Polycrystalline Cloth

In a typical polycrystalline material with ∼100–1000μm grains, private grains will contain ∼105 magnetic domains whose moments are randomly oriented in the virgin country.

From: Concrete Metallurgy (5th Edition) , 2014

Diffusion in Polycrystalline Materials

John C. Mauro , in Materials Kinetics, 2021

Abstract

Polycrystalline materials have a spectrum of defects of varying dimensionality, including zero-dimensional signal defects (vacancies and interstitials), i-dimensional line defects (dislocations), two-dimensional surface defects (grain boundaries and free surfaces), and three-dimensional volume defects (pores). Higher-dimensional defects such as dislocations and surfaces lead to "short-circuit" pathways for dramatically accelerated improvidence. Hence, the diffusivity of a polycrystalline textile depends strongly on the concentrations of such defects. The Harrison "ABC" model provides a description of three diffusion regimes in a polycrystalline material, depending on the relative diffusivity through the grains versus along the grain boundaries. Depending on the regime, the diffusion in polycrystalline materials can be treated using either a single diffusion equation with an constructive diffusivity or a coupled set of diffusion equations on different time scales. While defects act to enhance diffusivity, there is no universal atomistic mechanism enabling this fast diffusion. The details of the enhancement in diffusivity depend on the construction and microstructure of the material in question.

Read full chapter

URL:

https://world wide web.sciencedirect.com/science/article/pii/B9780128239070000194

Strengthening of metal alloys

In Introduction to Aerospace Materials, 2012

iv.iv.3 Grain boundary strengthening

Polycrystalline materials are equanimous of a large number of grains. As mentioned, the lattice arrangement of atoms within each grain is well-nigh identical, just the orientation of the atoms is dissimilar for each bordering grain. The surface that separates neighbouring grains is the grain boundary ( Fig. 4.12). Grain boundaries impede the movement of dislocations and thereby have a strengthening effect.

The process of grain boundary strengthening tin can be explained past the following sequence of events. Dislocations move through a crystal lattice until they accomplish a grain boundary. The mismatch in the lattice orientation at the purlieus betwixt two grains disrupts the slip plane of the dislocation. The boundary besides creates a repulsive strain field that opposes the slip motion of the dislocation. The dislocation is forced to finish just ahead of the boundary. As more dislocations movement to the purlieus a process of 'pile-up' occurs as a growing cluster of dislocations are unable to move past the boundary (Fig. 4.18). Dislocations generate their own repulsive strain field and, with each successive pile-up of a dislocation at the grain boundary, the repulsive force acting on the dislocation nearest the boundary rises. Eventually, the repulsive stress is high plenty to force the dislocation closest to the boundary to cantankerous over to the adjoining grain. This procedure of restricting dislocation movement across grain boundaries is the ground of the strengthening mechanism.

4.18. Dislocation pile-up at points A, B and C along a grain boundary.

(from Yard. Zlateva and Z. Martinova, Microstructure of metals and alloys, CRC Press, Boca Raton, 2008)

The strength of polycrystalline materials is increased past reducing their grain size. The grain size of metals is oftentimes controlled by thermomechanical treatment, rut treatment or microalloying. Grains in aerospace alloys typically range from about ane   μm (coarse or big grains) to 1   μm (fine or pocket-size grains), and the strength increases as the grains become smaller. For case, Fig. four.19 shows the outcome of average grain size on the yield strength of steel. Reducing the grain size also has the additional beneficial effects of increasing the ductility, fracture toughness and fatigue life. Forcefulness increases because the altitude that dislocations must travel through a grain core to accomplish a grain purlieus decreases with the grain size. Dislocations are more than probable to reach a boundary, and thereby strengthen the metal, when the grain size is reduced.

4.19. Effect of grain size on the yield force of steel.

The grain size is related to the yield force according to the Hall–Petch relationship:

σ y = σ o + k d 1 / 2

where σy is the yield strength and d is the boilerplate diameter of the grains; g is a material abiding representing the gradient of the σy d  1/2 plot; southwardo is called the friction stress, and is the intercept on the stress axis (Fig. four.19). σ o is a material abiding that defines the stress required to move dislocations in a single crystal without a grain boundary (d   1/2  =   0). The Hall–Petch relationship is used by engineers to summate the grain size necessary to achieve a required level of forcefulness. The Hall–Petch human relationship is authentic for metals with a grain size betwixt about one   μm and 1   mm, simply is not valid for materials with larger (>   1   mm) or finer (<   1   μm) grains.

Theoretically, metals can be made infinitely potent if their grains are infinitely pocket-size. In reality, this is impossible because the lower limit on grain size is a single unit prison cell of the crystal. The finest grain size in about aerospace alloys is about i   μm. Smaller than this, the length of the dislocation approaches the size of the grain.

The grain size of aerospace metals is reduced to the size range of about one to several hundred micrometres using various techniques, including rapid cooling of the molten metal during solidification casting, improver of grain-refining elements (called inoculants), and thermomechanical processing. The command of grain size is explained in more detail in Chapters 6 and 7.

Read full chapter

URL:

https://www.sciencedirect.com/science/article/pii/B9781855739468500042

Neutron Scattering - Applications in Biology, Chemistry, and Materials Science

Wanchuck Woo , ... Xun-Li Wang , in Experimental Methods in the Physical Sciences, 2017

12.3.2.i Deformation in Fibroid Grained Materials

Polycrystalline materials consist of grains with various orientations. Nether practical load, the lattice strain response inside a grain, which tin be readily measured by neutron diffraction, depends on the grain orientation. In the elastic authorities, the lattice strain is linear and scales with the rubberband anisotropy. When plastic deformation occurs, yet, the lattice strain starts to deviate from the linear elastic response. The additional rubberband strain that develops in the grains to accommodate inhomogeneous plastic deformation during the macroscopic elasto-plastic transition, is termed the intergranular or type II strain, ɛ Two hkl , where the superscript denotes grains with (hkl) planes normal to the management of measurement (specified by the scattering vector) [103].

The presence of intergranular strains has long been recognized. In 1947, Greenough [104,105] examined the residue lattice strains of polycrystalline ferritic steel afterward plastic tensile deformation, and pointed out the effect of plastic anisotropy. In 1987, Pitschovious et al. [106] studied the intergranular strains in their study of cold-rolled ferric steel. By using a triple-axis spectrometer to eliminate the event of centering [107], the authors were able to determine the intergranular strains with high precision. Meanwhile, in a study of steam generator tubing, Holden et al. [nine] found at some locations, the axial strains measured with {111} and {200} reflections are of opposite signs, which cannot be explained at all by elastic anisotropy alone.

Examinations of intergranular strains have led to the development of a new subfield, i.e., fundamental agreement of the deformation mechanisms in polycrystalline materials at the microscopic level. For this purpose, uniaxial loading devices were built to fit onto neutron diffractometers and the elastic lattice strains for different reflections were recorded every bit a office of practical stress. Fig. 8 illustrates the evolution of lattice strain in a fibroid-grained fcc metals [108]. Nether tensile deformation, tensile and compressive intergranular strains develop in ‹200›//LD and ‹220›//LD grains, respectively [103], where LD designates the loading management. This is a direct result of the crystal rubberband and plastic anisotropy [109–111]. The ‹111›//LD grains, on the other hand, show a minimal intergranular strain. Meanwhile, a feature texture also develops during deformation as grains rotate toward ‹111›//LD and to a lesser extent toward ‹200›//LD while the population of ‹220›//LD grains is depleted [103,112].

Figure viii. Evolution of lattice strain in coarse grained stainless steel. The deviation from the linear human relationship indicates the presence of intergranualar strain. The lines are calculations from elasto-plastic self consistent model.

Figure reproduced with permission from Ref. [108].

A major breakthrough in understanding the intergranular strains came from the advances in elastic-plastic self-consistent (EPSC) modeling [110,113]. However, the essential physics tin can be explained within the framework of the Taylor model. By considering an ensemble of grains, Holden et al. [114] demonstrated that the ‹111›//LD grains (i.e., all grains with (111) plane parallel to LD) yielded first, despite the fact that the ‹111›   management has the highest yield strength. This unusual behavior is a result of a combination of elastic and plastic anisotropy. For stainless steel, the ‹111›   management is the stiffest elastically. Every bit a result, in a polycrystalline environment, ‹111›//LD grains conduct the highest load, causing ‹111›//LD grains to accomplish the yield force faster than other grains. Similarly, ‹100›//LD grains yield concluding (see Fig. 8), despite having the lowest yield strength, because the ‹100›   is the softest direction. Unloading from the maximum tensile load thus leaves a tensile residual strain for ‹100›.

Signature intergranular strains develop due to the activation of select slip planes. For this reason, intergranular strains have been used every bit a fingerprint to gain insight in to elastic and plastic anisotropy in polycrystalline metals. For case, by measuring the intergranular strains and using polycrystal plasticity modeling to interpret the results, Wollmershauser et al. [115] demonstrated that the intergranular strain development of CeAg is nearly identical to that of NiAl, indicating that they share a common main plastic deformation mechanism, in this case the ‹100›{011} "cube" slip.

Read full chapter

URL:

https://www.sciencedirect.com/science/article/pii/B9780128053249000121

Cyclic Deformation, Scissure Initiation, and Low-Bicycle Fatigue

J. Polák , in Reference Module in Materials Science and Materials Engineering, 2016

3.ii Surface Relief at Emerging PSBs in Polycrystals

Polycrystalline materials are equanimous of individual grains oriented in a random manner to the loading axis when no texture is nowadays in the material. The size of the grains is commonly much smaller than that of unmarried crystalline specimens, which results in surface relief that is less pronounced than in single crystals. The surface relief development in polycrystals has been studied for decades using optical microscopy; the replica technique combined with transmission electron microscopy ( Hempel, 1956; Thompson and Wadsworth, 1958); SEM (Wang et al., 1982; Polák, 1991; Bayerlin and Mughrabi, 1991; Polák and Kruml, 1998; Polák et al., 2009); and AFM (Man et al., 2002; Cretegny and Saxena, 2001; Villechaise et al., 2002; Man et al., 2003, 2004, 2009b, 2012, 2015).

PSMs in polycrystals are like to individual PSMs in single crystals, consisting of extrusions and intrusions. Macrobands have not been observed. Unremarkably extrusions are more frequent and more than adult as intrusions. Figure 20(a) (Polák and Kruml, 1998) shows a grain of 316L steel cycled to 20% of the fatigue life (N f=2×ten5 cycles). Parallel PSMs run throughout the grain and carry on into the neighboring grains. The high magnification micrographs in Figure xx show details of the surface relief. In Figure 20(b) the thin PSM consists of alternating extrusions and intrusions, similar to the case of copper single crystals. Effigy 20(c) shows the section of a developed PSM later tilting the specimen, so that the electron beam became parallel to the slip aeroplane. An extrusion of a complicated shape is accompanied by two shallow intrusions at the interface with the matrix.

Figure xx. Surface of the grain of 316L steel (SEM) cycled with ɛap=v×x−4: (a) overall view; (b) detail of the PSM with alternating extrusions and intrusions; (c) section of the PSM inclined and then that primary slip plane of the PSB became parallel to the electron beam.

Man et al. (2009a) summarized number of experimental observations of extrusions and intrusions in fatigued materials and presented the most typical forms of PSMs. They concluded that in majority of unproblematic metals the typical well adult PSM consist of extrusion accompanied by parallel intrusions on both sides of the extrusion.

Considerable variation exists in the surface relief of PSMs in private materials and specimens, and even for different grains in one specimen. Figure 21 shows 1 end of the PSM on the etched surface of a large-grain of a cast polycrystalline INCONEL 713LC superalloy (Obrtlík et al., 2002). Both the matrix and γ′ precipitates are cut by a thin PSM. The surface relief of the PSM is formed by very sparse extrusion.

Figure 21. PSM on the etched surface of INCONEL 713LC fatigued with ɛa=6×10−iii (SEM).

Quantitative data on the surface relief of fatigued materials can be obtained using AFM, provided the finite dimension of the tip does non innovate distortion. Figure 22(a) shows an AFM micrograph of an expanse of the surface grain of drawn polycrystalline copper, cycled at room temperature with plastic strain amplitude of 5×x−iv to xx% of the fatigue life. The PSMs are formed by extrusions and accompanying intrusions. When the summit of an extrusion and the depth of an intrusion are small, the true profiles could be assessed without distortion using AFM. The profiles of extrusions and intrusions along the two lines perpendicular to the direction of PSM are shown in Figure 22(b). The shape of the cross-section of both extrusion and intrusion is approximately triangular, and the slopes of the intrusion and extrusion sides are equal.

Effigy 22. (a) AFM micrograph of the grain of drawn polycrystalline copper (ɛap=five×10−4 for 2.5×x4 cycles); and (b) two surface profiles forth the lines shown in (a).

Figure 23(a) shows the characteristic relief on the grain of a large-grain INCONEL 713LC cycled at a high strain aamplitude (εa=i.two×ten−2) until fracture (N f=85). The surface profile along a line perpendicular to the management of PSMs is shown in Effigy 23(b). The majority of PSMs are formed by extrusions, merely parallel intrusions are likewise nowadays. The peak of the extrusions is ~0.2   µm.

Effigy 23. (a) AFM micrograph of the grain of fatigued INCONEL 713LC; and (b) surface profile along the line shown in (a).

However higher extrusions and deeper intrusions as imaged by AFM are commonly distorted (Polák et al., 2003; Human et al., 2003). Not only the shapes of both extrusions and intrusions could exist distorted but an extrusion egressing from a crystal at 45° can cover the parallel intrusion. Therefore both directly observation of the metallic surface and plastic replicas are used to reveal a true shape of PSMs using AFM.

Human et al. (2002, 2003, 2012, 2015), using SEM and AFM, systematically studied the evolution of the surface relief in in 316L stainless steel fatigued at a constant plastic strain amplitude of 2×10−3. Figure 24(a) shows the sequence of the surface profiles during the fatigue life every bit observed by direct ascertainment metal surface. At the cease of the fatigue life as well the surface profile using plastic is shown. Here the intrusions are also detected. Figure 24(b) shows the evolution of the extrusion height with the number of loading cycles. The height of the extrusion at the end of the fatigue life was found to be proportional to the thickness of the PSB, while the extrusion width quickly reached a stable value that was non inverse significantly past further circadian straining. This is in agreement with observations on fatigued copper single crystals, where the protrusions grow in the direction of the master Burgers vector and the height of the protrusion in a item location is approximately proportional to the thickness of the specimen in that management. Most extrusions could be identified using both the SEM and AFM techniques. Since the orientations of all grains could be determined using EBSD, the principal slip plane and slip vector were assessed (Homo et al., 2002). From the surface markings, the thickness t and the peak h b in the management of the agile Burgers vector of the PSBs were evaluated.

Figure 24. Development of the surface relief of a grain of 316L steel during cycling with ɛap=2×x−3: (a) surface profile and (b) extrusion height (Human et al., 2003).

Temperature has an of import effect on the growth of extrusions (Homo et al., 2015). The average extrusion height in fatigued 316L steel versus number of loading cycles is shown in Figure 25. At depression temperature (93   M) simply initial growth is detected and later extrusion height saturates. This behavior corresponds to the growth of static extrusion. At temperature approximately parabolic growth is observed and in cycling at 573   K rapid initial growth of intrusions was found. These finding correspond to the dynamic growth of extrusions due to migration of point defects.

Effigy 25. Average extrusion height versus number of cycles in 316L steel cycled at three temperatures with plastic strain amplitude one×10−3 (Homo et al., 2015).

The study of intrusions using AFM is more hard than that of extrusions owing to the finite geometry of the AFM tip (Polák et al., 2003). The only possibility for obtaining the true depth of an intrusion is to employ replica techniques. The AFM image of the metallic surface in Figure 26(a) is compared with the AFM image of the same location using a plastic replica in Figure 26(b). Effigy 26(c) compares the profiles obtained from the observations of the metallic surface and the plastic replica. Direct observation of the metal surface yields the truthful height of an extrusion, but the width and the shape can be distorted. Intrusions are not imaged. The apply of the plastic replica allows the truthful depth of an intrusion to exist obtained, but its width and shape can exist distorted. Extrusions are not imaged accurately using a replica.

Effigy 26. Comparison of the AFM images of the aforementioned location on the surface of fatigued 316L steel: (a) metallic surface, (b) plastic replica, and (c) profiles of extrusions (E) and intrusions (I).

Early on stages of the surface relief development in a grain of 316L steel in constant plastic strain amplitude cycling using plastic replica technique which allows detecting also intrusions is shown in Figure 27. Initially (N<0.8%N f) but extrusions are detected, after (N>0.viii%Northward f) also intrusions accompany the extrusions. Intrusions are very thin and arise initially on 1 side and later on on both sides of an extrusion (Man et al., 2009b).

Figure 27. Early stages of surface relief evolution in a grain of 316L steel cycled with ɛab=1×ten−3 every bit obtained by AFM using plastic replicas. (a) Due north=350 cycles (0.8% N f), (b) Due north=500 cycles (1.1% N f), (c) N=750 cycles (1.6% Due north f), (d) N=2000 cycles (4.iii% N f).

The kinetics of extrusion and intrusion growth in constant plastic strain amplitude cycling of 316L steel shows Figure 28. The extrusion height and intrusion depth is plotted versus number of cycles in two grains of 316L steel. In both cases the start of intrusion growth is significantly delayed relative to the start of extrusion growth. The intrusion growth rate is in bulk of cases higher than the growth charge per unit of the parallel extrusion. Typical extrusion and intrusion growth rates are in the interval twob−10b but rates smaller than b are quite mutual.

Figure 28. Simultaneous growth of extrusions and intrusions in the grains of 316L steel at the beginning of cycling with ɛap=one×10−iii. (a) Extrusion and 1 parallel intrusion, (b) an extrusion and two parallel intrusions.

Read full chapter

URL:

https://world wide web.sciencedirect.com/science/article/pii/B9780128035818008900

Ultrasonic techniques for materials label

G. Hübschen , in Materials Characterization Using Nondestructive Evaluation (NDE) Methods, 2016

7.2.4 Scattering

In polycrystalline material ultrasonic grain boundary scattering exists due to the grain structure and the crystallographic orientation inside the single grains, which causes acoustical impedance mismatches at the grain boundaries. For quasi-isotropic polycrystalline materials three handful regions be ( Bhatia, 1967; Bhatia and Moore, 1959; Goebbels, 1980), which tin be distinguished by unlike functionalities between the scattering coefficient α S and the mean value of the grain diameter d and the ultrasonic wavelength λ:

[7.20] Rayleigh scattering, d λ α S = S 1 d three f 4

[7.21] Stochastic scattering, d λ α S = South 2 d f 2

[7.22] Diffuse scattering, d λ α S = S 3 / d

For the Rayleigh handful on grains the interaction of the ultrasonic wave takes place with a (theoretical globular) obstruction, which is minor in comparison to the ultrasonic wavelength. Hereby longitudinal and shear waves are generated by mode conversion.

In the example of stochastic scattering the grain diameter and the ultrasonic wavelength are in the same dimension so that resonance miracle can occur (generation of standing waves in the single grains) and therefore as well with quantitatively larger scattering losses compared to the two other handful mechanisms.

In current nondestructive testing (NDT) applications mainly the Rayleigh scattering case plays an important part. Here the scattering is proportional in a first approximation to the volume of the scattering obstacle and to the 4th ability of the frequency f. For longitudinal and shear waves the scattering parameter S1 is given by the known properties: ρ, mass density; 5 L, velocity of longitudinal wave; v T, velocity of shear wave and the elastic anisotropy factor A  = c 11  c 12    iic 44:

[7.23] South 1 Fifty = ( 8 π 3 / 375 ) ( A / ρ five L ii ) 2 ( 1 / v L four ) ( 2 + iii ( 5 L / five T ) 5 )

[7.24] S ane T = ( six π 3 / 375 ) ( A / ρ v T 2 ) two ( 1 / v T 4 ) ( 3 + 2 ( five T / 5 L ) 5 )

During the scattering procedure mode conversions occur. For an incident longitudinal wave, scattered longitudinal and shear waves are generated and also an incident shear wave is scattered in both wave types. The terms in the Eqs. [7.23] and [seven.24] including (v L/5 T) and (v T/five L) represent the mode conversion contribution.

Read full affiliate

URL:

https://www.sciencedirect.com/science/article/pii/B9780081000403000079

Irradiation Growth

Malcolm Griffiths , in Comprehensive Nuclear Materials (Second Edition), 2020

1.11.five.four.3 Effect of texture and grain structure

A polycrystalline textile consists of an bunch of private crystallites (grains) that take a particular shape and orientation relative to the specimen/component axes. The net strain in a given direction is comprised of the sum of the strains from each individual grain resolved in the direction of involvement with, perhaps, some modification due to the inter-granular stress interactions arising from differential growth of neighboring grains. Grain boundaries are also sinks for point defects and the grain boundary sink forcefulness has to exist a function of the grain purlieus orientation when there is biased, directional diffusional menstruation of one type of point defect, i.e., DAD. 53,54 For these reasons the effect of texture and grain structure on irradiation growth are inter-related, i.e., they are not independent variables. For a large (30 µm diameter) equiaxed grain structure the upshot of texture is illustrated for Zircaloy-two irradiated at ~60°C in Fig. 30. The Kearns f parameter 95 is a mensurate of tendency for grains to have a c-axis orientation that is close to the measurement direction. A small f-parameter means that most grains take their c-centrality oriented at a large angle from the measurement direction. A large f-parameter means that most grains tend to take their c-axes oriented close to the measurement direction.

Fig. 30

Fig. xxx. Outcome of crystallographic texture on irradiation growth for annealed Zircaloy-ii at 651–669K (~400°C).

Reproduced from Fidleris, V., 1988. The irradiation creep and growth phenomena. J. Nucl. Mater. 159, 22–42.

Results on the effect of grain size on irradiation growth are often confounded by the fact that other microstructural variables, such as texture and dislocation structure also vary. Data on grain size for annealed Zircaloy-two and iodide Zr with similar textures were compiled past Fidleris and are reproduced in Fig. 31. 8 The results are somewhat ambiguous. Considering the iodide Zr data solitary, higher growth strains (chief strain and secondary rates) are exhibited by the material with a smaller grain size. The outcome of grain size becomes more credible when the grain dimensions are small and the grain boundary sink strengths become comparable with the sink strengths of other microstructural features such as network dislocations and dislocation loops. As is often the case, it is sometimes difficult to isolate the outcome of one variable when other variables change at the same time. In such circumstances charge per unit-theory modeling helps to isolate and understand the contributions for dissimilar variables (run across Department 1.11.five.5).

Fig. 31

Fig. 31. Effect of grain size on irradiation growth of recrystallized zirconium and Zircaloy-2 at 330K.

Read full chapter

URL:

https://www.sciencedirect.com/science/article/pii/B9780128035818116467

Factors influencing mechanical backdrop

Jean P. Mercier , ... Wilfried Kurz , in Introduction to Materials Science, 2002

12.4.9 Polycrystal effect

For a polycrystalline material, the grains have diverse crystal orientations. The initiation of slip preferentially occurs in the grains having slip systems with a Schmid factor shut to 0.5 earlier actualization in the other grains ( figure 12.20.). To obtain a macroscopic plastic deformation, skid must exist activated in an adjoining ready of grains which exercise not all have slip planes and gliding directions oriented at 45 ° to the tensile axis. The cooperative slip of randomly distributed grains requires a college applied stress relative to that needed to plastically deform a monocrystal oriented in such way that Schmid factor is equal to 0.5. For a polycrystalline material, only an average yield strength τ ¯ due east determined which takes account of the random orientation of the grains, this involving diverse values of Schmid gene. The following approximate relations are then obtained:

Effigy 12.xx. Conditions of progressive plastic deformation in a polycrystalline material.

(12.27) τ ¯ e one , five τ e

In expression (12.27.), τe is the critical shear stress of the monocrystal. This enables it to be understood why a polycrystalline fabric has an yield force larger than that of the whole of its grains which would all be oriented with a Schmid cistron equal to 0.5. No account is taken here of stresses created past the delayed deformation of grains. This hardening issue depends only on the orientation of the grains and differs from the same effect obtained past the refining of grains discussed in the previous chapter. This latter type of hardening depends on the grain diameter and resistance of boundaries to the passage of dislocations.

Read total chapter

URL:

https://www.sciencedirect.com/science/commodity/pii/B9782842992866500184

Orthopedical and biomedical applications of titanium and zirconium metals

Gunarajulu Renganathan , ... Suguna Lakshmi Madurai , in Fundamental Biomaterials: Metals, 2018

10.three.3 Ceramics

Ceramics are polycrystalline materials. The compounds are fabricated upwards of metallic as well every bit nonmetallic elements. They are bonded past ionic bonds with some covalent bonds. Certain properties of ceramics like low ductility and brittleness have express the use of ceramics. The main ceramic in orthopedic surgery and their application are:

Al2O3 (alumina)—acetabular and femoral components.

ZrOii (zirconium)—femoral components.

Ca10(PO4)6(OH)2 (hydroxyapatite)—coating femoral stalk components to integrate the surface textile to the bone.

Read full chapter

URL:

https://www.sciencedirect.com/science/article/pii/B9780081022054000106

Characterization techniques in energy generation and storage

Northward. Fleck , ... Waqar Ahmed , in Emerging Nanotechnologies for Renewable Free energy, 2021

xi.two.iii.2 Bones principles

Crystals and polycrystalline materials consist of crystal planes from which diffraction can occur. This is the basis of XRD which can exist seen in Fig. 11.4. X-rays impinge on a sample at a sure bending (θ). A portion of the radiation is reflected at the upper surface while rays also penetrate into the crystal and reflect from the aeroplane beneath, and so on. The number of planes involved depends on the density of the cloth and energy of the incident rays. Constructive interference will occur when the rays recombine at the sample interface in phase. Bragg'south law (Eq. 11.2) describes this status. The wavelength of the incident rays is given by λ, d hkl is the interplane spacing, θ is the angle subtending the surface and incident 10-rays, and n is a multiplying integer which represents the order of diffraction. Interference maxima occur when the path difference between the recombining rays is an integer multiple of the incident wavelength. The equation is derived by simple geometric analysis of the path difference between the X-rays reflected from successive planes.

Figure 11.four. Planar diffraction schematic.

(11.2) n λ = two d h k l sin θ

Read full chapter

URL:

https://world wide web.sciencedirect.com/scientific discipline/commodity/pii/B9780128213469000031