Sie sind auf Seite 1von 3

CREEP creep is the tendency of a solid material to move slowly or deform permanently under the influence of stresses.

It occurs as a result of long term exposure to high levels of stress that are below the yield strength of the material. Creep is more severe in materials that are subjected to heat for long periods, and near melting point. Creep always increases with temperature. STAGES OF CREEP In the initial stage, or primary creep, the strain rate is relatively high, but slows with increasing time. This is due to work hardening. The strain rate eventually reaches a minimum and becomes near constant. This is due to the balance between work hardening and annealing (thermal softening). This stage is known as secondary or steady-state creep. This stage is the most understood. The characterized "creep strain rate" typically refers to the rate in this secondary stage. Stress dependence of this rate depends on the creep mechanism. In tertiary creep, the strain rate exponentially increases with stress because of necking phenomena.

General creep equation

where is the creep strain, C is a constant dependent on the material and the particular creep mechanism, m and b are exponents dependent on the creep mechanism, Q is the activation energy of the creep mechanism, is the applied stress, d is the grain size of the material, k is Boltzmann's constant, and T is the absolute temperature.

CREEP AT DIFFERENT TEMPERATURE

The shape of creep curve will slightly change according to the applied stress at a constant temperature. STRESS RELAXATION CURVE

CREEP OF POLYMERS
Creep can occur in polymers and metals which are considered viscoelastic materials. When a polymeric material is subjected to an abrupt force, the response can be modeled using the Kelvin-Voigt model. The creep strain is given by the following convolution integral:

where:

= applied stress C0 = instantaneous creep compliance C = creep compliance coefficient = retardation time = distribution of retardation times

When subjected to a step constant stress, viscoelastic materials experience a time-dependent increase in strain. This phenomenon is known as viscoelastic creep.

Creep of concrete
The creep of concrete, which originates from the calcium silicate hydrates (C-S-H) in the hardened Portland cement paste (which is the binder of mineral aggregates), is fundamentally different from the creep of metals as well as polymers. Unlike the creep of metals, it occurs at all stress levels and, within the service stress range, is linearly dependent on the stress if the pore water content is constant. Unlike the creep of polymers and metals, it exhibits multi-months aging, caused by chemical hardening due to hydration which stiffens the microstructure, and multi-year aging, caused by long-term relaxation of self-equilibrated micro-stresses in the nano-porous microstructure of the C-S-H. If concrete is fully dried, it does not creep, but it is next to impossible to dry concrete fully without severe cracking.

MECHANISM OF CREEP
The mechanism of creep depends on temperature and stress. The various methods are:

Bulk diffusion (Nabarro-Herring creep) Climb here the strain is actually accomplished by climb Climb-assisted glide here the climb is an enabling mechanism, allowing dislocations to get around obstacles Grain boundary diffusion (Coble creep) Thermally activated glide e.g., via cross-slip

Nabarro-Herring creep
Nabarro-Herring creep is a form of diffusion creep. In Nabarro-Herring creep, atoms diffuse through the lattice causing grains to elongate along the stress axis; k is related to the diffusion coefficient of atoms through the lattice, Q = Q(self diffusion), m = 1, and b = 2. Therefore NabarroHerring creep has a weak stress dependence and a moderate grain size dependence, with the creep rate decreasing as grain size is increased. Nabarro-Herring creep is strongly temperature dependent. For lattice diffusion of atoms to occur in a material, neighboring lattice sites or interstitial sites in the crystal structure must be free. A given atom must also overcome the energy barrier to move from its current site (it lies in an energetically favorable potential well) to the nearby vacant site (another potential well). The general form of the diffusion equation is D = D0exp(E/KT) where D0 has a dependence on both the attempted jump frequency and the number of nearest neighbor sites and the probability of the sites being vacant. Thus there is a double dependence upon temperature. At higher temperatures the diffusivity increases due to the direct temperature dependence of the equation, the increase in vacancies through Schottky defect formation, and an increase in the average energy of atoms in the material. Nabarro-Herring creep dominates at very high temperatures relative to a material's melting temperature.

Coble creep
Coble creep is a second form of diffusion controlled creep. In Coble creep the atoms diffuse along grain boundaries to elongate the grains along the stress axis. This causes Coble creep to have a stronger grain size dependence than Nabarro-Herring creep. For Coble creep k is related to the diffusion coefficient of atoms along the grain boundary, Q = Q(grain boundary diffusion), m = 1, andb = 3. Because Q(grain boundary diffusion) < Q(self diffusion), Coble creep occurs at lower temperatures than Nabarro-Herring creep. Coble creep is still temperature dependent, as the temperature increases so does the grain boundary diffusion. However, since the number of nearest neighbors is effectively limited along the interface of the grains, and thermal generation of vacancies along the boundaries is less prevalent, the temperature dependence is not as strong as in Nabarro-Herring creep. It also exhibits the same linear dependence on stress as Nabarro-Herring creep.

Dislocation climb

Dislocations can slip in planes containing both the dislocation and the Burgers Vector. For a screw dislocation, the dislocation and the Burgers vector are parallel, so the dislocation may slip in any plane containing the dislocation. For an edge dislocation, the dislocation and the Burgers vector are perpendicular, so there is only one plane in which the dislocation can slip. There is an alternative mechanism of dislocation motion, fundamentally different from slip, that allows an edge dislocation to move out of its slip plane, known as dislocation climb. Dislocation climb allows an edge dislocation to move perpendicular to its slip plane. The driving force for dislocation climb is the movement of vacancies through a crystal lattice. If a vacancy moves next to the boundary of the extra half plane of atoms that forms an edge dislocation, the atom in the half plane closest to the vacancy can "jump" and fill the vacancy. This atom shift "moves" the vacancy in line with the half plane of atoms, causing a shift, or positive climb, of the dislocation. The process of a vacancy being absorbed at the boundary of a half plane of atoms, rather than created, is known as negative climb. Since dislocation climb results from individual atoms "jumping" into vacancies, climb occurs in single atom diameter increments.

Factors affecting creep


Factors affecting creep are: *Material properties *exposure temperature *Exposure time and *Applied load. The curves will also highlight another important distinction between creep and stress relaxation: the failure mode. In creep, the failure is obvious to the naked eye. The part will continuously deform until fracture, leaving behind several very distorted pieces. In stress relaxation, the failure mode is a loss of contact force, and an inability for a deflected contact to return to its original position. A part with virtually no stress left in it will look no different than it did when the load was first applied. However, the normal force necessary to maintain electrical contact will no longer be present. Therefore, stress relaxation can be manifested as a constant increase in contact resistance across the contact interface, eventually leading to an open circuit. It takes some amount of energy for the crystals to permanently distort. This energy comes from load placed on the metal. At stress levels below the elastic limit, there is not enough energy available for the atoms to rearrange themselves. At higher stress levels, there is more energy available for the atoms to move. Therefore, deformation is more severe at higher stress levels.Additionally, elevated temperatures add thermal energy to the system, making the crystalline reorganization more likely. This is why there is more creep and stress relaxation at higher temperatures.At high enough temperatures, there is enough thermal energy available to induce creep and stress relaxation, even if the initial stress level is well below the yield strength of the material.

Applications
1.Though mostly due to the reduced yield stress at higher temperatures, the Collapse of the World Trade Center was due in part to creep from increased temperature operation.[5] 2.The creep rate of hot pressure-loaded components in a nuclear reactor at power can be a significant design constraint, since the creep rate is enhanced by the flux of energetic particles. 3.Creep was blamed for the Big Dig tunnel ceiling collapse in Boston, Massachusetts that occurred in July 2006. 4.An example of an application involving creep deformation is the design of tungsten light bulb filaments. 5.In steam turbine power plants, pipes carry steam at high temperatures (566C or 1050F) and pressures (above 24.1 MPa or 3500 psi). In jet engines, temperatures can reach up to 1400C (2550F) and initiate creep deformation in even advanced coated turbine blades. Hence, it is crucial for correct functionality to understand the creep deformation behavior of materials. Creep deformation is important not only in systems where high temperatures are endured such as nuclear power plants, jet engines and heat exchangers, but also in the design of many everyday objects.

Das könnte Ihnen auch gefallen