Grain boundary sliding

Last updated

Grain boundary sliding (GBS) is a material deformation mechanism where grains slide against each other. This occurs in polycrystalline material under external stress at high homologous temperature (above ~0.4 [1] ) and low strain rate and is intertwined with creep. Homologous temperature describes the operating temperature relative to the melting temperature of the material. There are mainly two types of grain boundary sliding: Rachinger sliding, [2] and Lifshitz sliding. [3] Grain boundary sliding usually occurs as a combination of both types of sliding. Boundary shape often determines the rate and extent of grain boundary sliding. [4]

Contents

Grain boundary sliding is a motion to prevent intergranular cracks from forming. Keep in mind that at high temperatures, many processes are underway, and grain boundary sliding is only one of the processes happening. Therefore it is not surprising that Nabarro Herring and Coble creep is dependent on grain boundary sliding. During high temperature creep, wavy grain boundaries are often observed. We can simulate this type of boundary with a sinusoidal curve, with amplitude h and wavelength λ. Steady-state creep rate increases with rising λ/h ratios. At high λ and high homologous temperatures, grain boundary sliding is controlled by lattice diffusion (Nabarro-Herring mechanism). On the other hand, it will be controlled by grain boundary diffusion (Coble Creep). Additionally, when λ/h ratios are high, it may impede diffusional flow, therefore diffusional voids may form, which leads to fracture in creep. [5] [6]

Many people have developed estimations for the contribution of grain boundary sliding to the total strain experienced by various groups of materials, such as metals, ceramics, and geological materials. Grain boundary sliding contributes a significant amount of strain, especially for fine grain materials and high temperatures. [1] It has been shown that Lifshitz grain boundary sliding contributes about 50-60% of strain in Nabarro–Herring diffusion creep. [7] This mechanism is the primary cause of ceramic failure at high temperatures due to the formation of glassy phases at their grain boundaries. [8]

A simple schematic of grain boundary sliding in a polycrystalline sample (adapted from ). When a tensile load is applied to the materials, the grains stretch along that direction. This leads to the creation of voids/cavities and a loss of coherency. To prevent void formation, the grains slide relative to each other to fill in these unfavorable gaps. GBsliding.tif
A simple schematic of grain boundary sliding in a polycrystalline sample (adapted from ). When a tensile load is applied to the materials, the grains stretch along that direction. This leads to the creation of voids/cavities and a loss of coherency. To prevent void formation, the grains slide relative to each other to fill in these unfavorable gaps.

Rachinger sliding

Rachinger sliding is purely elastic; the grains retain most of their original shape. [7] The internal stress will build up as grains slide until the stress balances out with the external applied stress. For example, when a uniaxial tensile stress is applied on a sample, grains move to accommodate the elongation and the number of grains along the direction of applied stress increases.

Lifshitz sliding

Lifshitz sliding only occurs with Nabarro–Herring and Coble creep. [7] The sliding motion is accommodated by the diffusion of vacancies from induced stresses and the grain shape changes during the process. For example, when a uniaxial tensile stress is applied, diffusion will occur within grains and the grain will elongate in the same direction as the applied stress. There will not be an increase in number of grains along the direction of applied stress.

Accommodation mechanisms

When polycrystalline grains slide relative to each other, there must be simultaneous mechanisms that allow for this sliding to occur without the overlapping of grains (which would be physically impossible). [10] Various accommodation mechanisms have been proposed to account for this issue.

Grain boundary sliding accommodated by diffusional flow:

Grain boundary sliding accommodated by diffusional flow takes place by grain-switching while preserving grain shape. This type of mechanism is synonymous to Nabarro Herring and Coble creep but describes the grain at superplastic conditions. This concept was originally proposed by Ashby and Verral. During grain switching, we can describe the process through three steps: a) Initial state b) Intermediate stage c) Final state. During the intermediate stage, there must first be an applied stress exceeding the “threshold” stress so that there is an increase in grain boundary area which is provided by the diffusional flow that occurs once the threshold stress is achieved. Under the assumption that the applied stress is much greater than the threshold stress, the strain rate is greater than conventional diffusional creep. The reason for this is that for grain switching diffusion, the distance is about 1/7 the distance of diffusional creep and there are two more paths to grain switching in comparison with diffusional creep. Thus, this will lead to about an order magnitude higher strain rate than diffusional creep.

Grain boundary sliding accommodated by dislocation flow:

At superplastic temperature, strain rate and stress conditions, dislocations are really observed because they are quickly emitted and absorbed at grain boundaries. However, careful studies have been conducted to verify that dislocations are indeed emitted during superplastic deformation. During dislocation flow, the shape of the grain must be ensured to not change. Based on models of super plasticity, transitioning from dislocation creep to super plasticity occurs when the sub grain size is less than the grain size. The sub grain size: often denoted as d’ can be described in the equation below:

d’/b =10G/𝜏, Where it has an inverse relationship with shear stress. [12]

Deformation rate from grain boundary sliding

Generally speaking, the minimum creep rate for diffusion can be expressed as: [13] [7]

where the terms are defined as follows:

In the case where this minimum creep rate is controlled by grain boundary sliding, the exponents become , , and the diffusion coefficient becomes (the lattice diffusion coefficient). [13] [7] Thus, the minimum creep rate becomes:

Estimating the contribution of GBS to the overall Strain

The total strain under creep conditions can be denoted as εt , where:

εt  =   εg + εgbsdc  

εg =   Strain associated with intragranular dislocation processes

εgbs =   Strain due to Rachinger GBS associated with intragranular sliding

εdc   =   Strain due to Lifshitz GBS associated with diffusion creep

During practice, experiments are normally performed in conditions where creep is negligible, therefore equation 1 will reduce to:

εt  =   εg + εgbs

Therefore the contribution of GBS to the total strain can be denoted as:

Ⲝ =   εgbs / εt  

First, we need to illustrate the three perpendicular displacement vectors: u, v, and w, with a grain boundary sliding vector: s. It can be imagined as the w displacement vector coming out of the plane.  While the v and u vectors are in the plane. The displacement vector u is also the tensile stress direction. The sliding contribution may be estimated by individual measurements of  εgbs through these displacement vectors. We can further define the angle at the u v plane of displacements as Ѱ, and the angle between the u w planes as Θ. u can then be related by the tangents of these angles through the equation:

U = vtan Ѱ + wtanΘ

A common and easier way in practice is to use interferometry to measure fringes along the v displacement axis. The sliding strain is then given by:

εgbs = k’’nr vr

Where k’’ is constant, nr is the number of measurements, and vr is the average of n measurements.

Thus we can calculate the percentage of GBS strain. [14]

Experimental evidence

A simple schematic showing how experimentalists observe grain boundary sliding between two adjacent grains. Initially, a polycrystalline material is scratched with a marker line (shown here as a thick dashed line). If these two grains slide relative to one another, there will be an offset in this marker line occurring at the grain boundary. This can be observed using various microscopy techniques. Grainboundarysliding marker.tif
A simple schematic showing how experimentalists observe grain boundary sliding between two adjacent grains. Initially, a polycrystalline material is scratched with a marker line (shown here as a thick dashed line). If these two grains slide relative to one another, there will be an offset in this marker line occurring at the grain boundary. This can be observed using various microscopy techniques.

Grain boundary scattering has been observed experimentally using various microscopy techniques. It was first observed in NaCl and MgO bicrystals in 1962 by Adams and Murray. [15] By scratching the surface of their samples with a marker line, they were able to observe an offset of that line at the grain boundary as a result of adjacent grains sliding with respect to each other. Subsequently this was observed in other systems as well including in Zn-Al alloys using electron microscopy, [16] and octachloropropane using in situ techniques. [10]

Nanomaterials

Nano-crystalline materials, or nanomaterials, have fine grains which helps suppress lattice creep. This is beneficial for relatively low temperature operations as it impedes dislocations motion or diffusion due to high volume fraction of grain boundaries. However, fine grains are undesirable at high temperature due to the increased probability of grain boundary sliding. [17]

Prevention

Grain shape plays a large role in determining the sliding rate and extent. Thus, by controlling the grain size and shape, the amount of grain boundary sliding can be limited. Generally, materials with coarser grains are preferred, as the material will have less grain boundaries. Ideally, single crystals will completely suppress this mechanism as the sample will not have any grain boundaries.

Another method is to reinforce grain boundaries by adding precipitates. Small precipitates located at grain boundaries can pin grain boundaries and prevent grains from sliding against each other. However, not all precipitates are desirable at boundaries. Large precipitates may have the opposite effect on grain boundary pinning as it allows more gaps or vacancies between grains to accommodate the precipitates, which reduces the pinning effect.

Modeling effects of GBS in high strength steel

The application of high-strength steel is ubiquitous in the engineering world today. To provide a substantial engineering basis for real-world construction, the modeling of high-strength steel is very important.

By inputting parameters such as elastic modulus, yield strength, Poisson’s ratio, and specific heat of high strength steel from two temperatures, we can derive the related GBS energy as a function of temperature and thus its yield strength as a function of temperature. [18]

Experimental Study: Superplastic Forming Technique via GBS

The superplastic forming technique is a technique where materials are deformed beyond the yield stress to form a complex shaped lightweight construction. This phenomenon is possible through grain boundary sliding that is enabled by dislocation slip/creep and diffusional creep.

An example would be for commercial fine-grained Al-Mg alloys, unusually weak grain boundary sliding is observed during the initial stage of superplastic deformation. Through a tensile test, grains were elongated along the tensile direction to 50~70%. The deformation was orchestrated by increased precipitation depletion zone fractions, particle segregation on the longitudinal grain boundaries, dislocation activity, and subgrains. Increased Mg content leads to increased GBS.  Increasing Mg content from 4.8 to 6.5~7.6% aids grain size stability during the increased temperature process, simplified the GBS and decreased diffusion creep contribution, and increased the failure strain from 300% to 430%. [19]

Application to tungsten filaments

The operation temperature for tungsten filaments used in incandescent lightbulbs is around 2000K to 3200K which is near the melting point of tungsten (Tm = 3695 K). [20] As lightbulbs are expected to operate for long periods of time at a homologous temperature up to 0.8, understanding and preventing creep mechanism is crucial to extending their life expectancy.

Researchers found that the predominant mechanism for failure in these tungsten filaments was grain boundary sliding accommodated by diffusional creep. [21] This is because tungsten filaments, being as thin as they are, typically consist of only a handful of elongated grains. In fact there is usually less than one grain boundary per turn in a tungsten coil. [21] This elongated grain structure is generally called a bamboo structure, as the grains look similar to the internodes of bamboo stalks. During operation, the tungsten wire is stressed under the load of its own weight and because of the diffusion that can occur at high temperatures, grains begin to rotate and slide. This stress, because of variations in the filament, causes the filament to sag nonuniformly, which ultimately introduces further torque on the filament. [21] It is this sagging that inevitably results in a rupture of the filament, rendering the incandescent lightbulb useless. The typical lifetime for these single coil filaments is approximately 440 hours. [21]

To combat this grain boundary sliding, researchers began to dope the tungsten filament with aluminum, silicon and most importantly potassium. This composite material (AKS tungsten) is unique as it is composed of potassium and tungsten, which are non-alloying. [22] This feature of potassium results in nanosized bubbles of either liquid or gaseous potassium being distributed throughout the filament after proper manufacturing. [22] These bubbles interact with all defects in the filament pinning dislocations and most importantly grain boundaries. Pinning these grain boundaries, even at high temperatures, drastically reduces grain boundary sliding. This reduction in grain boundary sliding earned these filaments the title of "non-sag filaments" as they would no longer bow under their own weight. [22] Thus, this initially counter-intuitive approach to strengthening tungsten filaments began to be widely used in almost every incandescent lightbulb to greatly increase their lifetime.

Related Research Articles

Plasticity (physics) Non-reversible deformation of a solid material in response to applied forces

In physics and materials science, plasticity, also known as plastic deformation, is the ability of a solid material to undergo permanent deformation, a non-reversible change of shape in response to applied forces. For example, a solid piece of metal being bent or pounded into a new shape displays plasticity as permanent changes occur within the material itself. In engineering, the transition from elastic behavior to plastic behavior is known as yielding.

In materials science, superplasticity is a state in which solid crystalline material is deformed well beyond its usual breaking point, usually over about 600% during tensile deformation. Such a state is usually achieved at high homologous temperature. Examples of superplastic materials are some fine-grained metals and ceramics. Other non-crystalline materials (amorphous) such as silica glass and polymers also deform similarly, but are not called superplastic, because they are not crystalline; rather, their deformation is often described as Newtonian fluid. Superplastically deformed material gets thinner in a very uniform manner, rather than forming a "neck" that leads to fracture. Also, the formation of microvoids, which is another cause of early fracture, is inhibited.

Creep (deformation) Tendency of a solid material to move slowly or deform permanently under mechanical stress

In materials science, creep is the tendency of a solid material to move slowly or deform permanently under the influence of persistent mechanical stresses. It can occur as a result of long-term exposure to high levels of stress that are still below the yield strength of the material. Creep is more severe in materials that are subjected to heat for long periods and generally increases as they near their melting point.

Grain boundary Concept in materials science

A grain boundary is the interface between two grains, or crystallites, in a polycrystalline material. Grain boundaries are 2D defects in the crystal structure, and tend to decrease the electrical and thermal conductivity of the material. Most grain boundaries are preferred sites for the onset of corrosion and for the precipitation of new phases from the solid. They are also important to many of the mechanisms of creep. On the other hand, grain boundaries disrupt the motion of dislocations through a material, so reducing crystallite size is a common way to improve mechanical strength, as described by the Hall–Petch relationship. The study of grain boundaries and their effects on the mechanical, electrical and other properties of materials forms an important topic in materials science.

Superalloy Alloy with higher durability than normal metals

A superalloy, or high-performance alloy, is an alloy with the ability to operate at a high fraction of its melting point. Several key characteristics of a superalloy are excellent mechanical strength, resistance to thermal creep deformation, good surface stability, and resistance to corrosion or oxidation.

Hardening is a metallurgical metalworking process used to increase the hardness of a metal. The hardness of a metal is directly proportional to the uniaxial yield stress at the location of the imposed strain. A harder metal will have a higher resistance to plastic deformation than a less hard metal.

The Portevin–Le Chatelier (PLC) effect describes a serrated stress–strain curve or jerky flow, which some materials exhibit as they undergo plastic deformation, specifically inhomogeneous deformation. This effect has been long associated with dynamic strain aging or the competition between diffusing solutes pinning dislocations and dislocations breaking free of this stoppage.

Embrittlement

Embrittlement is a significant decrease of ductility of a material, which makes the material brittle. Embrittlement is used to describe any phenomena where the environment compromises a stressed material's mechanical performance, such as temperature or environmental composition. This is oftentimes undesirable as brittle fracture occurs quicker and can much more easily propagate than ductile fracture, leading to complete failure of the equipment. Various materials have different mechanisms of embrittlement, therefore it can manifest in a variety of ways, from slow crack growth to a reduction of tensile ductility and toughness.

Diffusion creep refers to the deformation of crystalline solids by the diffusion of vacancies through their crystal lattice. Diffusion creep results in plastic deformation rather than brittle failure of the material.

Nanocrystalline material

A nanocrystalline (NC) material is a polycrystalline material with a crystallite size of only a few nanometers. These materials fill the gap between amorphous materials without any long range order and conventional coarse-grained materials. Definitions vary, but nanocrystalline material is commonly defined as a crystallite (grain) size below 100 nm. Grain sizes from 100–500 nm are typically considered "ultrafine" grains.

Coble creep

Coble creep, a form of diffusion creep, is a mechanism for deformation of crystalline solids. Contrasted with other diffusional creep mechanisms, Coble creep is similar to Nabarro–Herring creep in that it is dominant at lower stress levels and higher temperatures than creep mechanisms utilizing dislocation glide. Coble creep occurs through the diffusion of atoms in a material along grain boundaries. This mechanism is observed in polycrystals or along the surface in a single crystal, which produces a net flow of material and a sliding of the grain boundaries.

A deformation mechanism, in geotechnical engineering, is a process occurring at a microscopic scale that is responsible for changes in a material's internal structure, shape and volume. The process involves planar discontinuity and/or displacement of atoms from their original position within a crystal lattice structure. These small changes are preserved in various microstructures of materials such as rocks, metals and plastics, and can be studied in depth using optical or digital microscopy.

Methods have been devised to modify the yield strength, ductility, and toughness of both crystalline and amorphous materials. These strengthening mechanisms give engineers the ability to tailor the mechanical properties of materials to suit a variety of different applications. For example, the favorable properties of steel result from interstitial incorporation of carbon into the iron lattice. Brass, a binary alloy of copper and zinc, has superior mechanical properties compared to its constituent metals due to solution strengthening. Work hardening has also been used for centuries by blacksmiths to introduce dislocations into materials, increasing their yield strengths.

Grain boundary strengthening

Grain-boundary strengthening is a method of strengthening materials by changing their average crystallite (grain) size. It is based on the observation that grain boundaries are insurmountable borders for dislocations and that the number of dislocations within a grain has an effect on how stress builds up in the adjacent grain, which will eventually activate dislocation sources and thus enabling deformation in the neighbouring grain, too. So, by changing grain size one can influence the number of dislocations piled up at the grain boundary and yield strength. For example, heat treatment after plastic deformation and changing the rate of solidification are ways to alter grain size.

Viscoplasticity Theory in continuum mechanics

Viscoplasticity is a theory in continuum mechanics that describes the rate-dependent inelastic behavior of solids. Rate-dependence in this context means that the deformation of the material depends on the rate at which loads are applied. The inelastic behavior that is the subject of viscoplasticity is plastic deformation which means that the material undergoes unrecoverable deformations when a load level is reached. Rate-dependent plasticity is important for transient plasticity calculations. The main difference between rate-independent plastic and viscoplastic material models is that the latter exhibit not only permanent deformations after the application of loads but continue to undergo a creep flow as a function of time under the influence of the applied load.

Oxide dispersion strengthened alloys (ODS) are alloys that consist of a metal matrix with small oxide particles dispersed within it. They have high heat resistance, strength, and ductility. Alloys of nickel are the most common but includes iron aluminum alloys.

Dislocation creep is a deformation mechanism in crystalline materials. Dislocation creep involves the movement of dislocations through the crystal lattice of the material, in contrast to diffusion creep, in which diffusion is the dominant creep mechanism. It causes plastic deformation of the individual crystals, and thus the material itself.

In materials science, the yield strength anomaly refers to materials wherein the yield strength increases with temperature. For the majority of materials, the yield strength decreases with increasing temperature. In metals, this decrease in yield strength is due to the thermal activation of dislocation motion, resulting in easier plastic deformation at higher temperatures.

Nabarro–Herring creep is a mode of deformation of crystalline materials that occurs at low stresses and held at elevated temperatures in fine-grained materials. In Nabarro–Herring creep, atoms diffuse through the crystals, and the creep rate varies inversely with the square of the grain size so fine-grained materials creep faster than coarser-grained ones. NH creep is solely controlled by diffusional mass transport. This type of creep results from the diffusion of vacancies from regions of high chemical potential at grain boundaries subjected to normal tensile stresses to regions of lower chemical potential where the average tensile stresses across the grain boundaries are zero. Self-diffusion within the grains of a polycrystalline solid can cause the solid to yield to an applied shearing stress, the yielding being caused by a diffusional flow of matter within each crystal grain away from boundaries where there is a normal pressure and toward those where there is a normal tension. Atoms migrating in the opposite direction account for the creep strain. The creep strain rate is derived in the next section. NH creep is more important in ceramics than metals as dislocation motion is more difficult to effect in ceramics.

Microstructurally stable nanocrystalline alloys are alloys that are designed to resist microstructural coarsening under various thermo-mechanical loading conditions.

References

  1. 1 2 Bell, R.L., Langdon, T.G. An investigation of grain-boundary sliding during creep. J Mater Sci 2, 313–323 (1967). https://doi.org/10.1007/BF00572414
  2. W. A. Rachinger, J. Inst. Metals 81 (1952-1953) 33.
  3. I. M. Lifshitz, Soviet Phys. JETP 17 (1963) 909.
  4. 1 2 3 Raj, R., Ashby, M.F. On grain boundary sliding and diffusional creep. MT 2, 1113–1127 (1971). https://doi.org/10.1007/BF02664244
  5. Bhaduri, Amit (2018). "Mechanical Properties and Working of Metals and Alloys". Springer Series in Materials Science. doi:10.1007/978-981-10-7209-3. ISSN   0933-033X.
  6. Raj, R.; Ashby, M. F. (April 1971). "On grain boundary sliding and diffusional creep". Metallurgical Transactions. 2 (4): 1113–1127. doi:10.1007/bf02664244. ISSN   0360-2133.
  7. 1 2 3 4 5 Langdon, T.G. Grain boundary sliding revisited: Developments in sliding over four decades. J Mater Sci 41, 597–609 (2006). https://doi.org/10.1007/s10853-006-6476-0
  8. Joachim Rösler, Harald Harders, Martin Bäker, Mechanical Behaviour of Engineering Materials, Springer-Verlag Berlin Heidelberg, 2007, p 396. ISBN   978-3-540-73446-8
  9. Courtney, Thomas H. (2000). Mechanical behavior of materials (2 ed.). Boston: McGraw Hill. ISBN   0-07-028594-2. OCLC   41932585.
  10. 1 2 "Grain boundary sliding and development of grain boundary openings in experimentally deformed octachloropropane". Journal of Structural Geology. 16 (3): 403–418. 1994-03-01. doi:10.1016/0191-8141(94)90044-2. ISSN   0191-8141.
  11. Gifkins, R. C. (August 1976). "Grain-boundary sliding and its accommodation during creep and superplasticity". Metallurgical Transactions A. 7 (8): 1225–1232. doi:10.1007/bf02656607. ISSN   0360-2133.
  12. Courtney, Thomas H. (2013). Mechanical Behavior of Materials (2nd ed. Reimp ed.). New Delhi: McGraw Hill Education (India). ISBN   1-259-02751-1. OCLC   929663641.
  13. 1 2 Yang, Hong; Gavras, Sarkis; Dieringa, Hajo (2021), "Creep Characteristics of Metal Matrix Composites", Reference Module in Materials Science and Materials Engineering, Elsevier, pp. B9780128035818118223, doi:10.1016/b978-0-12-803581-8.11822-3, ISBN   978-0-12-803581-8 , retrieved 2021-05-11
  14. Langdon, Terence G. (February 2006). "Grain boundary sliding revisited: Developments in sliding over four decades". Journal of Materials Science. 41 (3): 597–609. doi:10.1007/s10853-006-6476-0. ISSN   0022-2461.
  15. Adams, M. A.; Murray, G. T. (June 1962). "Direct Observations of Grain‐Boundary Sliding in Bi‐Crystals of Sodium Chloride and Magnesia". Journal of Applied Physics. 33 (6): 2126–2131. doi:10.1063/1.1728908. ISSN   0021-8979.
  16. Naziri, H.; Pearce, R.; Brown, M.Henderson; Hale, K.F. (April 1975). "Microstructural-mechanism relationship in the zinc/ aluminium eutectoid superplastic alloy". Acta Metallurgica. 23 (4): 489–496. doi:10.1016/0001-6160(75)90088-7. ISSN   0001-6160.
  17. Sergueeva, A.V., Mara, N.A. & Mukherjee, A.K. Grain boundary sliding in nanomaterials at elevated temperatures. J Mater Sci 42, 1433–1438 (2007). https://doi.org/10.1007/s10853-006-0697-0
  18. Zhang, Xi; Li, Weiguo; Ma, Jianzuo; Li, Ying; Zhang, Xin; Zhang, Xuyao (January 2021). "Modeling the effects of grain boundary sliding and temperature on the yield strength of high strength steel". Journal of Alloys and Compounds. 851: 156747. doi:10.1016/j.jallcom.2020.156747. ISSN   0925-8388.
  19. Mikhaylovskaya, A.V.; Yakovtseva, O.A.; Irzhak, A.V. (January 2022). "The role of grain boundary sliding and intragranular deformation mechanisms for a steady stage of superplastic flow for Al–Mg-based alloys". Materials Science and Engineering: A. 833: 142524. doi:10.1016/j.msea.2021.142524.
  20. Wright, P. K. (1978-07-01). "The high temperature creep behavior of doped tungsten wire". Metallurgical Transactions A. 9 (7): 955–963. doi:10.1007/BF02649840. ISSN   1543-1940.
  21. 1 2 3 4 Raj, R.; King, G. W. (1978-07-01). "Life Prediction of Tungsten Filaments in Incandescent Lamps". Metallurgical Transactions A. 9 (7): 941–946. doi:10.1007/BF02649838. ISSN   1543-1940.
  22. 1 2 3 "100 years of doped tungsten wire". International Journal of Refractory Metals and Hard Materials. 28 (6): 648–660. 2010-11-01. doi:10.1016/j.ijrmhm.2010.05.003. ISSN   0263-4368.