Processing math: 100%

Residual stress measurement and analysis of siliceous slate-containing quartz veins

Tao Wang, Weiwei Ye, Yemeng Tong, Naisheng Jiang, Liyuan Liu

图(12)  /  表(2)

分享

计量
  • 文章访问数:  2249
  • HTML全文浏览量:  483
  • PDF下载量:  69
  • 被引次数: 8

目录

Cite this article as:

Tao Wang, Weiwei Ye, Yemeng Tong, Naisheng Jiang, and Liyuan Liu, Residual stress measurement and analysis of siliceous slate-containing quartz veins, Int. J. Miner. Metall. Mater., 30(2023), No. 12, pp.2310-2320. https://dx.doi.org/10.1007/s12613-023-2667-9
Tao Wang, Weiwei Ye, Yemeng Tong, Naisheng Jiang, and Liyuan Liu, Residual stress measurement and analysis of siliceous slate-containing quartz veins, Int. J. Miner. Metall. Mater., 30(2023), No. 12, pp.2310-2320. https://dx.doi.org/10.1007/s12613-023-2667-9
引用本文 PDF XML SpringerLink
研究论文

含石英脉硅质板岩封闭应力测量研究

文章亮点

(1) 使用电子背散射衍射和光学显微镜对岩石微观结构进行了表征,并确定了满足X射线衍射封闭应力测量要求的区域 (2) 使用携带X射线衍射仪的sin2ϕ方法测量和计算了含硅片岩的石英脉的封闭应力 (3) 根据显微观察和封闭应力测试的结果对封闭应力的机制进行了分析
冲击地压等工程地质灾害,一直是影响煤矿生产安全的关键因素,而封闭应力被认为是解释这些地质力学现象的可行方法。本研究通过EBSD及光学显微镜进行岩石微观表征,确定存在满足XRD封闭应力测量要求测区;以岩体正交异性弹性理论为基础,利用X射线衍射仪,对含石英脉硅质板岩样品采用sin2ϕ法测量并计算封闭应力的量值,确定封闭应力的主应力;依据微观测试与封闭应力测试结果相结合分析封闭应力产生机制。主要结论如下:依据微观测试结果分析可得岩石在毫米范围内存在均质、粒径较小区域,满足XRD应力测定统计量要求;含石英脉板岩样品确定石英为标定矿物,获取不同角度ϕφ下(324)晶面衍射图谱,其衍射峰偏移方向具有一致性,证明待测样品存在封闭应力;石英脉XRD测得封闭应力主应力均为压应力,大小在10到33 MPa之间,最大主应力平行脉体走向,最小主应力垂直脉体走向;石英脉中小角晶界及孪晶界含量均较高,分析岩石中应力封存与岩石中矿物晶体非均质特性、塑性变形相关。本文所提出的封闭应力的测量方法可为后续开展不同类型岩石封闭应力的观测和相关研究提供参考。

 

Research Article

Residual stress measurement and analysis of siliceous slate-containing quartz veins

Author Affilications
    Corresponding author:

    Yemeng Tong      E-mail: tongyemeng977@163.com

    Liyuan Liu      E-mail: liuliyuan@ustb.edu.cn

  • Received: 20 February 2023; Revised: 09 April 2023; Accepted: 26 April 2023; Available online: 28 April 2023
Engineering geological disasters such as rockburst have always been a critical factor affecting the safety of coal mine production. Thus, residual stress is considered a feasible method to explain these geomechanical phenomena. In this study, electron backscatter diffraction (EBSD) and optical microscopy were used to characterize the rock microcosm. A measuring area that met the requirements of X-ray diffraction (XRD) residual stress measurement was determined to account for the mechanism of rock residual stress. Then, the residual stress of a siliceous slate-containing quartz vein was measured and calculated using the sin2ϕ method equipped with an X-ray diffractometer. Analysis of microscopic test results showed homogeneous areas with small particles within the millimeter range, meeting the requirements of XRD stress measurement statistics. Quartz was determined as the calibration mineral for slate samples containing quartz veins. The diffraction patterns of the (324) crystal plane were obtained under different ϕ and φ. The deviation direction of the diffraction peaks was consistent, indicating that the sample tested had residual stress. In addition, the principal residual stress within the quartz vein measured by XRD was compressive, ranging from 10 to 33 MPa. The maximum principal stress was parallel to the vein trend, whereas the minimum principal stress was perpendicular to the vein trend. Furthermore, the content of the low-angle boundary and twin boundary in the quartz veins was relatively high, which enhances the resistance of the rock mass to deformation and promotes the easy formation of strain concentrations, thereby resulting in residual stress. The proposed method for measuring residual stress can serve as a reference for subsequent observation and related research on residual stress in different types of rocks.

 

  • As the global population and economy continue to grow, the demand for energy and raw materials is increasing [1], resulting in the deepening of coal mining activities. Consequently, many geological disasters have become prominent because of the complexity, suddenness, and diversity of the influencing factors of these disasters. Although it has been extensive researched, the occurrence mechanism of such disasters has no universally accepted explanation.

    Rock residual stress is the self-balanced stress that remains in the Earth crust after rock formation, metamorphism, and tectonic movements [2]. Such residual stress is also referred to as closed stress, “locked-in” stress, or freezing stress [3]. Residual stress is usually concentrated in deep rock mass areas with geological structures, such as joints, faults, and folds [4], which are often subject to rockburst, core fractures, and zonal fragmentation [56].

    Numerous studies have demonstrated the significance of studying residual stress to explain the mechanical behavior of deep rock masses. Varnes et al. [7] proposed the dynamic hypothesis of rock residual stress, indicating that residual stress in rock is a potential cause of problems in the foundation of engineering structures. Holzhausen et al. [2] systematically summarized the concept of residual stress in rock and clarified the confusion caused by misidentifying residual stress as crustal stress. Müller [8] emphasized the importance of studying residual stress in rocks, whereas Tan et al. [9] proposed the hypothesis of “locked-in” stresses and the concept of stress pockets, discussing the importance of stress pockets in practical engineering. The latter study has laid the foundation for subsequent research on “locked-in” stress. Yue [10] explained the rockburst hysteresis through the pocket theory of “locked-in” stress. An [11] suggested that residual stress and its strain energy are superimposed on stress caused by tectonic movement and released to a high degree when the rock mass fractures, resulting in rapid crack expansion and increasing energy released from rock mass failure, which affects seismic activity. Liu et al. [12] created a rock-like material containing residual stress with nitrile rubber and evaluated the effect of residual stress on the mechanical properties of rock-like materials through rock mechanics tests. Residual stress in rock contains rich information on rock structure as well as tectonic and mechanical evolution [13].

    Based on recent studies, residual stress can serve as an indicator of paleostress. Sekine et al. [14] estimated paleostress at the time of vein formation using residual stress magnitude and fluid-inclusion temperature measurement data. Kai et al. [15] and Wenk et al. [16] used a residual lattice strain in quartz as a paleotectonic stress indicator to explain stress field behavior during tectonic deformation. In addition, the release of residual stress could cause several special geological structures. Based on the characteristics of rock joints, Weinberger et al. [17] hypothesized that schist was formed because of the release of residual elastic strain energy. The microstructure of rock plays a vital role in its mechanical properties. Liu et al. [18] developed a novel dual-damage thermal-mechanical model to define the evolution of the thermal and mechanical properties of rock during thermal treatment, indicating that the thermal treatment caused a realignment in the rock microstructure and resulted in a change in the ultimate failure pattern. The existence of residual stress could also change the physical and mechanical properties of rocks to a certain extent. Lu et al. [19] found that closed stress was the most influential factor in the conductivity of undeveloped fractures in shale.

    The excavation or tectonic movement of a deep rock mass can lead to changes in its initial stress and energy state, resulting in a state that cannot be maintained at equilibrium. The release of internal residual stress can lead to disasters such as rock bursts and special geological phenomena such as core fractures and zone fragmentation. Research on residual stress measurement indicates that rock residual stress is similar to that in other materials. Therefore, methods for measuring residual stress in materials can be used in rock residual stress measurement. These methods are of two types, namely, destructive and nondestructive. Destructive measurement methods include drilling and ring-core methods. Gentry [20] suggested that the stress direction obtained by destructive measurement had some reference significance. However, the results are not successful in accounting for the generation mechanism of rock residual stress and its corresponding impact on structures and the environment because of the large disparity in measurements.

    Nondestructive measurement methods include ultrasonic, magnetic, neutron diffraction, and X-ray diffraction (XRD). The ultrasonic method calculates the change in residual stress by measuring the change in ultrasonic propagation velocity in the material. However, accurately determining rock residual stress from changes in ultrasonic wave velocity can be challenging when the ultrasonic wave velocity is disturbed, and the rock is a heterogeneous material. The magnetic measurement method determines the magnitude and direction of residual stress based on the permeability change of ferromagnetic materials, but this method is often inaccurate and energy-consuming. By contrast, XRD is widely used in quantitative residual stress measurement for metals, polymers, ceramics, and other materials [2125], which measures variations in crystal plane spacing to obtain strain. Geologist Friedman [26] was the first researchers to apply XRD to perform stress analysis for geological materials. Sekine et al. [14] evaluated the residual stress in quartz veins in metamorphic rock using XRD. Some scholars also utilized synchrotron radiation-based X-ray diffraction [1516] and neutron diffraction to measure rock residual stress [27], and they proposed the theoretical feasibility of using X-ray neutron diffraction and other nondestructive measurement methods [28]. Despite some exploratory studies conducted on the measurement of rock residual stress, the unified understanding of the test methods and principles of residual stress, the influence of residual stress on rock engineering, and the generation mechanism is still lacking. Therefore, in this work, electron backscatter diffraction (EBSD) and optical microscopes were used to characterize the rock microcosm to account for the generation mechanism of rock residual stress. XRD, a comprehensive method for rock analysis, was used to measure the residual stress of specific rocks.

    The sample used in this study was obtained from a siliceous slate-containing quartz vein located in Hongling Mine, Inner Mongolia. The sample has a complex matrix composition and random geological conditions, meeting the requirements for residual stress generation. In addition, the quartz vein in the siliceous slate is an intrusion. During formation, the vein body may have been constrained by the surrounding matrix, with rockburst often occurring at the intersection of the veins [29]. Based on the known mechanisms of rock formation, this type of rock is predisposed to residual stress, making it an ideal candidate for processing, testing, and analysis in this study. The sample consists of transverse quartz veins that are roughly parallel to the matrix bedding as well as longitudinal and some small quartz veins that are approximately perpendicular to the matrix bedding. The transverse and longitudinal quartz veins exhibit a vertical crossover state, with the longitudinal quartz veins running through the entire rock with a width of 1–4 mm, whereas the transverse quartz veins are 1–3 mm wide.

    Rock is a complex material consisting of multiple phases and crystals. XRD requires knowledge of the calibration of the target mineral; thus, the diffraction pattern of a specific mineral crystal was used to determine the stress state in the test area. Mineral composition, particle size, and preferred orientation of the siliceous slate-containing quartz veins were analyzed using EBSD.

    As shown in Fig. 1, the sample surface was mechanically polished and argon ion polished, and the siliceous slate-containing quartz vein was prepared into 20 mm × 20 mm × 10 mm EBSD plate samples.

    Figure  1.  Polished EBSD plate sample: (a) experimental sample; (b) arrangement of measuring points; (c) orientation contrast diagram.

    Data were collected using a JSM-IT800 field-emission scanning electron microscope (FESEM), Oxford Instrument EDS, and CMOS EBSD Instrument, with the matrix lineation direction as the X axis, vertical foliation direction as the Z axis, and vertical X axis within the foliation as the Y axis, at a low vacuum pressure of 30 Pa, an acceleration voltage of 20.00 kV, and a sample tilt angle of 70°.

    A forward scattering detector by Oxford Instrument was used to obtain the orientation contrast diagram shown in Fig. 1. Three measuring points with evident differences in brightness and darkness were selected in the scanning area to determine the potential mineral composition of the matrix.

    Based on the matrix energy spectrum, the minerals in the matrix were initially categorized. As shown in Fig. 2, the Kikuchi diffraction pattern measured by EBSD was compared with known Kikuchi diffraction patterns in the crystal library to determine the crystal mineral structure. The matrix minerals were feldspar, quartz, and chlorite, whereas the main gangue mineral of the quartz veins was quartz.

    Figure  2.  Matrix Kikuchi diffraction pattern calibration diagram of (a) quartz, (b) feldspar, and (c) chlorite.

    Fig. 3(a)–(b) shows the statistical map of grain size in the EBSD test area of the matrix, with an average grain size of 7.4 μm. Fig. 3(c)–(d) displays the statistical graph of grain size in the EBSD test area of the quartz veins, with an average grain size of 225 μm. Fig. 4 shows an orthogonal polarizing microscopic image of the sample. The quartz veins have relatively uniform small grains with a particle size ranging from 30 to 40 μm and relatively nonuniform large grains with a particle size ranging from 70 to 400 μm. The small-grained quartz has enough grains in the 2-mm XRD test area for statistical analysis.

    Figure  3.  Grain size statistics of the matrix and quartz veins test area: (a) matrix grain size content; (b) matrix grain size distribution; (c) quartz veins grain size content; (d) quartz veins grain size distribution.
    Figure  4.  Orthogonal polarizing microscopic image of the sample.

    Fig. 5 depicts the inverse pole figures (IPF) of XYZ in the test area of the matrix and quartz vein, respectively. The colors in the test area of matrix and quartz veins are diverse and similarly distributed, indicating the irregular arrangement of crystal particles. The primary mineral phases in the test areas of the matrix and quartz vein were analyzed with the corresponding pole figures to examine the characteristics of the rocks. Fig. 6 shows the pole figures of the main mineral phases of the matrix and quartz vein, corresponding to <001>, <100>, and <110>, respectively. No distinct preferred orientation of the minerals in the test area of the matrix and quartz veins is observed compared with the standard pole figures. Therefore, the EBSD test area of the rock has no specific preferred orientation.

    Figure  5.  Inverse pole figure of (a) matrix and (b) quartz veins at different directions.
    Figure  6.  Pole figure of (a) matrix feldspar, (b) matrix quartz, and (c) quartz veins.

    The quartz veins contain small and uniform areas that satisfy the statistical requirements for XRD testing, as well as areas without any specifically preferred orientation and homogeneous areas of several millimeters within the rock, which also meet the XRD testing requirements. Therefore, residual stress measurements can be performed.

    Samples were cut along the direction of the matrix and vertical planes to establish the X and Y coordinate systems parallel to the matrix foliation and the Z coordinate system perpendicular to the matrix foliation (Fig. 7). Plate samples measuring 20 mm × 20 mm × 10 mm were prepared using linear cutting, and the measuring points were set at Z-1 and Z-2 in the YZ plane of the longitudinal vein, H-1 in the XZ plane on the transverse vein, and point H-2 in the YZ plane.

    Figure  7.  Arrangement of measuring points and establishment of the coordinate system.

    Bragg’s law was applied to obtain XRD patterns. When an electron beam with a wavelength (λ) irradiates a crystal with a regular atomic arrangement and when the crystal plane spacing is d at an angle θ, the scattered electron beams will be superposed and reinforced by one another, resulting in diffraction. Diffraction occurs when the optical path difference is equal to an integer multiple of the wavelength. Bragg’s law [30], as shown in Eq. (1), describes this phenomenon.

    2dsinθ=nλ (1)

    where d is the distance between parallel atomic planes, λ is the wavelength of the incident beam, θ is the Bragg angle, which is half of the diffraction angle, and n is the integer representing the reflection series.

    Considering that X-rays can penetrate the rock to a depth of approximately 10 μm, the surface of the rock sample is in a state of plane stress, along with zero stress in the thickness direction. The azimuth angle (ϕ) of the diffracted crystal plane is the angle between the normal of the diffracted crystal plane and the normal of the sample surface. Using the sin2ϕ method, the stress is measured using various forces within the crystal at different ϕ. With an increase in ϕ, the internal stress of the crystal increases, the spacing between crystal planes decreases, and compressive stress is generated within the rock. Conversely, a decrease in ϕ leads to a decrease in the internal stress of the crystal and generates tensile stress. If the crystal plane spacing is equal to or within the measurement error range, then no stress occurs in the mineral.

    Based on the principle that the strain caused by a change in crystal plane spacing is equal to the macroscopic strain, as well as Bragg’s law and elasticity theory, the stress measurement equation of sin2ϕ was deduced under the assumption of a plane stress state. Eqs. (2)–(4) are the basic equations for macroscopic stress measurement [31]. The value of diffraction angle θ0 corresponding to the true stress-free rock sample was unknown, making it difficult to determine the error related to X-ray measurements. However, an accurate evaluation of stress depends on the relative position of the center of the diffraction peak (i.e., peak displacement) than on the absolute peak position. Thus, θ0 was considered as the radius corresponding to ϕ = 0° in each direction.

    σ=KM (2)
    K=E2(1+v)cotθ0π180 (3)
    M=Δ2θΔsin2ϕ (4)

    where σ is the macroscopic stress, K is the stress coefficient, M is the 2θ–sin2ϕ fitting line slope, E is the elastic modulus of the crystal surface of a mineral crystal, ν is the Poisson’s ratio of the crystal surface of a mineral crystal, θ0 is the diffraction angle without stress, and ϕ is the angle between the normal diffracted crystal plane and the normal sample surface.

    The experiment was conducted using a Bruker D8 Advance X-ray diffractometer equipped with Euler rings and a Cu-target X-ray tube with a print focal spot and one-dimensional detection mode. The working voltage and current were set at 40 kV and 40 mA, respectively, and stress was measured by side inclination. The X-ray generated by the interaction between accelerating electrons and the target passed through a Ni filter to remove Kβ radiation and then collimated to a diameter of 2.0 mm to irradiate the sample surface and obtain the diffraction pattern. The experiment was conducted in the following steps:

    (1) The samples were measured in full spectrum with a scanning step length of 0.1° and a dwell time of 0.15 s for each step to obtain the diffraction pattern within a diffraction angle range of 10°–158°.

    (2) The phase of the obtained diffraction pattern was identified and matched with the PDF card.

    (3) The diffraction peak with the highest intensity and without interference from random peaks was selected, and its corresponding diffraction crystal plane was determined as the characteristic crystal plane for residual stress calibration.

    (4) The scanning range of the diffraction angle was determined on the basis of the width of the diffraction peak. Step scanning was used, with a scanning step length of 0.1° and a dwell time of 0.6 s for each step.

    (5) Stress measurements were performed using side inclination. The angle of ϕ was selected as 0°, 9°, 18°, 27°, 33°, 39°, and 45°.

    The variation in crystal plane spacing in the three directions was measured at the same measurement point, and the diffraction patterns in the directions φ = 0, φ = π/4, and φ = π/2 were measured sequentially.

    The diffraction patterns of the quartz veins are shown in Fig. 8. Phase identification of the test results was conducted on the basis of the XRD spectra obtained from the three measuring points. The results of phase identification indicate that the XRD spectra of the three measuring points matched with the standard PDF No. 99-0088. The (324) crystal plane with the highest diffraction intensity and independence in high-angle diffraction was selected as the stress calibration crystal plane. The scanning range of diffraction angle 2θ ranged from 152° to 156°.

    Figure  8.  XRD patterns of full spectrum measurement of the quartz veins.

    For the experiment, residual stress measurements were conducted by selecting one coordinate axis and another coordinate axis at 45° with two coordinate axes. The direction of the first measured coordinate axis was recorded as the φ = 0 direction, whereas the measured direction with a 45° interval between the two coordinate axes was recorded as the φ = π/4 direction. The direction of the other measured coordinate axis was marked as the φ = π/2 direction, and the magnitude and direction of the principal stress of the residual stress were calculated.

    Fig. 9 shows the diffraction patterns obtained in the φ = 0, φ = π/4, and φ = π/2 directions at measurement points Z-1, Z-2, H-1, and H-2. The diffraction peak measured from 0° to 45° in ϕ shifted gradually to the right.

    Figure  9.  Diffraction pattern of measurement points (a) Z-1, (b) Z-2, (c) H-1, and (d) H-2.

    For each diffraction pattern measured at each measurement point in the directions of φ = 0, π/4, and π/2, Gaussian fitting was performed to determine the peak position (2θ) corresponding to Kα1 radiation. The scattergram was drawn with 2θ as the ordinate, the data after “±” as the error value, and sin2ϕ as the abscissa, and the least square method was used for linear fitting. Fig. 10 illustrates the 2θ–sin2ϕ fitting line of measurement points Z-1, Z-2, H-1, and H-2 on the (324) crystal plane at φ = 0, φ = π/4, and φ = π/2, respectively. The slope of the lines is denoted by M. When M > 0, the rock stress is compressive, and when M < 0, the rock stress is tensile. R2 is the determination coefficient of the linear fitting. When R2 ≥ 0.8, 2θ–sin2ϕ is linearly correlated, indicating that the rock has internal residual stress. As shown in Fig. 10, 2θ increases gradually with the increase of ϕ, indicating the presence of compressive stress inside the rock. The fitting determination coefficient R2 is higher than 0.8, which indicates that 2θ is linearly correlated with sin2ϕ, thereby confirming the presence of residual stress.

    Figure  10.  2θ–sin2ϕ fitting line of measuring points (a) Z-1, (b) Z-2, (c) H-1, and (d) H-2.

    The 2θ–sin2ϕ linear fit of the internal stress in three directions at the measuring points Z-1, Z-2, H-1, and H-2 is remarkable. Using Eq. (2), the residual stress was calculated. The elastic modulus E of the quartz mineral (324) crystal plane was set to 10,300 MPa, and Poisson’s ratio ν was set to 0.31. The radius corresponding to φ = 0 in each direction is presented as θ0. The results of the linear fitting of 2θ–sin2ϕ and the corresponding calculated residual stress for each measuring point are shown in Table 1.

    Table  1.  Calculated results of residual stress
    PointsDirectionφLine fitting parametersMaximum strainStress
    R2MStress value / MPaStress error / MPa
    Z-1Z00.984510.00230 ± 0.00012−0.000123−21.2263 1.1805
    π/40.926090.00149 ± 0.00021−0.000111−13.8468 1.9406
    Yπ/20.841950.00133 ± 0.00026−0.000087−12.2716−2.3700
    Z-2Z00.868190.00151 ± 0.00026−0.000105−13.9286−2.4344
    π/40.983270.00170 ± 0.00011−0.000092−15.6844−1.0225
    Yπ/20.927040.00129 ± 0.00016−0.000079−11.9942−1.5614
    H-1Z00.958240.00170 ± 0.00016−0.000127−15.6934−1.4674
    π/40.830760.00217 ± 0.00044−0.000092−20.0349−4.0491
    Yπ/20.896190.00356 ± 0.00054−0.000079−32.8748−4.9996
    H-2X00.849630.00218 ± 0.00041−0.000147−20.1569−3.8004
    π/40.874500.00207 ± 0.00035−0.000113−18.7669−3.4519
    Zπ/20.935760.00196 ± 0.00023−0.000079−18.1229−2.1764
    下载: 导出CSV 
    | 显示表格

    In determining the surface stress state of the rock samples, the magnitude and direction of the principal stress at the measured points were calculated for the three stress directions. The principal stress of residual stress is denoted as σ1 and σ2. The angle between the direction of σ1 and directions of φ = 0, φ = π/4, and φ = π/2 are presented as Φ, Φ + π/4, and Φ + π/2, respectively. Therefore, the measured stresses in the directions of φ = 0, φ = π/4, and φ = π/2 were recorded as σΦ, σΦ + π/4, and σΦ + π/2, respectively.

    Therefore, the stress values in three directions and the principal stress at each point were obtained in accordance with the principal stress calculation shown in Eqs. (5)–(7), which are described as follows:

    Φ=12arctanσΦ+σΦ+π/22σΦ+π/4σΦσΦ+π/2, (5)
    σ1=σΦ+π/2σΦcot2Φ1cot2Φ, (6)
    σ2=σΦ+π/2σΦtan2Φ1tan2Φ, (7)

    where Φ is the angle between the direction of σ1 and the direction of φ = 0, σΦ is the measured stresses in the direction of φ = 0, σΦ + π/4 is the measured stress in the direction of φ = π/4, σΦ + π/2 is the measured stress in the direction of φ = π/2, σ1 and σ2 are the principal stress of residual stress.

    The calculated results of the residual stress of the test points are shown in Table 2.

    Table  2.  Calculated results of residual principal stress
    PointsPlaneDirectional stress / MPaPrincipal stress / MPaAngle between σ1 and σΦ (Φ) / (°)
    σΦσΦ + π/4σΦ + π/2σ1σ2
    Z-1ZY−21.2263−13.8468−12.2716−22.08−11.41 16.48
    Z-2ZY−13.9286−15.6844−11.9942−15.85−10.07−35.22
    H-1ZY−15.6934−20.0349−32.8748−14.70−33.87−13.16
    H-2XZ−20.1569−18.7669−18.1229−20.23−18.05 10.50
    下载: 导出CSV 
    | 显示表格

    Fig. 11 illustrates the distribution of the residual stress on the rock sample surface. The principal stresses within the quartz vein are all compressive, with stress values ranging from 10 to 30 MPa at the measuring points. The principal stress along the transverse and longitudinal veins parallel to their respective dikes was significantly higher than that perpendicular to their respective dikes. Moreover, the principal stress in the transverse quartz vein, which is roughly parallel to the bedding direction, is greater than the stress locked within the longitudinal quartz vein, which is roughly perpendicular to the bedding direction.

    Figure  11.  Schematic diagram of the principal stress direction and relationship of measuring points: (a) longitudinal vein Z-1; (b) longitudinal vein Z-2; (c) transverse vein H-1; (d) transverse vein H-2.

    During the formation of the quartz vein, SiO2 evaporated from magmatic rock, and water vapor filled the fractures in the matrix. The fractures initially attached to the matrix wall, then coagulated and grew continuously with the quartz, and finally filled the cracks in the bedrock. In places where the raw materials were abundant, they were mutually constrained with siliceous slate to achieve balance, and the internal pressure was locked in. The siliceous slate is a metamorphic rock formed after compression and deformation of the protolith. As the maximum stress of the longitudinal and transverse veins is parallel to the direction of the dike, the metamorphic transformation of the siliceous slate might occur prior to the formation of the quartz vein. Residual stress measurements provide valuable information regarding rock formation.

    The EBSD experiment data were processed to obtain the grain boundary diagram, kernel average misorientation (KAM) diagram, and geometrically necessary dislocation (GND) diagram to analyze the mechanism for residual stress generation in rocks.

    High-angle grain boundaries are defined as having an orientation difference of more than 15°, whereas low-angle grain boundaries are defined as having an orientation difference between 2° and 15°. As shown in Fig. 12, the grain boundary diagram of the quartz vein test area is composed of high- (51.9vol%) and low-angle grain boundaries (48.1vol%). The defects formed by dislocation, arrangement, and combination are concentrated in the low-angle boundaries. The content of the low-angle grain boundaries is more prone to recovery and recrystallization in quartz veins.

    Figure  12.  Grain boundary map of the quartz vein area.

    The maximum KAM value of the quartz pulse is 4.52°, whereas the maximum GND value is 5.36 cm−2. Combined with the grain boundary map observation, the region with a higher geometric dislocation density is located at the low-angle grain boundaries, indicating that stress concentration likely occurs at low-angle grain boundaries of mineral crystals.

    In addition to the low- and high-angle grain boundaries, the quartz vein test area contains twin boundaries of the <001> 60° type, which account for 40.7vol% of the total area. When twins form in the crystal, the material experiences internal stress because the natural change in the shape of the twin zone is limited by the surrounding environment [32]. The <001> 60° twins in α-quartz are Dauphiné twins, which rotate 180° (60° in crystallography) around the C-axis, resulting in the reversal of the positive and negative forms of the crystal structure [3334]. Twins can switch between the positive and negative <a> directions [35]. During plastic deformation, Dauphiné twins create the main active slip system, and their impact on local strain strongly depends on the deformation time relative to the movement of the evolutionary shear band. Research has shown that Dauphiné twins have a strong ability to adapt to strain [34].

    While twin boundaries have resistance to large deformation, low-angle boundaries easily induce strain concentration in quartz veins. The mineral crystals inside the rocks deform with pressure or temperature changes in the Earth crust. The ability of mineral crystals to resist deformation differs, and grains squeeze one another because of different deformation states. This process produces dislocation and plastic deformation as well as changes the internal mechanical properties of the rocks. When the temperature changes or when an external force is applied, the internal dislocation and grain boundary hinder the deformation of the crystal, and the stress is locked inside. Consequently, the external state remains balanced and stable.

    In this work, EBSD and optical microscopy were performed to characterize the rock microcosm and explained rock residual stress.

    (1) The siliceous slate-containing quartz veins in the Hongling Mine were primarily composed of feldspar, quartz, and small quantities of chlorite. The statistics generated were sufficient to meet the requirements of XRD tests. Moreover, a homogeneous region in the millimeter scale was identified, which met the threshold required for residual stress measurement using XRD.

    (2) The results showed that the residual stresses measured in rocks were all compressive stresses, with magnitudes ranging from 10 to 33 MPa. The residual stress in transverse quartz veins was greater than that in longitudinal quartz veins. The maximum stress in the vein was roughly parallel to the direction of the vein, and the minimum stress was roughly perpendicular to the direction of the vein. The main stress direction was closely related to the vein. Therefore, the formation and development of residual stress were closely related to the rock microcosm. Thus, studying residual stress from the microscopic perspective of rocks is of great significance.

    (3) Plastic strain analysis of the rock microzone was performed by examining the difference in mineral crystal orientation. The mechanism for generating rock residual stress was analyzed by examining the grain boundaries, orientation difference, and dislocation density obtained from grain boundary maps, KAM maps, and geometric dislocation density maps. The results showed that several low-angle boundaries that are prone to strain concentration and twin boundaries could adapt to large strains in the quartz veins. Furthermore, differences in the ability of mineral crystals in rocks to resist deformation were observed. Temperature changes or the application of external forces resulted in incompatible deformation, and the stress was “locked inside” because of internal dislocations and grain boundaries.

    This work was funded by the National Natural Science Foundation of China (Nos. 51874014, 52004015, and 52311530070), the fellowship of China National Postdoctoral Program for Innovative Talents (No. BX2021033), the fellowship of China Postdoctoral Science Foundation (No. 2021M700389), the Fundamental Research Funds for the Central Universities of China (Nos. FRF-IDRY-20-003 and QNXM20210001), and State Key Laboratory of Strata Intelligent Control and Green Mining Co-founded by Shandong Province and the Ministry of Science and Technology, China (No. SICGM202108). These supports are gratefully acknowledged.

  • [1]

    L.Y. Liu, H.G. Ji, X.F. Lü, et al., Mitigation of greenhouse gases released from mining activities: A review, Int. J. Miner. Metall. Mater., 28(2021), No. 4, p. 513. DOI: 10.1007/s12613-020-2155-4

    [2]

    G.R. Holzhausen and A.M. Johnson, The concept of residual stress in rock, Tectonophysics, 58(1979), No. 3-4, p. 237. DOI: 10.1016/0040-1951(79)90311-1

    [3]

    M. Friedman, Residual elastic strain in rocks, Tectonophysics, 15(1972), No. 4, p. 297. DOI: 10.1016/0040-1951(72)90093-5

    [4]

    T.T. Kie, Rockbursts, case records, theory and control, [in] Proceedings of the International Symposium on Engineering in Complex Rock Formations, Pergamon, 1988, p. 32.

    [5]

    S.H. Tang, Z.J. Wu, and X.H. Chen, Approach to occurrence and mechanism of rockburst in deep underground mines, Chin. J. Rock Mech. Eng., 22(2003), No. 8, p. 1250.

    [6]

    H. Zhou, F.Z. Meng, C.Q. Zhang, D.W. Hu, F.J. Yang, and J.J. Lu, Analysis of rockburst mechanisms induced by structural planes in deep tunnels, Bull. Eng. Geol. Environ., 74(2015), No. 4, p. 1435. DOI: 10.1007/s10064-014-0696-3

    [7]

    D.I. Varnes and F.T. Lee, Hypothesis of mobilization of residual stress in rock, Geol. Soc. Am. Bull., 83(1972), No. 9, art. No. 2863. DOI: 10.1130/0016-7606(1972)83[2863:HOMORS]2.0.CO;2

    [8]

    L. Müller, Rock Mechanics, Springer-Verlag, Berlin, 1974.

    [9]

    T.K. Tan and W.F. Kang, Locked in stresses, creep and dilatancy of rocks, and constitutive equations, Rock Mech., 13(1980), No. 1, p. 5. DOI: 10.1007/BF01257895

    [10]

    Z.Q. Yue, Expansion power of compressed micro fluid inclusions as the cause of rockburst, Mech. Eng., 37(2015), No. 3, p. 287.

    [11]

    O. An, Forecast method for risk area and risk time of large earthquake in the Xianshuihe Fault Zone by superimposing the residual and present stress field, Bull. Inst. Crustal Dynamics, 1(1996), No. 0, p. 59.

    [12]

    X. Liu, H.S. Geng, H.F. Xu, Y.H. Yang, L.J. Ma, and L. Dong, Experimental study on the influence of locked-in stress on the uniaxial compressive strength and elastic modulus of rocks, Sci. Rep., 10(2020), No. 1, art. No. 17441. DOI: 10.1038/s41598-020-74556-1

    [13]

    W.C. Chen, S.J. Wang, and H.R. Fu, Study advance on basic characteristics and formation causes of rock inner stress, J. Eng. Geol., 26(2018), No. 1, p. 62.

    [14]

    K. Sekine and K. Hayashi, Residual stress measurements on a quartz vein: A constraint on paleostress magnitude, J. Geophys. Res., 114(2009), No. B1.

    [15]

    K. Chen, M. Kunz, N. Tamura, and H.R. Wenk, Residual stress preserved in quartz from the San Andreas Fault Observatory at Depth, Geology, 43(2015), No. 3, p. 219. DOI: 10.1130/G36443.1

    [16]

    H.R. Wenk, B.C. Chandler, K. Chen, Y. Li, N. Tamura, and R. Yu, Residual lattice strain in quartzites as a potential palaeo-piezometer, Geophys. J. Int., 222(2020), No. 1, p. 1363.

    [17]

    R. Weinberger, Y. Eyal, and N. Mortimer, Formation of systematic joints in metamorphic rocks due to release of residual elastic strain energy, Otago Schist, New Zealand, J. Struct. Geol., 32(2010), No. 3, p. 288. DOI: 10.1016/j.jsg.2009.12.003

    [18]

    L.Y. Liu, H.G. Ji, D. Elsworth, S. Zhi, X.F. Lv, and T. Wang, Dual-damage constitutive model to define thermal damage in rock, Int. J. Rock Mech. Min. Sci., 126(2020), art. No. 104185. DOI: 10.1016/j.ijrmms.2019.104185

    [19]

    C. Lu, Y. Lu, X.H. Gou, Y. Zhong, C. Chen, and J.C. Guo, Influence factors of unpropped fracture conductivity of shale, Energy Sci. Eng., 8(2020), No. 6, p. 2024. DOI: 10.1002/ese3.645

    [20]

    D.W. Gentry, Horizontal residual stresses in the vicinity of a Breccia Pipe, Int. J. Rock Mech. Min. Sci. Geomech. Abstr., 10(1973), No. 1, p. 19. DOI: 10.1016/0148-9062(73)90057-0

    [21]

    A. De Noni, D. Hotza, V.C. Soler, and E.S. Vilches, Analysis of the development of microscopic residual stresses on quartz particles in porcelain tile, J. Eur. Ceram. Soc., 28(2008), No. 14, p. 2629. DOI: 10.1016/j.jeurceramsoc.2008.04.009

    [22]

    S.R. Kiahosseini and H. Ahmadian, Effect of residual structural strain caused by the addition of Co3O4 nanoparticles on the structural, hardness and magnetic properties of an Al/Co3O4 nanocomposite produced by powder metallurgy, Int. J. Miner. Metall. Mater., 27(2020), No. 3, p. 384. DOI: 10.1007/s12613-019-1917-3

    [23]

    T. Sarkar, A.K. Pramanick, T.K. Pal, and A.K. Pramanick, Development of a new coated electrode with low nickel content for welding ductile iron and its response to austempering, Int. J. Miner. Metall. Mater., 25(2018), No. 9, p. 1090. DOI: 10.1007/s12613-018-1660-1

    [24]

    X.M. Yuan, J. Zhang, Y. Lian, et al., Research progress of residual stress determination in magnesium alloys, J. Magnesium Alloys, 6(2018), No. 3, p. 238. DOI: 10.1016/j.jma.2018.06.002

    [25]

    Z.H. Tang, X. Dong, Y.X. Geng, et al., The effect of warm laser shock peening on the thermal stability of compressive residual stress and the hot corrosion resistance of Ni-based single-crystal superalloy, Opt. Laser Technol., 146(2022), art. No. 107556. DOI: 10.1016/j.optlastec.2021.107556

    [26]

    M. Friedman, X-ray analysis of residual elastic strain in quartzose rocks, [in] The 10th U.S. Symposium on Rock Mechanics, Texas, 1968, p.68.

    [27]

    A. Frischbutter, D. Neov, C. Scheffzük, M. Vrána, and K. Walther, Lattice strain measurements on sandstones under load using neutron diffraction, J. Struct. Geol., 22(2000), No. 11-12, p. 1587. DOI: 10.1016/S0191-8141(00)00110-3

    [28]

    W.C. Chen, S.D. Li, X. Li, S.J. Wang, L.H. He, and S.M. Li, A novel method for determination of rock inner stress based on X-ray and neutron scaterring, J. Eng. Geol., 30(2022), No. 1, p. 223.

    [29]

    C.Y. Liu, H.L. Luo, H.J. Li, and X.X. Zhang, Formation mechanism and control technology of vein rockburst - A case study of Uzbekistan Kamchik tunnel, Rock Soil Mech., 42(2021), No. 5, p. 1413.

    [30]

    W.L. Bragg, The structure of some crystals as indicated by their diffraction of X-rays, Proc. R. Soc. London Ser. A, 89(1913), No. 610, p. 248. DOI: 10.1098/rspa.1913.0083

    [31]

    I.C. Noyan and J.B. Cohen, Residual Stress: Measurement by Diffraction and Interpretation, Springer-Verlag, Berlin, 1987.

    [32]

    C. Palache, J.D. Dana, E.S. Dana, H. Berman, and C. Frondel, The System of Mineralogy of James Dwight Dana and Edward Salisbury Dana, John Wiley and Sons, New York, 1944.

    [33]

    C. Frondel, Secondary Dauphiné twinning in quartz produced by sawing, Am. Mineral., 31(1946), No. 1-2, p. 58.

    [34]

    C. McGinn, E.A. Miranda, and L.J. Hufford, The effects of quartz Dauphiné twinning on strain localization in a mid-crustal shear zone, J. Struct. Geol., 134(2020), art. No. 103980.

    [35]

    G.E. Lloyd, Microstructural evolution in a mylonitic quartz simple shear zone: The significant roles of dauphine twinning and misorientation, Flow Process. Faults Shear Zones, 224(2004), p. 39.

Relative Articles

黄煜博, 徐宁, 卢怀乐, 任洋, 李时磊, 王沿东. 基于同步辐射X射线与中子衍射技术的高熵合金微观组织与微观变形行为研究综述 [J]. 矿物冶金与材料学报(英文版). DOI: 10.1007/s12613-024-2840-9

View details

Elisa Lam, Qinfen Gu, Peter J. Swedlund, Sylvie Marchesseau, Yacine Hemar. X-ray diffraction investigation of amorphous calcium phosphate and hydroxyapatite under ultra-high hydrostatic pressure [J]. 矿物冶金与材料学报(英文版). DOI: 10.1007/s12613-015-1189-5

View details

Int. J. Miner. Metall. Mater., 2010, 17(6): 777-781.

PDF View details

Chun Feng, Jing-yan Zhang, Jiao Teng, Fu-ming Wang. Study on NiO/Fe interface with X-ray photoelectron spectroscopy [J]. 矿物冶金与材料学报(英文版). DOI: 10.1007/s12613-010-0388-3

View details

Hejie Li, Jingtao Han, Huachun Pi, Zhengyi Jiang, Dongbin Wei, A. Kiet Tieu. Modeling texture development during cold rolling of IF steel by crystal plasticity finite element method [J]. 矿物冶金与材料学报(英文版). DOI: 10.1016/S1005-8850(08)60273-X

View details

Int. J. Miner. Metall. Mater., 2008, 15(4): 484-488.

PDF View details

Kezhi Li, Jian Wei, Hejun Li, Chuang Wang, Gengsheng Jiao. Silicon assistant carbothermal reduction for SiC powders [J]. 矿物冶金与材料学报(英文版). DOI: 10.1016/S1005-8850(08)60091-2

View details

Huachun Pi, Jingtao Han, Chuanguo Zhang, A. Kiet Tieu, Zhengyi Jiang. Modeling uniaxial tensile deformation of polycrystalline Al using CPFEM [J]. 矿物冶金与材料学报(英文版). DOI: 10.1016/S1005-8850(08)60009-2

View details

Int. J. Miner. Metall. Mater., 2004, 11(2): 161-164.

PDF View details

Jun Luo, Qijie Zhai, Pei Zhao, Xubo Qin. Liquid structure of pure iron by X-ray diffraction [J]. 矿物冶金与材料学报(英文版).

View details

Wenhuai Tian, Hongye Gao, Ailing Fan, Xiaoou Shan, Qi Sun. Dynamic recrystallization of electroformed copper liners of shaped charges in high-strain-rate plastic deformation [J]. 矿物冶金与材料学报(英文版).

View details

Jing Shi, Xianjin Wang. Annealing Texture Evolution of Ti-IF Steel during Batch Annealing [J]. 矿物冶金与材料学报(英文版).

View details

Citing articles(8)

Liyuan Liu, Juan Jin, Jiandong Liu, et al. Mechanical properties of sandstone under in-situ high-temperature and confinement conditions. International Journal of Minerals, Metallurgy and Materials, 2025, 32(4): 778. 必应学术
Peng Li, Meifeng Cai, Shengjun Miao, et al. Correlation between the rock mass properties and maximum horizontal stress: A case study of overcoring stress measurements. International Journal of Minerals, Metallurgy and Materials, 2025, 32(1): 39. 必应学术
Tao Wang, Weiwei Ye, Liyuan Liu, et al. Denoising of acoustic emission signals from rock failure processes through ICEEMDAN combined with multiple criteria and wavelet transform. Discover Applied Sciences, 2025, 7(4) 必应学术
More >

/

返回文章
返回