Processing math: 100%

Low cycle fatigue behavior of high strength gun steels

Maoqiu Wang, Han Dong, Qi Wang, Changgang Fan

分享

计量
  • 文章访问数:  380
  • HTML全文浏览量:  111
  • PDF下载量:  34
  • 被引次数: 29

目录

Cite this article as:

Maoqiu Wang, Han Dong, Qi Wang, and Changgang Fan, Low cycle fatigue behavior of high strength gun steels, J. Univ. Sci. Technol. Beijing , 11(2004), No. 3, pp.268-272.
Maoqiu Wang, Han Dong, Qi Wang, and Changgang Fan, Low cycle fatigue behavior of high strength gun steels, J. Univ. Sci. Technol. Beijing , 11(2004), No. 3, pp.268-272.
引用本文 PDF XML SpringerLink
Materials

Low cycle fatigue behavior of high strength gun steels

    通信作者:

    Maoqiu Wang E-mail: maoqiuwang@hotmail.com

The low cycle fatigue (LCF) behavior of two high strength steels, with nominal chemical compositions (mass fraction, %)of 0.40C-1.5Cr-3Ni-0.4Mo-0.2V (PCrNi3MoV) and 0.25C-3Cr-3Mo-0.8Ni-0.1Nb (25Cr3Mo3NiNb), was investigated by using the smooth bar specimens subjected to strained-controlled push-pull loading. It is found that both steels show cyclic softening, but 25Cr3Mo3NiNb steel has a lower tendency to cyclic softening. 25Cr3Mo3NiNb steel has higher fatigue ductility, and its transition fatigue life is almost three times that of PCrNi3MoV. 25Cr3Mo3NiNb steel also shows higher LCF life either at a given total strain amplitude above 0.5% or at any given plastic strain amplitude, despite its lower monotonic tensile strength than that of PCrNi3MoV.It also means that 25Cr3Mo3NiNb steel can endure higher total strain amplitude and plastic strain amplitude at a given number of reversals to failure within 104. 25Cr3Mo3NiNb steel is expected to be a good gun steel with high LCF properties because only several thousand firings are required for gun barrel in most cases.

 

Materials

Low cycle fatigue behavior of high strength gun steels

Author Affilications
  • Received: 17 March 2003;
The low cycle fatigue (LCF) behavior of two high strength steels, with nominal chemical compositions (mass fraction, %)of 0.40C-1.5Cr-3Ni-0.4Mo-0.2V (PCrNi3MoV) and 0.25C-3Cr-3Mo-0.8Ni-0.1Nb (25Cr3Mo3NiNb), was investigated by using the smooth bar specimens subjected to strained-controlled push-pull loading. It is found that both steels show cyclic softening, but 25Cr3Mo3NiNb steel has a lower tendency to cyclic softening. 25Cr3Mo3NiNb steel has higher fatigue ductility, and its transition fatigue life is almost three times that of PCrNi3MoV. 25Cr3Mo3NiNb steel also shows higher LCF life either at a given total strain amplitude above 0.5% or at any given plastic strain amplitude, despite its lower monotonic tensile strength than that of PCrNi3MoV.It also means that 25Cr3Mo3NiNb steel can endure higher total strain amplitude and plastic strain amplitude at a given number of reversals to failure within 104. 25Cr3Mo3NiNb steel is expected to be a good gun steel with high LCF properties because only several thousand firings are required for gun barrel in most cases.

 

  • Lead halide perovskites have received increasing attentionrecently because of their high performance as light absorbers. For example, their photoconversion efficiency can reach 22.1%, which suggests their great potential to be applied to the optoelectronic field [12]. However, hybrid perovskites have poor long-term stability due to the decomposition and volatilization of the organic components within them [34]. All-inorganic perovskites are more stable than organic–inorganic hybrid perovskites, and they exhibit broad chemical/physical tunability and excellent charge transport. Therefore, all-inorganic perovskites have been widely applied in the optoelectronic field in various forms, such as lasers [5], photodetectors (PDs) [67], light-emitting diodes [8], and photovoltaic solar cells [910], PDs are particularly regarded as a hot topic because of their wide applications. For instance, Li et al. [11] reported CsPbBr3 microparticles with a detectivityof 6.1 × 1010 Jones. Saidaminov et al. [12] reported CsPbBr3 bulk single crystals with a detectivity of 1.7 × 1011 Jones. Waleed et al. [13] reported CsPbI3 nanoarray PDs with a high detectivity of up to 1.57 × 1012 Jones. However, three-dimensional perovskite bulks and polycrystalline films possess low photoconversion efficiencies because of the undesired charge recombination at grain boundaries and the low carrier mobility (<10 cm2·V−1·s−1) [1415]. Improving perovskite PDs remains a great challenge today.

    Compared with three-dimensional perovskite bulks and polycrystalline films, one-dimensional single-crystalline perovskite nanobelts with well structures possess longer carrier diffusion lengths, larger carrier mobility, and higher photoluminescence quantum yields [1618]. Moreover, one-dimensional perovskite nanobelts can provide relatively direct charge transport pathways, which can benefit the collection of carriers. In this work, we report the exploration of PDs based on a single CsPbI3 nanobelt. First, the high-purity single-crystalline CsPbI3 nanobelts were synthesized via a solution process using different amounts of PbI2. Second, individual CsPbI3 nanobelt PDs were constructed via photolithography, magnetron sputtering, and lift-off process. Third, response repeatability, switching ratio, response time, external quantum efficiency, spectral responsivity, and detectivity of PDs were measured. At last, the mechanism of the outstanding performance of the nanobelt PDs was discussed.

    All chemicals were used without any further purification. Hexane (anhydrous, 95%), oleylamine (OlAm, 70%), 1-octadecene (ODE, 90%), toluene (anhydrous, 95%), hexanoic acid (≥99%), octanoic acid (99%), octylamine (OctAm, 99%), and oleic acid (OA, 90%) were purchased from Sinopharm Chemical Reagent Beijing Co., Ltd., China. Cesium carbonate (Cs2CO3, 99%) and lead(II) iodide (PbI2, 99%) were purchased from Aladdin Reagent Co., Ltd., China.

    Briefly, 1.2 mL of OA, 0.4 g of Cs2CO3, and 15 mL of ODE were added into a three-neck flask, degassed, and dried under vacuum at 120°C for 60 min. The mixture was then heated under N2 to 150°C until all Cs2CO3 dissolved into OA.

    Exactly 8 mL of ODE, 0.069 g of PbI2, and proper amounts of OlAm and OctAm were added into a 25 mL three-neck flask. The mixture was dried at 100°C for 45 min under vacuum to form a cloudy solution. Thereafter, the mixture was heated under N2 to 120°C and held for 10 min. Exactly 0.6 mL of Cs-oleate solution was injected in a three-neck flask. The nanobelt was then allowed to grow for 50 min at 120°C. Immediately following the synthesis, the reaction was quenched by an ice water bath and centrifuged at 10000 r/min for 5 min. The nanobelt was isolated through centrifugation at 7000 r/min for 3 min. The obtained nanobelt was redispersed in hexane/toluene for later use. All procedures were conducted under ambient conditions.

    Individual CsPbI3 nanobelt PDs were fabricated. Through photolithography, magnetron sputtering, and lift-off process, interdigitated Au electrodes (100 nm) with 4 µm separation were patterned on top of a Si substrate with a 300 nm SiO2 layer. Then, the CsPbI3 nanobelt in hexane was deposited dropwise on the electrodes.

    All experiments were conducted under ambient conditions at room temperature. X-ray diffraction (XRD, D8 Advance, Bruker, Germany) with Cu Kα radiation (λ = 0.15406 nm), transmission electron microscopy (TEM, JEM-2010, JEOL, Japan), and field-emission scanning electron microscopy (S-4800, Hitachi, Japan) together with energy-dispersive X-ray spectroscopy were used to characterize the nanobelt. A UV–vis scanning spectrophotometer (U-3900, HITACHI, Japan) was used to obtain the UV–vis spectrum of the nanobelt. A four-probe station with a semiconductor characterization system (Keithley 4200-CSC) was used to obtain the electrical and optoelectronic data of the PDs. A 500 W Xenon arc lamp coupled to an Acton Research monochromator with order-sorting filters was used as the light source. Light intensity was measured with an OAI-306 power meter. The laser had a wavelength of 405 nm, maximum power of 200 mW, and a spot size measuring 4 mm in diameter.

    In our recent work, we reported that the amount of PbI2 plays a fundamental role in the preparation of defect-free and high-quality one-dimensional CsPbI3 nanobelts [7]. Herein, the microstructure of the as-prepared CsPbI3 was revealed by the scanning electron microscope (SEM) techniques. When 96 mg of PbI2 was introduced (Figs. 1(a) and 1(b)), nanorods with a uniform diameter of 150 nm and an average length of 2 µm were obtained. When 87 mg of PbI2 was introduced (Figs. 1(c) and 1(d)), the resultant products were pure nanobelts with a mean width, length, and thickness of 100 nm, 5 µm, and 20 nm, respectively. The typical thickness and width measured by an atomic force microscope were approximately 20 nm and 100 nm, respectively (inset of Fig. 1(d)). The thickness/width (t/w) ratio was above 0.2. However, when the amount of introduced PbI2 was increased to 105 mg (Figs. 1(e) and 1(f)), the as-synthesized nanorods began to aggregate and conglutinate.

    Figure  1.  SEM images of the obtained CsPbI3 nanocrystal with varying amounts of introduced PbI2: (a, b) 96 mg; (c, d) 87 mg; (e, f) 105 mg. Inset of (d) is the typical thickness and width measurement result of nanobelts from an atomic force microscope.

    As shown in Fig. 2, strong diffraction peaks ascribed to the orthorhombic phase of CsPbI3 (JCPDS Card No. 18-0376) were observed in the XRD patterns of the obtained samples. Generally, CsPbI3 undergoes a one-phase transition under a reduced temperature [19], that is, its color changes from dark to yellow, and it exhibits a cubic-orthorhombic structure (328°C) [20]. As shown in Fig. 2, the CsPbI3 nanobelt was confirmed to be in the orthorhombic phase on the basis of the XRD pattern of the crystal and the yellow color.

    Figure  2.  XRD pattern of nanobelt. The inset is a photograph of the CsPbI3 nanobelt dispersed in hexane.

    The single-crystalline nature and chemical composition of the nanobelt were confirmed using TEM and energy dispersive X-Ray (EDX). As shown in Figs. 3(a)3(c), the morphology of the nanobelt indicated a typical width of 100 nm. A d spacing of ~0.479 nm, which corresponded to (100) planes, was observed in the high resolution transmission electron microscope (HRTEM) result (Fig. 3(d)). The selected area electron diffraction (SAED) pattern (Fig. 3(e)) can be indexed to an orthorhombic structure, which indicated its single-crystalline nature. The EDX pattern of the single nanobelt revealed Cs, Pb, and I elements with a quantified molar ratio of 1:1:3, which further confirmed that the nanobelt was CsPbI3. The single-crystalline nature and suitable one-dimensional morphology of the nanobelt provide PDs with capabilities for high-quality charge carrier transmission and complete light absorption.

    Figure  3.  (a–c) TEM images of CsPbI3 nanobelt; (d) a representative HRTEM image of CsPbI3 nanobelt; (e) SAED pattern of CsPbI3 nanobelt; (f) typical EDX spectrum of the nanobelt.

    The optoelectronic properties of the CsPbI3 nanobelt were investigated on the basis of the UV–vis absorption spectrum. As shown in Fig. 4(a), the absorbance peak of the CsPbI3 nanobelt was located at 405 nm. Fig. 4(b) shows the (α)2 versus Eg plot (α is absorption coefficient, h is Planck’s constant, v is frequency of light, Eg is band gap) for the CsPbI3 nanobelt. A direct bandgap of about 2.60 eV was observed. The CsPbI3 nanobelt can thus be used as a violet PD due to the location of the absorbance at 405 nm.

    Figure  4.  (a) UV–vis spectrum of CsPbI3 nanobelt; (b) (αhν)2 vs. Eg plot; (c) schematic of single CsPbI3 nanobelt PD; (d) SEM image of individual nanobelt PD.

    Utilizing the as-grown CsPbI3 nanobelt, we fabricated PDs with a low-dimensional structure. Figs. 4(c) and 4(d) present the schematic of the individual CsPbI3 nanobelt PD and its SEM image. Fig. 5(a) shows the comparative current–voltage (IV) characteristics of the CsPbI3 nanobelt PD under dark conditions and a light of 405 nm with an average power of 10 mW/cm2. The dark current of the PD was lower than 0.19 nA at 2.0 V. When the device was illuminated under a light above the bandgap Eg of 3.06 eV, the photocurrent magnitude of the CsPbI3 nanobelt PD was two orders higher than that of the dark current. The photocurrent approached 7.90 nA at a bias of 2.0 V. The asymmetric and nonlinear IV curves (Fig. 5(b)) indicated that Schottky contact occurred between the electrodes and the CsPbI3 nanobelt.

    Figure  5.  (a) Logarithmic IV under 405 nm light and dark conditions; (b) IV characteristics under 405 nm light and dark conditions; (c) reproducible on/off switching under 405 nm light; (d) response time (recovery time (τoff) and response time (τon)) under 405 nm light pulse chopped at a 0.05 Hz frequency.

    Response repeatability is a key parameter for PDs. The time–response of the CsPbI3 nanobelt PD was measured by periodically turning on and off the 405 nm light at a voltage of 2.0 V. As shown in Fig. 4(c), when the light irradiation was on and off, the current exhibited two distinct states, namely, dark current of 0.19 nA and increased photocurrent of 7.90 nA, respectively. In switching the light on/off for more than 200 cycles, the CsPbI3 nanobelt PDs exhibited excellent stability and reproducibility. The switching ratio (δSR) was calculated as:

    δSR=IphIdark=IonIoffIoff (1)

    where Idark is the dark current, Ioff is the current of turning off the light, Iph is the photocurrent under 405 nm light (10 mW/cm2) and Ion is the current of turning on the light. The δSR of the nanobelt PD reached 41. The huge gap in δSR was related to light absorptivity, which is influenced by the types of materials.

    Response time is another key parameter for PDs. Herein, the response time was measured by using 405 nm continuous laser triggers with a pulse width of 0.05 Hz. The sharp current from one state to another indicated an extremely fast response time. The recovery time (τoff) and response time (τon) were defined as the time for the maximum photocurrent to reach 10% of the dark current or vice versa (90%). As shown in Fig. 5(d), the recovery time and response time of the nanobelt PD were measured as ~0.5 s and 0.5 s, respectively. The single-crystalline nature and one-dimensional morphology of the nanobelt that favor carrier transport led to the fast response time.

    External quantum efficiency (ηEQE) and spectral responsivity (Rλ) are two important parameters for PDs [2122], and they are respectively calculated as

    Rλ=ΔILlight=IonIoffPA (2)
    ηEQE=hceλRλ (3)

    where A is the effective area of the detector, P is the light power intensity, Llight is the incident light intensity, and ΔI is the difference between the photocurrent and the dark current. λ is the exciting wavelength, e is the electron charge, c is the velocity of light, and h is Planck’s constant. Under 405 nm light with 10 mW/cm2 under an applied voltage of 2.0 V, the calculated Rλ and ηEQE of the PD (Figs. 6(a) and 6(b)) were as high as 770.65 A/W and 2.39 × 105%, respectively.

    Figure  6.  (a) Spectral response of PD from 250 to 600 nm; (b) detectivity and ηEQE of PD at different wavelengths; (c) time–response curves of nanobelt PD under different light power intensities; (d) relationship between light power intensity and photocurrents.

    The wavelength selectivity of a PD determines the applied wavelength range of the device. Fig. 6(a) shows the spectral response of the nanobelt PD to the wavelength changing from 250 to 600 nm at a bias of 2.0 V. The ratio between Rλ (405 nm, 770.65 A/W) and Rλ (500 nm, 0.46 A/W) was approximately 1675.32, which indicated the high spectral selectivity and sensitivity of the PD. Hence, as shown in Fig. 6(b), the nanobelt PD developed herein can be used as a typical ultraviolet (UV) and blue light PD.

    Detectivity (D*) is given by

    D=(Af)12RλIn (4)

    where In is the noise current and f is the electrical bandwidth. D* can be expressed as follows when the shot noise dominates the dark current:

    D=Rλ(2eIoffA)12 (5)

    Evidently, the smaller the dark current is, the better the detection of weak optical signals will be. Avoiding any leakage current during operation is extremely important to obtain a small Ioff. The effective ways to obtain a small Ioff include maintaining good single-crystalline quality, low thermal emission (recombination) rates, and low trap density of semiconductors. For the nanobelt PD in this work, the specific detectivity was calculated to be 3.12 × 1012 Jones (Jones = cm·Hz1/2·W−1) at 405 nm. The detectivity of the nanobelt PD approached 1012 Jones from 250 to 450 nm at 2.0 V and is thus at par with the detectivity of Si PDs [2324].

    In addition to the wavelength of light, light power intensity is another key influencing factor for the photocurrent of PDs. The time–response curves of the nanobelt PD are plotted with light power intensity in Fig. 6(c). As the 405 nm light power intensity increased from 3 to 16 mW/cm2 at a voltage of 2.0 V, the photocurrent of the nanobelt PD increased from 3.84 to 9.33 nA. This result agreed with the fact that the absorbed photon flux was proportional to the photoinduced carrier efficiency. Even after being subjected to the largest photocurrent for a long period, the photocurrent of the PD remained stable and exhibited good repeatability. The dependence of the photocurrent on light power intensity can be expressed by a power law: Iph = BPθ [25], where θ is the exponent (0.5 < θ < 1), B is a constant for a given wavelength; θ = 0.53 was obtained (Fig. 6(d)) by fitting the curve in Fig. 6(c). The non-unity exponent suggested a complex process, which included electron-hole generation and trapping and recombination within the CsPbI3 nanobelt [2628].

    Under 405 nm light illumination at a bias of 2.0 V, Rλ, D*, and ηEQE reached 770 A/W, 3.12 × 1012 Jones, and 2.39 × 105%, respectively. In the range of 300–450 nm, the detectivity of the PD exceeded 1012 Jones and is thus at par with the detectivity of Si PDs (i.e., 1012 Jones) [2324]. As listed in Table 1, the detectivity is comparable to the best detectivities of pristine perovskite PDs ever reported.

    Table  1.  Typical perovskite PDs reported in the literature
    PhotodetectorBias / VResponsivity / (A·W−1)Detectivity / JonesRef.
    CH3NH3PbI3 film33.49[29]
    CH3NH3PbI3 network100.11.02 × 1012[30]
    CsPbBr3 microparticles100.186.1 × 1010[11]
    CsPbBr3 nanoparticles/Au nanocrystals20.011.68 × 109[31]
    CsPbBr3 thin films6559 × 1012[32]
    CsPbI3 nanoarrays10.00671.57 × 1012[13]
    CsPbBr3 nanosheets/carbon nanotubes1031.1[33]
    CsPbBr3 nanoplatelets1.5347.5 × 1012[34]
    CsPbBr3 bulk single crystals00.0281.7 × 1011[12]
    CsPbI3 nanobelt2770.653.12 × 1012This work
    下载: 导出CSV 
    | 显示表格

    The outstanding performance of the nanobelt PD is mainly attributed to the following reasons. First, the CsPbI3 nanobelt has a low recombination of charge carriers, low density of defects, short paths for carrier transfer, and high absorption coefficient; these properties could result in a strong photoelectric effect. Second, the single-crystal CsPbI3 nanobelt possesses high single crystallinity. Therefore, the recombination of charge carriers is limited due to the low density of defects of the single-crystalline structure [35]. Third, the absorption coefficient of the perovskite reaches the order of 104 cm−1 because of the direct bandgap nature of the electronic transition. Therefore, almost all light can be absorbed by the CsPbI3 nanobelt [23].

    In summary, we demonstrated the growth of an all-inorganic CsPbI3 perovskite nanobelt via a solution process. And the amount of introduced PbI2 played a fundamental role in morphological regulation of the obtained single-crystalline nanobelt. When 87 mg of PbI2 was introduced, the resultant products were pure nanobelts with a mean width, length, and thickness of 100 nm, 5 µm, and 20 nm, respectively. The PDs based single CsPbI3 nanobelt showed an outstanding performance with an external quantum efficiency of 2.39 × 105%, a responsivity of 770 A/W and a detectivity of 3.12 × 1012 Jones. They are at par with the detectivity of Si PDs. The excellent performance of the nanobelt PDs are mainly attributed the intrinsic properties of CsPbI3, high crystallinity of CsPbI3, and special morphology of nanobelt.The overall excellent performance of the CsPbI3 nanobelt makes it an excellent candidate material for various optoelectronic areas.

    This work was financially supported by the National Natural Science Foundation of China (Nos. 51974021 and 51902020), the Fundamental Research Funds for the Central Universities (Nos. FRF-TP-18-045A1 and FRF-TP-19-004B2Z), the National Postdoctoral Program for Innovative Talents (BX20180034), and the China Postdoctoral Science Foundation (Grant No. 2018M641192).

Relative Articles

Chao Gu, Yan-ping Bao, Peng Gan, Min Wang, Jin-shan He. Effect of main inclusions on crack initiation in bearing steel in the very high cycle fatigue regime [J]. 矿物冶金与材料学报(英文版). DOI: 10.1007/s12613-018-1609-4

View details

Zhi-yang Lü, Ao-shuang Wan, Jun-jiang Xiong, Kuang Li, Jian-zhong Liu. Effects of stress ratio on the temperature-dependent high-cycle fatigue properties of alloy steels [J]. 矿物冶金与材料学报(英文版). DOI: 10.1007/s12613-016-1362-5

View details

Yu-bing Guo, Chong Li, Yong-chang Liu, Li-ming Yu, Zong-qing Ma, Chen-xi Liu, Hui-jun Li. Effect of microstructure variation on the corrosion behavior of high-strength low-alloy steel in 3.5wt% NaCl solution [J]. 矿物冶金与材料学报(英文版). DOI: 10.1007/s12613-015-1113-z

View details

Hong-xiang Yin, Ai-min Zhao, Zheng-zhi Zhao, Xiao Li, Shuang-jiao Li, Han-jiang Hu, Wei-guang Xia. Influence of original microstructure on the transformation behavior and mechanical properties of ultra-high-strength TRIP-aided steel [J]. 矿物冶金与材料学报(英文版). DOI: 10.1007/s12613-015-1070-6

View details

Yan-jun Zhao, Xue-ping Ren, Wen-chao Yang, Yue Zang. Design of a low-alloy high-strength and high-toughness martensitic steel [J]. 矿物冶金与材料学报(英文版). DOI: 10.1007/s12613-013-0791-7

View details

Qi-hang Han, Yong-lin Kang, Xian-meng Zhao, Lu-feng Gao, Xue-song Qiu. High-temperature properties and microstructure of Mo microalloyed ultra-high-strength steel [J]. 矿物冶金与材料学报(英文版). DOI: 10.1007/s12613-011-0454-5

View details

Zhi-qiang Cao, Yan-ping Bao, Zheng-hai Xia, Deng Luo, Ai-min Guo, Kai-ming Wu. Toughening mechanisms of a high-strength acicular ferrite steel heavy plate [J]. 矿物冶金与材料学报(英文版). DOI: 10.1007/s12613-010-0358-9

View details

Xue-xia Xu, Yang Yu, Wen-long Cui, Bing-zhe Bai, Jia-lin Gu. Ultra-high cycle fatigue behavior of high strength steel with carbide-free bainite/martensite complex microstructure [J]. 矿物冶金与材料学报(英文版). DOI: 10.1016/S1674-4799(09)60051-0

View details

Citing articles(29)

Linlin Zhou, Tao Yang, Chunyu Guo, et al. PVDF/6H-SiC composite fiber films with enhanced piezoelectric performance by interfacial engineering for diversified applications. Journal of Materials Science & Technology, 2024, 170: 238. 必应学术
Sheng Huang, Husheng Shan, Yafei Xu, et al. Preparation of CsPbBr3@Fe2O3 heterojunction nanocrystals for ppb-level H2S sensing. Ceramics International, 2024, 50(14): 25607. 必应学术
Linhao Li, Yixun He, Tingjun Lin, et al. MXene/AlGaN van der Waals heterojunction self-powered photodetectors for deep ultraviolet communication. Applied Physics Letters, 2024, 124(13) 必应学术
More >

/

返回文章
返回