Anda di halaman 1dari 18

CHAPTER 2

This chapter deals with three major topics. The first topic is the basic properties of each phase of an unsaturated soil. This information is used in later chapters when describing the behavior of the soil as a system of phases. The second topic deals with the understanding of the interaction between air and water. The third topic deals with establishing useful volume-mass relations for solving engineering problems. An unsaturated soil has commonly been referred to as a three-phase system. However, more recently, the realization of the important role of the air-water interface (Le., the contractile skin) has warranted its inclusion as an additional phase when considering certain physical mechanisms. When the air phase is continuous, the contractile skin interacts with the soil particles and provides an influence on the mechanical behavior of the soil. An element of unsaturated soil with a continuous air phase is idealized in Fig. 2.1. When the air phase consists of occluded air bubbles, the fluid becomes significantly compressible. The mass and volume of each phase can be schematically represented by a phase diagram. Figure 2.2(a) shows a rigorous four-phase diagram for an unsaturated soil. The thickness of the contractile skin is in the order of only a few molecular layers. Therefore, the physical subdivision of the contractile skin is unnecessary when establishing volume-mass relations for an unsaturated soil. The contractile skin is considered as part of the water phase without significant error. A simplified three-phase diagram, depicted in Fig. 2.2(b), can be used in writing the volumemass relationships. The term soil solids is used when refemng to the summation of masses and volumes of all the soil particles. 2.1 PROPERTIES OF THE INDIVIDUAL PHASES An understanding of the basic properties of the soil particles, water, air, and contractile skin should precede the consideration of the behavior of the soil system. This sec20

Co py rig hte dM ate ria l


phase.
Volume
( b)

Phase Properties and Relations

Figure 2.1 An element of unsaturated soil with a continuous air

Mass

Figure 2.2 Rigorous and simplified phase diagrams for an unsaturated soil. (a) Rigorous four phase unsaturated soil system; (b) simplified three phase diagram.

Copyright 1993 John Wiley & Sons

Retrieved from: www.knovel.com

2.1 PROPERTIES OF THE INDIVIDUAL PHASES

21

tion discusses the gravimetric and volumetric properties pertinent to each phase. The most important property of the contractile skin is its ability to exert a tensile pull. This property is called surface tension and is discussed later in this chapter.

Volume of individual phase

Mass
of individual phase

Soil Particles

The density of the soil particles, ps, is defined as follows (Fig. 2.3):

The density of the soil particles is commonly expressed as a dimensionless variable called specific gravity, G,. The specific gravity of the soil particles is defined as the ratio of the density of the soil particles to the density of water at a temperature of 4C under atmospheric pressure conditions (Le., 101.3 kPa). In the S I system of units, this variable is now referred to as the relative density of the soil particles.
PI G, = -.
P w

The density of water at 4C and 1 0 1 . 3 kPa is lo00 kg/m3. Table 2 . 1 presents typical values of specific gravity for several common minerals.

Water Phase The density of water, pw, is defined as follows:


P w =

Water is essentially a homogeneous substance the world over, except for variations produced by salts and isotopes of hydrogen and oxygen (Dorsey, 1940). Distilled water under the pressure of its saturated vapor is called pure, saturated water. The density of pure, saturated water must be measured experimentally. Figure 2.4 shows the density of pure water under various applied pressures and temperatures. For soil mechanics problems, the variation in the density of water due to temperature differences is more significant

Co py rig hte dM ate ria l


Total volume v = v, + vw+ v,

2.1.1 Density and Specific Volume Density, p , is defined as the ratio of mass to volume. Each phase of a soil has its own density. The density of each phase can be formulated from the phase diagram shown in Fig. 2.3. Specific volume, vo, is generally defined as the inverse of density; therefore, specific volume is the ratio of volume to mass. Unit weight, y, is a useful term in soil mechanics. It is the product of density, p , and gravitational acceleration, g (Le., 9.81 m/s2).

Total mass M = M. + M, + M,

Figure 2 . 3 Phase diagram for an unsaturated soil.

Table 2.1 Specific Gravity of Several Minerals (from Lambe and Whitman, 1979)
Mineral Specific Gravity, G,

Quartz K-Feldspars Na-Ca-Feldspars Calcite Dolomite Muscovite Biotite Chlorite Pyrophyllite Serpentine Kaolinite Halloysite (2H20) Illite Montmorillonite Attapulgite

2.65 2S4-2.57 2.62-2.76 2.72 2.85 2.7-3.1 2.8-3.2 2.6-2.9 2.84 2.2-2.7 2 . 6 1 ; 2.64 f 0.02 2.55 2 . 8 4 ; 2.60-2.86 2.7 4 ; 2.75-2.78 2.30

Calculated from crystal structure.

than its variation due to an applied pressure. For isothermal conditions, the density of water is commonly taken as lo00 kg/m3.

-.

M W
V W

(2.3)

Air Phase The density of air, pa, can be expressed as

Ma

pa =

v,
va

(2.4)

The specific volume of air, vd, is


V d

=Ma

Air behaves as a mixture of several gases (Table 2.2)and also varying amounts of water vapor. The mixture is called

Copyright 1993 John Wiley & Sons

Retrieved from: www.knovel.com

22

2 PHASE PROPERTIES AND RELATIONS

Nitrogen (N,) Oxygen ( 0 2 ) Argon (Ar) Carbon dioxide (CO,) Neon (Ne) Helium (He) Krypton (Kr) Hydrogen (H2) Xenon (Xe) Ozone (03) Air

dry air when no water vapor is present, and is called moist air when water vapor is present. Dry and moist air can be considered to behave as an ideal gas under pressures and temperatures commonly encountered in geotechnical engineering. The ideal gas law can be written
Ma RT ii, V, = wa

Co py rig hte dM ate ria l


Density of water, p. (kg/rn)

Figure 2.4 Density of pure water for various applied pressures and temperatures (from Dorsey,
1940).

Table 2.2 Composition of Dry Ail.8

Percentage by Volume

Density (kg/m3)

Molecular Mass (on Basis of Natural Scale, 0 = 16) (kg /kmol)


28.016 32.000 39.944 44,010 20.183 4.003 83.800 2.016 131.300 48.000 28.966

78.08 20.95 0.934 0.03 1 1.82 x 1 0 - ~ 5.24 x 1 0 - ~ 1.14 x 1 0 - ~ -5.0 x 10-5 8.70 X 1 x 10-6 to 1 x 10-5 100.0

1.25055 1.42904 1.7837 1.9769 0.90035 0.17847 3.708 0.08988 5.851 2.144 1 .2929

aUnder standard conditions (Le., 101.3 kPa and OOC) with no water vapor

u,,, = atmospheric pressure (i.e., 101.3 kPa or 1 atm)

V, = volume of air (m3) Ma = mass of air (kg)

wa = molecular mass of air (kg/kmol) R = universal (molar) gas constant [i.e., 8.31432 J/(mol K ) ] T = absolute temperature (Le., T = t o 273.16) (K) t o = temperature (c).

where

-, sign inti, = u, + dicates an absolute pressure, i.e., uatm) (kN/m* or kPa) u, = gauge air pressure (kN/m2 or kPa)
u, = absolute air pressure (note that a bar,

The right-hand side of Eq. (2.6) (Le., (M,/wa)RT) is a constant for a gas in a closed system with a constant mass and temperature. Under these conditions, Eq. (2.6) can be rewritten as Boyles law:

Copyright 1993 John Wiley & Sons

Retrieved from: www.knovel.com

2.1 PROPERTIES OF THE INDIVIDUAL PHASES

23

absolute pressure and volume of air, respectively, at condition 1 absolute pressure and volume of air, respectively, at condition 2. Rearranging the ideal gas equation (Le., Eq. 2.6) gives
M= a *a

Table 2.3 Density of Air at Different Absolute Pressures, Temperatures, and Relative Humidities (from Kaye and Laby, 1973)
Density of Air, pa (kg/m3) Temperature, r (C) Absolute Air Pressure, Ti, (kPa) 80 85 10 0.982 1.043 1.105 1.167 1.228 1.240 1.290
~

20 0.946 1.005 1.065 1.124 1.184 1.196 1.243


~~

30 0.910 0.968 1.025 1.083 1.140 1.152 1.198


-~

Substituting Eq. (2.4) into Eq. (2.8) gives an equation for the density of air: (2.9)

The molecular mass of air, oa,depends on the composition of the mixture of dry air and water vapor. The dry air has a molecular mass of 28.966 kg/kmol, and the molecular mass of the water vapor (H20) is 18.016 kg/kmol. The composition of air, namely, nitrogen ( N , ) and oxygen ( O , ) ,are essentially constant in the atmosphere. The carbon dioxide (CO,) content in air may vary, depending on environmental conditions, such as the rate of consumption of fossil fuels. However, the constituent of air that can vary most is water vapor. The volume percentage of water vapor in the air may range from as little as 0.000002% to as high as 4-596 (Harrison, 1965). The molecular mass of air is affected by the change in each constituent. This consequently affects the density of air. The concentration of water vapor in the air is commonly expressed in terms of relative humidity: (2.10)
U0

Co py rig hte dM ate ria l


90
95 100 101 105 Relative Humidity, RH (96) 20 25 30 35 40 45 50 55 60 65 70 75 80 10 +0.003 +0.003 +om2 +0.002 +0.001 +0.001

Va

R T a

Density Adjustments for Humidity (kg/m3) Temperature, to (C) 20

+0.006
+0.005

+0.004
+0.003 +om2 +0.001 0 -0.001 -0.002 -0.003 -0.004
-0.005
~

-0.001 -0.001

where

-0.002

-0.002

RH = relative humidity (96) uu = partial pressure of water vapor in the air (kPa) -

-0.003 -0.003

-0.006

ud = saturation pressure of water vapor at the same temperature (kPa).

Table 2.3 presents the values of air density for different absolute air pressures (Tia) and temperatures, t o . The figures in the top portion of Table 2.3 were computed for air with a relative humidity of 50% and 0.04% carbon dioxide by volume. For air having relative humidities other than 5096, a correction should be applied as shown in the bottom portion of Table 2.3. Although the corrections are small, it should be noted that the density of air decreases as the relative humidity increases. This indicates that the moist air is lighter than the dry air.

2.1.2 Viscosity All fluids resist a change of form or the action of shearing. This resistance is expressed by the property called viscos-

ity. The absolute (dynamic) viscosity, p , of a fluid is defined as the resistance of the fluid to a shearing force applied by sliding one plate over another with the fluid placed in between. The absolute viscosity depends on the pressure and temperature. However, the influence of pressure is negligible for the range of pressures commonly encountered in typical civil engineering applications. The viscosities of water and air at atmospheric pressure (Le., 101.3 @a) and different temperatures a 8 given in Tables 2.4 and 2.5, respectively. Figure 2.5 presents the absolute viscosities of water, air, and several other materials at different temperatures. The viscosities of liquids are shown to decmse with an increase in temperature, while the viscosity of air increases as the temperature increases.

Copyright 1993 John Wiley & Sons

Retrieved from: www.knovel.com

24

2 PHASE PROPERTIES AND RELATIONS

Table 2.4 Viscosity of Water at 101.3 kPa (Modified from Tuma, 1976)
Absolute (Dynamic) Viscosity, Temperature, to
("C)
P
(X

Absolute (Dynamic) Viscosity, p Tempera( X 10-~ ture, to ("C) N * s/m2)

10-~ s/m2)

0 5
1 0 15 20 25 30 35 40 45 50

Table 2.5 Viscosity of Air at 101.3 kPa (Modified from Tuma, 1976; and Kaye and Laby, 1973)
Absolute (dynamic) Viscosity, p Temperature, ( X IO-' N * to ("C) s/m2)

- 10
0 1 0 20 30 40 50 100 200

-20

2.1.3 Surface Tension The air-water interface (Le., contractile skin) possesses a property called surface tension. The phenomenon of surface tension results f r o m the intennolecular forces acting on molecules in the contractile skin. These forces are different from those that act on molecules in the interior of . 6 ( a ) ] . the water [Fig. 2 A molecule in the interior of the water experiences equal forces in all directions, which means there is no unbalanced force. A water molecule within the contractile skin

Co py rig hte dM ate ria l


1.794 1.519 1.310 1 . 1 4 4 1.009 0.895 0.800 0.731 0.654 0.597 0.548

55 60 65 70 7 5 80 85 90 95 100

0.507 0.470 0.437 0.407 0.381 0.357 0.336 0.317 0.299 0.284

'xloI;O

20

40 60 80 100 120 Temperature, t ( O C )

Figure 2.5 Viscosity of fluids at different temperatures (fmm Streeter and Wylie, 1975).

at the air-water interface 1.e.. contractile skin)

Sources

1.604 1.667 1.705 1.761 1.785 1.864 1.909 1 . 9 6 2.20 2 . 6 1

Tuma, 1976 Tuma, 1976 Tuma, 1976 Tuma, 1976 Tuma, 1976 Tuma, 1976 Tuma, 1976 Kaye and Laby, 1973 Kaye and Laby, 1973 Kaye and Laby, 1973

(b)

Figure 2.6 Surface tension phenomenon at the air-water interface. (a) Intermolecular forces on contractile skin and water; @) pressures and surface tension acting on a curved two-dimensional surface.

experiences an unbalanced force towards the interior of the water. In order for the contractile skin to be in equilibrium, a tensile pull is generated along the contractile skin. The property of the contractile skin that allows it to exert a ten, .Surface tension is sile pull is called its surface tension, T measured as the tensile force per unit length' of the contractile skin (i.e., units of N/m). Surface tension is tangential to the contractile skin surface. Its magnitude de. 6 gives surface creases as temperature increases. Table 2 tension values for the contractile skin at different temperatures.

Copyright 1993 John Wiley & Sons

Retrieved from: www.knovel.com

2.2 INTERACTION OF AIR AND WATER

25

Table 2.6 Surface Tension of Contractile Skin (Le., Air-Water Interface) (from Kaye and Laby, 1973).
Temperature,
to (C)

Surface Tension, T s (mN/m) 75.7 74.2 73.5 72.75 72.0 71.2 69.6 67.9 66.2 64.4 62.6
58.8

60
70
80

100

The surface tension causes the contractile skin to behave like an elastic membrane. This behavior is similar to an inflated balloon which has a greater pressure inside the balloon than outside. If a flexible two-dimensional membrane is subjected to different pressures on each side, the membrane must assume a concave curvature towards the larger pressure and exert a tension in the membrane in order to be in equilibrium. The pressure difference acmss the curved surface can be related to the surface tension and the radius of curvature of the surface by considering equilibrium across the membrane [Fig. 2.6(b)]. The pressures acting on the membrane are u and (u + Au). The membrane has a radius of curvature of, R,, and a surface tension, T,. The horizontal forces along the membrane balance each other. Force equilibrium in the vertical direction requires that 2T, sin 6 = 2 AuR, sin 6 (2.11) where

Co py rig hte dM ate ria l


where
&=-* 2TS
RS

10 15 20 25 30 40 50

Figure 2.7 Surface tension on a warped membrane.

R , and R2 = radii of curvature of a warped membrane in two orthogonal principal planes.

If the radius of curvature is the same in all directions (i.e., R, and R2 are equal to R,),Eq. (2.13) becomes
(2.14)

In an unsaturated soil, the contractile skin would be subjected to an air pressure, u,, which is greater than the water pressure, u,. The pressure difference, (u, - u,), is referred to as matric suction. The pressure difference causes the contractile skin to curve in accordance with Eq. (2.14):
(u,

- u,)

2TS R,

(2.15)

where
(u,

- u,)

= matric suction or the difference between

pore-air and pore-water pressures acting on the contractile skin.

2Rs sin B = length of the membrane projected onto the horizontal plane. Rearranging Eq. (2.11) gives

(2.12)

Equation (2.12) gives the pressure difference across a two-dimensional curved surface with a radius, R,, and a surface tension, T,. For a warped or saddle-shaped surface (Le., three-dimensional membrane), Eq. (2.12) can be extended using the Laplace equation (Fig. 2.7)
Au = T ,

Equation (2.15) is referred to as Kelvins capillary model equation. As the matric suction of a soil increases, the radius of curvature of the contractile skin decreases. The curved contractile skin is often called a meniscus. When the pressure difference between the pore-air and pore-wakr goes to zero, the radius of curvature, R,, goes to infinity. Therefore, a flat air-water interface exists when the matric suction goes to zem.

2.2 INTERACTION OF AIR AND WATER

(d + i)

(2.13)

Air and water can be combined as immiscible and/or miscible mixtures. The immiscible mixture is a combination of free air and water without any interaction. The immiscible

Copyright 1993 John Wiley & Sons

Retrieved from: www.knovel.com

26

2 PHASE PROPERTIES AND RELATIONS

Solid Liquid
u1 J

e! a

Figure 2.8 State diagram for water (not to scale; from Van Have m and Brown, 1972).

mixture is characterized by the separation produced by the contractile skin. A miscible air-water mixture can have two forms, First, air dissolves in water and can occupy approximately 2% by volume of the water (Dorsey, 1940). Second, water vapor can be present in the air. All of the above types of mixtures are dealt with in the following sections. Consideration is also given to all possible states for the water.

2.2.1 Solid, Liquid, and Vapor States of Water Water can be found in one of three states: the solid state as ice, the liquid state as water, and the gaseous state as water vapor (Fig. 2.8). Throughout the text, the words water or water phase refer to the liquid state of water. The state of water depends on the pressure and temperature environment. Three lines are drawn on the water state diagram (Fig. 2.8). These are the vaporization curve, AB, the fusion

Table 2.7 Saturation Pressures of Water Vapor at Different Temperatures(from Kaye and Laby, 1973)
to (C)

Co py rig hte dM ate ria l


0
100 374
Temperature, t (C)

curve, AC, and the sublimation curve, AD. The vaporization curve, AB, is also called the vapor pressure curve of water. It gives combination values of temperature and pressure for which the liquid and vapor states of water can coexist in equilibrium. The fusion curve, A C , separates the solid and the liquid states of water, and the sublimation curve, AD, separates the solid and the vapor states of water. The solid state can coexist in equilibrium with the liquid state along the fusion curve, and with the vapor state along the sublimation curve. The vaporization, fusion, and sublimation curves meet at point A . This point is called the triple point of water where the solid, liquid, and vapor states of water can coexist in equilibrium. The triple point of water is achieved at a temperature of 0C and a pressure of 0.61 kPa.

2.2.2 Water Vapor The vaporization curve, AB, in Fig. 2.8 represents an equilibrium condition between the liquid and vapor states of water. In this state of equilibrium, evaporation and condensation processes occur simultaneously at the same rate. The rate of condensation depends on the pressure in the water vapor which reaches its saturation value on the vaporization line. On the other hand, the evaporation rate depends only on temperature. Therefore, a unique relationship exists between the saturation water vapor pressure and temperature, which is described by the vaporization curve. Saturation water vapor pressures, Zoo, are presented in Table 2.7. In the atmosphere, water vapor is mixed with air. However, the presence of the air has no effect on the behavior of the water vapor. This phenomenon is expressed by Daltons law of partial pressures. Daltons law states that the pressure of a mixture of gases is equal to the sum of the

9 1.1477 2.1974 4.0074 6.9967


18

0 10 20 30
to (C)

0.6107 1.2276 2.3384 4.2451

0.6566 1.3123 2.4872 4.4949 2

0.7055 1.4022 2.6443 4.7574 4

0.7576 1.4974 2.8099 5.0332 6

0.8130 1.5983 2.9846 5.3226 8

0.8720 1.7051 3.1686 5.6264 10

0.9348 1.8180 3.3625 5.9451 12

1.0015 1.9375 3.5666 6.2793 14

1.0724 2.0639 3.7814 6.6296 16

40 60 80 100 120

7.3812 19.933 47.375 101.325 198.49

8.2053 21.852 51.344 108.77 211.39

9.1075 23.926 55.857 116.67 224.97

10.094 26.164 60.121 125.03 239.26

11.171 28.578 64.960 133.88 254.27

12.345 31.177 70.120 143.25 270.03

13.623 33.974 75.617 153.14 286.58

15.013 36.980 81.468 163.59 303.95

16.522 40.206 87.691 174.61 322.16

18.160 43.667 94.304 186.24 341.23

Saturation water vapor pressures, E,,, are in kPa.

Copyright 1993 John Wiley & Sons

Retrieved from: www.knovel.com

2.2 INTERACTION OF AIR AND WATER

2 . 2 . 3 Air Dissolving in Water Water molecules form a lattice structure with openings re-

ferred to as a cage that can be occupied by a gas (Rodebush and Busswell, 1958), as illustrated in Fig. 2.9. Air dissolves into the water and fills the cages which have a volume of approximately 2% by total volume. The water lattice is relatively rigid and stable (Dorsey, 1940), and the density of water changes little as a consequence of the presence of the dissolved air. Figure 2.10 shows the effect of dissolved air on the density of water for various temperatures. An analogy using a cylinder with a piston and porous stone is useful in analyzing the behavior of air-water mixtures. Consider a cylinder with a porous stone at its base and a frictionless piston at the top (Fig. 2.11). The porous stone has pores equaling 2% of its volume. The porous stone is used to simulate the behavior of water. In this model, the cylinder contains free air above the porous stone. An imaginary valve is situated at the boundary be-

Co py rig hte dM ate ria l


Dissolved air

partial pressures that each individual gas would exert if it alone filled the entire volume. In other words, the behavior of a particular gas of a mixture of gases is independent of the other gases. Therefore, the partial pressure of water vapor in the atmosphere which is in equilibrium with water is the saturation pressure given in Table 2.7. Similarly the presence of air above water does not change the state equilibrium of water (Fig. 2.8). In nature, the water vapor in air is usually not in equilibrium with adjacent bodies of water. This means that the partial pressure of the water vapor in air, is,, is usually not the same as the saturation pressure of the water vapor, isivo, at the corresponding temperature. The water vapor in air at a given temperature is therefore said to be undersaturated, saturated, or supersaturated when the partial pressure i , is less than, equal to, or greater than of water vapor, S the saturation water vapor pressure. The saturated condition indicates an equilibrium between the water vapor and the water where evaporation and condensation take place at the same rate. On the other hand, the undersaturated and supersaturated states of water vapor are not equilibrium conditions. The supenaturated state indicates an excess of water vapor which eventually condenses. In this case, the rate of condensation exceeds the evaporation rate until the partial pressure of water vapor, Si, has been reduced to the saturation pressure, isuo. In the undersaturated state, there is a lack of water vapor relative to the equilibrium condition. Therefore, the rate of evaporation exceeds the rate of condensation until the partial pressure of the water vapor, u , , has reached the saturation water vapor pressure, Zu0. The partial pressure of the water vapor in air defines the degree to which the air is saturated with water vapor at a specific temperature. The degree of saturation with respect to water vapor is referred to as the relative humidity, RH [Eq.(2.10)l.

27

Air pressure regulator

Free air Cylinder

Channel and cage network for air dissolving in water (approximately2% by volume)

Water

Cage

Figure 2.9 Visualization aid for air dissolving in water.

measurable effect

20

0.001 0.002 0.003 0.004 0.006

Difference in density between air-free and air-saturated water (kg/mJ

Figure 2.10 E f f e c t of dissolved air on the density of water (from


Dorsey, 1940).

tween the free air and the porous stone to control the movement of air into the porous stone. The movement of air into the porous stone represents the movement of air into water. Let us suppose there is an initial pressure applied equally to the free air and to the air in the porous stone in the cylinder. The imaginary valve is then closed. If the load on the piston is increased, the free air above the porous stone will be compressed following Boyles law [Q. (2.7)]. The imaginary valve is then opened, and some air will move into the porous stone in accordance with Henrys law. Henr y s law states that the mass of gas dissolved in a fixed

Copyright 1993 John Wiley & Sons

Retrieved from: www.knovel.com

28

2 PHASE PROPERTIES AND RELATIONS


1

initial load Frictionless piston Initial volume of free air

Added Initial

1
I

Volume change of the free air

v o l u m e change due to air going into porous disk Imaginary valve Cylinder

Frictionless piston Final volume of free air

QD

quantity of liquid, at constant temperature, is directly proportional to the absolute pressure of the gas above the solution (Sisler et d . , 1953). This process will continue as the piston load is increased. Eventually, all the free air will move into the porous stone, and any additional applied load will be carried by the porous stone. The above analogy cannot totally simulate the situation in an unsaturated soil. In the presence of a solid, such as soil particles, the air and water pressures can have different magnitudes. The air and water pressures in a soil can also change at differing rates during a process. In the analogy, the free air and the water (i.e., porous stone) have the same pressure. The possibility of a difference between the air and water pressures is later shown to be of significance in the compressibility formulation. The mass of air going into or coming out of water is time dependent. This time dependency can either be ignored or taken into consideration, depending upon the engineering problem under consideration. The amount of air that can be dissolved in water is referred to as its solubility, and the rate of solution is referred to as its difisivity.

Co py rig hte dM ate ria l


Before After

Figure 2.11 Piston and porous stone analogy.

stone (i.e., water). After some time, an equilibrium condition will be reached where the pressure in the free air and the dissolved air are equal. If the piston load is then increased, the process will be repeated. The mass of dissolved air at equilibrium is dependent upon the corresponding absolute air pressure as stated in Henrys law. If the temperature remains constant throughout the process, the ratio of the mass and the absolute pressure of the dissolved air is constant:
(2.17)

where

Mdl , Zal= mass and absolute pressure of the dissolved air, respectively, at condition 1 M d z , Za2 = mass and absolute pressure of the dissolved air, respectively, at condition 2.

Solubility of Air in Water The volume of dissolved air in water is essentially independent of air or water pressures. This can be demonstrated using the ideal gas law and Henrys law. The ideal gas law [Q. (2.6)] can be rearranged and applied to a gas dissolving in water at a certain temperature and pressure:
M d RT Vd = -

(2.16)

@a

where
Vd Md

= volume of dissolved air in water = mass of dissolved air in water u, = absolute pressure of the dissolved air.

The volume of dissolved air in water, v d , is computed using Eq. (2.16) and by considering the relationship shown in (Eq. 2.17). At a constant temperature, the volume of dissolved air in water is a constant for different pressures. The ratio between the muss of each gas that can be dissolved in a liquid and the mass of the liquid is called the coefficient of solubility, H. Table 2.8 presents the coefficients of solubility of oxygen, nitrogen, and air in water, over a temperature range. All coefficients of solubility are referenced to a standard pressure of 101.3 kPa. The ratio of the volume of dissolved gas, v d , in a liquid to the volume of the liquid is called the volumetric coefficient of solubility, h, which varies slightly with temperature. Values for the volumetric coefficient of solubility for air in water under different temperatures are given in Table 2.8.

The absolute pressure of the dissolved air is equal to the absolute pressure of the free air under equilibrium conditions. Refemng to the piston and porous stone analogy, an increase in the piston load will increase the pressure in the free air, and therefore more free air will go into the porous

of Gases Through Water The rate at which air can pass through water is described by Ficks law of diffusion. The rate at which mass is transf e d across a unit area is equal to the product of the coefficient of diffusion, D, and the concentration gradient. In the diffusion of air through water, the concentrationdiffer&@sion

Copyright 1993 John Wiley & Sons

Retrieved from: www.knovel.com

2.3 VOLUME-MASS RELATIONS

29

Table 2.8 Solubility of Gases in Water (Under a Pressure of 101.3 kPa) (from h r s e y ,
1940)

Coefficient of Solubility, Ha Temperature, to ("C) Nitrogen, Argon, etc.

Volumetric Coefficient of Solubility, hb Air Air

Oxygen

ence is equal to the difference in density between the free air and the dissolved air in the water. Under constant temperature conditions, the density of air is a function of the air pressure [Q. ( 2 . 9 ) ] .An increase in pressure in the free air will develop a pressure difference between the free and dissolved air. This pressure difference becomes the driving potential for the free air to diffuse into the water. The gases composing air individually diffuse into water. The coefficients of diffusion, D, for each component of air through water are tabulated in Table 2.9. Combined gases forming air dissolve in water at a rate of approximately 2.0 x Research Council, 1933). m2/s (U.S. Barden and Sides (1967)measured the coefficient of diffusion for air through the water phase of both saturated and compacted clays. Their results are presented in Table 2.10. The study concluded that the coefficient of diffusion appears to decrease with decreasing water content of the soil. The coefficient of diffusion for air through the water in a soil appears to differ by several orders of magnitude from the coefficient of diffusion for air through free water.

Co py rig hte dM ate ria l


"At standard atmospheric pressure.

0 4 10 15 20 25 30

14.56 X 13.06 x 11.25 X 10.07 X 9.11 x 8.28 x 7.55 X

23.87 X 21.59 x 18.82 X 17.00 X 15.51 x 14.24 x 13.10 X

38.43 X 34.65 x 30.07 X 27.07 X 24.62 x 22.52 x 20.65 X

0.02918 0.02632 0.02284 0.02055 0.01868 0.01708 0.01564

bh = (PW/PM.

Table 2.9 Coefiients of Diffusion for Certain G a s e s in Water (from Kohn, 1965)
Temperature, to ("C) Coefficient of Diffusion, D (m2/s)

Gas

co2
N2 H 2

0 2

20 22 21 25

1.7 x 10-~ 2.0 x 10-9 5.2 x 10-9 2.92 x 10-9

Similarly, porosity type terns can be defined with respect to each of the phases of a soil:

(2.19) (2.20) (2.21) (2.22)

2.3 VOLUME-MASS RELATIONS

The volume-mass relations of the soil particles, water, and air phases are useful properties in engineering practice. The derivations combine the gravimetric and volumetric properties of a soil.

where

2.3.1 Porosity Porosity, n , in percent is defined as the ratio of the volume of voids, Vu, to the total volume, V (Fig. 2.12):

. n,

n, = soil particle porosity (%) = water porosity (%) n, = air porosity (%) n, = contractile skin porosity (%).

(2.18)

The volume associated with the contractile skin can be

Copyright 1993 John Wiley & Sons

Retrieved from: www.knovel.com

30

2 PHASE PROPERTIES AND RELATIONS

Table 2.10 CoefRcient of Mffusion for Air Through Different Materials (from Barden and Sides, 1967)
Water Content, w (%) Coefficient of Diffusion, D (m2/s)

Material Free water Natural rubber Kaolin consolidated at 414 kPa (oriented parallel to flow) Kaolin consolidated at 414 kPa (oriented perpendicular to flow) Kaolin consolidated at 483 kPa Kaolin consolidated at 34.5 kPa Denvent clay (illite) consolidated at 34.5 kPa Jackson clay and 4%bentonite consolidated at 34.5 kPa Compacted Westwater clay Saturated ceramic Saturated coarse stone

49 49

2.2 x 10-9 1 . 1 x lo-'' 4.5 x lo-'' 3.2 x lo-''

Volume relations

Co py rig hte dM ate ria l


47 75 53 39 3.0 x lo-'' 6.2 x lo-'' 4.7 x lo-'' c 1 . 0 x lo-'' 16 49 21 1.0 x 10-I' 1.6 x lo-"' 2.5 x 10-5
Mass relations
b

2.3.2 Void Ratio

Void ratio, e, is defined as the ratio of the volume of voids, Vu, to the volume of soil solids, V, (Fig. 2.12):

M =pV

(2.24)

Total volume, V = V, + V,

= v,

v, + v,

Total mass, M = M, + M,

M,

Figure 2.12 Volume-mass Elations.

The relationship between porosity and void ratio is obtained by equating the volume of voids, Vu,from the two equations [Le., Eqs. (2.18) and (2,24)]: e n=(2.25) 1 +e' Typical values for void ratio are shown in Table 2.11,

assumed to be negligible or part of the water phase. The water and air porosities represent their volumetric percentages in the soil. The soil particle porosity can be visualized as the percentage of the overall volume comprised of soil particles. The sum of the porosities of all phases must equal 100%. Therefore, the following soil porosity equation can be written as
n,

2.3.3 Degree o f Saturation The percentage of the void space which contains water is expressed as the degree of saturation, S (%):
(2.26)

+ n = n, + n, + n,

= loo(%).

(2.23)

The degree of saturation, S, can be used to subdivide into three groups. Dry soils (i-e., S = 0%): Dry soil consists of soil particles and air. No water is present. Saturated soils (i.e., S = 100%):All of the voids in the soil are filled with water. Unsaturated soils (i.e., 0% < S C 100%):An unsaturated soil can be further subdivided, depending upon whether the air phase is continuousor occluded.

The water porosity, n , , expressed in decimal forin, is e commonly referred to as the volumetric water content, , in soil science and soil physics literature. The volumetric water content notation is also used throughout this book. Typical values of porosity for some soils are given in Table 2.11.

Copyright 1993 John Wiley & Sons

Retrieved from: www.knovel.com

2.3 VOLUME-MASS RELATIONS

31

Table 2.11 Typical Values of Porosity, Void Ratio, and D r y Density ( M o d i f i e d from Hough, 1969)
Void Ratio,
e

Porosity,
n

Soil Type

maximum

minimum

maximum

minimum

Density, (kg/m3) maximum minimum


P

Co py rig hte dM ate ria l


0.40 50.0 29.0 0.40 52.0 29.0 0.30 0.20 0.40 0.14 47.0 49.0 55.0 46.0 23 .O 17.0 29.0 12.0 1.8 1.o 0.25 0.20

Granular Materials: 1) Uniform Materials a) Equal spheres 0.92 (theoretical values) b) Standard Ottawa 0.80 sand c) Clean, uniform sand 1.o (fine or medium) d) Uniform, inorganic 1.1 silt

0.35

47.6

26.0 33.0

1762 1890

1474 1330 1281

0.50

44.0

1890

Granular Materials: 2) Well-Graded Materials a) Silty sand 0.90 b) Clean, fine to coarse sand 0.95 c) Micaceous sand 1.20 d) Silty sand and gravel 0.85 Mixed Soils a) Sandy or silty clay b) Skip-graded silty clay with stones or rock fragments c) Well-graded gravel, sand, silt, and clay mixture

2034 2210 1922 2239

1394 1362 1217 1426

64.0 50.0

20.0 17.0

2162 2243

96 1 1346

0.70

0.13

41.0

11.0

237 1

1602

Clay Soils a) Clay (30-50% clay sizes) b) Colloidal clay (-0.002 mm


2 50%)

2.4 12.0

0.50 0.60

71.O 92.0

33.0 37.0

1794 1698

801 308

Organic Soils a) Organic silt b) Organic clay (30-50% clay sizes)

3.0 4.4

0.55 0.70

75.0 81.0

35.0 41.0

1762 1602

641 48 1

General Note: Tabulation is based on G, = 2.65 for granular soils, G, = 2.70 for clays, and G, = 2.60 for organic

soils.

This subdivision is primarily a function of the degree of saturation. An unsaturated soil with a continuous air phase genedly has a degree of saturation less than approximately 80% (Le., S < 80%). Occluded air bubbles commonly occur in unsaturated soils having a degree of saturation greater than approximately 90% (Le,, S > 90%). The transition zone between con-

tinuous air phase and occluded air bubbles occurs when the degree of saturation is between appmximately 80-9096 (i.e., 80% < S < 90%).

2.3.4 Water Content Water content, w , is defined as the ratio of the 'mass of water, M,, to the mass of soil solids, M, (Fig. 2.12). It is

Copyright 1993 John Wiley & Sons

Retrieved from: www.knovel.com

32

2 PHASE PROPERTIES AND RELATIONS

presented as a percentage [Le., w (W)]: (2.27) Water content, w, is also referred to as the gravimetric water content. The volumetric water content, ,e , is defined as the ratio of the volume of water, V,, to the total volume of the soil, E
V W 8, = V'

The mass of the water phase of a soil, M,, is the product of the volume and density of water (Fig. 2.3):

M, = P W V , .

(2.34)

The volume of water, V,, can also be computed from the volume relation given in Fig. 2.12 (i.e., left-hand side):
V , = SeV,.

(2.35)

Co py rig hte dM ate ria l


(2.28)

The relationship given in Eq. (2.35) is shown in Fig. 2.13 (i.e., left-hand side). Equation (2.35) can then be rewritten as

M,

= p,SeV,.

(2.36)

The volumetric water content can also be expressed in terms of porosity, degree of saturation, and void ratio (Fig. 2.12). The volumetric water content can be written as

The mass of the water, M,, can also be related to the mass of the soil solids, M,:

s vu 8, = -

M, = wM,.

(2.37)

v'

(2.29)

Since V u / V i s equal to the porosity of the soil, Eq. (2.29) becomes

The mass of the soil solids, M,, is obtained from the phase diagram in Fig. 2.3:

8, = Sn.

(2.30)

Substituting Eq. (2.25) into Eq. (2.30) yields another form for the volumetric water content equation:
Se 8, = 1 +e'

M, = G,P, V,. Substituting Eq. (2.38) into Eq. (2.37) yields M, = wGsp,V,.

(2.38) (2.39)

(2.31)

Equating Eqs. (2.39) and (2.36) results in a basic volume-mass relationship for soils:
Se = wG,.

(2.40)

V' The total density is also called the bulk density. The dry density of a soil, pd, is defined as the ratio of the mass of the soil solids, M,,to the total volume of the soil, V (Fig. 2.12):

2.3.5 Soil Density Two commonly used soil density definitions are the total density and the dry density. The total density of a soil, p, is the ratio of the total mass, M,to the total volume of the soil, V (Fig. 2.12): M p = (2.32)

The total and dry densities of a soil defined in Eqs. (2.32) and (2.33), respectively, can also be expressed in terms of the volume-mass properties of the soil (i.e,, S , e , w , and G,). Assuming that the mass of the air phase, Mu, is negligible, the total mass of the soil is the sum of the mass of the water, M,, and the mass of the soil solids, M,. The total volume of the soil, V , is given by the volume of the soil solids, V,, and the volume of the voids, V,. Therefore, the equation for the total density of a soil, p, can be rewritten using Eqs. (2.24), (2.38), and (2.39): (2.41)
Mass relations

(2.33)

Typical minimum and maximum dry densities for various soils are presented in Table 2.11. Other soil density definitions are the saturated density and the buoyant density. The saturated density of a soil is the total density of the soil for the case where the voids are filled with water (Le., Vu = 0 and S = 100%).The buoyant density of a soil is the difference between the saturated density of the soil and the density of water.
2.3.6 Basic Volume-Mass Relationship

Volume relations

V, = e V,

T -

The volume and mass for each phase can be related to one another using basic relations from the phase diagram (Fig. 2.3) and the volume-mass relations shown in Fig. 2.12.

1w G , = S e
Figure 2.13 Derivation of the basic volume-mass relationship.

Copyright 1993 John Wiley & Sons

Retrieved from: www.knovel.com

2.3 VOLUME-MASS RELATIONS

33

(2.42) (2.43)

Substituting the basic volume-mass relationship in Eq.


(2.40) into Eq. (2.43) gives the following equation for the

total density:
p=-

The dry density of a soil, pd, is obtained by eliminating the mass of the water, Mu, from Eq. (2.41):

Co py rig hte dM ate ria l


ew = SWG,
S

G, Se PW. 1+e

is computed using Eq. (2.45) provided the void ratio, e, or the porosity, n, of the soil are known. The dry density curve corresponding to a degree of saturation of 100% is called the zem air voids curve. The dry density curves for different degrees of saturation are commonly presented in connection with soil compaction data (Fig. 2.15). Compaction is a mechanical process used to increase the dry density of soils (Le., densification). The compaction p m s s is not discussed in this book, since numerous soil mechanics textbooks deal with this subject. The relationship between the gravimetricwater contents, w , and the volumetric water content, Ow, can be established by substituting the basic volume-mass relationship [i.e., Eq. (2.40)] into Eq. (2.31):

+ wG,

(2.46)

The relationship between total density, p, and dry density, pd, for different water contents is presented graphically in Fig. 2.14. The specific gravity of the soil solids, G,, and the density of the water, pw, are properties described in Section 2.1. If any two of the volume-mass properties of a soil (e.g., e, w, or S) are known, the total density of the soil, p , can be computed in accordance with Q. (2.43) or Eq. (2.44). The dry density of the soil, pd,

2.3.7

Changes in Volume-Mass Properties

The basic volume-mass relationship [Q. (2.40)] applies to any combination of S, e, and w. Any change in one of these volume-mass Properties (i.e., S, e, and w) may produce changes in the other two properties. Changes in two of the volume-mass quantities must be determined or measured in order to compute a change in the third quantity. If

0.1

0.2

0.2

0.3

0.4

.6

9)

.i

0.e

0.4

a 8

0.E

0.6

1 .c

1.;

1.d

0.6

l. 1.E
2.:

2.c

0.7
0.74

2.U

Water content, w (%)

Flgure 2.14 Volume-mass relations for unsaturated soils.

Copyright 1993 John Wiley & Sons

Retrieved from: www.knovel.com

34

2 PHASE PROPERTIES AND RELATIONS

Substituting Eqs. (2.49), (2.50), (2.51), and (2.47) into

Eq. (2.48) gives


SiAe

+ ASei + ASAe = AwG,.

(2.52)

The change in the degree of saturation, AS, can be written in terms of the change in void ratio, Ae, and the change in water content, Aw:

Figure 2.15 Standard and m o d i f i e d AASHTO compaction curves.

Co py rig hte dM ate ria l


Water content, w (%)

16006

%-i;-$

u25

(2.53) . , ef Similarly, the change in the void ratio, Ae, is obtained by substituting Eq. (2.49) into Eq. (2.52) and solving for Ae:

As=

(AwG, - SiAe)

Ae =

(AwG,

- ASei)

(2.54)

changes in the void ratio, e, and the water content, w, are known, the change in the degree of saturation, S, can be computed. Similarly, if the changes in S and e or in S and w are known, then the change in w or e , respectively, can be computed. The relationship between the changes in the volumemass properties can be derived from the baric volume-muss relationship expressed in Eq. (2.40). Consider a soil that undergoes a pmeess such that changes occur in the volume-mass properties of the soil. Prior to the process, the volume-mass properties of the soil have the following relationship:
Siei = wiGs
(2.47)

3 f

The change in water content, Aw, can be written as follows:


Aw = (SfAe

+ Mei)
I

(2.55)

G S

where

Si = initial degree of saturation


e, = initial void ratio wi = initial water content.

At the end of the process, the soil has final volume-mass properties which are also related by the basic volume-mass relation:

Sfe , = wf G,

(2.48)

where

Sf = final degree of saturation


ef = final void ratio wf = final water content.

2.3.8 Density of M i x t u r e s Subjected to Compression of the Air Phase Soil mixtures can occur in various forms in nature. A mixture of soil particles and air constitutes a dry soil, while a mixture of soil particles and water constitutes a saturated soil. Between these extremes lies the category of unsaturated soils which consist of soil particles, water, and air in differing volumetric percentages. A mixture of soil particles, water, and air has a total density which has been defined in Eqs. (2.43) and (2.44). The dry density, &, of the mixture is expressed in Eq. (2.45) by considering only the mass of the soil particles. The density of a soil mixture can also be formulated for situations where there is a change in the volume of air due to compression. The formulation can be visualized with the assistance of the piston-porous stone analogy described in Fig. 2.11. The density can be derived in a general form for a mixture of soil particles, water, and air. The derivations can then be extended to solve for the density of mixtures of soil particles and air, soil particles and water, and air and water, as well as the density of each phase.

The following relationships between initial and final conditions can be written: Sf = Si + As (2.49)

+ Ae W f = wi + AW
ef = ei

(2.50)

(2.51)

where
= change in the degree of saturation Ae = change in the void ratio Aw = change in the water content.

AS

Piston-Porous Stone Analogy Consider a cubic element of soil that consists of soil particles at the bottom, water in the middle, and free air at the top. All sides are impervious and fixed, with the exception of the top which is a sealed frictionless piston. Initial volume-mass properties for each phase are shown in Table 2.12. The water phase initially has an amount of dissolved air which is in equilibrium with the free air. The air pressure is then increased by placing an additional load on the piston. As a result of the pressure difference between the free

Copyright 1993 John Wiley & Sons

Retrieved from: www.knovel.com

2.3 VOLUME-MASS RELATIONS

35

Table 2.12 Symbols and Notations for the Initial and Final Stages of Each Phase in the o i l Solids, Water, and Air Mixture of S
Volume-Mass Properties of Each Phase Soil Solids Volume Mass Density

Initial Stage

Final Stage

Water Phase (Water and dissolved air) Volume of water and dissolved air Mass of water Density of water Mass of dissolved air Mass of water and dissolved air Density of water and dissolved air Degree of saturation Water content Air Phase (free &d dissolved air) Volume of free air Volume of dissolved air Absolute pressure of air Density of air Volumetric coefficient of solubility Voids (water and air) Void ratio

and dissolved air, some of the free air dissolves into the water in accordance with Henrys law. This process is time dependent. At the end of the p m s s , the pressure in the free and dissolved air are the same. The final volume-mass symbols and notations for water and air are summarized in Table 2.12. The soil particles are assumed to be incompressible (Le., Vs constant), and the mass of soil panicles, Ms, is also constant throughout the process. Similarly, the mass and volume of water (Le., Mu and Vu, respectively) are constant during the process. After an additional load is applied to the piston, the volume and pressure of the free air change. However, the total mass of air (Le., free air and dissolved air) remains constant. At the end of the process, the pressure in the free air and dissolved air has changed from iiai to Si,, and the volThe volume of ume of free air has changed from Vai to va. dissolved air, v d , was shown in Eq, (2.17) to be essentially constant for different air pressures. kyles law states that the p d u c t of the volume and

Co py rig hte dM ate ria l


V U

VU

MU

M W

Pw

=MW/VW
Mdi

Pw

=MW/VW
Md

pwi =

(Mw + M d i ) / v w
= vw/v, = Mw/Ms

+ Mdi

Mw

P,,,, =

(Mu + Md)/vw
S
W

+ Md

s i

wi

Vai vd

Va

#ai

Ua
v d

Poi

h=

vd/v,

Pa

ei = V,/Vs

absolute pressure of a fixed amount of gas is constant under constant temperature conditions [Eq. (2.7)]. Applying kyles law to the free air and dissolved air at a constant temperature gives the final volume of free air, V,:

(v, + vd)zai = (va + vd)Ea


va = (Vai

+ v d ) u.i - vd ua

(2.56)
(2.57)

where

Vai = initial volume of free air corresponding to the initial absolute air pressure, Eai uai = initial absolute air pressure v d = volume of dissolved air V,, = final volume of free air corresponding to the final absolute air pressure, Za u, = final absolute air pressure.

The density of the free and dissolved air also changes

Copyright 1993 John Wiley & Sons

Retrieved from: www.knovel.com

36

2 PHASE PROPERTIES AND RELATIONS

du, I the change in pressure [see Eq. (2.9)]. At a constant ter mture, the initial and final densities of air are related as (2.58)

Conservation of Mass Applied to a Mixture The final density of a mixture can be written by satisfying the conservation of mass of the element:
(2.62) where

where
Pai

= initial density of air (free and dissolved) corre-

Pa

The water phase can be assumed to be incompressible (Le., Vwis constant) during the process. However, the mass of water and dissolved air is changed at the end of the pro-

Co py rig hte dM ate ria l


(2.59)
P =

sponding to the initial air pressure = final density of air (fnx and dissolved) corresponding to the final air pressure

total density of the mixture (i.e., soil particles, water, and air) after the change in air pressure V = total volume of the mixture after the change in absolute air pressure (i.e., V, + Vw + V,) p. . " = density of soil solids V, = volume of soil solids.
p =

The density of the mixture can then be derived by substituting Eqs. (2.57), (2.58), and (2.61) into Eq. (2.62)

(2.63)

cess due to a change in the mass of dissolved air from Mdi to Md. This corresponds to the change in absolute air pressure from iai to Ea as described by Henry's law. The final density of the water phase (Le., water plus dissolved air), pwf, is computed as the ratio of the total mass of water and dissolved air to the total volume of water and dissolved air under final conditions:

This equation can be reduced to the following form:

The volume of dissolved air, Vd, is related to the volume of water, V , , by the volumetric coefficient of solubility, h (i.e., Vd = hVw). Equation (2.64) can therefore be rewritten as follows:
P,V~+
~ w V w

where
pwf =

final density of the water phase (i.e., water plus dissolved air) Mw = mass of water before and after the change in air pressure Md = final mass of dissolved air Vw = total volume of water and dissolved air before and after the change in air pressure The mass of dissolved air, Md, is given as the product of the dissolved air volume, Vd, and the density of air, pa [Eq. (2.58)]:

V, + Vw + Vai?

+ PaiVai + PaihVw -

+ hV,

ua

(2-

.
1)

(2.65)

The volume of each phase can be expressed in terms of the volume of soil particles, V,, using the initial void ratio, ei, and the initial degree of saturation, Si. Therefore, the volume of water, V , , can be written as
Vw = SieiV,

(2.66) (2.67)

and the volume of the initial free air, Vai, can be written as Vai = (1

- Si)eiV,.

Substituting Eqs. (2.66) and (2.67) into Eq. (2.65) and dividing the top and bottom of the equation by V, gives
P =

The final density of water, pd, can be obtained by substituting Eq. (2.60) into Eq. (2.59).

+ pwSiei + pai(l-- &)ei + pdhSiei 1 + Siei + (1 - &)ei T + hSiei


ps
"ai

The first term in Eq. (2.61) (Le., M w / V w )is the density of water without dissolved air, pw.

(2.68) Substituting the basic volume-mass relationship (Le., Siei = wiGs) into Eq. (2.68) yields another form for the

Copyright 1993 John Wiley & Sons

Retrieved from: www.knovel.com

2.3 VOLUME-MASS RELATIONS

37

density of the soil (Le., the soil particles-water-air mixture): 1 + wiGs + (ei - wiG,) T + hwiGs
Uai

ei, yields the following form:

p. + pwSi + pai(l - Si)


P = .

ei

+ p,hSi
I-

, . (2.72)

Ua

(2.69)

Soil Particles- Water-Air Mixture The general equation for the density of mixtures can be specialized to the density of particular mixtures. Consider the case of an unsaturated soil not subjected to a pressure change. Let us assume that the mass of the air phase is negligible (i.e., pai = 0), and that there is no change in the pore-air pressure (i.e., iS, equals Tiai). Under this condition, the water content, the degree of saturation, and the void ratio remain constant (Le., wi = w , Si = S, and e, = e). Equation (2.69) then specializes to the following form:
P'
Ps

Co py rig hte dM ate ria l


where
pm =

Equations (2.68) and (2.69) are general mixture equations for the density after the soil has been subjected to a change in air pressure. The equations incorporate the effect of air going into solution due to the change in the air pressure. Where the air does not have time to dissolve in the water, the volumetric coefficient of solubility, h, can be set to zero. The equations use the initial degree of saturation and the initial void ratio as a reference. The density of the mixture before a change in air pressure can also be obtained from these equations by setting iiaequal to Tiai. This also applies to a mixture which does not experience any change in the air pressure (i.e., i;, equals Tiai).

Setting the initial void ratio to infinity gives the density of an air-water mixture after a change in air pressure from uai to Ea:

density of an air-water mixture after a change in air pressure from Zai to i s,.

The air-water mixture always has a density between that of air and water. As the final air pressure, ia, increases, the final density of the air-water mixture also increases. If the air pressure is continuously increased, the density of the air-water mixture will reach the density of the water phase (Le., water plus dissolved air). If the density of air is assumed to be negligible, the density of the air-water mixture reaches the density of water (Le., pm approaches p,,,). At this stage, the free air has been dissolved into the water and the degree of saturntion approaches 10096. The magnitude of the air pressure required to bring the air-water mixture to saturation can be found from Eq. (2.73) by substituting (p, = pWand pai = 0), and dividing the top and bottom portions of the equation by the initial degree of saturation, Si:
Pw

+ P w WG,
l + e

(2.70)

Noting that the density of the solids can be written in terms of the specific gravity of the solids (Le., ps = G , p w ) , Eq. (2.70) becomes
p=-

+ ($- 1)

2 + (2h

Pw

.
1)

(2.74)

G, Se Pw. 1+e

Solving Eq. (2.74) for the absolute pore-air pressure gives


(2.75)

(2.71)

Equation (2.71) gives the total density of a soil at a constant air pressure, and is the same as the expressions derived previously [Le., Eq. (2.4411.

Air- Water Mixture The general mixture equation can also be shown to specialize to the case of an air and water mixture. The volume of the soil paxticles, V,, is set to zero, and the initial void ratio, ei, becomes infinity. Rearranging Eq. (2.68) by dividing the top and bottom portions by the initial void ratio,

Equation (2.75) shows that the final absolute air pressure required to dissolve all of the free air depends upon the , ,and the initial initial degree of saturation of the mixture, S absolute air pressure, Z , . The lower the initial degree of saturation, the larger the air pressure required to saturate the air-water mixture. The absolute air pressure, is,, obtained from Eq. (2.75) is essentially the same as the maximum air pressure required to saturate a soil, as suggested by Bishop and Eldin (1950), and Schuurman (1966).

Copyright 1993 John Wiley & Sons

Retrieved from: www.knovel.com

Anda mungkin juga menyukai