Welding stresses and strains

A. Nedoseka , in Fundamentals of Evaluation and Diagnostics of Welded Structures, 2012

ii Coefficient of linear expansion of materials

The coefficient of linear expansion of materials α T is found from the relationship

(2.1) α T = ε / T = Δl / lT ,

where ε is the relative elongation (deformation); T is the temperature to which the cloth is heated; fifty is the length of the controlled section of the specimen; Δl is the elongation of this section. In approximate calculations. α T is usually assumed to exist independent of temperature and is equal to:

25⋅10−6 1/°C – for alloys of AMg type;

8⋅10−6 1/°C – for titanium alloys;

9⋅10  vi ane/°C – for regular steels;

sixteen⋅10  6 1/°C – for depression-alloyed steels.

In reality the coefficient of linear expansion of materials α T depends on temperature, fifty-fifty though slightly. This dependence can be derived by dilatometric testing of specimens. 1

Read full chapter

URL:

https://www.sciencedirect.com/science/article/pii/B9780857095312500023

Temperature Measurement

C. Hagart-Alexander , in Instrumentation Reference Book (Fourth Edition), 2010

Expansion of Solids

When a solid is heated, it increases in volume. It increases in length, breadth, and thickness. The increase in length of any side of a solid will depend on the original length l 0, the rise in temperature t, and the coefficient of linear expansion α.

The coefficient of linear expansion may be defined as the increase in length per unit of measurement length when the temperature is raised 1°C. Thus if the temperature of a rod of length l 0 is raised from 0°C to t°C, the new length, lt , will be given by

(21.1) l t = fifty 0 + l 0 · α t = l 0 ( one + α t )

The value of the coefficient of expansion varies from substance to substance. The coefficients of linear expansion of some mutual materials are given in Tabular array 21.1.

Table 21.1. Coefficients of linear expansion of solids, extracted from Tables of Physical and Chemical Constants by Kaye and Laby (Longmans); the values given are per Kelvin and, except where some temperature is specified, for a range of about 20 degrees

Substance α (ppm)
Aluminum 25.5
Copper 16.seven
Gold thirteen.9
Atomic number 26 (cast) x.two
Lead 29.1
Nickel 12.eight
Platinum eight.nine
Silver 18.8
Tin 22.4
Contumely (typical) xviii.9
Constantan (Eureka) 60 Cu, 40 Ni 17.0
Duralumin 22.6
Nickel steel,
10% Ni thirteen.0
30% Ni 12.0
36% Ni (Invar) −0.3 to +2.5
40% Ni half dozen.0
Steel 10.v to xi.half-dozen
Phosphor bronze, 97.6 Cu, two Sn, 0.two P xvi.eight
Solder, 2 Atomic number 82, 1 Sn 25
Cement and physical 10
Drinking glass (soda) 8.5
Glass (Pyrex) 3
Silica (fused) −80° to 0°C 0.22
Silica (fused) 0° to 100°C 0.50

The increase in area with temperature, that is, the coefficient of superficial expansion, is approximately twice the coefficient of linear expansion. The coefficient of cubic expansion is almost three times the coefficient of linear expansion.

Read full chapter

URL:

https://world wide web.sciencedirect.com/science/article/pii/B9780750683081000218

Applied science materials

J. Carvill , in Mechanical Engineer's Data Handbook, 1993

6.21.2 Thermal expansion

Allow:

α = coefficient of linear expansion (°C−1)

β = coefficient of superficial expansion (°C−1)

γ = coefficient of cubical expansion (°C−1)

θ = temperature change (°C)

Fifty = initial length

A = initial area

5= initial volume

L' = final length

A' = final expanse

V' = final volume

Then:

L' = 50(1 + αθ)

A' = A(1 +southward βθ)

V'= V(ane + γθ)

Approximately:

β = 2α

γ = 3α

Coefficients of linear expansion α(× 106 ° C−ane) at normal temperature (unless otherwise stated)

Textile α Textile α Material α
Aluminium 23 Gilded 14 Safe: natural, soft 150–220
29 (0–600 °C) 15 (0–500 °C) natural, hard 80
Antimony eleven Granite 8.3 nitrile 110
Brass 19 Graphite vii.9 silicone 185
Brick 5 Gunmetal 18 Sandstone 12
Bronze 18 Water ice fifty Argent 19
Cadmium 30 Fe: bandage xi 20.five (0–900 °C)
Cement eleven Wrought 12 Slate ten
Chromium 7 15 (0–700 °C) Solder (2 lead: 1 tin) 25
11 (0–900 °C) Lead 29 Steel: hardened 12.4
Cobalt 12 33 (0–320 °C) balmy 11
18 (23–350 °C) Magnesium 25 stainless x.4
Physical 13 30 (0–400 °C) Tin 21
Copper 16.7 Nickel 12.viii Titanium nine
twenty (0–yard °C) 18 (0–1000 °C) Tungsten 4.five (20 °C)
Diamond 1.three Phosphor bronze sixteen.7 6
Duralumin 23 Plaster 17 (600–1400°C)
Ebonite 70 Platinum 8.9 7
High german silvery eighteen.4 11 (0–800°C) (1400–2200 °C)
Glass 8.half-dozen (0–100 °C) Porcelain 4 Vanadium 8
9.nine (100–200 °C) Quartz 8–fourteen Zinc 30
11.9 (200–300 °C)

6.21.3 Freezing mixtures

Properties of some greyness irons (BS 1452)

Ammonium nitrate (parts) Crushed ice or snow in water (parts) Temperature (°C)
1 0.94 –4
1 1.twenty –14
i one.31 –17.5
ane 3.61 –eight
Calcium chloride (parts) Crushed water ice or snowfall in water (parts) Temperature (°C)
1 0.49 –20
ane 0.61 –39
ane 0.70 –55
ane 1.23 –22
i four.92 –4
Solid carbon dioxide with booze −72
Solid carbon dioxide with −77
chloroform or ether

6.21.4 Coefficients of cubical expansion of liquids at normal temperature (unless otherwise stated)

γ(× 10half-dozen °C−one)
Liquid γ Liquid γ
Acetic acid 107 Olive oil seventy
Aniline 85 Alkane series 90
Benzene 124 Sulphuric acid 51
Chloroform 126 (twenty%)
Ethanol 110 Turpentine 94
Ether 163 Water 41.five
Glycerine 53 (0–100 °C)
Mercury 18 100
(100–200 °C)
180
(200–300 °C)

half-dozen.21.6 Antifreeze mixtures

Freezing indicate (°C)
Concentration (%vol.) 10 20 xxx 40 50
Ethanol (ethyl alcohol) –3.3 –7.8 –14.4 –22.2 –30.6
Methanol (methyl alcohol) –5.0 –12.1 –21.ane –32.2 –45.0
Ethylene glycol –3.9 –eight.nine –fifteen.6 –24.iv –36.seven
Glycerine –ane.vii –5.0 –9.4 –15.half-dozen –22.8

Read total chapter

URL:

https://world wide web.sciencedirect.com/science/commodity/pii/B978008051135150011X

Slope AND DEFLECTION OF BEAMS

E.J. HEARN Ph.D., B.Sc. (Eng.) Hons., C.Eng., F.I.Mech.E., F.I.Prod.E., F.I.Diag.E. , in Mechanics of Materials i (Third Edition), 1997

5.13 Deflections due to temperature effects

It has been shown in §2.3 that a uniform temperature increase t on an unconstrained bar of length L volition produce an increase in length

where α is the coefficient of linear expansion of the material of the bar. Provided that the bar remains unconstrained, i.east. is free to expand, no stresses volition event.

Similarly, in the instance of a beam supported in such a fashion that longitudinal expansion can occur freely, no stresses are ready and there will exist no tendency for the beam to bend. If, even so, the beam is constrained and then stresses will result, their values being calculated using

the procedure of §2.iii provided that the temperature change is uniform across the whole beam section.

If the temperature is not constant across the axle then, once more, stresses and deflections will outcome and the post-obit procedure must be adopted:

Consider the initially directly, merely-supported axle shown in Fig. 5.32(a) with an initial uniform temperature T 0. If the temperature changes to a value T 1 on the upper surface and T two on the lower surface with, say, T 2 > T 1 so an element dx on the lesser surface volition expand to α(T 2T 0). dx whilst the same length on the meridian surface volition only aggrandize to α(T 2T 0).dx. As a result the beam will bend to accommodate the distortion of the element dx, the sides of the chemical element rotating relative to i another by the angle , as shown in Fig. 5.32(b). For a depth of axle d:

Fig. 5.32(a). Beam initially straight before application of temperature T one on the top surface and T 2 on the lower surface. (Beam supported on rollers at B to allow "gratuitous" lateral expansion).

Fig. 5.32(b). Beam later application of temperatures T 1 and T ii, showing distortions of element dx.

(5.31) d . d θ = α ( T 2 T 0 ) d x α ( T one T 0 ) d x o r d θ d x = α ( T 2 T one ) d

The differential equation gives the charge per unit of change of slope of the beam and, since θ = dy/dx,

and so d d x ( d y d x ) = ( d 2 y d x 2 ) = α ( T 2 T 1 ) d

Thus the standard differential equation for bending of the beam due to temperature gradient beyond the beam section is:

(5.32) d 2 y d 10 2 = α ( T 2 T 1 ) d

This is direct analogous to the standard deflection equation d two y d x ii = M EI so that integration of this equation in exactly the aforementioned way every bit previously for bending moments allows a solution for slopes and deflections produced by the thermal effects.

NB. If the temperature gradient across the beam section is linear, the average temperature 1 ii (T one + T 2) will occur at the mid-height position and, in add-on to the bending, the beam volition alter in overall length by an amount αL[ 1 2 (T ane + T 2) − T 0] in the absence of whatsoever constraint.

Application to cantilevers

Consider the cantilever shown in Fig. five.33 subjected to temperature T one on the top surface and T 2 on the lower surface. In the absenteeism of external loads, and considering the cantilever is free to bend, there will be no moment or reaction gear up at the congenital-in stop.

Fig. 5.33. Cantilever with temperature T ane on the upper surface, T 2 on the lower surface (T 2 > T 1).

Applying the differential equation (5.32) we have:

d 2 y d ten two = α ( T two T 1 ) d . I due north t eastward g r a t i northward g d y d x = α ( T 2 T 1 ) d x + C 1 But a t 10 = 0 , d y d x = 0 , C one = 0 and : d y d 10 = α ( T 2 T 1 ) d x = θ

∴ The gradient at the end of the cantilever is:

(5.33) θ max = α ( T 2 T one ) d L .

Integrating once again to find deflections:

y = α ( T 2 T 1 ) d x 2 2 + C 2

and, since y = 0 at ten = 0, so C 2 = 0, and:

y = α ( T two T 1 ) 2 d 10 2

At the end of the cantilever, therefore, the deflection is:

Application to built-in beams

Consider the congenital-in axle shown in Fig. 5.34. Using the principle of superposition the differential equation for the beam is given by the combination of the equations for applied angle moment and thermal effects.

Fig. 5.34. Built-in beam subjected to thermal slope with temperature T 1 on the upper surface, T 2 on the lower surface.

For bending

E I d 2 y d ten two = G A + R A x .

For thermal effects

d two y d x 2 = α ( T ii T 1 ) d E I d two y d x 2 = E I α ( T two T one ) d

∴ The combined differential equation is:

E I d 2 y d 10 2 = M A + R A x + E I α ( T 2 T 1 ) d .

Nonetheless, in the absenteeism of applied loads and from symmetry of the axle:

R A = R B = 0 , and Thousand A = Chiliad B = M Eastward I d two y d x two = M + E I α ( T 2 T one ) d . I n t e m r a t i n g E I d y d x = Chiliad x + Eastward I α ( T 2 T 1 ) d x + C ane N o w a t x = 0 , d y d x = 0 C 1 = 0 ,

and at

(5.35) ten = 50 d y d ten = 0 Grand = E I α ( T two T 1 ) d

Integrating once again to find the deflection equation we have:

E I y = Chiliad . ten 2 ii + E I α ( T two T 1 ) d . ten 2 2 + C 2

When x = 0, y = 0 ∴ C 2 = 0,

and, since M = East I α ( T 2 T 1 ) d .

Thus a rather surprising result is obtained whereby the beam volition remain horizontal in the presence of a thermal gradient. Information technology will, however, be bailiwick to residual stresses arising from the constraint on overall expansion of the beam under the average temperature 1 2 (T ane + T ii). i.due east. from §2.iii

(5.36) residual stress = Due east [ ane 2 ( T one + T ii ) ] = 1 2 E a ( T 1 + T 2 )

Read full chapter

URL:

https://world wide web.sciencedirect.com/scientific discipline/commodity/pii/B9780750632652500062

Heat

K.M. SMITH C.ENG., Yard.I.E.E. , P. HOLROYD C.ENG., Thou.I.MECH.E., A.M.I.STRUCT.E. , in Engineering Principles for Electric Technicians, 1968

7.6 Expansion of solids by heat

Near all materials aggrandize when heated although the corporeality of expansion for a given temperature rise varies with the type of material. The increase in dimensions takes place in all directions. Nosotros phone call the increase in length linear expansion, the increase in area superficial expansion, and the increase in volume cubical expansion. The ratio increment in length original length for 1 degree rise in temperature is called the coefficient of linear expansion, the ratio increase in area original area for 1 degree ascent in temperature is called the coefficient of superficial expansion, and the ratio increase in book original volume for 1 degree rise in temperature is called the coefficient of cubical expansion.

It is generally accepted that if the coefficient of linear expansion is the aforementioned in all directions for a given material, and so the coefficient of superficial expansion is approximately twice the coefficient of linear expansion and the coefficient of cubical expansion iii times the coefficient of linear expansion.

FIG. 7.5. Linear expansion.

FIG. 7.6. Superficial expansion.

Values of coefficients of linear expansion for various materials are given in Table 7.1. It should be noticed that the coefficient of expansion of a material should exist qualified by the degree of temperature increase, i.e. deg F or deg C.

Table 7.one. COEFFICIENTS OF LINEAR EXPANSION (α)

Textile per deg F per deg C
Aluminium 0·0000128 0·000023
Brass 0·0000104 0·0000188
Copper 0·0000093 0·0000167
Cast iron 0·0000061 0·000011
Lead 0·0000161 0·000029
Monel 0·0000076 0·0000137
Nickel 0·0000071 0·0000128
Platinum 0·00000495 0·0000089
Cast steel 0·0000061 0·000011
Can 0·0000117 0·000021
Glass 0·0000047 0·0000085
Graphite 0·0000044 0·0000079

FIG. 7.7. Cubical expansion.

The coefficient of expansion per deg F = five 9 coefficient of expansion per deg C.

Example

(a) At 0°C a contumely rod is 100 in. long. If the coefficient of linear expansion of brass is 0·0000188 per deg C, what will be its length at 80°C? (b) A foursquare canvas of brass 10 in. length of side at 0°C is heated to lxxx°C. What will then exist its area?

(a)

Coefficient of expansion

= increase in length original length  per 1 deg C temperature ascent . Increase in length = 0 · 0000188 × 100 × 80 in . = 0 · 1504 in . Last length = 100 + 0 · 1504 in . = 100 · 1504 in .

(b)

Coefficient of superficial expansion = ii × 0 · 0000188 = 0 · 0000376 . Increase in expanse = 0 · 0000376 × 10 × 10 × 80 in 2 = 0 · 3008 in 2 . Terminal area = 100 + 0 · 3008 in 2 = 100 · 3008 in ii .

EXAMPLE

What increment in length volition take place in a steam pipe 120 ft long, the temperature of the steam being 450°F, if the atmospheric temperature is l°F and the coefficient of linear expansion of the material is 0·0000061 per deg F?

Coefficient of expansion = increment in length original length per deg F temperature change . Increase in length = 0 · 0000061 × 120 × (450−l) ft = 0 · 0000061 × 120 × 400 ft = 0 · 2928 ft = 3 · 5136 in .

EXPERIMENTAL VERIFICATION OF COEFFICIENT OF LINEAR EXPANSION

Equipment required

Copper rod, steam jacket, boiler, thermometer, dial gauge, and bunsen burner.

The length of the rod is measured from the costless end to the heart of the fastening at the fixed end. The dial gauge measuring to i/m in. is zeroed at room temperature and this temperature is read off the thermometer. Steam produced in the boiler is then passed through the steam jacket in which the rod is enclosed. Later a few minutes the rod volition take reached the same temperature as the steam and this temperature is read off the thermometer. The expansion is measured on the dial gauge.

FIG. 7.8.

Original length of copper rod = 30 · 25 in . Room temperature = 65°F . Expansion of rod (from dial gauge reading) = 0 · 041 in . Concluding temperature = 212°F . Coefficient of linear expansion = increment in length original length per deg temperature rise = 0 · 041 30 · 25 ( 212 65 ) = 0 · 041 30 · 25 × 147 = 0 · 0000093 per deg F .

Read full chapter

URL:

https://www.sciencedirect.com/science/commodity/pii/B9780080129853500121

Finite chemical element assay

Carl T.F. Ross BSc, PhD, DSc, CEng, FRINA, MSNAME , in Advanced Applied Finite Chemical element Methods, 1998

6.vii TO CALCULATE ELEMENTAL NODAL FORCES DUE TO THERMAL EFFECTS

Equivalent load vectors {P T} due to thermal furnishings will be calculated for various one- and two-dimensional elements.

6.7.1 Rod element (run across Fig. half dozen.26)

Allow,

Fig. 6.26.

T = temperature rise α = coefficient of linear expansion East = rubberband modulus .

From (6.110)

{ P T } = [ B ] T [ D ] { ε o } d ( vol ) .

Substituting (6.13) and (6.27) into the above,

{ P T } = 1 l [ 1 one ] E α T A dx = E α T A { i 1 } u i u 2 .

vi.7.2

For a two-dimensional rod element chemical element in global coordinates,

(6.113) { P T } = [ Ξ ] T Due east α T A { 1 0 1 0 } u 1 υ 1 u 2 υ 2 .

Substituting [] from (4.7) into (half-dozen.113),

(half dozen.114) { P T } = E α T A [ c due south s c 0 2 0 2 c s s c ] { 1 0 one 0 } = E α T A { c s c s } .

6.7.iii

For a 3-dimensional rod in global coordinates,

(half-dozen.115) { P T } = [ Ξ ] T E α T A { i 0 0 one 0 0 } u one υ i w i u 2 υ 2 w ii .

Substituting [] from (4.xxx) into (6.115)

6.7.4

Similarly for a two-dimensional rigid-jointed frame,

{ P T } = Due east α T A { c s 0 c s 0 }

and a three-dimensional rigid-jointed frame,

{ P T } = E α T A { C x . ten C x . y C 10 . z 0 0 0 C x . x C 10 . y C ten . z 0 0 0 } .

6.vii.5

For an in-plate, in global coordinates,

{ P T } = [ B ] T [ D ] { ε 0 } d ( vol ) = 1 2 Δ [ b i 0 c i 0 c i b i b j 0 c j 0 c j b j b k 0 c g 0 c m b k ] Eastward 1 [ 1 μ 0 μ 1 0 0 0 γ ] { α T α T 0 } dx dy

= E one t 2 Δ [ b i 0 c i 0 c i b i b j 0 c j 0 c j b j b k 0 c chiliad 0 c k b 1000 ] { ( 1 + μ ) α T ( i + μ ) α T 0 } dx dy = E one t ( 1 + μ ) α T 2 { b i c i b j c j b 1000 c k } f o r p l a north east s t r e south southward

and

{ P T } = E 1 ( 1 + μ ) ( 1 + ν ) α T 2 { b i c i b j c j b m c thou } for p fifty a n east s t r a i n ,

where b i , c i , etc., are divers in Sec. vi.two.3 as

{ ε } = ( 1 + ν ) α T { ane 1 0 } .

North.B. It must be remembered that (6.107) must be used for computing 'stresses'.

Read total chapter

URL:

https://www.sciencedirect.com/science/article/pii/B9781898563518500147

Metal Matrix Composites

F. Delannay , in Comprehensive Blended Materials 2, 2018

4.8.4.4.2 Anisotropic composites

As the direction of plastic catamenia is very sensitively afflicted by the reinforcement orientation distribution, the linear expansion coefficient of a composite in a given direction will tend toward the rule of mixture in the fully plastic matrix regime only on the status of a perfectly isotropic distribution of the phases.

Many authors take reported the occurrence of marked hysteresis loops in the thermal expansion response of unidirectional continuous fiber MMCs subjected to thermal cycling (Garmong, 1974; Kural and Min, 1984; Tompkins and Dries, 1988; Dumant et al., 1988; Masutti et al., 1990; Lacom et al., 1990; Vaidya and Chawla, 1994; Böhm et al., 1995; Chun et al., 1995; Korb et al., 1998). Similar for isotropic composites, this beliefs tin exist quite simply viewed as resulting from the succession of elastic and plastic straining in alternating directions when temperature increases or decreases (east.grand., Rabinovitch et al., 1983; Kural and Min, 1984; Dumant et al., 1988; Clyne and Withers, 1993).

As an example, Fig. 11 presents the expansion curves in the axial direction during the starting time and second cycles between room temperature and 550   °C for a composite consisting of an Al-3%Mg matrix with 30   vol% of unidirectional continuous SiC fibers (Nicalon®) (Masutti et al., 1990). The blended had previously been cooled to liquid nitrogen temperature. The first thermal wheel ends with a negative residual strain because the centric stress in the fibers is more compressive at the end of cooling than in the pristine specimen that had been heated from low temperature. The centric expansion loop runs in the clockwise direction. As suggested past Masutti et al. (1990), the centric stage stresses can be quite straightforwardly estimated from the expansion curves of Fig. xi past considering that internal stresses are practically aught at 550   °C. Hence, during cooling (heating) from (to) this temperature, the departure of the blended axial strain, ε 3 c = ε three r , from the thermal expansion strain of the fibers, αrΔT, is equal to the rubberband strain of the fibers. Using the equilibrium relation

Fig. 11. Expansion curves in the axial management during the first and second cycles between room temperature and 550   °C for a composite consisting of an Al-three%Mg matrix with 30% of unidirectional continuous SiC fibers.

Masutti et al. (1990), reproduced past permission of Chapman and Hall from Journal of Materials Science Letters, 1990, 9, 340–342.

[56] σ 3 m V m + σ 3 r V r = 0

and neglecting the effect of the transverse stresses σ 1 r and σ 2 r on ε three r , this yields

[57] σ three one thousand = 5 r V m Due east r ( α r Δ T ε 3 r )

Fig. 12 shows the variation of the longitudinal mean stress σ three k as a office of temperature computed from the curves of Fig. 11 using eqn [57]. Equally suggested on this Fig., the thermal expansion of the fibers αrΔT was inferred past drawing a directly line extrapolating the composite expansion at high temperature (where the matrix flow forcefulness is nigh zero). Two regimes can be distinguished: the elastic regimes (dotted lines), during which the σ 3 1000 variation can be approximated by eqn [eighteen], and the fully plastic authorities. During the latter authorities, the difference σ 3 m σ one m = σ eff m follows the temperature dependence of the matrix yield stress σY. If σ 1 k 0 , σ 1 m σ Y and the maximum amplitude of the strain hysteresis in the centric direction, Δ ε iii max c , can be approximated as (Pedersen, 1990)

Fig. 12. Variation of the average longitudinal stress σ 3 m every bit a function of temperature computed from the curves of Fig. eleven using eqn [57].

Masutti et al. (1990), reproduced by permission of Chapman and Hall from Journal of Materials Science Messages, 1990, ix, 340–342.

[58] Δ ε 3 max c = ii σ Y V yard E r V r

Reference to Fig. 7 is a reminder that σ i m is not zero, equally it is equal to V r Five m σ r r . Yet, the neglect in eqn [58] of the outcome of transverse stresses on ε 3 r partly compensates for the approximation σ 1 m 0 .

Using a similar method, Nassini et al. (2001) accept traced the temperature dependence of the in-plane matrix stress during cycling between RT and 560   °C of Al-base composites containing a random planar distribution of Al2O3 (Saffil) fibers. A detailed insight into the roles of interfacial sliding and matrix creep during thermal cycling of continuous fiber composites has been provided by Dutta (2000) via a micro-mechanical model accounting for the operation of multiple matrix creep mechanisms at various stages of thermal cycling. The model incorporates the result of interfacial sliding by an interface improvidence-controlled mechanism. Simulation results were compared with experimental data on a unidirectional graphite fiber/6061 Al composite.

If the yield stress σY would not vary with temperature, the centric CTE α3c would exist equal to αr during the part of the expansion bend where the matrix is fully plastic. Considering volume conservation and applying the police of mixture to express the volume expansion α iic of the blended, the transverse CTE, α1c, is then (Böhm et al., 1995)

[59] α 1 c = α m + ane two ( α yard α r ) ( i 3 V r )

For low 5 r, the transverse expansion coefficient of the composite can thus be significantly larger than αm.

Unidirectional continuous fiber MMCs exhibit a counterclockwise hysteresis in the direction transverse to the fibers (e.chiliad., Böhm et al., 1995). This contrary beliefs with respect to the longitudinal direction results essentially from volume conservation associated with plastic yielding. Nonetheless, in the direction perpendicular to the fibers, the rubberband strain in the matrix acts in opposition to the plastic flow of the matrix. Hence, the higher the average mean stresses in the phases, the lower the amplitude of the transverse hysteresis. For the fully plastic government, no closed class expression of the radial stress σ r one thousand appears to accept been worked out in literature on the basis of the coaxial cylinders model. It is thus not possible to derive, for the transverse hysteresis amplitude, the expression of an offset stress εos like to eqn [54].

The earliest elasto-plastic model for predicting the CTEs of anisotropic composites on the ground on the Eshelby approach has been proposed by Wakashima et al. (1974) for the instance of a composite with continuous Westward fibers in a Cu matrix. This model correctly predicts the clockwise hysteresis in the axial management and provides an belittling expression for the amplitude of the loop. Several other models have been proposed for predicting the thermal expansion coefficients α3c and α1c of unidirectional composites while accounting for more complex situations of matrix strain-hardening and temperature dependence of σY. These models accept been reviewed past Hahn (1991). Like matrix plasticity, interface sliding may also contribute to the relaxation of internal stresses. A theoretical report of the effect of interface sliding on the CTE of a curt fiber composite with purely elastic fibers and matrix has been proposed by Jasiuk et al. (1988) on the footing of an Eshelby-type analysis.

The thermal expansion beliefs of composites containing a random planar distribution of fibers or whiskers can be analyzed on the basis of the same principles every bit for unidirectional continuous cobweb composites. Equally an case, Fig. thirteen(a) shows the thermal expansion curves in the directions parallel ( ε 1 c ) and perpendicular ( ε iii c ) to the plane of isotropy, measured for a composite consisting of a pure Al matrix reinforced with a random planar network of 20% of continuous fibers of Inconel 601 (Salmon et al., 1997; Boland et al., 1998). Thermal cycling was carried out betwixt 310 and 30   °C. An initial dwell had been made at 310   °C in order to allow phase stresses to be eliminated by creep of the matrix. Like results have been presented by Neite and Mielke (1991) and past Nassini et al. (2001) for Al-base composites containing a random planar distribution of Al2O3 (Saffil) fibers. Like in unidirectional composites, the in-plane beliefs is dominated past the constraint exerted by the cobweb network. The overall α1c is depression and the ε 1 c hysteresis runs in the clockwise management. In contrast, α3c is close to αm and the ε 3 c hysteresis runs in the counterclockwise direction. The strain hysteresis amplitude amounts to about 0.three×10−three and 0.6×ten−3 in the in-aeroplane management and perpendicular direction, respectively. The overall volume at a given temperature is thus the same during heating and cooling.

Fig. xiii. (a) Thermal expansion curves measured for pure Al, for Inconel 600, and for a composite consisting of a pure Al matrix reinforced with a random planar network of continuous fibers of Inconel 601, directions parallel ( ε one c ) and perpendicular ( ε iii c ) to the plane of isotropy; (b) non-thermal dilatation strain of the composite Δ non - therm .310 T c calculated from the expansion curves according to eqn [59].

If phase stresses are zero at 310   °C, the thermal expansion curves of Fig. 13(a) permit calculating the non-thermal dilatation strain of the composite, Δ non - therm . Δ T c , during cycling. Indeed,

[threescore] Δ non - therm . Δ T c = two ε i Δ T c + ε 3 Δ T c three ( ε Δ T r V r + ε Δ T m V g )

The result is shown in Fig. 13(b). During cooling from 310   °C, the blended exhibits a fairly linear increment of the non-thermal book expansion, which reaches 1.4×10−iii at 30   °C. No significant difference between the cooling and heating ramps can exist distinguished. According to eqn [10], if no change of porosity book fraction has occurred during cooling, the average matrix hateful stress at 30   °C could exist calculated as

[61] σ g m 310 30 = Δ not - therm .310 - xxx c V yard ( ane Grand m i K r )

Taking K Al=75.2   GPa and Yard In601=170   GPa, one obtains σ thou g 310 - 30 = 236 MPa . The corresponding average mean stress in the fibers would nearly 950   MPa. Obviously, these values are much likewise high as to be correct for the case of a pure Al matrix whose flow strength does not exceed 50   MPa. Plastic flow and creep driven by local deviatoric stresses does certainly reduce the average mean stage stresses to a much lower value. Equally a matter of fact, in literature, neutron diffraction and 10-ray diffraction studies of Al-base of operations composites have never measured such loftier thermal phase stresses in the matrix. The conclusion is that the presence of porosity must be taken into account for correctly interpreting the expansion beliefs of this type of composite. It is very difficult to discover by density measurements porosity volume fractions of the order of the measured volume expansion (0.1%).

Read full affiliate

URL:

https://www.sciencedirect.com/science/article/pii/B9780128035818038741

Statics

Richard Gentle , ... Nib Bolton , in Mechanical Engineering Systems, 2001

Temperature stresses

Telephone engineers, when suspending cables between telegraph poles, ever permit some slack in the cables. This is considering if the temperature drops then the cable will decrease in length and if it becomes taut then the stresses produced could be high enough to break cables. There are many situations in engineering science where we have to consider the stresses that can be produced equally a result of temperature changes.

When the temperature of a torso is inverse it changes in length and if this expansion or contraction is wholly or partially resisted, stresses are set up upwards in the body. Consider a bar of initial length L 0. If the temperature is at present raised past θ and the bar is free to expand, the length increases to L θ = L 0(1 + αθ), where αis the coefficient of linear expansion of the bar material. The change in length of the bar is thus L θL 0 = L 0(1 + αθ) −50 0 = 50 0αθ. If this expansion is prevented, it is as if a bar of length L 0(1 + αθ) has been compressed to a length L 0 and and so the resulting compressive strain εis:

ɛ = L 0 α θ L 0 ( 1 + α θ )

Since αθ is pocket-sized compared with 1, then:

ɛ = α θ

If the material has a modulus of elasticity E and Hooke's law is obeyed, the stress σproduced is:

(5.3.viii) σ = α θ E

The stress is thus proportional to the coefficient of linear expansion, the change in temperature and the modulus of elasticity.

Case 5.three.seven

Decide the stress produced per caste change in temperature for a fully restrained steel member if the coefficient of linear expansion for steel is 12 × 10−6 per °C and the modulus of elasticity of the steel is 200 GPa.

σ = α θ Eastward = 12 × 10 6 × 1 × 200 × 10 9 = 2 4 × 10 vi Pa = 2 4MPa

Example 5.iii.8

A steel wire is stretched taut between 2 rigid supports at 20°C and is nether a stress of 20 MPa. What volition be the stress in the wire when the temperature drops to 10°C? The coefficient of linear expansion for steel is 12 × 10−vi per °C and the modulus of elasticity is 200 GPa.

The effect of the drop in temperature is to produce a tensile stress of:

σ = α θ East = 12 × half dozen × x × 200 × ten 9 = 24 MPa

The total stress interim on the wire will be the sum of the thermal stress and the initial stress 24 + 20 = 44 MPa.

Composite confined

Consider a composite bar with materials in parallel, such as a circular bar A inside a circular tube B, with the two materials having different coefficients of expansion, α A and α B, and different modulus of elasticity values, Due east A and E B, just the same initial length L and attached rigidly together (Effigy 5.3.ten). The 2 materials are considered to exist initially unstressed. The temperature is so changed by θ.

Effigy 5.three.x. Composite bar: (a) initially; (b) when free to expand; (c) when the expansion is restrained by being stock-still together

If the two members had not been fixed to each other, when the temperature inverse they would have expanded; A would accept changed its length by Lα Aθ A and B its length by Lα Bθ B and there would be a deviation in length between the two members at temperature θof (αA − αBL (Effigy 5.3.ten(b)). All the same, the two members are rigidly fixed together and and then this divergence in length is eliminated past compressing member B with a forcefulness F and extending A with a force F (Figure 5.3.10(c)).

The extension eastward A of A due to the forcefulness F is:

e A = F L E A A A

where A A is its cross-sectional area. The contraction e B of B due to the forcefulness F is:

due east B = F L E B A B

where A B is its cross-exclusive expanse. Merely eA + eB = (αA − αBL and so:

(v.iii.9) ( α A α B ) θ 50 = F 50 ( one E A A A + i E B A B ) F = ( α A α B ) θ ( one E A A A + i Eastward B A B )

Example 5.3.9

A steel rod has a bore of thirty mm and is fitted centrally inside copper tubing of internal diameter 35 mm and external diameter 60 mm. The rod and tube are rigidly fixed together at each cease but are initially unstressed. What will be the stresses produced in each past a temperature increase of 100°C? The copper has a modulus of elasticity of 100 GPa and a coefficient of linear expansion of 20 × 10−half dozen/°C; the steel has a modulus of elasticity of 200 GPa and a coefficient of linear expansion of 12 × ten−half-dozen/°C.

The strength compressing the copper and extending the steel is:

F = ( α A α B ) θ ( 1 E A A A + 1 E B A B ) = ( 20 12 ) × ten 6 × 100 1 100 × ten 9 × 1 4 π ( 0 060 ii 0 035 two ) + i 200 × 10 9 × 1 four π 0 03 = 64 3 kN

The compressive stress interim on the copper is thus:

σ A = 64 iii × 10 3 ane 4 π ( 0 060 ii 0 035 2 ) = 34 5 MPa

The tensile stress acting on the steel is:

σ B = 64 three × x three i four π 0 030 two = 91 0 MPa

Read total chapter

URL:

https://www.sciencedirect.com/science/article/pii/B9780750652131500052

New Conductor Technology Summary

Yuan Chen , in Engineering Energy Aluminum Usher Composite Core (ACCC) and its Awarding, 2019

1.4 Summary

Amidst many new types of conductors, carbon fiber composite cadre aluminum conductor has the advantages of lightweight, low coefficient of linear expansion (transfer temperature above), high strength, high elastic modulus, high temperature resistance, fatigue resistance, aging resistance, chemic corrosion resistance, and low resistivity (combined with soft aluminum strands) and other technical advantages. It is a new blazon of wire that has evolution potential, non merely for utilise in improving manual capacity and reducing power loss, but also in mostly improving the line margin of safe operation. Information technology can play an important part in the construction and upgrading of the power grid. In view of the above technical advantages of carbon fiber composite core aluminum conductor, the national Power Filigree Corp in 2006 included the technology and products in its first directory of key applications of new technology.

Technology and price of the carbon cobweb blended core aluminum usher, equally a new type of high-tech wire production, have been monopolized by adult countries, and the imperfection of structure and maintenance techniques constitutes two major challenges for promoting the product widely. In the above context, the "Carbon fiber composite core aluminum usher of contained research and development and application" science and technology project was established by the Country Grid Corporation of Cathay in 2008. This project was included in the national loftier-tech development plan in 2010 to completely break the bottleneck of carbon cobweb composite cadre aluminum conductor, and promote the evolution of the national loftier-performance carbon cobweb industry.

Read total chapter

URL:

https://world wide web.sciencedirect.com/scientific discipline/article/pii/B9780128156117000016

Moulding

John Campbell , in Consummate Casting Handbook (Second Edition), 2015

15.2.4 Zircon Sand

Zircon sand is a light colour and generally in the class of fine rounded grains. Its extremely low and linear coefficient of expansion allows it to make extremely accurate cores and even whole moulds. Its high bulk density nearly exactly matches the density of liquid aluminium, so that cores have close to neutral buoyancy which makes zircon the ideal amass for frail and poorly supported cores in Al alloy castings.

Zircon has a adequately high thermal diffusivity compared to silica, giving it greater chilling power and faster freezing of castings. However, this effect is not bully, and in any case, greater chilling power is a two-edged sword; although conferring slightly improved properties for the casting, the achievement of thin-walled sections is fabricated more difficult.

The high density of zircon (nearly double that of silica sand) leads to significantly greater mechanical handling challenges in the foundry. Robots for conventional sand moulds will usually find zircon moulds too heavy to lift and sand silos will require strengthening. Vibrating equipment and pneumatic conveying systems crave twice the free energy.

The rate of habiliment of patternwork and tooling is a serious consideration when using denser moulding materials. Zircon in item is found to be particularly abrasive toward tooling as a event of its hardness and its high density thus requiring high blow pressures on core blowing machines to fill cadre boxes. As the percentage of broken zircon grains increases with cycling around the sand organization, the increasing number of sharp edges profoundly increases the rate at which it tin inflict wearable damage on the core boxes and other expensive tooling and establish. Although founders that employ zircon mutter about its high cost, the real hidden cost is the rate of devastation of the tooling.

Read full chapter

URL:

https://www.sciencedirect.com/science/article/pii/B9780444635099000157