1307–

1325.

[55]Y. Xu, H.J. Yang, M.A. Meyers, Dynamic recrystallization in the shear bands of Fe–Cr–

Ni monocrystal: Electron backscatter diffraction characterization, Scr. Mater. 58 (2008)

691–694.

- 24 -
[56]K.-I. Sugimoto, T. Iida, J. Sakaguchi, T. Kashima, Retained austenite characteristics and

tensile properties in a TRIP type bainitic sheet steel, ISIJ Int. 40 (2000) 902–908.

[57]J. Chiang, J.D. Boyd, A.K. Pilkey, Effect of microstructure on retained austenite stability

and tensile behaviour in an aluminum-alloyed TRIP steel, Mater. Sci. Eng. A 638 (2015)

132–142.

[58]K.-I. Sugimoto, M. Murata, T. Muramatsu, Y. Mukai, Formability of C–Si–Mn–Al–Nb–

Mo ultra high-strength TRIP-aided sheet steels, ISIJ Int. 47 (2007) 1357–1362.

[59]R.J. Moat, S.Y. Zhang, J. Kelleher, A.F. Mark, T. Mori, P.J. Withers, Work hardening

induced by martensite during transformation-induced plasticity in plain carbon steel, Acta

Mater. 60 (2012) 6931–6939.

[60]N. Ishikawa, K. Yasuda, H. Sueyoshi, S. Endo, H. Ikeda, T. Morikawa, K. Higashida,

Microscopic deformation and strain hardening analysis of ferrite–bainite dual-phase

steels using micro-grid method, Acta Mater. 97 (2015) 257–268.

[61]C. Garcia-Mateo, F.G. Caballero, T. Sourmail, M. Kuntz, J. Cornide, V. Smanio, R. Elvira,

Tensile behaviour of a nanocrystalline bainitic steel containing 3 wt% silicon, Mater. Sci.

Eng. A 549 (2012) 185–192.

[62]Y. Wang, K. Zhang, Z. Guo, N. Chen, Y. Rong, A new effect of retained austenite on

ductility enhancement in high strength bainitic steel, Mater. Sci. Eng. A 552 (2012) 288–

294.

[63]Y. Huang, Q. Li, X. Huang, W. Huang, Effect of bainitic isothermal transformation plus

Q&P process on the microstructure and mechanical properties of 0.2C bainitic steel,

Mater. Sci. Eng. A 678 (2016) 339–346.

[64]H. Luo, H. Dong, New ultrahigh-strength Mn-alloyed TRIP steels with improved

formability manufactured by intercritical annealing, Mater. Sci. Eng. A 626 (2015) 207–

212.

- 25 -
[65]J.A. Jiménez, M. Carsí, O.A. Ruano, G. Frommeyer, Effect of testing temperature and

strain rate on the transformation behaviour of retained austenite in low-alloyed

multiphase steel, Mater. Sci. Eng. A 508 (2009) 195–199.

[66]T.W. Wright, Approximate analysis for the formation of adiabatic shear bands, J. Mech.

Phys. Solids 38 (1990) 515–530.

[67]T.W. Wright, The physics and mathematics of adiabatic shear bands, Cambridge press,

NY, USA, (2002).

[68]B.B. Singh, G. Sukumar, P.P. Rao, K.S. Kumar, V. Madhu, R.A. Kumar, Superior ballistic

performance of high-nitrogen steels against deformable and non-deformable projectiles,

Mater. Sci. Eng. A 751 (2019) 115–127.

[69]J.-K. Hwang, Deformation behaviors of flat rolled wire in twinning-induced plasticity

steel, Met. Mater. Int. 26 (2020) 603–616.

[70]Q. Xue, G.T. Gray III, B.L. Henrie, S.A. Maloy, S.R. Chen, Influence of shock

prestraining on the formation of shear localization in 304 stainless steel, Metall. Mater.

Trans. A 36 (2005) 1471–1486.

[71]J. He, F. Yuan, M. Yang, S. Jiao, X. Wu, Superior mechanical properties and deformation

mechanisms of heterogeneous laminates under dynamic shear loading, Mater. Sci. Eng. A

756 (2019) 492–501.

[72]N. Li, Y.D. Wang, R. Lin Peng, X. Sun, P.K. Liaw, G.L. Wu, L. Wang, H.N. Cai,

Localized amorphism after high-strain-rate deformation in TWIP steel, Acta Mater. 59

(2011) 6369–6377.

[73]Y. Me-Bar, D. Shechtman, On the adiabatic shear of Ti-6Al-4V ballistic targets, Mater.

Sci. Eng. 58 (1983) 181–188.

- 26 -
Table 1. Room-temperature Vickers hardness and quasi-static tensile test results of the QT

and AT steels.

Quasi-static Tension

Vickers Yield Tensile


Elongation Yield
Steel Hardness Strength Strength
(%) Ratio
(HV) (MPa) (MPa)

QT 692 ± 6 1817 ± 19 2336 ± 5 9.6 ± 0.3 0.78

AT 672 ± 8 1327 ± 13 2295 ± 4 12.3 ± 0.2 0.58

Table 2. Room-temperature quasi-static and dynamic compressive test results of the QT and

AT steels.

Quasi-static Compression Dynamic Compression

True True
True True
True Maximum Yield Maximum
Steel Yield Fracture
Strength (MPa) Strength Strength
Strength (MPa) Strain
(MPa) (MPa)

QT 2420 ± 14 2839 ± 9 2620 ± 41 3088 ± 29 0.25 ± 0.01

AT 2140 ± 11 2851 ± 13 2301 ± 34 3025 ± 21 0.28 ± 0.01

- 27 -
Table 3 Average length (per unit area of 600 µm distant from the penetrated surface) and

width of ASBs and cracks formed in the ballistically-impacted QT and AT steels.

Length of ASB Width of ASB Length of Major Crack


Steel
(µm) (µm) (µm)
QT 142 10 350

AT 46 7 -

- 28 -
Figure Caption

Fig. 1. Optical and SEM micrographs of the (a,c) QT and (b,d) AT steels. The QT and AT

steels consist mainly of tempered martensite and bainite, respectively, which cannot be

clearly differentiated in these micrographs.

Fig. 2. TEM BF and DF images with SAD patterns of the (a,b) QT and (c,d) AT steels.

Fig. 3. EBSD inverse pole figure (IPF) and phase maps for the (a,b) QT and (c,d) AT steels.

The QT and AT steels consist mainly of lath-type tempered martensite and bainite,

respectively. The austenite is observed at lath boundaries, and its fraction (Vγ) is measured to

be 0.8 and 2.7% in the QT and AT steels, respectively. According to the XRD analysis data in

(e), the Vγ is 1.9% and 3.9% in the QT and AT steels, respectively, which corresponds to the

EBSD data, although the Vγ is slightly underestimated for the EBSD analyses.

Fig. 4. (a) Quasi-static tensile stress-strain curves, (b) quasi-static true compressive stress-

strain curves, and (c) dynamic true compressive stress-strain curves, and (d) dynamic strain-

hardening-rate curves of the QT and AT steels. Dynamic curves were obtained from the

dynamic compressive test under an impact momentum of 12.4 or 15.2 kg∙m∙s-1 (air pressure

of 0.2 or 0.3 MPa, respectively).

Fig. 5. Low-magnification photographs of the dynamically compressed cylindrical specimens

of the (a) QT and (b) AT steels.

Fig. 6. Optical micrographs of the half-sectioned area of the dynamically compressed

specimen of the (a,c) QT and (b,d) AT steels. Adiabatic shear bands (ASBs) form at

specimen edge areas.

Fig. 7. (a) Volume fraction of martensitic transformation as a function of compressive true

strain for the QT and AT steels. (b–d) EBSD phase maps of the dynamically compressed

- 29 -
specimens at the true strains of 0, 0.11, and 0.16 for the AT steel, showing the transformation

from austenite to martensite.

Fig. 8. Montages of nital-etched high-magnification optical micrographs of the half-

sectioned area near the penetrated region of the ballistically-impacted target specimen of

the (a) QT and (b) AT steels. The penetrated surface is covered with ASBs (yellow-arrow

marks), and cracks propagate along already-embrittled ASBs (blue-arrow marks).

Fig. 9. TEM BF images with SAD patterns of the interior and exterior areas of ASB after

the dynamic compression under the impact momentum of 15.2 kg∙m∙s-1 for the (a,b) QT

and (c,d) AT steels.

Fig. 10. TEM BF images with SAD patterns of the interior and exterior areas of ASB

after the ballistic impact for the (a,b) QT and (c,d) AT steels.

Fig. 11. (a,b) Quasi-static true compressive stress-strain curves and (c) variations of true

flow stress at the strain of 0.1 as a function of temperature for the QT and AT steels. The AT

steel has the about twice higher n value than the QT steel (0.070 vs. 0.038).

- 30 -
(a) QT (b) AT

Band Structure

Band Structure


50 μm ㅋ
50 μm

(c) QT (d) AT


4 μm ㅋ
4 μm

Fig. 1. Optical and SEM micrographs of the (a,c) QT and (b,d) AT steels. The QT and AT steels consist mainly of tempered martensite and bainite, respectively,
which cannot be clearly differentiated in these micrographs.
(a) (b) Austenite

111 200

111


0.1 μm ㅋ
0.1 μm
[011]FCC

(c) (d) Austenite

111 200

111


0.1 μm ㅋ
0.1 μm
[011]FCC

Fig. 2. TEM BF and DF images with SAD patterns of the (a,b) QT and (c,d) AT steels.
(e)
(a) QT (b) QT (c) AT (d) AT

Austenite

(110)α

(200)α

(211)α

(220)α
Intensity

(111)γ

(200)γ

(220)γ

(311)γ

(222)γ
Austenite
AT Vγ: 3.9%

QT Vγ: 1.9%

2 µm Vγ : 0.8% 2 µm Vγ : 2.7%
40 50 60 70 80 90 100
BCC FCC
θ (Degree)

Fig. 3. EBSD inverse pole figure (IPF) and phase maps for the (a,b) QT and (c,d) AT steels. The QT and AT steels consist mainly of lath-type tempered
martensite and bainite, respectively. The austenite is observed at lath boundaries, and its fraction (Vγ) is measured to be 0.8 and 2.7% in the QT and AT
steels, respectively. According to the XRD analysis data in (e), the Vγ is 1.9% and 3.9% in the QT and AT steels, respectively, which corresponds to the EBSD
data, although the Vγ is slightly underestimated for the EBSD analyses
(a) (b)
3500 3500

Engineering Stress (MPa)


3000 3000

True Stress (MPa)


2500 2500

2000 2000

1500 1500

1000 1000

500 QT_0.001 s-1 500 QT_0.002 s-1


AT_0.001 s-1 AT_0.002 s-1
0 0
0 5 10 15 20 25 30 35 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35
Engineering Strain (%) True Strain
(c) (d)
3500 12000
QT_1780 s-1
3000 10000 AT_1870 s-1
QT_3030 s-1

σ/dεε (MPa)
True Stress (MPa)

2500
8000 AT_3100 s-1
2000
6000
1500

dσ/
σ/
4000
1000 QT_1780 s-1
AT_1870 s-1
500 QT_3030 s-1 2000
AT_3100 s-1
0 0
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35
True Strain True Strain

Fig. 4. (a) Quasi-static tensile stress-strain curves, (b) quasi-static true compressive stress-strain curves, and (c) dynamic true compressive stress-strain
curves, and (d) dynamic strain-hardening-rate curves of the QT and AT steels. Dynamic curves were obtained from the dynamic compressive test under an
impact momentum of 12.4 or 15.2 kg∙m∙s-1 (air pressure of 0.2 or 0.3 MPa, respectively).
(a)
Before 12.4 kg∙m∙s-1 15.2 kg∙m∙s-1
1780 s-1 3030 s-1
QT

1 mm

12.4 kg∙m∙s-1 15.2 kg∙m∙s-1


(b) Before
1870 s-1 3100 s-1
AT

1 mm

Fig. 5. Low-magnification photographs of the dynamically compressed cylindrical specimens of the (a) QT and (b) AT steels.
(a) QT_1780 s-1 (b) AT_1870 s-1

ASB


20 μm ㅋ
20 μm

(c) QT_3030 s-1_20% (d) AT_3100 s-1_22.5%

Crack
ASB

Crack
ASB


50 μm ㅋ
50 μm

Fig. 6. Optical micrographs of the half-sectioned area of the dynamically compressed specimen of the (a,c) QT and (b,d) AT steels. Adiabatic shear bands
(ASBs) form at specimen edge areas.
(a)
5 (b) 0 (c) 0.11 (d) 0.16
QT
Volume Fraction of Martensitic

AT
4
Transformation (%)

2 Austenite

0.00 0.03 0.06 0.09 0.12 0.15 0.18 2 µm Vγ : 2.7% Vγ : 0.4% Vγ : 0%

True Strain BCC FCC

Fig. 7. (a) Volume fraction of martensitic transformation as a function of compressive true strain for the QT and AT steels. (b–d) EBSD phase maps of the
dynamically compressed specimens at the true strains of 0, 0.11, and 0.16 for the AT steel, showing the transformation from austenite to martensite.
(a) ASB QT

Crack
ASB ASB


100 μm

(b) AT

ASB

100 μm

Fig. 8. Montages of nital-etched high-magnification optical micrographs of the half-sectioned area near the penetrated region of the ballistically-impacted
target specimen of the (a) QT and (b) AT steels. The penetrated surface is covered with ASBs (yellow-arrow marks), and cracks propagate along already-
embrittled ASBs (blue-arrow marks).
(a) QT_Interior of ASB (b) QT_Exterior of ASB


100
0.1 μm ㅋ
100
0.1 μm

(c) AT_Interior of ASB (d) AT_Exterior of ASB


100
0.1 μm ㅋ
100
0.1 μm

Fig. 9. TEM BF images with SAD patterns of the interior and exterior areas of ASB after the dynamic compression under the impact momentum of 15.2
kg∙m∙s-1 for the (a,b) QT and (c,d) AT steels.
(a) QT_Interior of ASB (b) QT_Exterior of ASB

110
011

101


100
0.1 μm [111]BCC ㅋ
100
0.1 μm

(c) AT_Interior of ASB (d) AT_Exterior of ASB

110
011

101


100
0.1 μm [111]BCC ㅋ
100
0.1 μm

Fig. 10. TEM BF images with SAD patterns of the interior and exterior areas of ASB after the ballistic impact for the (a,b) QT and (c,d) AT steels.
(a) (b) (c)
QT AT
200 °C

True Stress (MPa)


True Stress (MPa)

True Stress (MPa)


300 °C

400 °C

500 °C
QT n=0.038 QT / -5.74 (MPa/K)
AT n=0.070 AT / -5.37 (MPa/K)

True Plastic Strain True Strain Temperature (K)

Fig. 11. (a,b) Quasi-static true compressive stress-strain curves and (c) variations of true flow stress at the strain of 0.1 as a function of temperature for the
QT and AT steels. The AT steel has the about twice higher n value than the QT steel (0.070 vs. 0.038).
Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

You might also like