Soil mechanics

  

The science that applies laws and principles of physics, especially continuum and particulate mechanics and hydraulics of porous media, to solve problems encountered in civil engineering that are related to force-deformation behavior of soil and the flow of fluids through soil. Civil engineers use the term soil for a naturally occurring multiphase material consisting of gas or liquid and solid particles, whether or not the material supports plant life. The term geotechnics or geotechnical engineering refers to a subspecialty of civil engineering that combines the disciplines traditionally known as soil mechanics, soil dynamics, rock mechanics, earthwork engineering, and foundation engineering for the design or analysis of natural or constructed facilities of, in, or on rock or soil. See also: Rock mechanics

 

Fundamental principles

 

Soil mechanics rests on five fundamental relationships. The effective stress principle formulated by K. Terzaghi in the early 1920s in relation to work on the consolidation of clays, is given by Eq. (1),

 

 

 (1)                  Eq 1 

 

 

 

where σ = total normal stress, σ = effective normal stress, and u = pore water pressure. The effective stress principle unifies and extends the applicability of other fundamental relationships such as Darcy's law, Eq. (2),

 

 

 (2)               Eq 2

 

 

where Q is flow rate, k permeability, i hydraulic gradient, and A cross-sectional area; Coulomb friction, Eq. (3),

 

 

 (3)                Eq 3

 

where τ is shear stress, σn normal stress, and φ friction angle; Newton's second law of motion, Eq. (4),

 

 

 (4)                Eq 4

 

where F is force, m mass, and a acceleration; and Hooke's law, Eq. (5),

 

 

 (5)           

 

 

where σ is stress, E modulus of elasticity, and ε strain. Most of the advances that have been made in soil mechanics use these fundamental principles as their starting point. See also: Friction; Hooke's law; Newton's laws of motion

 

 

 

Fig. 3  Relationship between pore water and heads. (a) Relationship between heads in multilayered soil. (b) Placement of the piezometers and well [he = elevation head; hp = pressure head (pore pressure); ht = total head; zw = depth below the phreatic surface (water table)].

 

 

 

fig 3

 

 

 

 

Nature of soil

 

Soil consists of a multiphase assemblage of mineral or organic particles, as shown in the diagram in Fig. 1. The voids within this solid soil skeleton contain gas (generally air), liquid (generally water), or both. A saturated soil has its voids completely filled with water. Various mathematical equations are used to determine the relationships that exist among the three phases within a soil element, an arbitrarily small volume within a soil mass.

 

 

Fig. 1  Diagram showing the three phases in a thin section of a soil element.

 

 

 

fig 2

 

 

 

When an element of dry soil experiences a load (force F in Fig. 2), normal (N) and tangential (T) forces develop immediately at the contact points of the soil particles. The soil mass deforms as the air in the voids compresses or escapes. If the soil voids contain water (Fig. 2b), an essentially incompressible liquid, the spring–dashpot analogy helps visualize the load transfer process. Upon application of the load, the pore fluid (water in chamber) cannot compress, and the small size of the soil pores (valve orifice) restricts the escape of water. The pressure in the pore water increases to counteract the applied force F. Immediately after time t = 0, water slowly flows out through the tortuous path formed by the microscopic pores. The soil skeleton (spring) begins to feel an increased load, deforming as a load transfer from the pore water to the soil skeleton occurs. After sufficient time has elapsed, pressure within the pore water returns to zero, flow through the soil pores ceases, and the system reaches an equilibrium position, with a compressed soil skeleton resisting the applied force. The applied force imposes a stress F/A (force F per unit area A) on the system, and a strain d/L in the vertical direction (deformation d per unit length L in a given direction) in the equilibrium position. Geotechnical engineers consider stresses and strains positive when compressive. A system with large pores reaches equilibrium quickly, while a system with small pores requires a long time to reach equilibrium.

 

 

Fig. 2  Load transfer and consolidation; the effects of the application of a load (F) to soil. (a) Dry soil (N = normal force; T = tangential force). (b–e) Soil containing water and the corresponding spring and dashpot analogy from time = 0 until establishment of equilibrium.

 

 

 

fig 2

 

 

 

The spring–dashpot system exhibits linearly elastic behavior. Doubling the force F doubles the deformation. Upon removing the force F from the spring–dashpot system, the compressed spring pushes the piston back to its original position as water flows back into the chamber. In soils, crushing and slippage of particles results in nonlinear behavior and unrecoverable deformations that prevent a total rebound of the system to its original position. This nonlinear inelastic behavior means that soils do not obey Hooke's law relating stress to strain.

Since the early days of modern soil mechanics, Hooke's law served as a benchmark to evaluate how far soil behavior deviated from its simple generalization. Complex nonlinear stress–strain relationships attempt to model more closely the deformation characteristics of soil, but linear elasticity continues to provide a useful and sufficiently accurate simplification for the solution of many engineering problems.

 

Soil classification and index properties

 

The discipline of soil mechanics relies on the quantification of observed soil behavior. The geotechnical engineer utilizes soil classification to organize soils that exhibit similar behavior into groups. The table shows a simplified soil classification based on the widely accepted Unified Soil Classification System (USCS). Simple index tests such as grain size and plasticity help classify soils.

Among the most common index tests, the Atterberg limits that use the water content at the boundary between states of consistency to correlate soils. The water content of a remolded soil at which a ½-in. (13-mm) section of a 5/√64-in.-wide (2-mm) groove closes from 25 successive shocks in a standard device corresponds to the liquid limit. In effect, the test measures the water content required for an arbitrary but constant shear strength of the remolded soil sample. Therefore, all soils have approximately the same shear strength at the liquid limit. The plastic limit corresponds to the water content at which a thread of soil with a diameter of ⅛ in. (3.2 mm) crumbles when pressed together and rerolled. The shrinkage limit corresponds to the minimum water content required to completely fill all the voids of a dried soil pat (or similarly, the water content at which further drying will not cause additional shrinkage of the sample).

Correlations between fundamental soil properties (such as strength, permeability, compressibility, and expansivity) and index tests help the engineer obtain approximate solutions. The actual soil properties measured in laboratory or field tests can vary greatly from the values suggested by the index property correlations.

 

Pore pressures and heads

 

Water in the soil pores exists at a measurable pressure—positive, zero, or negative. Zero pore pressure corresponds to atmospheric pressure in the pore water. Figure 3 shows the relationship between pore pressure and head. An instrument known as a piezometer measures total head near its sensor. The elevation of the piezometer sensor equals the elevation head, he. The dimensions of a piezometer sensor generally do not exceed 4 in. (10 cm) in diameter and 3.3 ft (1 m) in length. The total head, ht, equals the elevation to which water rises in a piezometer. The height of water in the piezometer equals the pressure head, hp. The pore pressure, u, is given by Eq. (6), where γw equals the unit

 

 

 (6)               Eq 6

 

weight of water. Wells can have long collection zones that penetrate different soil layers. This characteristic makes total head data from wells difficult to interpret and use, as shown in Fig. 3b.

Positive pore pressures decrease effective stresses, which results in increased deformations and decreased soil shearing resistance. In general, the higher the pore pressure, the higher the possibility of encountering unsatisfactory performance.

For hydrostatic conditions, the total head remains constant everywhere, no flow occurs and the pore pressure, u, is given by Eq. (7),

 

 

 (7)                    Eq 7

 

where zw equals the depth below the phreatic surface, the imaginary line where hp = 0. Most soil deposits do not exist under hydrostatic conditions. A steady flow of water through the soil results from a constant pattern of pore pressures, known as the steady-state pore pressures or heads. Pore pressures different from steady-state values, known as excess pore pressures, result from loading, unloading, rainfall, increase in river or reservoir levels, leaks from pipelines or tanks, and drying of the soil, among other causes.

 

Flow

 

Total heads determine the direction of flow in soil. Water flows from a zone of higher total head toward one of lower total head. Darcy's law [Eq. (2)] describes accurately the flow of water through the soil pores. Figure 4a illustrates flow through soil under conditions similar to those used by Darcy for his experiments. Rewriting Darcy's law as Eq. (8), yields a

 

 

 (8)             Eq 8

 

flow rate equal to 8.5 × 10−3 gal/s (3.2 × 10−5 m3/s). [Here Q is flow rate, k permeability or hydraulic conductivity, i hydraulic gradient, ht1 total head at entrance end of soil, ht2 total head at exit end of soil, L length of soil sample, and A cross-sectional area of sample.] Darcy's coefficient of permeability varies more than any other engineering soil property, as much as 10 orders of magnitude (or 109 times) between a clean gravel and a plastic clay. Furthermore, discontinuities such as cracks or very pervious thin layers can greatly affect the permeability of a soil deposit.

 

 

Fig. 4  Diagrams of flow through soil. (a) Flow through a permeameter. (b) Simple flow net.

 

 

 

fig 4

 

 

 

Fig 5.  Analysis of flow through soil of an earth dam. (a) Profile of flow through the dam. (b) Flow parallel to the slope. Piezometric line is a line located at a vertical distance hp from a surface of interest; uppermost flow line represents the phreatic surface (water table). (c) Flow toward a fully penetrating well.

 

 

fig 5

 

 

 

A flow net provides a graphical representation of flow through soil, particularly useful for evaluating complex conditions. For two-dimensional flow through isotropic soil, the flow net consists of a set of flow lines and orthogonal equipotential lines that divide the soil into curvilinear squares. A flow line traces the flow path through the soil for a drop of water. While an infinite number of flow lines exist, a flow net utilizes an arbitrary selection to divide the soil into a convenient number of flow channels, nf. Equipotential lines, drawn through points of equal total head, divide the flow net into a number of uniform head drops, nd. The flow per unit length, l, perpendicular to the page is given by

 

 

 (9)           Eq 9

 

Eq. (9), where Δh = total head drop = ht1 − ht2, and sint = shapefactor = nf/nd

The flow net for the Darcy permeameter (Fig. 4b) has a shape factor equal to 4/5, which yields a flow equal to that computed by using Darcy's equation. The gradient between two points in the net equals the head drop divided by the distance between them. Since the net in Fig. 4 has equally spaced equipotentials, the gradient in the direction of flow remains constant.

Figure 5a shows a flow net for an earth dam with the following characteristics: Δht = 88 ft(27 m), nf = 9.9, nd = 18, sint = 0.55 and Q/l = 0.014 ft2/day (0.0013 m2/day) for k =4 × 10−8 in./s (1 × 10−7 cm/s). The gradient in the direction of flow varies throughout the flow region, increasing as the flow approaches the drain. (Note that the distance between equipotentials decreases in Fig. 5a.) Areas of uncontrolled high gradients can experience instability problems from low effective stress and internal erosion. The graded filter shown on the upstream side of the drain ensures that the fine particles of the compacted fill do not get washed into the drain. Filters generally consist of layers of selected soil with particle sizes chosen to prevent migration of fines. Geosynthetic filters, rapidly growing in acceptance, consist of one or more layers of fabric engineered to retain soil fines. See also: Dam; Geosynthetic

A flow net provides information on the pore pressure throughout the flow region. Rather than using the entire flow net for some analyses, engineers can obtain the information applicable to a surface of interest (for example, a potential shear surface) and portray the heads obtained from the net in terms of a piezometric line. Figure 5b shows that the piezometric line does not necessarily coincide with the phreatic surface. Under conditions of horizontal flow or no flow, the piezometric line and phreatic surface coincide.

Figure 5c shows a simplification that permits evaluation of the flow into a fully penetrating well using Darcy's law. The Dupuit assumption to Darcy's law states that the hydraulic gradient remains constant from top to bottom in the pervious formation and equal to the slope of the water surface, as in Eq. (10).

 

 

 (10)              Eq 10

 

Then, the flow rate, q, out of the well using any set of consistent units is given by formula (11), where

 

 

 (11)              Eq 11 

 

q is water quantity per unit time (Q/t). The values for h1, h2, r1, and r2 are measured (Fig. 5c) at observation wells 1 and 2, respectively. Pumped wells have many uses in projects involving civil engineering, including water supply, reduction of pore pressures to increase effective stresses and thus strength, and control of contaminants in the groundwater.

 

Stress in soil

 

Figure 6 illustrates the concept of stress in soil. Geotechnical engineers frequently work with normal stresses acting perpendicular to a surface of interest, shear stresses acting parallel to a surface of interest, and pore pressures (sometimes known as neutral stresses) acting equally in all directions, expressed as stress vectors (Fig. 6b). See also: Shear

 

 

Fig. 6  Stress in soil represented by (a) soil element beneath loaded surface and (b) free-body diagram with stress vectors. τ = shear stress. Friction angle (θ) is measured counterclockwise from direction of normal stress (σ1).

 

 

 

fig 6

 

 

 

The effective stress principle [Eq. (1)] states that the total normal stress experienced by an element of soil consists of the effective normal stress, which controls strength and deformation, and the pore water pressure. Since water can not sustain shear, pore water pressures do not affect shear stresses. A bar over the symbol identifies a stress as effective or a property as related to effective stresses (for example σ, φ, c.

The stress history of a soil greatly influences its behavior. Two powerful mathematical tools used to understand and make use of stress history are Mohr's circle of stress and stress path (Fig. 7). Mohr's circle provides a graphical representation of the stresses in any direction acting at a point. A distance equal to the magnitude of the pore pressure separates the circle for total stresses from that representing effective stresses. Each point along the circle represents the normal and shear stresses acting on a particular plane. The pole of the Mohr circle relates the stresses to the physical planes on which they act. A line through the pole and point σθ, τθ will lie parallel to the plane on which σθ and τθ act.

 

 

Fig. 7  Mohr's circle (a) for state of stress at a point and (b) stress path to failure for the soil element shown in Fig. 6.

 

 

 

Fig 7

 

 

 

Two particular points prove very helpful in portraying stress history—the point for stresses on the failure plane, often taken as the point of maximum obliquity (maximum value of τ/σ), and that for maximum shear stress τ (identified by coordinates p and q). The total stress path (TSP) and effective stress path (ESP) provide a locus of stress points representing successive states of stress in the history of the soil. For clarity, Mohr circles usually do not appear in stress path plots, although one could reconstruct the circle at any point along the stress path. Failure occurs when the effective failure-plane stress path reaches the strength envelope (also known as yield surface) or when the p-q stress path reaches the Kf line as in Fig. 7.

The stress path can reach the strength envelope because of a loading, an unloading, an increase in pore pressure, or a combination of these. Stress paths help engineers identify mechanisms and select methods and soil parameters for solving a problem. See also: Stress and strain

 

Strength

 

Soil behaves as a frictional material; that is, it resists shear stresses by the frictional action between particles. Figure 8 shows a strength test where a sample under a constant normal force gets sheared along a predetermined horizontal plane. The displacement-versus-shear stress plot (Fig. 8b) shows two important characteristics—a peak shear strength and a residual value obtained after large deformations. A series of these tests at different values of normal stress yields the peak and residual strength envelopes shown. These strength envelopes, unique for practical purposes, remain valid for a wide range of effective stress paths to failure. Some soils exhibit a monotonically ascending displacement–versus–shear stress curve where the peak strength equals that at large deformations.

 

شکل ۸

fig 8

 

 

Fig 9.  Consolidation settlement. (a) Clay layer with double drainage, no lateral strains. (b) Oedometer test to obtain soil parameters. (c) Oedometer test results; each data point represents the end of a load increment; and left broken line represents 10 metric tons/m2 and right broken line represents 100 metric tons/m2. 1 metric ton = 1.1 ton; 1 m2 = 10.8 ft2.

 

 

 

 

fig 9

 

Soils generally do not have true cohesion, that is, when σ = 0, strength = 0. Since strength envelopes can be curved, use of the linear Coulomb friction equation (3) requires care in the selection of parameters that are valid for the stress range of interest.

 

Consolidation and swelling

 

Consolidation refers to a process during which a saturated soil experiences a decrease in water content due to a decrease in the volume of voids. The opposite process, swelling (heave) results in an increase in water content due to an increase in the volume of voids. Consolidation and heave involve flow of water into or out of the soil, and a corresponding change in volume. Changed pore pressures (for example, from a change in load or heads) establish a hydraulic gradient, since the pore pressure near the drainage boundary remains essentially unchanged. In soils with low permeability, the water flows slowly, and a time lag exists between the event that changed the pore pressures and the completion of consolidation. The compression of a partially saturated soil involves compression, solution, or expulsion of air until the soil becomes nearly saturated. The spring–dashpot example in Fig. 2 illustrates the mechanism of consolidation.

Figure 9a shows a layer of clay with double drainage subjected to a uniform load. After application of the load, pore pressures near the center of the layer increase by an amount equal to the magnitude of the load. The high permeability of the sand permits rapid dissipation of excess pore pressures, and deformations occur almost instantaneously. The total head difference between the center of the clay and the sand layers creates a hydraulic gradient, which produces a flow from the center outward in two directions. Figure 9b shows a consolidation test (known as an oedometer test) that models the one-dimensional compression experienced by the clay layer. The expression for final settlement shown in Fig. 9d remains valid for fine- as well as coarse-grained soils.

The oedometer tests provide information to predict both the magnitude and rate of deformation. In coarse-grained soils, deformations occur very rapidly. In fine-grained soils, the theory of consolidation helps predict the rate of deformation using the coefficient of consolidation, a parameter influenced mostly by the permeability of the soil. The theory of consolidation permits computation of more detailed results, including distribution of pore pressures and deformations throughout the deposit at any time.

 

Soil improvement

 

For many centuries, builders have tried to make soil stronger, denser, lighter, less compressible, less pervious, more pervious, or more resistant to erosion. One of the most common soil improvement techniques compacts the soil with a heavy roller or tamper. Compaction of fine-grained soils generally takes place in layers no more than 12 in. (30 cm) thick. Compaction densifies the soil, and it can increase its strength and decrease its compressibility and permeability. The key to obtaining the desired dry density lies in achieving a preselected water content in the soil undergoing compaction. Compaction at the optimum water content yields the highest dry density for a given compactive effort. Soil compacted at or near the optimum water content has negative pore pressures.

Other techniques use to improve engineering properties of soil include preloading; dewatering; stabilization with chemical admixtures; and reinforcement with rods, strips, grids, and fabrics made of natural of synthetic materials.

 

Slope stability

 

In earth slopes, the force of gravity pulls the soil mass downward. The internal friction between soil particles resists this force. As long as the resisting forces exceed the driving forces, the slope remains standing. A stability assessment of a slope involves a comparison of the stresses acting on the sliding mass and the strength of the soil along which sliding occurs. This assessment requires evaluation of the geometry and pore pressures, which together determine the stresses, and the strength of soils involved in a potential slide.

Geotechnical engineers use the term factor of safety (FS) to indicate the level of stability of a slope. Using forces, FS is the ratio of the shear resistance on a shear surface to the net shear force along a shear surface; or, using stresses, FS is the ratio of the shear strength along a shear surface to the average shear stress on a shear surface. When FS decreases to 1, failure occurs. Figure 10 shows the forces acting on a wedge of soil. To increase FS would require an increase in shear resistance, S, or a decrease in shear force, T.

 

 

Fig. 10  Free-body diagram showing forces on a slope wedge.

 

 

 

fig 10

 

 

 

The following actions would increase S: (1) Increase W. (2) Increase c (densify, replace soil chemicals, temperature). (3) Increase φ. (4) Decrease Us (drain).

The following actions would decrease T: (1) Decrease UL (drain). (2) Decrease W (excavate, replace with lighter soil). (3) Decrease PL (unload above slide). (4) Increase UR (raise reservoir level). (5) Increase R (install tiebacks, piles, berm retaining wall).

Slope stability analyses generally use the principles of force and moment equilibrium, a technique known as limit equilibrium. In general, a free body bounded by the slope surface and a potential failure surface is isolated, and the stresses are determined by solving the static equilibrium equations. The ratio of the determined stresses and the shear strength of the soil yields the factor of safety. Repetition of the analysis with other possible failure surfaces yields different factors of safety because both the driving and resisting forces can vary from surface to surface. The most critical surface has the lowest factor of safety. Computer programs allow the rapid investigation of hundreds of potential failure surfaces along cross sections with different soil layers. See also: Statics

More sophisticated numerical techniques consider the stress–strain properties of the soil, and allow the analysis of slopes composed of soils with complicated stress–strain–strength properties. Finite-element stress–deformation analyses can provide a detailed description of the stresses acting throughout the slope. The engineer calculates the factor of safety by integrating the acting shear stress along a potential failure surface and comparing the result to the available shear strength along the surface. See also: Finite element method

The principles of stability analysis outlined above apply for natural slopes as well as for earth structures such as earth dams and highway embankments. With constructed facilities, the engineer has more detailed knowledge and some degree of control over the material properties. An earth structure must satisfy all the criteria needed to perform its intended function. These may include overall stability, limited deformation, weight limitations, specific flow characteristics, or particular dimensions. See also: Highway engineering

 

Foundations

 

Nearly every large structure derives its support from contact with the earth. The foundation, whether of a building, tower, highway, retaining wall, bridge, or storage tank, transfers live and dead loads from the structure to the ground. Besides providing support, or adequate bearing capacity, the foundation must keep settlements within an acceptable range for the structure. Foundations vary from the simple preparation of a patch of ground to elaborate arrays of long steel piles driven to great depths. Other foundation types include rafts, footings, caissons, many types of piles, stone columns, grouted soil, frozen soil, mechanically reinforced soil, and hybrid combinations of these.

The bearing capacity of a foundation element depends on its ability to resist a shear failure, that is, a loss of stability. The principles of static equilibrium that govern stability of slopes also control foundation stability. The sum of the resisting forces must exceed the sum of the driving forces. See also: Foundations

 

Shallow

 

In practice, engineers design most shallow foundations by using semiempirical formulas that combine the concepts embodied in the theories with the experience gained by measuring the performance of actual foundations. Most shallow foundations have a large factor of safety against a general shear failure (3 or above) because restrictions on allowable settlement usually prove much more stringent and result in low allowable bearing stresses.

For either drained or undrained conditions, elastic theory predicts the stress change at any depth below a loaded area. The change in stress produces a strain in accordance with Hooke's law. The sum of the individual strains from the footing base to some depth where the stress change becomes negligible equals the settlement of the footing. In the case of shallow foundations on fine-grained soils, additional settlement will occur because of time-dependent consolidation under the applied load. Difficulties arise in selecting the appropriate elastic parameters, Young's modulus and Poisson's ratio, for the anticipated drainage and stress conditions. The stress path method, which attempts to simulate the anticipated changes in the state of stress with soil samples in the laboratory, can aid the engineer in selecting correct parameters. See also: Elasticity; Young's modulus

 

Deep

 

Where soils of adequate strength do not lie sufficiently close to the ground surface, deep foundations can transfer the loads to stronger soil strata or distribute the loads throughout a greater mass of soil. Deep foundation elements, such as caissons or piles, can develop support along their entire length as well as at their base or tip. See also: Caisson foundation

Friction piles develop resistance primarily through shear resistance (skin friction) along their length. End-bearing piles develop resistance primarily at their tip by bearing on firmer soil or rock. Most piles combine the characteristics of the two pile types. Pile equipment has evolved that can drive the piles into the ground with powerful hammers, vibrate the piles through soil, or place the piles in predrilled holes. Pile widths or diameters range from about 4 in. (10 cm) for minipiles to several feet for offshore oil platforms, with lengths up to several hundred feet. Common pile materials include steel, concrete, and wood.

Caissons or drilled piers usually have greater diameters than piles and rely on end-bearing support. Their construction involves excavation instead of driving, and it often allows visual inspection of the ultimate bearing surface.

 

 

Tests and properties

 

Field and laboratory tests are used to determine soil properties. Those conducted in the field measure density, permeability, shear strength, penetration resistance, compressibility, and deposit boundaries. Those done in the laboratory measure water content, expansion index, grain size distribution, permeability, shear strength, dynamic stress, resonant frequency, and mineral composition.

The broad range of geotechnical instrumentation includes piezometers for measuring ground-water pressure, earth pressure cells for measuring total stress in soil, settlement plates for measuring vertical displacements, extensometers for determining deformation, tiltmeters for measuring inclination, inclinometers for measuring lateral deformation versus depth, piezoelectric transducers for measuring acoustic emissions, load cells for measuring load in a structural member, strain gages for measuring strain in a structural member, and various devices for measuring temperature.

 

Other soil engineering problems

 

A naturally occurring material such as soil presents geotechnical engineers with several particularly difficult problems. In addition to the complexities of working with a nonhomogeneous material with non­linear anisotropic stress–strain properties, the strength varies from point to point, may vary with time or deformation, and remains difficult to measure precisely. The simple examples that have been discussed illustrate the fundamentals of soil mechanics. Engineers routinely extend and apply these fundamentals to the complicated conditions encountered in the field. Some of these applicationsinclude two- and three-dimensional problems; dynamic loading of soil; construction of earth­quake-resistant facilities; earth pressures and retaining walls, tunnels, and excavations; subsidence; problems involving expansive soils; field exploration; probability and reliability; and the use of synthetic materials such as fabrics, membranes, and superlight fills. See also: Clay; Construction methods; Engineering geology; Retaining wall; Sand; Soil; Tunnel

 

 


 

ENCYCLOPEDIA ARTICLE: Soil mechanics

A simplified soil classification*
Material Symbol Grain size USCA symbol Name
Coarse-grained
       
 Boulder
None
12 in. + (30 cm +)
   
 Cobbles
None
3–12 in. (7.6–30 cm)
   
 Gravel
G
 
GW
Well-graded gravel
 Coarse
 
¾–3 in. (1.9–7.6 cm)
GP
Poorly graded gravel
 Fine
 
No. 4, ¾ in. (1.9 cm)
GM
Silty gravel
     
GC
Clayey gravel
Sand
S
 
SW
Well-graded sand
 Coarse
 
No. 10–No. 4
SP
Poorly graded sand
 Medium
 
No. 40–No. 10
SM
Silty sand
 Fine
 
No. 200–No. 40
SC
Clayey sand
Fine-grained
 
Passing No. 200
   
 Silt
M
 
ML
Silt
   
MH
Elastic silt
 Clay
C
 
CL
Low-plasticity clay
   
CH
High-plasticity clay
 Organic
O
 
OL
Organic silt
   
OH
Organic clay
   
PT
Peat

 

Bibliography

 

 

American Society for Testing and Materials, Annual Book of ASTM Standards, sect. 4: Construction, vol. 04.08: Soil and Rock; Building Stones, Geotextiles, 1989

H. R. Cedergren, Seepage, Drainage and Flow Nets, 3d ed., 1997

R. F. Craig, Soil Mechanics, 6th ed., 1997

B. M. Das, Fundamentals of Soil Dynamics, 1983

J. Dunnicliff, Geotechnical Instrumentation for Monitoring Field Performance, 1988

T. W. Lambe and R. V. Whitman, Soil Mechanics, 1969

J. K. Mitchell, Fundamentals of Soil Behavior, 2d ed., 1993

K. Terzaghi and R. B. Peck, Soil Mechanics in Engineering Practice, 3d ed., 1996

U.S. Navy, Soil Mechanics, Design Man. 7.01, 1986