Sie sind auf Seite 1von 35

!

"#$
Materials & NDE Foundation Course

Introduction to Fracture
22 June 2009 to 26 June 2009

h:\data/a8/wp/training1/fracture Rev.2

FRACTURE LECTURE NOTES Index


1. Introduction
1.1 Overview............................................................................................... 3 1.2 Control of Fracture............................................................................... 6

2. Mechanisms of Fracture
2.1 Stress Concentrations........................................................................... 7 2.2 Fracture Mechanisms 2.2.1 Brittle Fracture........................................................................8 2.2.2 Ductile Failure........................................................................ 10 2.3 Ductile to Brittle Transition................................................................. 11 2.4 Structural Thickness and Constraint..................................................12

3. Toughness Testing
3.1 The Charpy Test 3.1.1 Introduction............................................................................ 13 3.1.2 Test Procedure ....................................................................... 15 3.2 Fracture Toughness Testing................................................................ 17 3.3 Crack Arrest Tests.................................................................................19

4. Material Effects
4.1 Definition of the Ductile to Brittle Transition Temperature........... 21 4.2 Chemical Composition.........................................................................22 4.3 Microstructure Effects.......................................................................... 23 4.4 Strain Rate.............................................................................................. 24 4.5 Service Embrittlement.......................................................................... 24

5. Toughness Of LR Materials
5.1 Normal Strength Ship Steels............................................................... 25 5.2 Higher Strength Ship Steels.................................................................25

6. Material Selection For Fracture Control


6.1 General LR Rule Requirements for Ships..........................................27 6.2 Ice-breaking Ships.................................................................................28 6.3 Liquefied Gas Ships.............................................................................. 29

7. Fracture Mechanics
7.1 Introduction........................................................................................... 31 7.2 Design against Failure..........................................................................33

8. Summary......................................................................................................... 35

Page 2 of 35

1. Introduction
1.1 Overview
Fracture is defined as the separation of a solid body into two or more parts under the action of stress. This process consists of two parts: crack formation (initiation) and crack growth. Obviously for marine applications, it is important to design against fracture as even a localised structural failure can have catastrophic consequences for the integrity of a vessel. Fracture can occur rapidly under constant load, but slow crack growth can occur under the influences of alternating load (fatigue) or chemical attack (stress corrosion cracking). In this paper, we shall consider the rapid fracture of metals under constant load. There are two general types (modes) of fracture: brittle and ductile. Brittle fracture involves hardly any deformation, and often occurs in materials subjected to applied stresses far below the yield strength. The crack growth can be very rapid (up to 1km/s) and unstable, so failure is often catastrophic and without warning. Ferritic steel structures are particularly susceptible to brittle fracture at low temperatures. Brittle fracture has only become a major problem this century with the introduction of fusion-welded structures. With old-style riveted constructions, a crack in a plate would usually be arrested at the plate edge. However, with an all-welded fabrication, the crack can propagate through the entire structure. In a ship for example, this can result in the vessel breaking in two. The classic examples of this are the Liberty ships, built in the USA during World War II (Figure 1). These failures and others led to the introduction of toughness requirements for ship steels into LR Rules (1958).

Figure 1 - Brittle fracture of the Liberty Ship USS Schenectady (1943)

Page 3 of 35

However, there have been more recent examples. In 1979, the bulk carrier MV Kurdistan broke in two (Figure 2). Here the crack was shown to have started at a defective butt weld in the ground bar attaching the bilge keel to the side shell. The crack then grew into the bilge keel itself, through the side shell, and propagated round the entire vessel. Since then LR Rules have been introduced prohibiting the bilge keel from being directly welded to the side shell and introducing rat holes into bilge keel butt welds, to prevent any cracks from propagating into the main structure (LR Rules Part 3 Chapter 10 Section 5.6).

crack initiated at various sites in ground bar weld crack propagated via intermittent weld into bulb bar

ground bar butt weld

crack propagated via ground bar to shell weld into bilge strake and shell plate

Figure 2 - Fracture initiation of Kurdistan failure

Page 4 of 35

Catastrophic brittle fractures have also occurred in bridges, pipelines, pressure vessels (Figure 3) and electrical turbo generators.

Figure 3 - Brittle fracture of a pressure vessel, initiated from a defective weld (1958)

Ductile fracture is always accompanied by a large amount of (plastic) deformation. Failure, which occurs by slow and stable crack growth, will only occur when the yield stress of the material is exceeded locally. Under normal circumstances, structural metals are designed to fail by ductile mode, but this will only happen if the structure is overloaded (Figure 4).

Figure 4 - Ductile fracture following explosion of a fuel pipeline (1997)

Page 5 of 35

1.2 Control Of Fracture


The common feature of all of these failures is the total loss of the structure, which inevitably has significant safety and economic consequences. There are generally common features between incidents of brittle fracture, typified by the Liberty ship failures: cracks or notches in the structure (e.g. weld defects) high local tensile stress (e.g. hatch cover corners) low ambient temperature (e.g. Arctic waters) Therefore, to combat the incidence of brittle fracture, the following precautionary measures can be taken: avoid excessive tensile stress by attention to the construction details use qualified welding procedures and carry out NDE on welds to ensure that they do not contain significant defects. select a steel that will not behave in a brittle manner at the service temperature. It was the need for adequate resistance to cracking, defined as the toughness of steel, that was not properly understood at the time of construction of the Liberty ships. As a direct result of these failures, a number of toughness tests were developed to ensure that ship steels had adequate resistance to fracture. It is important to note that as cracks in service developed from cracks or notches, the laboratory test testpieces had to contain similar discontinuities.

Page 6 of 35

2. Mechanisms Of Fracture
In order to understand how the fracture of a structure can be prevented, it is useful to briefly consider the mechanisms of the fracture process.

2.1 Stress Concentrations


A critical factor in determining whether or not fracture will occur is the level of tensile stress within the structure. However, the stress level is not constant throughout a structure. Geometrical discontinuities, such as holes and notches, produce a local stress concentration that can far exceed the general stress level applied. This concentration tends to increase with decreasing radius of the discontinuity. Thus a volumetric defect, such as a gas pore in a weld, will only provide a slight stress concentration. However a planar defect, such as a crack, can produce a stress of yield magnitude with only a small applied stress (Figure 5).

stress concentration

2.5

1 0 0.2
discontinuity radius (mm)

crack tip

1.5

0.4

0.6

0.8

gas pore

Figure 5 - Variation of stress concentration with radius of discontinuity (note that this is only an example - the actual level of stress concentration varies according to discontinuity orientation and structural geometry)

Welds are a common source of discontinuities which, depending on their orientation, may initiate brittle fracture. If the weld profile itself is not smooth (e.g. undercut), a severe stress concentration can arise. On a larger scale, other common stress raisers in ships include hatch cover openings and ratholes at the heel of stiffener brackets. Note that fracture is most likely to initiate in the midships of a vessel, where the applied tensile stresses are at a maximum. Therefore, higher toughness steels tend to be used in this area. Page 7 of 35

It must be remembered that in addition to the applied stress, residual stresses will be present within a fabricated structure, which will reinforce the applied stress. Contraction of the weld metal during cooling will set up significant stresses (of up to yield strength in magnitude) in the weld and adjacent parent material. These can be decreased by a suitable post weld heat treatment, which will, therefore, reduce the susceptibility of the structure to fracture.

2.2 Fracture Mechanisms


2.2.1 Brittle Fracture When a material fails by brittle fracture, very little deformation occurs within the structure. Instead the material literally cleaves in two. Failure is produced by tensile stresses, thus the fracture surface is very flat and perpendicular to the tensile axis. In plate materials, brittle fractures show a characteristic chevron pattern (Figure 6). These chevron markings have a useful property, in that they show the direction of crack propagation by pointing back to the origin of fracture. Therefore, the possible cause of the brittle crack initiation may be traced.

direction of crack growth Figure 6 - Chevron markings on brittle fracture surface of grade A plate

Brittle fracture is initiated by a microcrack, which can form at a brittle intermetallic inclusion or other microstructural discontinuity. Once a microcrack has formed, the large stress concentration ahead of the crack tip allows the crack to rapidly propagate through the structure, cleaving grains in its path. The initiation of a microcrack requires only a small amount of energy, and, as there is little plastic deformation, there is only a small resistance to crack growth. Thus a brittle material has a low toughness. As there is little ductile deformation at the crack tip, significant crack tip blunting does not occur, thus a brittle crack may grow right through a structure with little chance of crack arrest (providing that the load is maintained). A brittle fracture surface has a brittle, shiny appearance. Under an electron microscope, the fracture surface can be seen to be composed of a number of angular flat facets, corresponding to the fracture faces of the cleaved grains (Figure 7). Page 8 of 35

Figure 7 - Brittle (cleavage) fracture of mild steel

(x250)

The general brittle fracture mechanism described above is that of transgranular (across the grains) fracture. However, brittle fracture may also occur by an intergranular (between the grains) mechanism (Figure 8). An example of this is a result of temper embrittlement, which may occur in low alloy steels if they are heat treated at an incorrect temperature. Residual elements segregate along the grain boundaries resulting in local embrittlment. This provides a path along which brittle fracture can occur, even though the bulk of the material retains a high toughness.

intergranular crack

Figure 8 - Intergranular brittle fracture in low carbon steel steel

x50

Page 9 of 35

2.2.2 Ductile Failure Ductile fracture (or shear) occurs as a result of extensive deformation of the microstructure, and is promoted by shear stresses within the material. These stresses reach a maximum at an angle of 45 to the tensile axis, and ductile crack growth will tend to take place along this plane. This results in the characteristic angular appearance of a ductile fracture surface (Figure 9).

ductile fracture surface

Figure 9 - Ductile fracture surface showing shear at 45 to tensile axis in grade A plate

When the material is locally stressed above the yield point, shear (plastic) deformation occurs. As the stress in increased, the deformation becomes more extensive until small holes (microvoids) become nucleated at microstructural discontinuities (such as inclusions). These microvoids then grow together (coalesce), resulting in fracture (Figure 10). The important thing to note about ductile fracture is that because a large amount of plastic deformation occurs, a great deal of energy is required to produce failure. Thus there is a large resistance to crack growth, so the toughness of a ductile material is high. Once crack growth has initiated, ductile deformation tends to blunt the crack tip, thus reducing the local stress concentration. Therefore, ductile fractures often stop (arrest) with relatively little crack growth.
plastic deformation ahead of crack tip nucleation of large void at inclusion linkage of crack with void linkage of crack with other voids

increasing stress

Figure 10 - Schematic of ductile fracture mechanism

Page 10 of 35

At low magnifications, a ductile fracture surface appears dull and fibrous. However, under a microscope the characteristic fracture surface of ductile microvoid coalescence can be clearly seen (Figure 11).

Figure 11 - Ductile fracture surface of C-Mn steel

(x250)

2.3 Ductile To Brittle Transition


As the ambient temperature is lowered, it becomes progressively more difficult for ductile fracture to occur, as plastic deformation requires thermal energy to operate. Eventually a point will be reached at which there is insufficient thermal energy for ductile failure to occur, and brittle fracture will result. This is known as the ductile to brittle transition temperature (DBTT). Plain carbon manganese steels typically have a DBTT of between -60 and +20C. Note that the fracture mode does not instantly switch at the transition temperature, but changes gradually over a temperature range of 20-30C. Thus fracture surfaces are often of mixed-mode appearance, and contain both ductile and brittle regions (Figure 12).

ductile tear brittle facet

Figure 12 - Mixed mode fracture surface of C-Mn steel

(x500)

Page 11 of 35

The ductile to brittle transition will usually only occur in materials with a body-centred cubic (bcc) structure, such as ferritic steels. Materials with a face-centred cubic (fcc) structure, such as aluminium and copper alloys and some stainless steels are much more ductile, and will remain so even at very low temperatures. However that these materials can become embrittled if brittle compounds are precipitated within the structure, such as by an incorrect heat treatment.

2.4 Structural Thickness And Constraint


The fracture mode can also be influenced by the thickness of a structure. The stress condition within a material is not uniform throughout the depth of a section (Figure 13). In the central region, the stress is distributed in three dimensions (triaxial), a condition known as plane strain. Here deformation of the microstructure is restricted by the triaxial stress state, and it is difficult for ductile fracture to occur (i.e. stresses above the uniaxial yield stress are required to cause yield). This is known as a region of high constraint. Towards the surface of a material, the constraint is reduced, as it is not possible for a triaxial stress state to exist. Here the stress state is two dimensional or biaxial (known as plane stress) and it is much easier for ductile failure to occur. This accounts for the characteristic shear lips which can be found at the edge of (even brittle) fracture surfaces.
stress,

plane strain (triaxial stresses x, y, z)

plane stress (biaxial stress x,z)

Figure 13 - Change in stress state through depth of section

It should be noted that constraint is not the only reason that a thick section is particularly susceptible to brittle fracture. During processing of the material, it is not possible to produce a uniform microstructure throughout a thick section, thus the central region often has poorer mechanical properties than the surface. Page 12 of 35

3. Toughness Testing
In order to avoid brittle fracture in a structural material, it is necessary to perform tests to measure the materials resistance to fracture, i.e. the toughness. There is no single method of measuring the toughness of a material. Tests can be grouped into three main categories, which measure different aspects of the material toughness: Charpy tests fracture toughness testing crack arrest tests It is important to note that as cracks in service developed from cracks or notches, the laboratory test testpieces must contain similar discontinuities.

3.1 The Charpy Test


3.1.1 Introduction
2mm deep V-notch

10mm

55mm 10mm standard specimen

IMPACT LOAD

Figure 14 - The Charpy test

By far the most widely used test within industry for assessing the toughness of metals is the Charpy V-notch impact test (Figure 14). This measures the energy required to fracture a standard testpiece. The testpiece is loaded by a weighted pendulum at a very high strain rate (approximately 103/sec) and the energy required to produce total fracture (crack initiation and propagation) is measured. Note that similar tests with other shaped notches (such as the U-notch test) are sometimes used, but the V-notch is by far the most common and is the test required by LR Rules. Page 13 of 35

Test machines normally measure the energy absorbed during fracture in Joules (J) over a range of 0-300J. If the energy absorbed during impact is relatively high, a large amount of plastic (ductile) deformation will have occurred within the testpiece, thus the fracture mode is ductile. However, if the impact energy is relatively low, little ductile failure has occurred and the fracture mode will be mostly brittle. Standards (including LR Rules) usually quote a requirement for a minimum impact energy at a set temperature to guard against brittle fracture. As well as impact energy, the fracture mode can be deduced from the appearance of the testpiece (Figure 15). A ductile testpiece will have a fibrous (dull) fracture surface, and will be heavily deformed, thus the lateral expansion of the testpiece is high. A brittle testpiece will have a crystalline (shiny) fracture surface, and the lateral expansion will be low. Minimum lateral expansion and maximum percentage fracture surface crystallinity values are sometimes required by materials standards, but not by LR Rules. DUCTILE MIXED-MODE BRITTLE

189J impact energy 0% crystalline 2.1mm lateral expansion

83J impact energy 50% crystalline 1.3mm lateral expansion

21J impact energy 90% crystalline 0.2mm lateral expansion

Figure 15 - Fracture surface appearance according to failure mode (grade A material)

The overwhelming advantage of a Charpy test is that it is a quick, simple and cheap test to conduct. It is also easy to perform tests at a range of sub-ambient temperatures. Therefore, the test is widely used for the batch testing of materials for quality control and acceptance purposes. However, a problem with the Charpy test is the small size of the testpiece. The testpiece thickness may be much smaller than the structure for which material is being used, thus the constraint of the testpiece is lower. Therefore, a more ductile fracture tends to occur in the testpiece than would happen in practice. This makes it impossible to predict the behaviour of a cracked structure, and thus design against failure. Also, the high rate of loading of the testpiece is not usually representative of that in a real structure, and a machined notch does not truly reflect a crack in a real structure.

Page 14 of 35

3.1.2 Test Procedure Charpy testing should be performed to BSEN 10045-1 (Charpy Test for Metallic Materials, 1990) or an equivalent National Standard. Note that the test is particularly sensitive to the size and shape of the machined notch, so this should be checked before testing, using a suitable gauge. Details of the test procedure requirements are contained within the LR Rules Chapter 2 Section 3. There is a degree of scatter in Charpy test data, particularly within the ductile to brittle transition temperature region. Therefore, it is standard practice to conduct the test upon three testpieces and calculate the average impact energy of the set. This average must meet the minimum impact energy requirement of the specification. However, according to LR Rules, one individual value less than this level can be accepted, provided that it is no less than 70% of the required energy. Further explanation of the acceptance and re-testing procedures for Charpy testing are contained within the Survey Procedures Manual Part B Chapter 1. If the material being tested is less than 10mm thick, it is not possible to produce a full-sized Charpy testpiece. In this case a sub-sized testpiece (7.5x10x55mm or 5x10x55mm) can be used for testing. Obviously the constraint of these thin testpieces is even less than that of a standard testpiece, thus the minimum impact energy requirement is not directly reduced accordingly to fracture surface area. LR Rules specify the following minimum energy requirement: testpiece 5x10x55mm testpiece 7.5x10x55mm 2/3 of tabulated energy 5/6 of tabulated energy

For rolled materials, such as plates and sections, the microstructure is directionally deformed (anisotropic) and so the fracture properties vary according to notch orientation (Figure 16).
125 Charpy impact energy (J)
longitudinal

100 75 50 25 0 -60

transverse

rolling direction

transverse

longitudinal
-40 -20 0 test temperature (C) 20

Figure 16 - Variation of Charpy impact energy with testpiece orientation for mild steel (note - this is an example only and should not be used as a direct comparison)

Page 15 of 35

A longitudinal testpiece (notch perpendicular to the rolling direction) will generally have a higher Charpy impact energy than a transverse (notch parallel to rolling direction) testpiece. LR Rules now contain different impact energy requirements according to the testpiece orientation. Note that for a cross-rolled plate, the direction of the final pass is taken as the principal rolling direction. Here the anisotropy in the Charpy values is much reduced. When witnessing an impact test, it is important to note the type of Charpy machine used. The pendulum hammer radius (2mm) of BSEN/ISO standard machines is smaller than that of ASTM type machines (8mm), which affects the loading geometry at the notch. This can give different test results for identical material, particularly if the impact energy is large, as shown in Figure 17. LR Rule requirements are for BSEN/ISO machines.

350 Charpy impact energy (J) 300 250 200 150 100 50 0 -80 -60 -40 -20 0 test temperature (C) 20 40
ISO ASTM

Figure 17 - Variation in measured Charpy impact energy for ASTM and ISO standard impact test machines for a structural steel (note - this is an example only and should not be used for energy conversions between the two tests)

To counter the limited constraint of a Charpy testpiece, plates over 50mm thick, which are particularly susceptible to brittle fracture, usually have an increased Charpy energy for acceptance purposes.

Page 16 of 35

3.2 Fracture Toughness Testing


The fracture toughness of a material is a measure of its resistance to crack initiation. Unlike Charpy toughness, which is limited by the constraint of a testpiece, the fracture toughness of a material may be a material property. Therefore, the fracture toughness of a material may be used for design purposes to predict the failure of a cracked structure (design against failure). The theory behind toughness and structural integrity (known as fracture mechanics) will be discussed later in the paper. There are several different parameters which can be used to assess the fracture toughness of a material, which include: KIC - plane strain fracture toughness JIC - J-integral fracture toughness test CTOD - crack tip opening displacement The test procedures for these are all described in BS 7448-1 (Fracture Mechanics Toughness Testing, 1991). Generally, tests are conducted using very slow (quasi-static) loading rates, although high loading rate testing is occasionally required for high strain rate applications. KIC and JIC are absolute material properties independent of thickness. However, this is only true for KIC above a minimum thickness, where the testpiece is of high constraint (i.e. a plane strain state predominates), as shown in Figure 18.
250 200

B C
A - plane stress B - transition C - plane strain

KC 150 (MPam) 100


50 0 1

KIC

10 specimen thickness (mm)

100

Figure 18 - Typical variation of KC with testpiece thickness

As well as testpiece thickness, the constraint of a testpiece is affected by its ductility. Thus for structural steels, which are normally partially ductile at ambient temperature, the minimum testpiece thickness at which plane strain conditions are predominant (which can be calculated from the test standard) can be in excess of 1 metre.

Page 17 of 35

This thick section size is impractical for testing and not relevant to real structures, thus CTOD testing is generally used for ductile materials (Figure 19). This is the only fracture toughness test that is mentioned within LR requirements, such as for the approval of a stud link chain cable manufacturer (LR MQPS Book H Procedure 10-1 Section 6), weldability tests (LR MQPS Book A Procedure 0-3 Section 3), and approval of controlled rolling procedures (LR MQPS Book C Procedure 3-3 Section 5).

Figure 19 - A CTOD fracture toughness test

The CTOD test is described in the LR SPM Part B Chapter 2 Section 6. A graphical record of force versus clip gauge extension is measured. The point of crack initiation can be determined from a discontinuity in the forceextension curve (Figure 20), and the amount by which the crack tip has opened (, measured in mm) is calculated. The resistance to crack initiation (i.e. fracture toughness) increases with increasing .
critical force

A - elastic strain B - stable (ductile) crack growth C - unstable crack growth

A
extension
Figure 20 - Typical force-displacement curve for CTOD test of a structural steel

Page 18 of 35

Unlike KIC and JIC, the CTOD is not a true material property, thus the test should always be conducted on a full thickness testpiece with regard to the structural application. Testpieces are usually of an SENB (single edge notch bend) or CT (compact tension) geometry (Figure 21). Prior to testing, a fatigue pre-crack is grown from the machined notch under alternating load, in order to provide a sharp notch (thus high stress concentration) to initiate fracture.
fatigue precrack

SENB specimen CT specimen

Figure 21 - Typical fracture toughness testpieces

3.3 Crack Arrest Tests


Fracture toughness tests are used for predicting the susceptibility of a cracked structure to fracture. However, they only relate to the initiation of a crack, thus do not predict whether a growing crack will arrest. If a brittle fracture initiates in a material, it is important to try to ensure that it does not grow through the entire structure. The crack arrest test can be used to determine this will occur. There are several crack arrest tests, the most common being the Drop Weight Tear or Pellini test which is described in ASTM E208 (Standard Method for Drop Weight Test of Ferritic Steels, 1995). There are more sophisticated fracture mechanics crack arrest tests, but these are beyond the scope of this paper.

Page 19 of 35

The Pellini test was developed by the US Navy Research Establishment following the Liberty ship failures. In order to initiate a brittle crack, a deliberately brittle weld bead is deposited upon a plate of full thickness and subsequently notched. A weight with a known kinetic energy is then dropped onto the opposite face under controlled conditions (Figure 22), and the test plate is deformed by a fixed displacement.
machined notch

weld bead

Pellini testplate

Figure 22 - The drop weight tear (Pellini) test

Following the test, two criteria for acceptance/rejection exist, subject to visual examination of the propagated crack (Figure 23). The first is the break condition. Here the crack must reach at least one edge of the plate. The second is the no-break condition, in which the crack has been found to reach neither edge.
no-break break

Figure 23 - Schematic of break and no-break condition Pellini testpieces

Tests are conducted upon at least six successive specimens to gauge the critical break/no break test temperature. An iterative process is to increase or decrease the test temperature of each subsequent test (according to a National standard procedure), in order to hone in upon the critical temperature. The highest temperature at which a break result occurs is defined as the Nil Ductility Transition Temperature (NDTT). Provided that the design temperature does not fall below some margin above the NDTT, it may be assumed that a brittle crack will not propagate through a structure. Determination of the NDTT is required for the approval of certain steelworks, (e.g. LR MQPS Book C Procedure 3-1 Section 5),

Page 20 of 35

4. Material Effects
It has already been shown that structural steels are susceptible to embrittlement below a certain temperature. This ductile to brittle transition temperature (DBTT) is affected by a number of material parameters including: chemical composition grain size and microstructure strain rate The DBTT must be low enough to ensure an adequately small risk of brittle fracture at normal operating temperatures. Firstly, it is important to define what the DBTT is.

4.1 Definition Of The Ductile To Brittle Transition Temperature


There are various methods of defining the ductile to brittle transition temperature. The most common method used in industrial specifications is to specify the transition as occurring at some set Charpy impact energy value. For structural steels, this is usually taken as the temperature at which an impact energy of 27J is achieved (Figure 24), and this is widely used within LR Rules. This implies that structural steels above this energy level will fracture in a largely ductile mode, and hence be resistant to brittle fracture at the test temperature.

Charpy impact energy 27J

ductile

brittle DBTT test temperature Figure 24 - Typical Charpy impact energy transition temperature for a structural steel

Page 21 of 35

An alternative method of estimating the DBTT is by plotting a transition curve in terms of percentage crystallinity (Figure 25). Here the transition temperature (known as the fracture appearance transition temperature, FATT) is defined as that at which a fracture surface of 50% ductile, 50% crystalline appearance is achieved. Above this temperature, ductile failure predominates and hence the material will be more resistant to brittle failure
CX Charpy impact energy (CV) brittle fracture surface 50% crystallinity (CX) ductile CV DBTT test temperature Figure 25 - Transition temperature for a structural steel, based upon crystallinity

4.2 Chemical Composition


The DBTT of a material can be altered by varying the chemical composition of a material. Figure 26 summarises the effects of various alloying elements upon the transition temperature of a mild steel.

brittle Charpy impact energy

ductile

Mn Ni Al
DBTT

C P Si Mo O
test temperature

Figure 26 - Variation of Charpy transition curve of a mild steel with alloying additions

Carbon, which is added to strengthen the metal, and certain other elements raise the transition temperature and can, therefore, embrittle the material. Elements that lower the transition temperature such as manganese improve the toughness. Carbon raises the DBTT by 15C for every 0.1wt% added, Page 22 of 35

whereas manganese lowers it by 5C for every 0.1wt.% added. Therefore, to counter the effect of increased carbon content manganese should be added in a ratio of 3 to 1. LR Rules for Grade A rolled steel plates specify a manganese to carbon ratio of at least 2.5 to 1 in order to achieve adequate toughness (LR Rules Part 2 Chapter 3 Section 2.2). Other elements, such as nickel, can be used to produce further reductions in the DBTT. Note, however, that care should be taken with such highly alloyed steels. An incorrect heat treatment can embrittle the microstructure and drastically increase the DBTT. An example of this is temper embrittlement, which can occur in ferritic steels containing chromium, molybdenum or nickel, during slow cooling between 600 and 300C. If impurity elements (especially antimony, phosphorous, tin or arsenic) are present, these segregate to the grain boundaries and cause intergranular fracture. Tests are sometimes required (LR Rules Part 2 Chapter 2 Section 5.1) to ensure that a material is not susceptible to temper embrittlement.

4.3 Microstructure Effects


The processing method of a rolled material, which can alter the microstructure, can also affect the DBTT. Materials with a fine grain size generally have a lower DBTT and therefore improved toughness. Decreasing the grain size of a mild steel from ASTM 5 to ASTM 10 can lower the DBTT by up to 60C (Figure 27). Note that an additional advantage of decreasing the grain size is that the yield strength will be increased.
0 DBTT -10 DBTT (C) -20 -30 -40 -50 5 6 7 8 ASTM grain size 9 10 yield strength

400
yield strength (MPa)

350 300 250 200 150

Figure 27 - Typical variation of yield strength and toughness of a mild steel with grain size (note - this is an example only and will vary according to the material)

Page 23 of 35

The factors that can be used to decrease the grain size of steels include: adding grain refining alloy additions (e.g. Al, Nb, V) lowering the finishing temperature in as-rolled steels normalising a finished product using a thermo-mechanical controlled process (TMCP) This can be a problem in the heat affected zone (HAZ) of a weld, where grain growth may occur (depending on the heat input of the process), which will thus raise the transition temperature.

4.4 Strain Rate


The toughness of a material will also be affected by the rate at which it is loaded. For a material to fracture in a ductile manner, plastic deformation is required. This takes a finite amount of time, whereas brittle fracture (cleavage) can occur almost instantly. Therefore, if a ductile material is shock loaded (such as the wave slamming experienced against a ship hull in a storm), there may be insufficient time for it to undergo ductile deformation, and it may show a tendency towards brittle fracture. It is not possible to alter the loading rate of a Charpy test, although this is a high loading rate test, and thus will provide a conservative measure of the transition temperature. However, for fracture toughness testing, the DBTT has been shown to increase by up to 40C by changing the loading rate from slow (quasi-static) to shock loading.

4.5 Service Embrittlement


It is important to realise that the toughness of a material can degrade during service. An example of this is neutron embrittlement, which can be experienced in nuclear reactor vessels. However, the most important service embrittlement mechanism is strain ageing. Strain ageing may occur in ferritic steels with a high free nitrogen content. If the material is cold strained (such as during fabrication processes such as cold forming or line heating), nitrogen atoms diffuse within the structure to these strained areas and produce an embrittling effect. This can increase the DBTT of the material by as much as 60C. To counter this, LR recommends that the nitrogen level of the ladle analysis should not exceed 90ppm or 0.009wt.% (SPM Part B Chapter 3 Section 1.2), and that the cold strain applied to steel sections should not exceed 5% without some form of stress relieving heat treatment (LR Rules Part 2 Chapter 13 Section 7.3). In addition, strain age embrittlement tests (LR Rules Part 2 Chapter 2 Section 5.2) are required during the works approval of steel plates and sections, to ensure the materials are not susceptible to strain ageing.

Page 24 of 35

5. Toughness Of LR Materials
This section will deal with the LR Rule requirements for the toughness (Charpy) testing of materials. As discussed in earlier sections, compositional requirements have been introduced into LR Rules to ensure adequate material toughness. Minimum ductility levels of tensile test specimens are also required to guard against brittle fracture.

5.1 Normal Strength Ship Steels


There are four toughness grades for LR normal strength ship steels, which are classified according to achieving a minimum Charpy impact energy at a specified test temperature (LR Rules Part 2 Chapter 3 Section 2.4): Grade A tested at +20C Grade B tested at 0C Grade D tested at -20C Grade E tested at -40C It is important to realise that this is a test temperature only, and does necessarily reflect the minimum temperature to which a structure can be exposed. This is dealt with in the next chapter (materials selection). For steels with a thickness of less than 50mm, an impact energy of 27J is considered sufficient to avoid brittle fracture at the specified test temperature. However, if the section thickness of a steel exceeds 50mm, the required minimum Charpy energy is increased accordingly, to order counter the effects of increased constraint in the structural application. Although only longitudinal oriented Charpy testpieces are usually required, transverse Charpy requirements (which are sometimes specified in a design) have a lower specified minimum energy to reflect the anisotropic toughness properties of a rolled material. Note that impact tests are not directly required for grade A material less than 50mm thick. However, regular in-house checks are required every 250t of material, to ensure that the material meets the required toughness.

5.2 Higher Strength Ship Steels


As the strength (UTS) of a steel is increased, in general its ductility and resistance to brittle fracture (toughness) are decreased. LR Rules sometimes limit the UTS, or yield to UTS ratio, in order to ensure that high strength materials have adequate ductility (LR Rules Part 2 Chapter 3 Section 3.6). Page 25 of 35

Higher strength steels also generally have a higher minimum Charpy energy requirement, which can be approximated (in J) as the minimum yield strength (in MPa or N/mm2) divided by ten: AH 32 grade (minimum yield strength 315MPa) - energy requirement 31J AH 36 grade (minimum yield strength 355MPa) - energy requirement 34J AH 40 grade (minimum yield strength 390MPa) - energy requirement 41J These steels are similarly graded to normal strength steel according to test temperature. However, there is no BH grade and an additional FH grade exists. Grade AH tested at 0C Grade DH tested at -20C Grade EH tested at -40C Grade FH tested at -60C Impact energy requirements are also varied according to section thickness and Charpy testpiece orientation.

Page 26 of 35

6. Material Selection For Fracture Control


Material selection for fracture control is the task of the naval architect designing a vessel, thus the requirements are contained within Part 3 of the LR Rules. Although this section only details materials selection according to toughness, it is important to note that other material properties (such as strength and corrosion resistance) must also be considered.

6.1 General LR Rule Requirements For Ships

Under the general LR Rule requirements for materials selection regarding fracture control, each region of a vessel is assigned one of five different material classes (LR Rules Part 3 Chapter 2 Section 2.1). For each hull member, this class reflects the: level of stress, for example the midships of a vessel tend to experience higher stresses than the bow or stern exposure to weather, as exposed material will experience lower temperatures and may corrode at a higher rate than within a sheltered region, thus may eventually experience a higher stress and be more susceptible to brittle fracture need to arrest cracks, as a brittle crack could be catastrophic for the integrity of the vessel in some hull members, e.g. a sheerstrake, but low risk in others This class is then related to a toughness grade of steel, a higher class requiring a higher grade. The required grade also increases with section thickness, reflecting the increase in constraint for thick materials. Higher toughness grade materials are also required (for certain members) if a ship operates for a extended periods in low air temperatures or contains refrigerated spaces.

Page 27 of 35

6.2 Ice-Breaking Ships

Due to the ductile to brittle transition of structural steels at lower temperatures, particular care has to be taken for ice-breaking ships. The LR requirements for ice-class vessels (LR Rules Part 3 Chapter 2 Section 2.3) allow for specification of air temperatures of between 0 and -45C. Material selection for lower air temperatures must be given special consideration. As for the normal LR Rule requirements, the hull members are classed according to location. The forward part of the vessel (which will experience the most severe ice conditions) has the most stringent material requirements. As well as stress however, the expected operating temperatures are specifically considered. These are calculated from the minimum design air temperature, making allowance for the degree of exposure to the elements. Thus the operating temperature of sheltered sections may be higher than the minimum design air temperature. The grade of material is then determined from one of two graphs (Figure 28) according to minimum air temperature and structural thickness.
0 operating temperature (C) -5 -10 -15 -20 -25 -30 -35 -40 -45 0 10 20 30

D DH E, EH FH
Boundary lines form part of the lower grade 40 50 60

material thickness (mm)

Figure 28 - Selection of steel grade for exposed regions of ice-breaking ships

Page 28 of 35

6.3 Liquefied Gas Ships

A similar situation to that of ice-breakers exists in the design of liquefied gas ships, where even lower temperatures are encountered in the containment and transport of liquid fuel gases. Examples of temperatures for common liquid gases are: propane (LPG) ethylene (LEG) methane (LNG) -45C -104C -163C

The material requirements for vessels required to transport such gases are given in the LR Rules for Ships for Liquefied Gases Chapter 6 Section 2. The problem of using C-Mn structural steels is that even with grade FH material, the lower temperature is limited to -60C. Therefore, special low temperature materials need to be employed. Figure 29 details the Charpy impact energy of various candidate materials at low temperatures.
40 35
Charpy impacy energy

30 25 20 15 10 5 0
-250
9%Ni 5%Ni 3.5%Ni stainless steel Invar

Al alloys

LT60 C-Mn

LT40 C-Mn

LT20 C-Mn

-200

-150

-100

-50

temperature (C)

Figure 29 - Typical variation of Charpy impact energy with temperature for various materials

Page 29 of 35

It can be seen that high nickel alloys (such as Invar, Fe-35%Ni), stainless steels and aluminium alloys remain ductile even at very low temperatures. However, these materials are very costly and may exhibit other problems such as poor weldability or corrosion susceptibility. A cheaper solution is to use ferritic steels with a low to medium nickel content (sufficient to provide a transition temperature suitable for liquid gas cargo) which are detailed in LR Rules Part 2 Chapter 3 Section 6. If an operating temperature of between 0 and -55C is required, special (LT designated) high strength carbon-manganese grades can be used. However, these are subject to special processing and compositional limits, and more extensive impact testing than standard high strength materials. If the required operating temperature is between -60 and -165C, nickel grade ferritic steels can be used, with a nickel content of between 1.5 to 9.0wt.%. The required material is selected according to design temperature and section thickness (Figure 30).

minimum design temperature (C)

0 -40

1.5Ni
-80 -120 -160 -200 0 5 10 15 20 25 30 35 40 material thickness (mm)
Figure 30 - Minimum design temperature for nickel grade steels

3.5Ni 5Ni

9Ni

Page 30 of 35

7. Fracture Mechanics
It has been shown that Charpy tests can be used for acceptance purposes to ensure adequate toughness in service. However, such tests cannot be used to analyse to assess the integrity of a cracked structure. The theory behind the mathematical analysis of fracture is known as fracture mechanics. This enables a designer to conduct engineering critical assessments, to demonstrate the fitness for purpose or safety of a structure with crack-like defects. It is important to realise that a cracked structure will not necessarily fail, and it is sometimes safer to leave a crack within a structure than attempting a repair, which could embrittle the microstructure or introduce residual stresses, for instance. Fracture mechanics is a highly complex subject, so this section is only intended as a brief introduction to the subject. Further information can be obtained from the LRTA paper:
Engineering Fracture Mechanics, A.Cameron and V.Pomeroy, Paper no.2, session 1985-6

7.1 Introduction
When a body containing a crack (of length a) is subjected to an applied stress (), a stress concentration () is produced ahead of the crack tip, which varies with distance from the crack tip (x). Elasticity theory suggests that as the crack tip is approached (i.e. x0), the stress concentration approaches infinity (), as shown in Figure 31.

Figure 31 - Theoretical variation of stress concentration with distance from a crack tip

However, in practice no material can withstand an infinite stress, and material close to the crack tip yields by ductile deformation, resulting in the formation of a plastic zone (Figure 32). Note that the local stress at which crack tip yielding occurs may be greater than the uniaxial (tensile) yield stress, and varies according to the specimen constraint. The yield stress is greater for a

Page 31 of 35

triaxial stress state (plane strain) than a biaxial one (plane stress). Note that even in a brittle material, some ductility always occurs at the crack tip
f plastic zone x

Figure 32 - Actual variation of stress concentration with distance from the crack tip

The stress concentration at the crack tip is quantified by a value K, the stress intensity factor. This parameter can be used to scale the stress distribution ahead of the crack tip according to the applied stress, crack length and geometry of the crack and surrounding structure (geometrical factor Y):

K = Y a
As an increasing stress is applied to a cracked structure, the stress intensity factor increases, until one of three possible failure mechanisms occurs: Brittle failure If the stress intensity factor reaches a critical value, which is known as KIC, the stress ahead of the crack tip is high enough to nucleate a brittle crack. This crack propagates in a rapid, unstable manner through the structure, causing it to fail catastrophically. The critical stress intensity factor to produce brittle fracture is called the fracture toughness of a material. KIC is a material property which can be measured in the laboratory, and can be used to predict the occurrence of brittle fracture in engineering structures. This theory is known as linear elastic fracture mechanics (LEFM), and generally holds true as long as the size of the plastic zone is small compared to the crack, and the material is thick enough to experience plain strain conditions. Stable (ductile) crack growth For a ductile material, such as a structural steel at ambient temperature, the plastic zone size is much larger than that of a brittle material and the stress distribution ahead of the crack tip is distorted. The size of the plastic zone is so large, that before sufficient stress is generated ahead of the crack tip to

Page 32 of 35

produce brittle fracture, local ductile failure occurs at the crack tip. Crack growth tends to be slow and stable, controlled by the plastic strain at the crack tip region. This local strain is measured by the CTOD test, and hence can be used to quantify the resistance to crack extension (fracture toughness). This theory is known as elastic plastic fracture mechanics (EPFM) and can be applied to ductile or brittle materials, as local ductility always occurs at the crack tip. plastic collapse In a very ductile material, the size of the plastic zone is so large that it reaches the edge of a structure without any local ductile crack growth happening. In this event, the whole cross section of the structure fails uniformly by necking (reduction of area) or shear, a mechanism known as plastic collapse.

7.2 Design Against Failure


It is relatively straightforward to design against brittle fracture in a structure. This can be achieved by ensuring that the stress intensity factor (K) from the maximum design stress is below that of the fracture toughness (KIC). Similarly, a structure will only fail by plastic collapse if the design stress exceeds the flow stress of the material. In practice, however, the failure mode of a structure will be between these two extremes. One way of assessing the integrity of a structure against these intermediate failure modes is to use a failure assessment diagram (FAD) of a material, such as shown in Figure 33.

1.0

FAIL KR =

KR
0.5

stress intensity factor fracture toughness load on structure plastic collapse load

SAFE

SR =

0.5

SR

1.0

Figure 33 - Typical failure assessment diagram

The two extremes of behaviour are linked by a curve separating a SAFE region from a FAIL region. The shape of the curve can be derived from the stress/strain curve of the material. Both KR and SR can then be calculated for a particular defect in a loaded structure, and plotted on the FAD. Inside the

Page 33 of 35

SAFE region, the structure will be safe, but in the FAIL region there will be a risk of failure. Failure assessment diagrams and fitness for purpose applications are outlined within BS7910 (British Standard for Assessing the Acceptability of Flaws in Metallic Structures, 1999) Further information on the application of fracture mechanics to engineering structures can be obtained from the LRTA paper:
Fracture Mechanics Applications, D.Howarth and P.Pumphrey, Paper no.6, Session 1997-8

Page 34 of 35

8. Summary
This paper has introduced the basic aspects of fracture in metallic structures. In summary: there are two basic fracture modes: ductile and brittle. Brittle fracture can be particularly catastrophic for a structure, as it propagates in a rapid, unstable manner. carbon-manganese steel structures are subject to brittle fracture at low temperatures. Therefore for low temperature service (<-55C), special high nickel grade steels and non-ferrous materials should be used. the susceptibility to brittle fracture increases with section thickness. Charpy testing is used as a quality control check to ensure the adequate toughness of structural materials. fracture toughness testing can be used to predict the integrity of a cracked structure.

Page 35 of 35

Das könnte Ihnen auch gefallen