Você está na página 1de 25

Advanced Mechanical Design

by
Dr. Syed Ismail
Strength failures of load bearing structures can be either of the yielding-dominant or
fracture-dominant types. Defects are important for both types of failure, but those of
primary importance to fracture differ in an extreme way from those influencing yielding
and the resistance to plastic flow.

Types of structural failure


For yielding-dominant failures the significant defects are those which tend to warp and
interrupt the crystal lattice planes, thus interfering with dislocation glide and providing a
resistance to plastic deformation that is essential to the strength of high strength metals.
Examples of such defects are interstitial and out-of-size substitutional atoms, grain
boundaries, coherent precipitates and dislocation networks.

Larger defects like inclusions, porosity, surface scratches and small cracks may influence
the effective net section bearing the load, but otherwise have little effect on resistance to
yielding.

For fracture-dominant failures, i.e. fracture before general yielding of the net section, the
size scale of the defects which are of major significance is essentially macroscopic, since
general plasticity is not involved but only the local stress-strain fields associated with the
defects. The minute lattice-related defects which control resistance to plastic flow are not of
direct concern. They are important in so far as the resistance to plastic flow is related to the
materials susceptibility to fracture.
Fracture mechanics, which is the subject of this course, is concerned almost entirely with
fracture-dominant failure

WHAT IS FRACTURE MECHANICS


Fracture mechanics is mechanics of solids containing planes of displacement
discontinuities (cracks) with special attention to their growth
Fracture mechanics is a failure theory that
1. Determines material failure by energy criteria, possibly in conjunction with strength (or
yield) criteria
2. Considers failure to be propagating throughout the structure rather than simultaneous
throughout the entire failure zone or surface
The presence of a crack in a component of a machine, vehicle, or structure may weaken it so
that it fails by fracturing into two or more pieces. This can occur at stresses below the
materials yield strength, where failure would not normally be expected.

Where cracks are difficult to avoid, a special methodology called fracture mechanics can
be used to aid in selecting materials and designing components to minimize the possibility
of fracture.

In addition to cracks themselves, other types of flaws that are crack-like in form may
easily develop into cracks, and these need to be treated as if they were cracks.

Examples:
deep surface scratches or gouges, Voids in welds,
Inclusions of foreign substances in cast and forged materials, and
Delaminations in layered materials
The study and use of fracture mechanics is of major engineering importance simply because
cracks or crack-like flaws occur more frequently than we might at first think.

For example, the periodic inspections of large commercial aircraft frequently reveal cracks,
sometimes numerous cracks, that must be repaired.

Cracks or crack-like flaws also commonly occur in ship structures, bridge structures,
pressure vessels and piping, heavy machinery, and ground vehicles.

Prior to the development of fracture mechanics in the 1950s and 1960s, specific analysis
of cracks in engineering components was not possible.

Engineering design was based primarily on tension, compression, and bending tests,
along with failure criteria for nominally uncracked material.

Such methods automatically include the effects of the microscopic flaws that are inherently
present in any sample of material. But they provide no means of accounting for larger
cracks, so their use involves the implicit assumption that no unusual cracks are present
Notch-impact tests, do represent an attempt to deal with cracks. These tests provide a
rough guide for choosing materials that resist failure due to cracks, and they aid in
identifying temperatures where particular materials are brittle.

But there is no direct means of relating the fracture energies measured in notch-impact
tests to the behavior of an engineering component.

In contrast, fracture mechanics provides materials properties that can be related to


component behavior, allowing specific analysis of strength and life as limited by various
sizes and shapes of cracks

Hence, it provides a basis for choosing materials and design details so as to minimize the
possibility of failure due to cracks.
Effective use of fracture mechanics requires inspection of components, so that there is
some knowledge of what sizes and geometries of cracks are present or might be present.

For example, periodic inspections are commonly performed on large aircraft and bridges
so that a crack cannot grow to a dangerous size before it is found and repaired.

Methods of inspection for cracks include not only simple visual examination, but also
sophisticated means such as X-ray photography and ultrasonic.

Repairs necessitated by cracks may involve replacing a part or modifying it, as by


machining away a small crack to leave a smooth surface, or by reinforcing the cracked
region in some manner
The process of fracture can, in most cases, be subdivided into the following categories:
1. Damage accumulation.
2. Nucleation of one or more cracks or voids.
3. Growth of cracks or voids. (This may involve a coalescence of the cracks or voids.)

Damage accumulation is associated with the properties of a material, such as its atomic
structure, crystal lattice, grain boundaries, and prior loading history.

When the local strength or ductility is exceeded, a crack (two free surfaces) is formed.

On continued loading, the crack propagates through the section until complete rupture
occurs.

Linear elastic fracture mechanics (LEFM) applies the theory of linear elasticity to the
phenomenon of fracture -- mainly, the propagation of cracks.

If we define the fracture toughness of a material as its resistance to crack propagation,


then we can use LEFM to provide us with a quantitative measure of fracture toughness
Various standardization bodies, such as the American Society for Testing and Materials
(ASTM), British Standards Institution (BSI), and Japan Institute of Standards (JIS), have
standards for fracture toughness tests

we will develop a quantitative understanding of cracks. It is very important to calculate


the stresses at the tip (or in the vicinity of the tip) of a crack, because these calculations
help us answer a very important practical question

At what value of the external load will a crack start to grow?


The microstructure of a material has
a great influence on its fracture
behavior.

Figure shows, schematically, some


important fracture modes in a variety
of materials.

Metals fail by two broad classes of


mechanisms: ductile and brittle
failure.
Ductile failure occurs by (a) the nucleation, growth, and coalescence of voids, (b)
continuous reduction in the metals cross-sectional area until it is equal to zero, or (c)
shearing along a plane of maximum shear.

Ductile failure by void nucleation and growth usually


starts at second-phase particles. If these particles are
spread throughout the interiors of the grains, the
fracture will be transgranular (or transcrystalline).
If these voids are located preferentially at grain boundaries,
fracture will occur in an intergranular (or intercrystalline) mode.

The appearance of a ductile fracture, at high magnification (500


or higher) is of a surface with indentations, as if marked by an ice
cream scooper. This surface morphology is appropriately called
dimpled
Rupture by total necking is very rare, because most metals contain second-phase
particles that act as initiation sites for voids. However, high-purity metals, such as copper,
nickel, gold, and other very ductile materials, fail with very high reductions in their areas
Brittle fracture is characterized by the
propagation of one or more cracks through
the structure.

For metals and ceramics, two modes


of crack propagation: transgranular
fracture (or cleavage) and intergranular
fracture are observed. For energy-
related reasons, a crack will tend to
take the path of least resistance. If this
path lies along the grain boundaries,
the fracture will be intergranular.
Often, a crack also tends to run along specific crystallographic planes, as is the case for
brittle fracture in steel. Upon observation at high magnification, transgranular brittle fracture is
characterized by clear, smooth facets that have the size of the grains. In steel, brittle fracture
has the typical shiny appearance, while ductile fracture has a dull, grayish aspect.

In addition to brittle fracture, polymers undergo a


mode of fracture involving crazing, in which the
polymer chains ahead of a crack align
themselves along the tensile axis, so that the
stress concentration is released. Another mode
of deformation that is a precursor to fracture is
the phenomenon of shear banding in a polymer.
If one stretches the polymer material, one
observes the formation of a band of material
with a much higher flow stress than exists in the
unstretched state. Shear banding (or
localization) is also prevalent in metals.
Composites -- especially fibrous ones -- can
exhibit a range of failure modes that is
dependent on the components of the material
(matrix and reinforcement) and on bonding. If
the bond strength is higher than the strength of
the matrix and reinforcement, the fracture will
propagate through the latter. If the bonding is
weak, one has debonding and fiber pullout. In
compression, composites can fail by a kinking
mechanism, also shown in the figure; the fibers
break, and the entire structure rotates along a
band, resulting in a shortening of the composite.
This mechanism is known as plastic
microbuckling.
Griffith proposed a criterion based on a thermodynamic energy balance

He pointed out that two things happen when a crack propagates:


Elastic strain energy is released in a volume of material, and two new
crack surfaces are created, which represent a surface-energy term.
Thus, according to Griffith, an existing crack will propagate if the elastic
strain energy released by doing so is greater than the surface energy
created by the two new crack surfaces

Figure 7.7(a) shows an infinite plate of thickness t that contains


a crack of length 2a under plane stress. As the stress is
applied, the crack opens up. The shaded region denotes the
approximate volume of material in which the stored elastic
strain energy is released (Figure 7.7(b)). When the crack
extends a distance da on the extremities, the volume over
which elastic energy is released increases as shown in Figure
7.7(c).
The Griffith concept was first related to brittle fracture of metallic materials by Zener and
Hollomon in 1944. Soon after, Irwin pointed out that the Griffith-type energy balance must
be between (i) the stored strain energy and (ii) the surface energy plus the work done in
plastic deformation. Irwin defined the energy release rate or crack driving force, G, as
the total energy that is released during cracking per unit increase in crack size.
In a manner similar to that of Griffith, Irwin made a fundamental contribution to
the mechanics of fracture when he proposed that fracture occurs at a stress that
corresponds to a critical value of the crack extension force

G is sometimes called the strain energy release rate.

Now, Ue = a22/E, the energy released by the advancing crack per unit of plate
thickness. This is for plane stress. For plane strain, a factor of (1 -- 2) is introduced in
the denominator. Thus,
fracture is givenby Eq. (7-17a), but t 3a0 the fracture stress is given by Eq. (7-10).

In the above sentence, it is not the Eq. (7-15) actually it is Eq. (7-17a)

Você também pode gostar